Chemical Engineering Science 63 (2008) 3923 -- 3931

Contents lists available at ScienceDirect

Chemical Engineering Science journal homepage: w w w . e l s e v i e r . c o m / l o c a t e / c e s

One-equation sub-grid scale (SGS) modelling for Euler--Euler large eddy simulation (EELES) of dispersed bubbly flow B. Niceno a,∗ , M.T. Dhotre a , N.G. Deen b a Thermal-Hydraulics b Faculty

Laboratory, Nuclear Energy and Safety Department, Paul Scherrer Institut, 5232 Villigen PSI, Switzerland of Science and Technology, University of Twente, P.O. Box 217, 7500AE Enschede, The Netherlands

A R T I C L E

I N F O

Article history: Received 25 July 2007 Received in revised form 3 February 2008 Accepted 11 April 2008 Available online 4 May 2008 Keywords: Euler--Euler model Large eddy simulation Sub-grid scale model One-equation model Bubbly flows Bubble column

A B S T R A C T

In this work, we have presented a one-equation model for sub-grid scale (SGS) kinetic energy and applied it for an Euler--Euler large eddy simulation (EELES) of a bubble column reactor. The one-equation model for SGS kinetic energy shows improved predictions over the state-of-the-art dynamic procedure. With grid refinement, the amount of modelled SGS turbulent kinetic energy diminishes, as one would expect. Bubble induced turbulence (BIT) at the SGS level was modelled with two approaches. In the first approach an algebraic model was used, while in the other approach extra source terms were added in the transport equation for SGS kinetic energy. It was found that the latter approach improved the quantitative prediction of the turbulent kinetic energy. To the best of authors knowledge, this is the first use of a transport equation for SGS kinetic energy in bubbly flows. © 2008 Elsevier Ltd. All rights reserved.

1. Introduction In spite of the importance of multiphase flows in many technical applications, the complexity of these flows still leaves a lot of room for improvement of simulation methods. The complexity stems from the range of turbulent scales present in such flows. Bubbly flows, in particular, usually feature large scales of motion induced by buoyant forces on the global scale, which range down to the Kolmogorov scales of motion. Another feature, which makes the modelling of bubbly flow difficult, is the presence of the interface between the liquid and gas phases. The influence of the bubble motion and the flow topology on the turbulence and vice versa is generally complex and hard to predict. A number of simulation strategies for multiphase flows have been proposed which differ in complexity of the physical models involved. An approach which uses no modelling at all, is coupled direct numerical simulation (DNS) for turbulence, with surface tracking algorithm for resolving all interface details (Tryggvason et al., 2006). In these approaches all scales of motion are resolved explicitly. This is the most accurate approach, but unfortunately, in spite of the increase of the computational power, it is suitable only for low Reynolds numbers and a limited number of bubbles. On the other end of modelling complexity spectrum, lie Reynolds averaged Navier--Stokes (RANS)



Corresponding author. Tel.: +41 563104149. E-mail address: [email protected] (B. Niceno).

0009-2509/$ - see front matter © 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.ces.2008.04.050

equations, with ensemble averaged Euler--Euler equations for interpenetrating continua. It is a computationally affordable and often used approach, but features complex modelling of interfacial forces, in addition to turbulence modelling. A number of tuning constants in this approach does not encourage confidence in applying it to new problems. Because of the necessity to simulate bubbly flows, intermediate approaches have been proposed, that lie in between DNS with surface tracking and two-fluid model RANS. Since bubbly flows almost invariantly feature large buoyant regions which are inheritably unsteady, transient RANS (TRANS) can also be sought as a suitable modelling approach (Smith and Milelli, 1998; Pfleger et al., 1999; Dhotre et al., 2008). However, it inherits the modelling complexity of the steady RANS, while increasing the cost of simulations substantially. As another example (Alajbegovic, 2001) proposed large scale simulation (LSS) of multiphase flow, which simulates the most important details of turbulence and interface explicitly and models the smaller details. Although this approach is much more affordable than DNS, it is still too expensive for general bubbly flow simulations of practical relevance. Tracking the most important details of each bubble separately in a large reactor would lead to a very high number of computational cells and prohibitively expensive computations. Another possible intermediate approach combines interpenetrating continua (Euler--Euler) for modelling interface details, with large eddy simulation (LES) for simulating the largest scales of turbulent motions (Milelli, 2002; Deen et al., 2001; Zhang et al., 2006; Dhotre et al., 2008). Such an approach inherits complexity of interpenetrating

3924

B. Niceno et al. / Chemical Engineering Science 63 (2008) 3923 -- 3931

continua modelling, but at least does not rely on the steady-state assumption implicitly imposed by RANS turbulence models, thus allowing us the analysis of unsteady behaviour of bubbly flows. It is particularly attractive for buoyant bubbly flows, since the largest and most energetic buoyancy driven structures are explicitly computed. However, such an approach suffers from two principal drawbacks. For the sake of consistency, the largest interface details should be smaller than the grid size, which means that the cell size must be larger than the bubble size. That implies that the grid might not be fine enough to capture all turbulence details, hence making the basic assumption of LES invalid. This is particularly pronounced for bubble induced turbulence (BIT), which is, by the definition of this simulation strategy, moved to sub-grid scale (SGS). Moreover, concerning turbulent dispersion, the two-fluid model LES can account only for large scale effects. It is conceivable that, if the bubbles are small compared to SGS eddy size, the SGS turbulent dispersion should be accounted for. Traditional mixing length type SGS models are not able to model these terms, since the information on SGS energy is implicitly added to pressure and is therefore lost. As a solution to these problems, we propose to use the transport equation for SGS kinetic energy (Davidson, 1997), adapted for interpenetrating continua modelling. To our knowledge, only Yuu et al. (2001) have used a transport equation for Euler--Euler large eddy simulation (EELES), albeit for gas-particle flows. In the transport equation for SGS kinetic energy, we add a source term for BIT (Simonin and Viollet, 1988). We also envisage the usage of the modelled SGS energy to estimate the SGS turbulent dispersion force. The information on SGS kinetic energy helps us to model BIT more accurately and is also more robust in the regions of the flow where the grid is coarse. To show the performance of this approach, we have implemented the one-equation SGS model into CFX 4.4 commercial package and carried out LES simulations for a bubble column. We have used one-equation SGS model and dynamic procedure (Germano et al., 1991), based on Smagorinsky model Smagorinsky (1963). The results obtained with one-equation model are superior to the others, with additional benefit that the value of SGS energy is known and can be used to model SGS interface forces more accurately. Prospects for future developments are also discussed. 2. Mathematical models 2.1. Euler--Euler description of the flow field In the two-fluid model description of a bubbly flow, we have one set of governing equations per phase, plus the expressions for the interface exchange terms. To keep the number of unknowns equal to the number of equations, we assume the pressure to be the same for both phases. The governing equations for the conservation of mass, with no mass exchange between the phases, takes the form j (r) + ∇(ru) = 0, jt

(1)

in which t is time,  is the phase indicator, r, , and u are the volume fraction, density, and resolved (filtered) velocity of phase . Momentum conservation is expressed by the following equation: D(ru) = − ∇(rS ij ) − r ∇p Dt + (r) g + M − ∇(rij ) ,

(2)

where S ij =∇u+(∇u)T is the strain rate of the resolved velocity field, p

is the pressure, ij is the SGS stress tensor, g is the gravity vector, and M is the momentum exchange between the phases due to interface forces. The D/Dt operators denote the substantial derivatives. The

SGS stress tensor and momentum exchange terms from Eq. (2) require modelling, and will be discussed below. Eqs. (1) and (2) are usually derived by ensemble averaging. However, the same form of equations are obtained if one performs filtering (volume averaging) of the governing equations (Drew and Passman, 1999). This is of practical importance for LES, because it means that the same numerical tools developed for ensemble averaged Euler--Euler equations, can be used for LES. The Boussinesq hypothesis is used in all turbulence models used in this work, so Eq. (2) can conveniently be written as D(ru) = −∇(reff S ij ) − r ∇p + (r) g + M , Dt

(3)

where eff is effective viscosity, defined as eff =  + T + BIT .

(4)

Here T is turbulent (eddy) viscosity, obtained from a turbulence model, and BIT is viscosity due to BIT. 2.2. LES of turbulence Turbulence is modelled for the continuous (liquid) phase. The dispersed gas phase is modelled as laminar. For the sake of brevity, the index of the liquid phase is dropped from the equations describing turbulence models. Velocities in Eqs. (1) and (2) represent the part of the velocity field resolved by the numerical method and are defined as follows: u = u − u ,

(5)

where u is the true velocity and u is the SGS part, not resolved by the numerical simulation. These SGS parts of velocity components, gives rise to additional stress terms and interface forces. The SGS stress terms of phase  are related to the resolved scale strain tensor (S ij ) described as ij − 13 kk = 2T S ij .

(6)

Eq. (6) shows that we are modelling only the deviatoric part of the SGS, while its trace (kk ) is implicitly added to the pressure. This means that the information on the amount of SGS kinetic energy (ksgs ) is lost in algebraic models, unless one introduces a model equation to account for it. In this work, we use two different SGS models to compute T . One is the dynamic procedure proposed by Germano, and the other is a one-equation transport equation for SGS kinetic energy. 2.2.1. Dynamic procedure The dynamic procedure used in this work is based on the Smagorinsky model. The turbulent viscosity is calculated from the Smagorinsky model as follows: T,L = L (CS )2 |S ij |,

(7)

where CS is a model constant and |Sij | is the magnitude of the strain rate tensor, and  is filter width, in this work taken as the grid spacing, which is equal in all coordinate directions. A dynamic procedure is used to calculate model constant CS from the expression: (CS )2 =

1 Lij Mij , 2 Mij Mij

(8)

where   Lij = u i uj − ui uj

(9)

B. Niceno et al. / Chemical Engineering Science 63 (2008) 3923 -- 3931

and Mij =

3925

(Clift et al., 1978):    |S ij |S ij − |S ij |S ij 

(10)

and the wide hat indicates a second filter, usually called the test filter, which is twice the mesh size in the present study. This is a standard procedure, and details can be found elsewhere (Germano et al., 1991; Germano, 1992; Lilly, 1992). An advantage of the dynamic procedure is that it calculates the model constant, in our case the Smagorinsky constant CS . It should, theoretically model the backscatter as well (energy transfer from small to large scales), but in practical simulations this means that expression (CS )2 from Eq. (8) becomes negative, leading to negative eddy viscosity, making the solution procedure unstable. To avoid this problem, we trim the negative values of eddy viscosity. 2.2.2. One-equation model Although the dynamic procedure calculates CS , making the SGS modelling free of specifying constants, it does not provide information on the SGS kinetic energy (ksgs ). In this work, a transport equation for ksgs is introduced. The trace of ij defines the SGS kinetic energy: ksgs = 12 kk .

(11)

(13)

Eddy viscosity is calculated from 1/2

(14)

Model constants are C =1.05 and Ck =0.07, as suggested by Davidson (1997).

CD = 23 Eo1/2 ,

(17)

where Eo is the dimensionless Eötvös number (Eo = gdB2 /). In this work, a bubble size of 4.0 mm is used, giving Eo = 2.2 and CD = 1.0. The bubble size used in this work is in correspondence with the experimental observations of Deen et al. (2000), the test case used here for the validation of the model. 2.3.2. Lift force A bubble traveling through a fluid in shearing motion will experience a lift force transverse to the direction of motion. The effect of shearing motion in the liquid phase on the movement of the gas phase is modelled through the lift force as (Drew and Lahey, 1987; Zun, 1990): (18)

2.3.3. Virtual mass force The virtual mass force accounts for the additional work performed by the bubbles in accelerating the surrounding liquid (Jakobsen et al., 1997), and is given by   Dug Dul MVM,l = rg l CVM − , (19) Dt Dt where the virtual mass coefficient CVM is generally shape dependent and is taken to be 0.5 for individual spherical bubbles. At the start of each run, initially the virtual mass force was deselected and this effect was taken into account by simply using an enhanced gas density (Smith and Milelli, 1998): g = g + CVM l

2.3. Interface force modelling The term Ml in Eq. (2) represents the interface forces, and is given as Ml = −Mg = MD,l + ML,l + MVM,l + MTD,l ,

The drag coefficient CD depends on the flow regime and the liquid properties. The Reynolds number covered in the present work falls under the inertial range and as known from experimental visual observations bubbles become distorted in this regime. We calculate CD for distorted bubbles following (Ishii and Zuber, 1979):

(12)

where Pksgs is the production term, defined as

T = Ck ksgs .

(16)

where CL is a model coefficient, set to a constant value of 0.5 for all simulations reported in this work.

3/2

Pksgs = T |S ij |.

C 3 rg l D |ug − ul |(ug − ul ). 4 dB

ML,l = rg l CL (ug − ul ) × ∇ × ul ,

The transport equation for ksgs reads (Davidson, 1997) ksgs Dksgs , = ∇[( + T )∇ksgs ] + Pksgs − C Dt 

MD,l =

(15)

where subscripts l indicates the liquid phase and g the gas phase. The terms on the right-hand side of Eq. (15) are respectively: drag, lift, virtual mass, and turbulent dispersion force. All these forces have grid scale and SGS component. In previous works on EELES (Smith and Milelli, 1998; Deen et al., 2001; Zhang et al., 2006), the SGS components of these forces were neglected, but in this work we make an attempt to envisage a way model the SGS turbulent dispersion force. Note that the grid scale turbulent dispersion is resolved as the fluctuating component of the drag force (see the following section). 2.3.1. Drag force A drag force occurs due to the resistance experienced by a body moving in the liquid. Viscous stress creates skin drag, whereas the pressure distribution around the moving body creates form drag. The drag force per unit volume is written in the following form

(20)

in the acceleration term of the gas momentum conservation equation. 2.3.4. Turbulent dispersion force It should be noted that the drag and lift forces depend on the actual relative velocity between the phases, but the ensemble equations of motion for the liquid only provide information regarding the mean flow field. To account for the random influence of the turbulent eddies, the concept of a turbulent dispersion force has been advanced. By analogy with molecular movement, the force is set proportional to the local bubble concentration gradient (or void fraction), with a diffusion coefficient derived from the turbulent kinetic energy. In LES velocities are decomposed into a resolved and an SGS part, by filtering. We explicitly compute the resolved part of the turbulent dispersion. However, some transport is present at SGS level as well. The situation is illustrated in Fig. 1. The big squares represent the control volume, which is also the filter in LES and the averaging volume in derivation of Euler--Euler description of the flow. If the bubbles are small compared to the control volume (Fig. 1(a)), SGS motions, represented by circular arrows resembling eddies, do

3926

B. Niceno et al. / Chemical Engineering Science 63 (2008) 3923 -- 3931

Water level

H

L

Sparger

Fig. 1. Sub-grid scale motions, filter and bubble size: (a) bubbles are small and (b) large in comparison with SGS scales of motion.

contribute to the turbulent transport of bubbles. One might use the concept derived by Lopez de Bertodano et al. (1994) to approximate the turbulent diffusion of the bubbles by the liquid eddies. It is formulated as MTD,l = CTD l ksgs ∇rg ,

(21)

where CTD is the turbulent dispersion coefficient. In the above equation we replace the total kinetic energy by the SGS contribution, which is based on the assumption that only SGS motions contribute to SGS transport of bubbles (Fig. 1(a)). The exact shape of these motions is unknown, but their intensity is represented in ksgs . In this work, however, such an approach is not necessary, since the bubble size is relatively large in comparison with the filter width and SGS motions as illustrated in Fig. 1(b). Therefore, the inclusion of the modelled SGS turbulent dispersion, such as the one given by Eq. (21), does not have a sound physical justification for the cases considered in this work. 2.4. Bubble induced turbulence Although the turbulence is modelled in the liquid phase only, the dispersed gas bubbles traveling through the liquid induce turbulence. The influence of BIT can be accounted for by increasing the eddy viscosity, using the model of Sato and Sekoguchi (1975): BI = CBI l rg dB |ug − ul |.

(22)

This approach is suitable for algebraic SGS models and it was explored by Deen et al. (2001). In this work, in addition to the Sato model, we explore the influence of BIT by adding the extra production terms into equation, following the procedure described by Pfleger and Becker (2001). PBIT = Ck |MD,l | · |ug − ul |,

(23)

where PBIT is the extra production term in Eq. (12) and MD,l is defined by Eq. (16). Ck is defined following the rationale from (Pfleger and Becker, 2001), i.e. it corresponds to Ck in Eq. (12). 3. Results 3.1. Experimental data used for validation A schematic representation of the experimental set-up is shown in Fig. 2 (Deen et al., 2000). It consists of a column with a 0.15 × 0.15 m2 cross-section and filled with distilled water up to the height of 0.45 m. A distributor plate containing 49 holes with a diameter of 1 mm were placed in the middle of the column at the base with a square pitch of 6.25 mm.

W W Fig. 2. Schematic representation of the bubble column set-up. W = 0.15 m, H = 0.5 m, L = 0.45 m.

3.2. Method of solution A commercial CFD package, CFX version 4.4, was used to solve the governing equations of continuity and momentum. This package is a finite volume solver, using body-fitted grids. Two grids were used in this work: the coarse grid, consisting of 15×15×45 cells, and the fine grid with the resolution of 30 × 30 × 90. Both grids were orthogonal and uniform in all directions. The gas inlet area Ain (0.03 × 0.03 m2 ) was implemented in a central area of 3 × 3 grid cells for the coarse grid and 6 × 6 grid cells for the fine grid. The first node near the wall has y+ of 120 and 70 for the coarse and fine grids, respectively. The boundary conditions were defined as follows. At the inlet the gas velocity was calculated using the superficial gas velocity and the area inlet: Vg,in =

Vg Ac , g,in Ain

(24)

where Vg is the superficial gas velocity and Ac is the cross-sectional

area of the column. A superficial gas velocity of 4.9 × 10−3 m/s leads to a gas velocity at the inlet of 0.12 m/s. Along the walls, the noslip boundary condition was adopted. For the one-equation model, wall functions were applied, whereas in case of LES, this condition was implicitly imposed by the dynamic adjustment of the Smagorinsky constant. For present calculations the domain was limited to the liquid-filled region, to reduce the number of computational cells. It was observed that, if the domain was extended to include the gas region above the liquid level, we need to have a domain as high as 0.8 m. The change in results, with such an extended domain, is insignificant. At the outlet of the column, a pressure boundary condition was used. In the simulations, a bounded third-order accurate QUICK scheme was used for the discretization of the advection term and a second-order, fully implicit backward differencing scheme was used for time differencing. The mass residual was set to be 1 × 10−4 . 3.3. Simulation results In this section, two different SGS closures are compared: a oneequation SGS model and a model employing the dynamic procedure from Germano. Long time averaged vertical velocity components,

B. Niceno et al. / Chemical Engineering Science 63 (2008) 3923 -- 3931

3927

0.2

Axial Gas Velocity, [m/s]

Axial Liquid Velocity, [m/s]

0.25

0.15 0.1 0.05 Coarse Fine Experiment

0 −0.05

0.4 0.35 0.3 0.25 Coarse Fine Experiment

0.2 0.15 0.1

−0.1 0

0.05 0.1 Column Width, [m]

0.15

0

0.05 0.1 Column Width, [m]

0.15

Lateral Fluctuating Liquid Velocity, [m/s]

Axial Fluctuating Liquid Velocity, [m/s]

Fig. 3. Vertical velocity component for: (a) liquid and (b) gas, obtained on two grids, and experiments.

0.2 0.18 0.16 0.14 0.12 0.1 0.08 0.06 Coarse Fine Experiment

0.04 0.02 0 0

0.05

0.1

0.12 0.1 0.08 0.06 0.04 Coarse Fine Experiment

0.02 0

0.15

0

Column Width, [m]

0.05

0.1

0.15

Column Width, [m]

Fig. 4. Fluctuating liquid velocity profiles for: (a) vertical and (b) lateral components, obtained on two grids, and experiments.

0.12 0.11

0.025 SGS/Resolved Energy

Turbulent Kinetic Energy, [m2/s2]

0.03

0.02 0.015 0.01

Coarse Fine Experiment

0.005

0.1 0.09 0.08 0.07 0.06

Coarse Fine

0.05

0

0.04 0

0.05

0.1

0.15

Column Width, [m]

0

0.05

0.1

0.15

Column Width, [m]

Fig. 5. (a) Resolved (dashed) and total (continuous) liquid kinetic energy and (b) ratio of modelled and resolved part of the turbulent kinetic obtained on two grids.

fluctuating liquid velocities, and turbulent kinetic energy profiles are compared to experimental data of Deen et al. (2000). Different closures of the BIT are also assessed.

All simulations presented in this section were run for 400 s and the results obtained by averaging over the time period from 100 to 400 s have been compared with experimental results.

3928

B. Niceno et al. / Chemical Engineering Science 63 (2008) 3923 -- 3931

0.2

Axial Gas Velocity, [m/s]

Axial Liquid Velocity, [m/s]

0.25

0.15 0.1 0.05 Dynamic One eq. SGS Experiment

0

0.4 0.35 0.3 0.25 Dynamic One eq. SGS Experiment

0.2 0.15

−0.05 −0.1

0.1 0

0.05

0.1

0.15

0

0.05

Column Width, [m]

0.1

0.15

Column Width, [m]

Fig. 6. Vertical velocity component for: (a) liquid and (b) gas, obtained from dynamic procedure, one-equation model, and experiments.

0.12 Lateral Fluctuating Liquid Velocity, [m/s]

Axial Fluctuating Liquid Velocity, [m/s]

0.2 0.18 0.16 0.14 0.12 0.1 0.08 Dynamic One eq. SGS Experiment

0.06 0.04 0.02

0.1 0.08 0.06 Dynamic One eq. SGS Experiment

0.04 0.02 0

0 0

0.05

0.1

0

0.15

Column Width, [m]

0.05

0.1

0.15

Column Width, [m]

Fig. 7. Fluctuating liquid velocity profiles for: (a) vertical and (b) lateral components, obtained from dynamic model, one-equation model, and experiments.

The coarse grid was used in all simulations presented below, except for the grid refinement study (Section 3.3.1).

  1.5. dG

Turbulent Kinetic Energy, [m2/s2]

3.3.1. Grid refinement study It is important to show how the one-equation model will behave on a refined grid. In this section, we compare results obtained on two significantly different grids. Both grids are single-block and uniform in all directions. The coarse grid consists of 15 × 15 × 45 cells, resulting in 10 125 cells. The fine grid was obtained by halving the grid size in each coordinate direction, resulting in 30 × 30 × 90 cells and resulting in 81 000 cells. The filter widths for the coarse and fine grids are c = 0.01 m and f = 0.005 m, respectively, where subscript c denotes the coarse grid and f the fine grid. According to Milelli (2002), the ratio of filter width to bubble diameter should obey:

0.03 0.025 0.02 0.015 Dynamic One eq. SGS Experiment

0.01 0.005 0 0

(25)

In both simulations presented here, the bubble diameter was set at a constant value of 4 mm, as reported by Deen et al. (2000), meaning that the coarse grid satisfies the Milelli condition, while the fine grid does not. The time step for the coarse grid was set to 0.01 s which resulted in the CFL number approximately equal to 1. For

0.05 0.1 Column Width, [m]

0.15

Fig. 8. Resolved (dashed) and total (continuous) liquid kinetic energy obtained with one-equation model, compared to the resolved liquid kinetic energy obtained with dynamic model.

the fine grid, we halved the step size and set it to 0.005 s, which resulted in practically the same CFL number as for the coarse grid simulation.

B. Niceno et al. / Chemical Engineering Science 63 (2008) 3923 -- 3931

3929

0.2

Axial Gas Velocity, [m/s]

Axial Liquid Velocity, [m/s]

0.25

0.15 0.1 0.05 No BIT Sato Pfleger Experiment

0 −0.05

0.4 0.35 0.3 0.25 No BIT Sato Pfleger Experiment

0.2 0.15

−0.1

0.1 0.05

0

0.1

0

0.15

0.05

Column Width, [m]

0.1

0.15

Column Width, [m]

Fig. 9. Vertical velocity component for: (a) liquid and (b) gas, obtained with various BIT approaches, and experiments.

0.2 0.12 Lateral Fluctuating Liquid Velocity, [m/s]

Axial Fluctuating Liquid Velocity, [m/s]

0.18 0.16 0.14 0.12 0.1

No BIT Sato Pfleger Experiment

0.08 0.06 0.04 0.02 0

0.1 0.08 No BIT Sato Pfleger Experiment

0.06 0.04 0.02 0

0

0.05

0.1

0.15

Column Width, [m]

0

0.05

0.1

0.15

Column Width, [m]

Fig. 10. Fluctuating liquid velocity profiles for: (a) vertical and (b) lateral components, obtained with various BIT models, and experiments.

Fig. 3 shows the comparison of the vertical liquid and gas velocity components. Both grids show very similar quantitative results. The vertical and lateral liquid velocity fluctuations for both grids are given in Fig. 4. One can observe that the coarser grid gives higher vertical velocity fluctuations than the fine grid. On the other hand, the lateral velocity fluctuations are higher with the fine grid. Qualitatively, the coarse grid seems to give a better agreement, which may seem surprising. However, one must bear in mind the conflicting grid requirement from Euler--Euler and LES expressed by the Milelli condition. In this particular case, the coarse grid satisfies the Milelli condition, while the fine grid does not. In other words, the finer grid is for this case further from physical reality than the coarse grid. A comparison of the liquid turbulent kinetic energy for coarse and fine grids is shown in Fig. 5(a). The dashed lines represent the resolved part of the kinetic energy, while the continuous lines represent the total kinetic energy, i.e. the resolved part plus the modelled part. Both grids show very good comparison with the experimental data. The coarse grid has a pronounced double-peaked profile, while the fine grid, in addition to the double-peaked shape, features a slight increase in kinetic energy in the centre of the column, in-between the two peaks. The ratio of modelled to resolved turbulent kinetic energy is shown in Fig. 5(b). As one would expect, the amount of the SGS kinetic energy is roughly twice as high on the coarse grid than it is on the fine grid.

An important outcome from the grid refinement study is that the results do not differ much on the two grids, meaning that we can use the coarser grid for the rest of the present work. 3.3.2. One-equation and dynamic SGS model The one-equation model is tested against the dynamic model from Germano and experiments. At first BIT was modelled with Sato's model. Fig. 6 shows the long time averaged vertical liquid and gas velocity. It can be observed that both models underpredict the liquid velocity. This underprediction can only be attributed to insufficient drag force, but its analysis is beyond the scope of this work. The one-equation model shows very good agreement for the gas velocity, which is slightly overpredicted with dynamic model. The resolved parts of vertical and lateral liquid phase velocity fluctuations are presented in Fig. 7. Only the resolved parts are shown. The one-equation model compares better to experimental data for vertical fluctuating velocity and is able to reproduce the doublepeaked shape also observed in the experiments. For the lateral velocity fluctuations, the dynamic model performs quantitatively better than the one-equation model. The one-equation model predicts liquid turbulent kinetic energy better than the dynamic procedure, as shown in Fig. 8. The kinetic energy in an LES is partly resolved, while the unresolved part is modelled. The dotted line presents the resolved part of the kinetic energy

B. Niceno et al. / Chemical Engineering Science 63 (2008) 3923 -- 3931

0.03

0.24 0.22

0.025 SGS/Resolved Energy

Turbulent Kinetic Energy, [m2/s2]

3930

0.02 0.015 No BIT Sato Pfleger Experiment

0.01 0.005

0.2

No BIT Sato Pfleger

0.18 0.16 0.14 0.12 0.1 0.08 0.06

0

0.04 0

0.05

0.1

0.15

0

Column Width, [m]

0.05

0.1

0.15

Column Width, [m]

Fig. 11. (a) Resolved (dashed) and total (continuous) liquid kinetic energy and (b) ratio of modelled and resolved part of the turbulent kinetic for various BIT models.

and the continuous line show the sum of resolved and modelled kinetic energy. For the dynamic model, only the resolved part can be shown, since the information on the SGS part of kinetic energy is lost, as explained above. Although the dynamic SGS procedure represents the state of the art in LES of single phase flows, for the EELES of bubble column, the one-equation model performs arguably better. The axial velocity profiles of gas and liquid are predicted equally well with both models, but the transport equation for ksgs significantly improves the prediction of turbulent kinetic energy.

model is close to the extreme values. The ratio of the modelled SGS energy to the resolved energy is shown in Fig. 11(b). As one would expect, modelling no BIT results in the lowest ratio of modelled to resolved kinetic energy. The Sato model yields more SGS energy, while the Pfleger model gives a ratio that is roughly twice as high, which is particularly pronounced in the middle of the column. The Pfleger model shows the best potential to predict the kinetic energy quantitatively, but more elaborate evaluation of the model constant is needed in the future. 4. Conclusions

3.3.3. Bubble induced turbulence Three approaches for modelling BIT in the framework of EELES are considered in this section: (i) no BIT modelling, (ii) BIT by the algebraic model from Sato and Sekoguchi (1975), and (iii) by adding an extra production term in the equation for ksgs , following the rationale from Pfleger and Becker (2001). Fig. 9 presents vertical velocity components for all three approaches. If no BIT is used, the velocity profiles tend to be slightly higher than if the BIT is modelled. The most probable reason for that, is the increased eddy viscosity for the cases in which BIT is modelled. The vertical liquid velocity is better predicted without BIT model, whereas the gas velocity profile is predicted better with modelled BIT. This sheds some new light on the comparison of the vertical velocity components between the dynamic model and the one-equation model. One can observe that the quantitative and qualitative difference between velocity components simulated with the dynamic model and the one-equation model shown in Fig. 6, is the same as the difference between the simulations with and without BIT shown in Fig. 9. This may imply that the eddy viscosity obtained from the one-equation model is overestimated. This issue deserves more attention in the future investigation. The liquid fluctuating velocity components are shown in Fig. 10. The axial components are closest to the experimental findings for the Sato model. Modelling no BIT gives too low profile of the axial component, while the results of the Pfleger approach are quite close to those obtained with the Sato model. The lateral component of the liquid fluctuating velocity is well predicted by all approaches. Fig. 11(a) shows the comparison of the liquid kinetic energy. Following the underpredicted axial component of the fluctuating velocity, the simulation without BIT model underpredicts the kinetic energy as well. The Sato model is the only approach able to reproduce the double-shaped profile for the kinetic energy, while the Pfleger

In this work, we have presented a one-equation SGS model for Euler--Euler LES. Although a transport equation for SGS kinetic energy is known in single-phase LES, to the best of the authors' knowledge, it has never been applied to the Euler--Euler description of multiphase flow. The new modelling approach was applied on the square cross-sectional bubble column from Deen et al. (2000). The obtained results reveal that the one-equation SGS model gives superior results to the well-established dynamic model (Germano et al., 1991), with the additional benefit of giving information on the modelled SGS energy. Grid refinement study revealed that the model predicts smaller amounts of SGS kinetic energy with grid refinement. However, it also confirmed that in Euler--Euler LES one has to make special precautions to prepare the numerical grid, as expressed by Milelli (2002) condition. The information on the amount of SGS kinetic energy offers the possibility to more accurately model physical phenomena at the SGS level. In this work, we have applied it for BIT modelling. As far as BIT is concerned, the Sato model shows the best qualitative comparison of the computed results with the experiments. Pfleger-like models, however, shows best quantitative comparison. The modelled SGS kinetic energy for the Pfleger model is much higher than for the Sato model, meaning that the Pfleger model needs a more appropriate constant for LES. Having the information on SGS kinetic energy makes possible also the modelling of the SGS turbulent dispersion, but it is physically sound only if bubble size are much smaller than filter and, associated with it, the SGS eddy size. In this work, since the bubble size is comparable to the filter size, and much larger than small eddies, SGS turbulent diffusion does not have physical justification. Introducing a one-equation model for the SGS kinetic energy, allows a better insight into the phenomena taking place at the SGS

B. Niceno et al. / Chemical Engineering Science 63 (2008) 3923 -- 3931

level. In this work we showed some initial attempts to take advantage of this information to model the SGS BIT and SGS turbulent dispersion force. Although our modelling attempts did not automatically improve the results, they reveal that modelling of SGS interphase terms has to be reckoned with. References Alajbegovic, A., 2001. Large eddy simulation formalism applied to multiphase flows. In: Proceedings of FEDSM 01. ASME 2001 Fluids Engineering Division Summer Meeting, 29 May--1 June 2001, New Orleans, Louisiana, pp. 529--534. Clift, R., Grace, J.R., Weber, M.E., 1978. Bubbles, Drops and Particles. Academic Press, New York, USA. Davidson, L., 1997. Large eddy simulations: a note on derivation of the equations for the subgrid turbulent kinetic energies. Technical Report No. 97/12, 980904, Chalmers University of Technology, Gothenburg, Sweden. Deen, N.G., Hjertager, B.H., Solberg, T., 2000. Comparison of piv and lda measurement methods applied to the gas--liquid flow in a bubble column. In: Proceedings of the 10th International Symposium on Applications of Laser Techniques to Fluid Mechanics, 9--13 July 2000, Lisbon, Portugal. Deen, N.G., Solberg, T., Hjertager, B.H., 2001. Large eddy simulation of the gas--liquid flow in a square cross-sectional bubble column. Chemical Engineering Science 56, 6341--6349. Dhotre, M.T., Niceno, B., Smith, B.L., 2008. Large eddy simulation of a bubble column using dynamic sub-grid scale model. Chemical Engineering Journal 136 (2--3), 337--348. Drew, D.A., Lahey, R.T., 1987. The virtual mass and lift force on a sphere in a rotating and straining flow. International Journal of Multiphase Flow 13, 113--121. Drew, D.A., Passman, S.L., 1999. Theory of Multicomponent Fluids. Springer, New York, USA. Germano, M., 1992. Turbulence: the filtering approach. Journal of Fluid Mechanics 238, 325--336. Germano, M., Piomelli, U., Moin, P., Cabot, W.H., 1991. A dynamic subgrid-scale eddy viscosity model. Physics of Fluids A 3, 1760--1765.

3931

Ishii, M., Zuber, N., 1979. Drag coefficient and relative velocity in bubbly, droplet or particulate flows. A.I.Ch.E. Journal 25, 843--855. Jakobsen, H.A., Sannæs, B.H., Grevskott, S., Svendsen, H.F., 1997. Modeling of vertical bubble-driven flows. Industrial & Engineering Chemistry Research 36, 4052--4074. Lilly, D.K., 1992. A proposed modification of the Germano subgrid-scale closure method. Physics of Fluids A 4, 633--635. Lopez de Bertodano, M.A., Lahey, R.T., Jones, O.C., 1994. Development of a k −  model for bubbly two-phase flow. Journal of Fluids Engineering---Transactions of the ASME 116, 128--134. Milelli, M., 2002. A numerical analysis of confined turbulent bubble plume. Diss. EH. No. 14799, Swiss Federal Institute of Technology, Zurich, Switzerland. Pfleger, D., Becker, S., 2001. Modeling and simulation of the dynamic flow behaviour in a bubble column. Chemical Engineering Science 56, 1737--1747. Pfleger, D., Gomes, S., Gilbert, N., Wagner, H., 1999. Hydrodynamic simulations of laboratory scale bubble columns fundamental studies of Eulerian Eulerian modelling approach. Chemical Engineering Science 54, 5091--5099. Sato, Y., Sekoguchi, K., 1975. Liquid velocity distribution in two-phase bubble flow. International Journal of Multiphase Flow 2, 79--95. Simonin, O., Viollet, P.L., 1988. On the computation of turbulent two-phase in eulerian formulation. In: Proceedings of EUROMECH 234, Toulouse, France. Smagorinsky, J., 1963. General circulation experiments with the primitive equations. I. The basic experiment. Monthly Weather Review 91, 99--164. Smith, B.L., Milelli, M., 1998. An investigation of confined bubble plume. In: Proceedings of the 3rd International Conference on Multiphase Flow, ICMF '98, 8--12 June 1998, Lyon, France. Tryggvason, G., Esmaeeli, A., Lu, J., Biswas, S., 2006. Direct numerical simulations of gas/liquid multiphase flows. Fluid Dynamics Research 38, 660--681. Yuu, S., Ueno, T., Umekage, T., 2001. Numerical simulation of the high Reynolds number slit nozzle gas particle jet using subgrid-scale coupling large eddy simulation. Chemical Engineering Science 56, 4293--4307. Zhang, D., Deen, N.G., Kuipers, J.A.M., 2006. Numerical simulation of the dynamic flow behaviour in a bubble column: a study of closures for turbulence and interface forces. Chemical Engineering Science 61, 7593--7608. Zun, I., 1990. Mechanism of bubble non-homogeneous distribution in two-phase shear flow. Nuclear Engineering and Design 118, 155--162.

(SGS) modelling for Euler--Euler large eddy simulation

May 4, 2008 - One-equation sub-grid scale (SGS) modelling for Euler--Euler large eddy simulation. (EELES) of dispersed .... Prospects for future developments are also discussed. 2. ..... compared to experimental data of Deen et al. (2000).

424KB Sizes 3 Downloads 245 Views

Recommend Documents

Large Eddy Simulation in Hydraulics.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Large Eddy ...

(SGS) modelling for Euler
May 4, 2008 - One-equation sub-grid scale (SGS) modelling for Euler--Euler large eddy simulation. (EELES) of .... lowing us the analysis of unsteady behaviour of bubbly flows. It is ... means that the same numerical tools developed for ensemble aver-

Large eddy simulation of stably stratified open channel ...
with a fixed temperature difference T across the channel. In that study, the authors found that the turbulent momentum and buoyancy fluxes and the turbulent ...

Large eddy simulation of a bubble column using ...
CFD platform, in: accepted the 12th International Topical Meeting on. Nuclear Reactor Thermal Hydraulics (NURETH-12), Pittsburgh, Pennsyl- vania, USA ...

toward one-dimensional turbulence subgrid closure for large-eddy ...
Our new method combines large-eddy simulation (LES) with the ...... go on endlessly about his technical insight and his ability to truly make learning fun. ..... peak machine performance (like 20%, or worse) whereas ODT utilizes cache in close .... d

Modelling & Simulation (A Model Curriculum for High Schools ...
Whoops! There was a problem loading more pages. Main menu. Displaying Modelling & Simulation (A Model Curriculum for High Schools) - Norfolk, Virginia.pdf.

Distributed Modelling Techniques for System Simulation
The introduction of software tools for model specification has greatly ..... would be marginally more computationally demanding than an analytical solution.

Direct and large-eddy simulations of internal tide ...
3. 0. Figure 2. Contours of the kinetic energy,E, in a case simulated to illustrate the formation of a beam of internal ... Legg (2004) used the Massachusetts Institute of Technology (MIT) model to perform three- .... alternative to Re. We employ ...

System Modelling and Simulation by V.P Singh.pdf
This page. intentionally left. blank. Page 3 of 260. System Modelling and Simulation by V.P Singh.pdf. System Modelling and Simulation by V.P Singh.pdf. Open.

Work plan for the Modelling and Simulation Working Group for 2018
Feb 23, 2018 - Additional virtual meetings may be organised ad-hoc to respond to time-sensitive requests on products and to progress guidelines, as required. 2. Guidelines. 2.1. New EU Guidelines ... Assessor training on the Exposure Response Analysi

Agent Based Modelling and Simulation of the ... - Semantic Scholar
guage. Independently of the programming language, ImmSim has a very rigid ... C-ImmSim and the correspondent parallel variant, ParImm, are versions of Imm-.

Modelling and simulation of fluid-structure interactions ...
range of 20–320 Hz with a mean of 89 Hz by Brietzke and Mair (2006). 3 RESULTS & ..... 19–49. Springer. oomph-lib is available as open-source software at:.

System Modelling and Simulation by V.P Singh.pdf
PERATURAN DIRJEN DIKTI PEDOMAN OPERASIONAL. Desember 2014. Page 3 of 260. System Modelling and Simulation by V.P Singh.pdf. System Modelling ...

pdf-1294\mathematical-modelling-and-simulation-of-electrical ...
... the apps below to open or edit this item. pdf-1294\mathematical-modelling-and-simulation-of-ele ... roceedings-international-series-of-numerical-math.pdf.

Modelling and Simulation of Dry Sliding Wear
(1) University of Karlsruhe,(2) Technical University of Hamburg - Harburg,. (3) GKSS-Forschungszentrum Geesthacht, (4) Forschungszentrum Karlsruhe.

modelling and simulation of corrugated plate structures
500. Normalised pressure (qa4/Eh4 ). 0.00. 0.25. 0.50. 0.75. Norm alise d ce ... experiment has been carried out wherein a box structure with square plates at the top .... plates by spline finite strip method”, Computers and Structures, 76(3), pp.

Agent Based Modelling and Simulation of the ... - Semantic Scholar
Generally, analytic treatment does not yield a complex system's ... based solutions is performed by comparing its perfomance with other approaches to the same ...

SGS Overview-Updated.pdf
contracting of a business. task to another company. Categories: non-voice &. voice-based services. Preferably called as. Contact Center and not “Call. Center ...

WEEKLY SGS REVIEW & STRATEGY - BTInvest
Jan 10, 2014 - palm oil prices may recover in 2014. The key .... according to US DOE data, US crude oil production grew at a ...... and other investment or securities-related services for the corporations whose securities are mentioned in this ...

hppnetsim: a parallel simulation of large-scale ...
HPPNetSim is designed to simulate large/ultra-large interconnection networks and study the communication behavior of parallel applications. In the full system simulator HPPSim, network is abstracted as a single black-box switch, which simulates commu

weekly sgs review & strategy -
Dec 29, 2014 - The PBoC released the Document 387 on Sunday to redefine the ... This will give banks more room to lend to support the real economy.