+ All Categories
Home > Documents > AIAD_Hoehne (3)

AIAD_Hoehne (3)

Date post: 23-Nov-2015
Category:
Upload: dr-mohammed-azhar
View: 16 times
Download: 0 times
Share this document with a friend
Description:
AIAD_Hoehne (3)
Popular Tags:
13
Experiments and numerical simulations of horizontal two-phase flow regimes using an interfacial area density model Thomas Höhne, Christophe Vallée Forschungszentrum Dresden-Rossendorf e.V., P.O. Box: 510119, 01314 Dresden, Germany [Received date; Accepted date] – to be inserted later Abstract Stratified two-phase flow regimes can occur in the main cooling lines of Pressurized Water Reactors, Chemical plants and Oil pipelines. A relevant problem occurring is the development of wavy stratified flows, which can lead to slug generation. In the last decade, stratified flows have increasingly been modelled with computational fluid dynamics (CFD) codes. In CFD, closure models are required that must be validated. Recent improvements of the multiphase flow modelling in the ANSYS CFX code, now make it possible to simulate these mechanisms in detail. In order to validate existing and further developed multiphase flow models, a high spatial and temporal resolution of measurement data are required. For the experimental investigation of co- current air/water flows, the HAWAC (Horizontal Air/Water Channel) was built. The channel allows in particular the study of air/water slug flow under atmospheric pressure. Parallel to the experiments, CFD calculations were carried out. The two-fluid model was applied with a special turbulence damping procedure at the free surface. An Algebraic Interfacial Area Density (AIAD) model based on the implemented mixture model was introduced, which allows the detection of the morphological form of the two-phase flow and the corresponding switching via a blending function of each correlation from one object pair to another. As a result, this model can distinguish between bubbles, droplets and the free surface using the local value of the volume fraction of the liquid phase. The behaviour of slug generation and propagation was qualitatively reproduced by the simulation, while local deviations require a continuation of the work. INTRODUCTION Stratified two-phase flows occur in many industrial applications. The effects of the flow on the quantities (such as flow rate, pressure drop and flow regimes) have always been of engineering interest. Wallis (1973) analysed the onset of slugging in horizontal and near horizontal gas-liquid flows. A prediction of horizontal flow regime transitions in pipes was introduced by Taitel and Dukler (1976). They explained the formation of slug flow by the Kelvin-Helmholtz instability. They also proposed a model for the frequency of slug initiation (1977). The viscous Kelvin- Helmholtz analysis proposed by Lin & Hanratty (1986) generally gives better predictions for the onset of slug flow. A general overview of the phenomenological modelling of slug flow was given by Hewitt (2003). Various multidimensional numerical models were developed to simulate stratified flows: Marker and Cell (Harlow and Welch, 1965), Lagrangian grid methods (Hirt et al., 1974), Volume of Fluid method (Hirt and Nichols, 1981) and Level set method (Osher and Sethian, 1988). These methods can in principle capture accurately most of the physics of the stratified flows. However, they cannot capture all the morphological formations such as small bubbles and droplets, if the grid is not reasonably small enough. One of the first attempts to simulate mixed flows was presented by erne et al. (2001) who coupled the VOF method with a two-fluid model in order to bring together the advantages of the both analytical formulations. Issa
Transcript
  • Experiments and numerical simulations of horizontal two-phase flow regimes

    using an interfacial area density model

    Thomas Hhne, Christophe Valle Forschungszentrum Dresden-Rossendorf e.V.,

    P.O. Box: 510119, 01314 Dresden, Germany

    [Received date; Accepted date] to be inserted later

    Abstract Stratified two-phase flow regimes can occur in the main cooling lines of Pressurized

    Water Reactors, Chemical plants and Oil pipelines. A relevant problem occurring is

    the development of wavy stratified flows, which can lead to slug generation. In the last

    decade, stratified flows have increasingly been modelled with computational fluid

    dynamics (CFD) codes. In CFD, closure models are required that must be validated.

    Recent improvements of the multiphase flow modelling in the ANSYS CFX code,

    now make it possible to simulate these mechanisms in detail. In order to validate

    existing and further developed multiphase flow models, a high spatial and temporal

    resolution of measurement data are required. For the experimental investigation of co-

    current air/water flows, the HAWAC (Horizontal Air/Water Channel) was built. The

    channel allows in particular the study of air/water slug flow under atmospheric

    pressure. Parallel to the experiments, CFD calculations were carried out. The two-fluid

    model was applied with a special turbulence damping procedure at the free surface. An

    Algebraic Interfacial Area Density (AIAD) model based on the implemented mixture

    model was introduced, which allows the detection of the morphological form of the

    two-phase flow and the corresponding switching via a blending function of each

    correlation from one object pair to another. As a result, this model can distinguish

    between bubbles, droplets and the free surface using the local value of the volume

    fraction of the liquid phase. The behaviour of slug generation and propagation was

    qualitatively reproduced by the simulation, while local deviations require a

    continuation of the work.

    INTRODUCTION Stratified two-phase flows occur in many industrial applications. The effects of the flow on the

    quantities (such as flow rate, pressure drop and flow regimes) have always been of engineering

    interest. Wallis (1973) analysed the onset of slugging in horizontal and near horizontal gas-liquid

    flows. A prediction of horizontal flow regime transitions in pipes was introduced by Taitel and

    Dukler (1976). They explained the formation of slug flow by the Kelvin-Helmholtz instability.

    They also proposed a model for the frequency of slug initiation (1977). The viscous Kelvin-

    Helmholtz analysis proposed by Lin & Hanratty (1986) generally gives better predictions for the

    onset of slug flow. A general overview of the phenomenological modelling of slug flow was given

    by Hewitt (2003). Various multidimensional numerical models were developed to simulate

    stratified flows: Marker and Cell (Harlow and Welch, 1965), Lagrangian grid methods (Hirt et al.,

    1974), Volume of Fluid method (Hirt and Nichols, 1981) and Level set method (Osher and

    Sethian, 1988). These methods can in principle capture accurately most of the physics of the

    stratified flows. However, they cannot capture all the morphological formations such as small

    bubbles and droplets, if the grid is not reasonably small enough. One of the first attempts to

    simulate mixed flows was presented by erne et al. (2001) who coupled the VOF method with a two-fluid model in order to bring together the advantages of the both analytical formulations. Issa

  • Experiments and numerical simulations of horizontal two phase flow regimes using an interfacial area density model

    The Journal of Computational Multiphase Flows

    2

    (2003) worked on multi-fluid simulation of slugging phenomena in horizontal channels and

    showed a mechanistic approach to the prediction of hydrodynamic slug initiation, growth and

    subsequent development into continuous slug flow in pipelines. A systematic study of numerical

    simulation of slug flow in horizontal pipes using the two-fluid formulation was carried out by

    Frank (2003). It was shown that the formation of the slug flow regime strongly depends on the

    wall friction of the liquid phase. In simulations using inlet/outlet boundary conditions, it was

    found that the formation of slug flow regimes strongly depends on the agitation or perturbation of

    the inlet boundary conditions. Furthermore, Frank (2003) showed that the length of the

    computational domain plays an important role in slug formation. However, direct comparisons

    between CFD calculations and measurements of the slug generation mechanisms and its

    propagation in horizontal pipes were not performed.

    For the experimental investigation of air/water flows, the HAWAC (Horizontal Air/Water

    Channel) with a rectangular cross-section was built at Forschungszentrum Dresden-Rossendorf

    (FZD). Its inlet device provides defined inlet boundary conditions. The channel allows in

    particular the study of air/water slug flow under atmospheric pressure. Parallel to the experiments,

    CFD calculations were carried out (Valle and Hhne et al., 2008). The aim of the numerical

    simulations presented in this paper is the validation of the prediction of slug flow with newly

    developed and implemented multiphase flow models in the code ANSYS CFX.

    HAWAC The Horizontal Air/Water Channel (HAWAC) (Fig. 1) is devoted to co-current flow experiments.

    A special inlet device provides defined inlet boundary conditions by a separate injection of water

    and air into the test-section. A blade separating the phases can be moved up and down to control

    the free inlet cross-section for each phase, which influences the evolution of the two-phase flow

    regime. The cross-section of this channel is smaller than the channel used in an earlier study

    described in Valle and Hhne et al., 2008: its dimensions are 100 x 30 mm (height x width). The

    test-section is about 8 m long, and therefore the length-to-height ratio L/h is 80. In terms of the

    hydraulic diameter, the dimensionless length of the channel is L/Dh = 173.

    Fig. 1: Schematic view of the horizontal channel with inlet device for a separate injection

    of water and air into the test-section

    pump

    air outlet

    inlet device

    air inlet

    8 m

    GV

    LV

    wire mesh filters

  • Thomas Hhne, Christophe Valle

    Volume 1 Number 1 2009

    3

    Fig. 2: The inlet device

    The inlet device (Fig. 2) is designed for a separate injection of water and air into the channel. The

    air flows through the upper part and the water through the lower part of this device. As the inlet

    geometry introduces many perturbations into the flow (bends, transition from pipes to rectangular

    cross-section), 4 wire-mesh filters are mounted in each part of the inlet device. The filters are

    made of stainless steel wires with a diameter of 0.63 mm and have a mesh size of 1.06 mm. The

    wire-mesh filters are used provide homogenous velocity profiles at the test-section inlet.

    Moreover, the filters produce a pressure drop that attenuates the effect of the pressure surge

    created by slug flow on the fluid supply systems.

    Air and water come in contact at the final edge of a 500 mm long blade that divides both phases

    downstream of the filter segment. The free inlet cross-section for each phase can be controlled by

    inclining this blade up and down. In this way, the perturbation caused by the first contact between

    gas and liquid can be either minimised or, if required, a perturbation can be introduced (e.g.

    hydraulic jump). Both, filters and inclinable blade, provide well-defined inlet boundary conditions

    for the CFD model and therefore offer very good validation possibilities. Optical measurements

    were performed with a high-speed video camera.

    FREE SURFACE MODELING The CFD simulation of free surface flows can be performed using the multi-fluid Euler-Euler

    modelling approach available in ANSYS CFX. Detailed derivation of the two-fluid model can be

    found in the book of Ishii and Hibiki (2006). However, it requires careful treatment of several

    aspects of the model:

    o The interfacial area density should satisfy the integral volume balance condition. In the

    case of when surface waves are present, their contribution to the interfacial area density

    should be taken into account.

    o The turbulence model should address the damping of turbulence near the free surface.

    o The interphase momentum models should take into account the surface morphology.

    o Turbulence damping at the free surface must also be considered.

    Algebraic Interfacial Area Density (AIAD) Model

    Fig. 3 shows different morphologies at slug flow conditions. Separate models are necessary for

    dispersed particles and separated continuous phases (interfacial drag etc.). Two approaches are

    possible within the Euler-Euler methodology:

    Fig. 3: Different morphologies at slug flow conditions

    o Four phases: Bubble/Droplet generation and degassing have to be implemented as

    sources and sinks

    Droplets

    Bubbles

    Air (cont.)

    Water(cont.)

    inclinable blade

    control rod

    air inlet

    water inlet

  • Experiments and numerical simulations of horizontal two phase flow regimes using an interfacial area density model

    The Journal of Computational Multiphase Flows

    4

    o Two phases: Momentum exchange coefficients depend on local morphology

    For the second approach, Egorov (2004) proposed an Algebraic Interfacial Area Density (AIAD)

    Model. The basic idea of the model is:

    o The interfacial area density allows the detection of the morphological form and the

    corresponding switching of each correlation from one object pair to another.

    o It provides a law for the interfacial area density and the drag coefficient for full range

    0r1 (Fig. 4). o The model improves the physical modelling in the asymptotic limits of bubbly and

    droplet flows.

    o The interfacial area density in the intermediate range is set to the interfacial area density

    for free surface (Fig. 4).

    Fig. 4: Air volume fraction and corresponding morphologies/ models

    In an Euler-Euler simulation of horizontal slug,

    the air entrainment below the water surface can

    be caused by the drag force. The magnitude of

    the force density for the drag is

    2

    2

    1|| UACD D = (1)

    where CD is the drag coefficient, A the interfacial area density and the density of the continuous phase (if the other phase is a dispersed phase). U is the relative velocity between the two phases.

    The AIAD model applies three different drag coefficients, CD,B for bubbles, CD,D for the droplets

    and CD,FS for free surface. Non-drag forces (e.g. lift force and turbulent dispersion force) are

    neglected. The interfacial area density A also depends on the morphology of the phases. For

    bubbles it is

    B

    G

    d

    rA

    6= (2)

    where dB is the bubble diameter and rG is the gas void fraction. For a free surface the interfacial

    area density is

    (3)

    is applied as an average density is applied, i.e.

    (4)

    where rL and rG are the liquid and the gas phase density respectively. In the bubbly regime, where

    rG is low, the average density is close to the liquid phase density L. According to the flow regime (bubbly flow, droplet flow or stratified flow with a free surface) the corresponding drag

    coefficients and interfacial area densities have to be applied (Fig. 4).

    n

    rrA LLFS

    ==

    LLGG rr +=

    1=G

    0=G

    Free surface region

    Bubble region

    BBD AC ;,

    FSFSD AC ;,

    Droplet region DDD AC ;,

  • Thomas Hhne, Christophe Valle

    Volume 1 Number 1 2009

    5

    The simplest switching procedure for the interfacial area density uses the blending function f. Such

    functions introduce void fraction limits, the weights for flow regimes and length scales for bubbly

    and droplet flow (dB, dD), which are defined in the following equations:

    (5)

    (6)

    (7)

    (8)

    (9)

    Fig. 7 shows different blending functions f for different volume fraction limits and blending

    coefficients.

    For the simulation of slug flow the void fraction limits of rB,limit=0.3 resp. rD,limit=0.3 and blending

    coefficients of aB=aD=70 were used.

    Modelling the free surface drag

    In simulations of free surface flows, eq. (1) does not represent a realistic physical model. It is

    reasonable to expect that the velocities of both fluids in the vicinity of the interface are rather

    similar. To achieve this result, it is assumed that the shear stress near the surface behaves like a

    wall shear stress on both sides of the interface in order to reduce the velocity differences of both

    phases.

    A viscous fluid moving along a solid like boundary will incur a shear stress, the no-slip

    condition. It is assumed that the morphology region free surface is acting like a wall and a wall

    like shear stress is introduced at the free surface which influences the loss of gas velocity shown in

    Fig. 5:

    . (10)

    Figure 5: Air velocity near the free surface dependent on the normal vectors

    The components of the Normal vector at the free surface are taken from the gradients of the void

    fraction in all directions. As a result, the modified drag coefficient is dependent on the viscosities

    of both phases, the local gradients of gas/liquid velocities normal to the free surface, the liquid

    density and the slip velocity between the phases:

    ( ),limit1

    1 B G Ba r r

    Bf e

    = +

    ( ),limit1

    1 D L Da r r

    Df e

    = +

    1FS B Df f f=

    , , ,FS FS B B D DA f A f A f A = + +

    , , ,D FS D FS B D B D D DC f C f C f C= + +

    0=

    =

    yW y

    u

    Normal vectors at the free surface

  • Experiments and numerical simulations of horizontal two phase flow regimes using an interfacial area density model

    The Journal of Computational Multiphase Flows

    6

    [ ]

    2

    ,,2

    U

    rrC

    L

    GWGLWL

    D

    +=

    (11)

    Turbulence Damping

    As the goal of the CFD calculation was to induce surface instabilities, which generate waves and

    slugs, the interfacial momentum exchange and the turbulence parameters had to be modelled

    correctly. Without any special treatment of the free surface, the high velocity gradients at the free

    surface, especially in the gaseous phase, generate levels of turbulence that are too high throughout

    the two-phase flow when using the differential eddy viscosity models like the k- or the k- model. Therefore, a certain amount of damping of turbulence is necessary in the region of the

    interface, because the mesh is too coarse to resolve the velocity gradient in the gas phase at the

    interface. A few empirical models have been suggested, which address the turbulence anisotropy

    at the free surface, see among others Celik and Rodi (1984). However, no model is applicable for a

    wide range of flow conditions, and all of them are non-local: they require for example explicit

    specification of the liquid layer thickness, of the amplitude and period of surface waves, etc.

    Direct and Large-Eddy Simulation of turbulent multi-material flow have been applied to model

    surface waves. Specifically, the work of Reboux, Sagaut and Lakehal (2006) used DNS to quantify

    the damping of turbulence approaching the interface and incorporated this knowledge into the

    damping of LES turbulence models. Nourgaliev, Liou and Theofanous (2008), Boeck and Zaleski

    (2005), Coward, Renardy and Renardy (1997) reported representations of surface instabilities that

    were obtained by DNS.

    For the two-fluid formulation, Egorov (2004) proposed a symmetric damping procedure. This

    procedure provides a solid wall-like damping of turbulence in both gas and liquid phases. It is

    based on the standard -equation, formulated by Wilcox (1994) as follows:

    ( ) ( ) ( )[ ]++=+

    t

    2

    t Sk

    Ut

    (12)

    where = 0.52 and = 0.075 are the k- model closure coefficients of the generation and the destruction terms in the -equation, = 0.5 is the inverse of the turbulent Prandtl number for , t is the Reynolds stress tensor, and S is the strain-rate tensor. In order to mimic the turbulence damping near the free surface, a source term is introduced on the right hand side of the gas and

    liquid phase -equations. The factor A activates this source term only at the free surface, where it cancels the standard -destruction term of the -equation ( )2iiir and enforces the required high value of i and thus the turbulence damping.

    BOUNDARY CONDITIONS The HAWAC channel with rectangular cross-section was modelled using ANSYS CFX. The

    model dimensions are 8000 x 100 x 30 mm (length x height x width) (Fig. 6a). The grid consists

    of 1.2x106 hexahedral elements. A slug flow experiment at a superficial water velocity of 1.0 m/s

    and a superficial air velocity of 5.0 m/s was chosen for the CFD calculations. In the experiment,

    the inlet blade was in horizontal position. Accordingly, the model inlet was divided into two parts:

    in the lower half of the inlet cross-section, water was injected and air in the upper half. The inlet

    blade was modelled (Fig. 6a). An initial water level of y0 = 50 mm was assumed for the entire

    model length (Fig. 6b). In the simulation, both phases have been treated as isothermal and

    incompressible, at 25C and at a reference pressure of 1 bar. A hydrostatic pressure was assumed

    for the liquid phase. Buoyancy effects between the two phases are taken into account by the

    directed gravity term. At the inlet, the turbulence properties were set using the Medium intensity

    and Eddy viscosity ratio option of the flow solver. This is equivalent to a turbulence intensity of

    5% in both phases. The inner surface of the channel walls has been defined as hydraulically

    smooth with a non-slip boundary condition applied to both gaseous and liquid phases. The channel

    outlet was modelled with a pressure controlled outlet boundary condition. The parallel transient

    calculation of 15.0 s of simulation time on 4 processors took 10 CPU days. A high-resolution

  • Thomas Hhne, Christophe Valle

    Volume 1 Number 1 2009

    7

    discretization scheme was used. For time integration, the fully implicit second order backward

    Euler method was applied with a constant time step of dt = 0.001 s and a maximum of 15

    coefficient loops. Convergence was defined in terms of the RMS values of the residuals, which

    could assured to be less then 10-4

    most of the time. The implementation of the AIAD model and

    turbulence damping functions into CFX was done via the command language CCL.

    a) Fluid domain (channel with inlet

    blade in horizontal position)

    b) Air volume fraction, initial state (Zoom) [-]

    Fig. 6: Model and initial conditions of the volume fractions

    RESULTS: COMPARISON BETWEEN SIMULATION AND EXPERIMENT A simulated free surface at the HAWAC channel (Fig. 6) with small surface instabilities is shown

    in Fig. 8. Fig. 9 shows the resulting interfacial area density variable. The AIAD model uses the

    following three different drag coefficients: CD,B = 0.44 for bubbles, CD,D = 0.44 for the droplets

    and CD,S according to eq. 11 for the free surface (see Fig. 10).

    In the picture sequences (Fig. 11 and 12), a comparison is presented between a CFD calculation

    and an experiment: the calculated phase distribution is visualized and comparable camera frames

    are shown. In both cases, a slug is generated. The sequences show that the qualitative behaviour of

    the creation and propagation of the slug is similar in both the experiment and the CFD calculation.

    !"

    !"#

    # !"#

    Figure 7: Blending functions fB blending

    coeff.

    Figure 8: Air volume fraction [-]

  • Experiments and numerical simulations of horizontal two phase flow regimes using an interfacial area density model

    The Journal of Computational Multiphase Flows

    8

    Figure 9: Interfacial area density variable

    [m-1

    ]

    Figure 10: Drag coefficient [-]

    Fig. 11: Measured picture sequence at JL = 1.0 m/s and JG = 5.0 m/s with t = 50 ms

    (depicted part of the channel: 0 to 3.2 m after the inlet)

    Fig. 12: Calculated picture sequence at JL = 1.0 m/s and JG = 5.0 m/s (depicted part of the

    channel: 1.4 to 6 m after the inlet)

    In this CFD calculation the inlet device was not modelled. Fig. 12 shows the development of the

    slug. These slugs are induced only by instabilities. These single effects leading to slug flow that

    can be simulated here are:

    o Instabilities and small waves randomly generated by the interfacial momentum transfer.

    As a result, bigger waves are generated.

    o The waves can have different velocities and can merge together.

    o Bigger waves roll over smaller waves and can close the channel cross-section.

  • Thomas Hhne, Christophe Valle

    Volume 1 Number 1 2009

    9

    In order to extract quantitative information from the calculation as well as from the measurements,

    interface capture methods were developed. Moreover, a statistical treatment of the data is proposed

    for the comparison between both results.

    Interface capture in the high-speed video observations

    To capture the gas-liquid interface in the camera frames, an image-processing algorithm was

    developed. The capture method illustrated in Fig. 13 consists of the following steps:

    1. the synthesis of a background from the picture sequence by filtering out its dynamic

    component (Fig. 13-b);

    2. a background subtraction with the image generated in step 1 (Fig. 13-c);

    3. a pixel detection in each vertical line:

    a. of the darkest pixel;

    b. of the minimum of a grey-level time variation;

    4. selection of the pixels detected in step 3 that best fit onto a continuous interface line (Fig.

    13-d).

    The capture method allows the representation of the interface by a water level as function of the

    duct length z and the time t. The accuracy of the interface detection algorithm depends on the

    thickness of the interface in the images. In the stratified flow regions, the interface thickness is

    quite thin with at most 3 pixels. Therefore, the accuracy of the water level measurement is there

    about 1.5 pixels. This corresponds to 3.9 mm for a picture resolution of about 2.6 mm/pixel in

    this experiment. The accuracy is worse in the region of the slug front, where a two-phase mixture

    is generated (red circle in Fig. 13-c). This makes a definition of the water level difficult and

    sometimes induces unphysical fluctuations in the detected interface. In particular, at the end of the

    visualised region (further than 2.5 m from the inlet), where the slugs are developed the

    interpretation of the water level measurement is sometimes quite delicate.

    Interface capture in the CFD simulation

    Contrary to the interface tracking methods, the two-fluid model does not reproduce a sharp

    interface between air and water. Nevertheless, for the comparison between the CFD calculation

    and experimental results, a surface similar to the interface observed in the

    a

    b

    c

    d

    Fig. 13: Original picture (a), background picture (b), picture with subtracted background

    (c) and detected interface during slug flow (d)

    camera pictures should be defined. Therefore, an isosurface with a void fraction of 50% was

    chosen (Fig. 14) and the coordinates of its intersection with the vertical mid-plane was exported

    from ANSYS-CFX. With this simplification, the three-dimensional shape of the isosurface is not

    taken into account.

    For each time step, the exported data set was treated in order to determine the minimum and

    maximum water levels in each vertical cross-section. As shown in Fig. 15, this data set is

    composed of several domains representing the free surface, bubbles or droplets.

  • Experiments and numerical simulations of horizontal two phase flow regimes using an interfacial area density model

    The Journal of Computational Multiphase Flows

    10

    At first, the free surface has to be recomposed. The domains representing bubbles and droplets are

    recognised and separated from the domains corresponding to the free surface. Then, free surface

    domains are unified to a single one by connecting them in the empty spaces left by slugs on top of

    the channel. In the next step, the droplet and bubble domains are treated. In order to obtain similar

    results to the interface capture method used for the experimental data, only the bigger structures

    with a circumference longer than approximately 20 mm are selected. The small formations are also

    mostly excluded in the experiment because they do not fit into a continuous interface line (step 4

    of the method described in the previous section). Furthermore, droplets attached to the top of the

    channel are not taken into account. Finally, the minimum and maximum water levels are

    determined from the coordinates of the free surface domain as well as the selected bubble/droplet

    domains. This is done in each vertical cross-section with a resolution of 1 cm in the horizontal axis

    (Fig. 16).

    Statistical treatment of the measured water levels for comparison with CFD

    Since a direct comparison of the measured

    water levels with CFD results is difficult, a

    statistical approach is proposed. Therefore,

    a time-averaged water level was calculated

    and bounded by the standard deviation in

    each cross-section (Fig. 17). This results in

    a mean water level profile along the

    channel, which reflects the structure of the

    interface. Furthermore, the standard

    deviation quantifies the spread of the measured values, which originates from the

    dynamic change of the free surface. In the

    first part of Fig. 17, a slight increase of the

    mean water level from 50 mm at the inlet to 58 mm is observed as well as a low standard

    deviation. Both are characteristic for the supercritical flow (Fr 1) obtained at the test-section inlet. In fact, in a supercritical flow the pressure drop due to wall friction results in an increase of

    the water level. Furthermore, only small supercritical waves can propagate in such a flow. Around

    the maximum of the mean water level reached at about 0.9 m from the inlet, the standard deviation

    increases quickly up to about 18 mm. This data points out the rapid wave growth induced by the

    high air velocity in this zone. In the downstream region, where the slugs are generated and

    propagate the mean water level decreases to an asymptotic value of about 30 mm. It shows that

    there is an acceleration of the water flow, which can be attributed to the exchange of momentum

    between the phases.

    0,00

    0,05

    0,10

    4,00 4,25 4,50 4,75 5,00 5,25 5,50

    Length [m]

    Wate

    r le

    vel [m

    ]

    Fig. 15: Example of data set exported from ANSYS-CFX (in red: domains of the free

    surface; in blue: a bubble domain)

    Comparison with the CFD calculation

    The time-averaged water level profiles were calculated for the CFD results from the minimum and

    maximum water levels determined as described previously. These are shown in Fig. 18 for t =

    Fig. 1 Side view of the calculated void fraction and structure of the 50% isosurface (in green)

  • Thomas Hhne, Christophe Valle

    Volume 1 Number 1 2009

    11

    17.25020.250 s (i.e. from the passage of the first slug clearing the channel to the end of the

    calculation time) and are bounded by the standard deviation.

    Qualitatively, Fig. 18 shows that the trend obtained for the simulation is similar to the

    measurement (Fig. 17). Like in the experiment, the mean water level profile increases after being

    injected over a height of 50 mm at the inlet. Further downstream, the water levels decrease

    simultaneously with an increase of the standard deviation. Furthermore, a difference between

    minimum and maximum appears which is due to the development of waves that roll over smaller

    waves and slugs or the presence of bigger bubbles or droplets structures in the flow. In the last part

    of the channel, the mean water levels converge slightly together, probably due to collapsing slugs.

    The water levels tend to about 24 mm and 30 mm for the minimum and maximum water levels

    respectively.

    However, a detailed comparison shows quantitative deviations between simulation and

    measurement. In the interface rising region, the water level reaches a maximum of 68 mm, which

    is 10 mm more than in the experiment. This region is followed by a plateau between 2 and 3 m,

    which is only observed in the calculation. Furthermore, wave growth starts downstream of the 3 m

    mark, revealed by a rapid increase of the standard deviation, compared to about 0.9 m in the

    experiment.

    These quantitative differences can be explained with the flow regimes observed at the test-section

    inlet. In fact, the flow pattern has an import influence on the momentum exchange between the gas

    and liquid, especially with high velocity differences between the phases. Small disturbances of the

    interface provide a more efficient momentum transfer from the air to the water than in a stratified

    smooth flow. A high momentum transfer induces a rapid wave growth and therefore slug

    generation. In this case, the very low standard deviation ( 0.5 mm) observed over the 3 first meters in the simulation reveals a smooth interface, whereas in the experiment the flow is wavy

    from the inlet of the channel. This means that the boundary conditions chosen for the CFD model

    do not reproduce the small disturbances observed in the experiment. In the end, a quite long

    channel length is needed before waves appear spontaneously in the simulation, deferring

    accordingly the instable wave growth to slugs. This occurs downstream of the 3 m mark, inducing

    an increase of the standard deviation up to about 18 mm (at 4.00 m) and 23 mm (at 5.20 m) for the

    minimum and maximum water levels respectively. These values correspond to the order of

    magnitude measured about 3 m after the inlet.

    Finally, the quantitative differences noticed between simulation and experiment concern in

    particular the inlet boundary conditions. As the conditions have an important influence on the

    generation of the two-phase flow, future work should focus on the proper modelling of the small

    instabilities observed at the channel inlet.

    0

    20

    40

    60

    80

    100

    0,0 0,5 1,0 1,5 2,0 2,5 3,0 3,5

    Distance from the inlet blade [m]

    Wate

    r le

    ve

    l [m

    m]

    Fig. 16: Time-averaged water level bounded by the standard deviation

  • Experiments and numerical simulations of horizontal two phase flow regimes using an interfacial area density model

    The Journal of Computational Multiphase Flows

    12

    0,00

    0,02

    0,04

    0,06

    0,08

    0,10

    0 1 2 3 4 5 6 7 8

    Length [m]

    Wa

    ter

    level [m

    ]

    Min

    Max

    Fig. 17: Comparison between the mean water level profiles (ANSYS-CFX calculation)

    Legend: dark colours: average ; light colours: average standard deviation

    SUMMARY In the HAWAC test facility, a special inlet device provides well defined as well as variable

    boundary conditions, which allow very good CFD-code validation possibilities. A picture

    sequence recorded during slug flow was compared with the equivalent CFD simulation made with

    the code ANSYS CFX. The two-fluid model was applied with a special turbulence damping

    procedure at the free surface. An Algebraic Interfacial Area Density (AIAD) model on the basis of

    the implemented mixture model was introduced and implemented. It improves the physical

    modelling, detection of the morphological form and the corresponding switching of each

    correlation is now possible. The behaviour of slug generation and propagation at the experimental

    setup was reproduced, while deviations require a continuation of the work. Experiments are

    planned that include pressure and velocity measurements that will allow quantitative comparisons.

    The formation of waves and slugs at other superficial velocities will also be studied.

    ACKNOWLEDGEMENTS This work was carried out in the frame of a research project funded by the German Federal

    Ministry of Economics and Labour, project number 150 1329. Thanks to Yuri Egorov and Thomas

    Frank from ANSYS CFX for their fruitful cooperation.

    REFERENCES

    ANSYS CFX, 2008. User Manual. Ansys Inc.

    Baker, O. 1954. Simultaneous flow of oil and gas. Oil & Gas Journal, 53, 184-195.

    Boeck, T. and Zaleski, S. 2005. Viscous versus inviscid instability of two-phase mixing layers with continuous velocity

    profile, Phys. Fluids, 17, 032106-1-032106-11.

    Celik, I., and Rodi, W. 1984. A deposition-entrainment model for suspended sediment transport. Report SFB 210/T/6,

    Strmungstechnische Bemessungsgrundlagen fr Bauwerke, University of Karlsruhe, Germany.

    Cerne, G., Petelin, S., Tiselj, I., 2001. Coupling of the interface tracking and the two-

    fluid models for the simulation of incompressible two-phase flow. Journal of Computational Physics 171 (2), 776-

    804.

    Coward, A.V.,Renardy, Y., Renardy, M. and Richards, J., 1997. Temporal evolution of periodic disturbances in two-layer

    Couette flow, J. Comp. Phys. 132, pp. 346361

    Egorov, Y. 2004. Contact condensation in stratified steam-water flow, EVOL-ECORA D 07.

    Frank, T. 2003. Numerical simulations of multiphase flows using CFX-5. CFX Users conference, Garmisch-Partenkirchen,

    Germany.

    Harlow, F. H., Welch, J. E., 1965. Numerical calculation of time-dependent viscous incompressible flow of fluid with free

    surface. Physics of Fluids 8 (12), 2182-2189.

    Hewitt, G. F. 2003. Phenomenological modelling of slug flow. Short course modelling and computation of multiphase

    flows, ETH Zurich, Switzerland.

    Hirt, C. W., Nichols, B. D., 1981. Volume of fluid (VOF) method for the dynamics of free boundaries. Journal of

    Computational Physics 39 (1), 201-225.

    Ishii, M., Hibiki, T., 2006. Thermo-fluid Dynamics of Two-phase Flow. Springer-Verlag.

  • Thomas Hhne, Christophe Valle

    Volume 1 Number 1 2009

    13

    Issa, R., e.g., 2003. Simulation of slug flow in horizontal and nearly horizontal pipes with the two-fluid model, INT J

    MULTIPHAS FLOW, Vol: 29, Pages: 69 95.

    Liebert, J., Gaul, H.-P., Weiss, P. 1997. UPTF-Experiment: Durchfhrung von Versuchen zu Fragestellungen des

    anlageinternen Notfallschutzes (TRAM), Abschlubericht zum Forschungsvorhaben 1 500 876. Siemens AG,

    KWU, KWU NT31/97/58.

    Lin, P. Y., and Hanratty, T. J. 1986. Prediction of the initiation of slugs with linear stability theory. Int. J. Multiphase Flow,

    12, 79-98.

    Lun, I., Calay, R. K., and Holdo, A. E. 1996. Modelling two-phase flows using CFD. Applied Energy, 53, 299-314.

    Mandhane, J. M., Gregory, G. A., and Aziz, K. 1974. A flow pattern map for gas-liquid flow in horizontal pipes: predictive

    models, Int. J. Multiphase Flow, 1, 537-553.

    Manfield, P. D. 2000. Experimental, computational and analytical studies of slug flow. PhD thesis, Imperial College of

    Science, Technology and Medicine, London.

    Menter, F. R. 1993. Zonal two-equation k- turbulence models for aerodynamic flows. AIAA, 93-2906.

    Menter, F. R. 2002. CFD best practice guidelines for CFD code validation for reactor safety applications. EC project

    ECORA, report EVOL-ECORA-D01.

    Moe, R. 1992. Transient simulation of 2-3D stratified and intermittent two-phase flows. PhD thesis, University of Oslo.

    Nourgaliev, R., Liou, M.-S. and Theofanous, T.G. 2008. Numerical Prediction of Interfacial Instability: Sharp Interface

    Method (SIM), Journal of Computational Physics, 227, pp.3940-3970.

    Osher, S. and Sethian, J. A., 1988. Fronts propagating with curvaturedependent speed: algorithms based on Hamilton-

    Jacobi formulations. J. Comput. Phys., vol. 79, no. 1, pp. 1249.

    Reboux S., Sagaut P., Lakehal D., 2006. LES of Sheared Interfacial Two-Fluid Flow, Phys. Fluids, 18 (10).

    Scheuerer, M. 2003. Selection of PTS-relevant test cases, EC-report EVOL-ECORA D05a.

    Soleimani, A., and Hanratty, T. J. 2003. Critical liquid flows for the transition from the pseudo-slug and stratified patterns

    to slug flow. Int. J. of Multiphase Flow, 29, 51-67.

    Taitel, Y., and Dukler, A. E. 1976. A model for predicting flow regime transitions in horizontal and near horizontal gas-

    liquid flow. AIChE J. 22, 47-55.

    Taitel, Y., and Dukler, A.E., 1977. A model for slug frequency during gas-liquid flow in horizontal and near horizontal

    pipes. Int. J. Multiphase Flow, 3, 585-596.

    Ujang, P. M. et al. 2006. Slug initiation and evolution in two-phase horizontal flow. Int. J. Multiphase Flow, 32, 527-552.

    Valle, C., Hhne, T., Prasser, H.-M., Shnel, T. 2008. Experimental investigation and CFD simulation of horizontal

    stratified two-phase flow phenomena. NED, Volume 238, Issue 3, Pages 637-646

    Wallis, G. D., and Dobson, J. E. 1973. Onset of slugging in horizontal stratified air-water flow. Int. J. Multiphase Flow 1,

    173-193.

    Wilcox, D. C. 1994. Turbulence modelling for CFD. La Caada, California: DCW Industries Inc.

    Woods, B. D., and Hanratty, T. J. 1999. Influence of Froude number on physical processes determining frequency of

    slugging in horizontal gas-liquid flows. Int. J. of Multiphase Flow, 25, 1195-1223.


Recommended