investigation of shallow mixing layers by bgk finite...

16
This article was downloaded by: [University of California, Los Angeles (UCLA)] On: 05 January 2013, At: 00:55 Publisher: Taylor & Francis Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK International Journal of Computational Fluid Dynamics Publication details, including instructions for authors and subscription information: http://www.tandfonline.com/loi/gcfd20 Investigation of shallow mixing layers by BGK finite volume model Mohamed S. Ghidaoui a & Jun Hong Liang b a Hong Kong University of Science and Technology, Kowloon, Hong Kong, P.R. China b University of California, Los Angeles, CA, USA Version of record first published: 24 Jul 2008. To cite this article: Mohamed S. Ghidaoui & Jun Hong Liang (2008): Investigation of shallow mixing layers by BGK finite volume model, International Journal of Computational Fluid Dynamics, 22:7, 523-537 To link to this article: http://dx.doi.org/10.1080/10618560802238283 PLEASE SCROLL DOWN FOR ARTICLE Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions This article may be used for research, teaching, and private study purposes. Any substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to anyone is expressly forbidden. The publisher does not give any warranty express or implied or make any representation that the contents will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses should be independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.

Upload: others

Post on 25-Jan-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

  • This article was downloaded by: [University of California, Los Angeles (UCLA)]On: 05 January 2013, At: 00:55Publisher: Taylor & FrancisInforma Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,37-41 Mortimer Street, London W1T 3JH, UK

    International Journal of Computational Fluid DynamicsPublication details, including instructions for authors and subscription information:http://www.tandfonline.com/loi/gcfd20

    Investigation of shallow mixing layers by BGK finitevolume modelMohamed S. Ghidaoui a & Jun Hong Liang ba Hong Kong University of Science and Technology, Kowloon, Hong Kong, P.R. Chinab University of California, Los Angeles, CA, USAVersion of record first published: 24 Jul 2008.

    To cite this article: Mohamed S. Ghidaoui & Jun Hong Liang (2008): Investigation of shallow mixing layers by BGK finitevolume model, International Journal of Computational Fluid Dynamics, 22:7, 523-537

    To link to this article: http://dx.doi.org/10.1080/10618560802238283

    PLEASE SCROLL DOWN FOR ARTICLE

    Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

    This article may be used for research, teaching, and private study purposes. Any substantial or systematicreproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form toanyone is expressly forbidden.

    The publisher does not give any warranty express or implied or make any representation that the contentswill be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses shouldbe independently verified with primary sources. The publisher shall not be liable for any loss, actions, claims,proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or indirectly inconnection with or arising out of the use of this material.

    http://www.tandfonline.com/loi/gcfd20http://dx.doi.org/10.1080/10618560802238283http://www.tandfonline.com/page/terms-and-conditions

  • Investigation of shallow mixing layers by BGK finite volume model

    Mohamed S. Ghidaouia* and Jun Hong Liangb

    aHong Kong University of Science and Technology, Kowloon, Hong Kong, P.R. China; bUniversity of California, Los Angeles,CA, USA

    (Received 14 May 2008; final version received 29 May 2008 )

    Turbulent shallow mixing layers and their associated vortical structures are ubiquitous in rivers, estuaries and coasts.Examples of these flows can be found in compound/composite channels, at the confluence of two rivers, at harbour entrancesand at groyne fields. A finite volume 2D model, based on the averaging of the 3D shallow water equations with respect todepth and in which the numerical fluxes are obtained from the Bhatnagar–Gross–Krook (BGK) Boltzmann equation, isapplied to shallow mixing layers for which experimental results are available. This model is hereafter referred to as the BGKmodel or BGK scheme. The BGK scheme is explicit, second order in time and space and conserves bothmass andmomentum.The BGK relaxation time is locally evaluated from the classical turbulence model of Smagorinsky. The BGK modelaccurately represents the mean flow field such as mean velocity profile, mean spread of the mixing layer, mean position of themixing layer centreline and mean surface water profile. In addition, the Kelvin–Helmholtz (KH) instability includinginception, vortex roll up, vortex growth by pairing and the eventual decay of the vortices by bed shear is well represented bythe model. On the other hand, the magnitude of the turbulence intensity is over-predicted by the shallow water model. Thisdiscrepancy is partly due to the fact that the turbulence forcing assumed may not represent the actual random perturbationsthat may exist in the laboratory experiments and partly due to the inability of the depth-averaged shallow water equations toallow for the redistribution of turbulent energy along the vertical direction, since these governing equations do not model the3D turbulence. Thus, the depth-averaged shallow water equations are well suited for investigating the KH stability and forpredicting mean flow field including velocity profiles and transversal mixing of mass momentum in shallow environments.Accurate prediction of turbulence statistics would require resolving the small 3D scales with respect to water depth.

    Keywords: shallow mixing layers; stability of shallow shear flows; BGK modelling of shallow flows; coherent structures;quasi-2D turbulence

    1. Introduction

    Turbulent mixing layers are commonly observed in various

    engineering applications such as combustion, propulsion

    and environmental flows. Turbulent mixing layers are

    susceptible to Kelvin–Helmholtz (KH) instabilities and

    these instabilities are known to result in complex flow

    phenomena such as vortex roll up and pairing. Turbulent

    deep mixing layers, where the length scale perpendicular to

    the plane of the flow is much larger than the length scale in

    the plane of the flow, have been extensively investigated

    through stability theory as well as laboratory and numerical

    experiments. It is found that the growth of KH instabilities

    lead toKHvortices. The subsequent pairing of these vortices

    lead to organised nearly 2D vortical structures. These

    organised structures are themselves susceptible to secondary

    instabilities, which are responsible for the breakdown of the

    2D organised structures into 3D turbulence. The breakdown

    of organised structures into 3D turbulence is associated with

    the stretching and tilting of vortices and follows the classical

    energy cascade principle, where turbulent kinetic energy

    flows from large to small scales.

    Turbulent shallow mixing layers and their associated

    vortical structures are ubiquitous in rivers, estuaries and

    coasts. Shallow mixing layers in surface waters are

    characterised by having a large horizontal length scale in

    comparison to their vertical length scale. The spectacular

    Naruto vortical structures (whirlpools) generated by tidal

    exchange between the Pacific Ocean and the Seto Inland

    sea in Japan are a typical example of shallow mixing

    layers. The understanding of shallow mixing layers is

    important for modelling water quantity and quality, and

    for the analysis and prediction of the transversal exchange

    and spread of mass, momentum and energy in shallow

    water environments.

    The behaviour of vortical structures in shallow mixing

    layers is distinct from that in deep flows. Unlike in deep

    flows, vortical structures in shallow flows are subjected to

    bottom shear stresses and their vertical extent is limited by

    the water depth. Indeed, stability analysis and experimental

    investigations show that the dynamics of shallow mixing

    layers is controlled by the bottom shear and by the

    shallowness of the flow (e.g. Chu et al. 1991, Chen and Jirka

    ISSN 1061-8562 print/ISSN 1029-0257 online

    q 2008 Taylor & Francis

    DOI: 10.1080/10618560802238283

    http://www.informaworld.com

    *Corresponding author. Email: [email protected]

    International Journal of Computational Fluid Dynamics

    Vol. 22, No. 7, August 2008, 523–537

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • 1995, 1997, Ghidaoui and Kolyshkin 1999, Uijttewaal and

    Booij 2000, Kolyshkin and Ghidaoui 2002, van Prooijen and

    Uijttewaal 2002a, 2002b).

    The role of shear stresses and flow shallowness on

    the behaviour of turbulent shallow mixing layers is as

    follows. KH instabilities result in the formation of

    vortical structures extending from the flow bed to the

    water surface. The confinement of the flow by the water

    depth suppresses the vortex stretching mechanism, forcing

    the KH vortical structures to move in horizontal plane and

    to exhibit a strongly 2D behaviour. The presence of the

    vortex pairing mechanism and the absence of the vortex

    stretching mechanism result in a reverse energy cascade,

    which leads to a horizontal growth of the vortical

    structures along the downstream direction. The shallow

    flow experiments of Uijttewaal and Booij (2000) and

    Uijttewaal and Jirka (2003) show that there is a range

    of scales for which the slope of energy spectrum follows

    the -3 law. The -3 law provides clear evidence for the

    reverse energy cascade, also called enstrophy cascade and

    explains the flow of energy from small to large scales

    evidenced by the growth of the vortical structures in the

    downstream direction. The bottom friction, on the other

    hand, dissipates energy. This dissipation increases as the

    scale of the vortices increases. As a result, the bottom

    friction reduces the flow of energy from the small to the

    large scales and limits the extent of the enstrophy cascade.

    Indeed, the shallow flow experiments of Chu and

    Babarutsi (1988) and Uijttewaal and Booij (2000) show

    that the growth/spread rate of the mixing layer decreases

    with downstream distance and eventually becomes

    negligibly small.

    Most current channel models are still based on 1D or

    2D shallow water equations due to their relatively low

    computational cost. However, their ability in modelling

    complex flow structures governed by 3D dynamics, which

    features channels with complex topographical conditions,

    is still questionable. For example, Ghidaoui et al. (2006)

    found that the shallow water model underpredicts the

    bubble size in shallow recirculation zone in the wake of a

    circular cylinder by about 30% when compared to

    laboratory experiments. It is conjectured by Ghidaoui

    et al. (2006), that the difference is partly due to the failure

    of modelling unsteady flow conditions by friction

    coefficients derived from steady flows.

    The current paper investigates the ability of the

    shallow water equations in modelling shallow mixing

    layers using a finite volume numerical solution on the

    basis of the Boltzmann equation. The classical Smagor-

    insky turbulence model is used to estimate the collision

    term. The model is applied to the experimental set-up of

    Uijttewaal and Booij (2000). The study investigates the

    ability of the shallow water model to reproduce the mean

    flow field, the turbulence statistics and the general flow

    features such as vortex roll up and pairing.

    2. Governing equations

    Governing equations for shallow turbulent flow can be

    obtained by integrating the Navier–Stokes equations

    (Batchelor 1967) over the vertical dimension and averaging

    over resolved scales. The depth-averaged shallow water

    equations, obtained by integrating the incompressible 3D

    Navier–Stokes equations, invoking the kinematic condition

    at the free surface, the no slip condition at the bed and the

    hydrostatic pressure assumption gives

    ›h

    ›tþ ›huk

    ›xk¼ 0; ð1Þ

    ›hua›t

    þ›ðhuaukþghh=2Þ›xk

    ¼ ghSa2 tarþ y ›

    ›xkh›ua›xk

    � �;

    ð2Þ

    where t is time; x is spatial coordinates;k ¼ 1, 2 anda ¼ 1, 2with 1 indicating the streamwise direction and 2 indicating

    the cross-stream direction;g is the gravitational acceleration;

    h is water depth; u is the depth-averaged velocity; S is the bed

    slope; r is the density of water; n is the kinematic viscosity ofwater; and t is the bottom shear stress (shear between thefluid and the bed).

    Similar to compressible flows, it is convenient to

    introduce the mass-weighted (Favre) filtering as follows

    (e.g. Xu et al. 2005):

    ~f ¼ hf�h; ð3Þ

    where ~f is the Favre filtering of f. Applying the Favre

    filtering to systems (1)–(2) gives:

    ›�h

    ›tþ ›

    �h~uk

    ›xk¼ 0; ð4Þ

    ›�h~ua›t

    þ ››xk

    �h~ua ~ukþdak 12g�h2

    � �

    ¼ g�hSa2 �tarþ y ›

    ›xk�h›~ua›xk

    � �þ ››xk

    ð�h~ua ~uk2 �h~ua ~ukÞ:

    ð5Þ

    The last term on the right hand side of (5) requires a

    turbulence closure model. The effects of unresolved

    motions on resolved motions are conventionally related to

    resolved motions by an expression similar to that for

    viscous stress except that the kinematic viscosity is

    replaced by the eddy viscosity. Various turbulence models

    are devised to relate the eddy viscosity with the resolved

    flow field. In the current study, the Smagorinsky model

    (Pope 2002) is used,

    y e ¼ CSlm2ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2SkaSka

    p; Ska ¼ 1

    2

    ›~ua›xk

    þ ›~uk›xa

    � �; ð6Þ

    M. Ghidaoui and J.H. Liang524

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • where CS is the Smagorinsky constant; ye is eddyviscosity; lm is the mixing length, which is related to the

    grid size by lm ¼ffiffiffiffiffiffiffiffiffiffiffiDxDy

    p. Although there is no theoretical

    foundation for why the Smagrorinsky model can be

    applied to 2D flows, the adoption of this model outside its

    domain of applicability has generally led to encouraging

    results. For example, the model adopted in this study has

    been successfully applied by Zhou (2004) to 2D channel

    flows and by Zhou (2004) as well as Ghidaoui et al. (2006)

    to shallow wake flows. In addition, the Princeton ocean

    model (POM) uses the Smagorinsky relation to estimate

    the horizontal component of the turbulent diffusion

    (Mellor 2004). POM has been successfully applied to a

    wide range of problems in ocean circulations (e.g. Oey

    et al. 2005). It is worthwhile noting that the 2D version of

    the Smagorinsky model belongs to the general class of

    mixing-length models (Pope 2002), where the mixing-

    length in the Smagorinsky model is related to the

    computational grid-size. The linkage between the

    Boltzmann equation and the filtered shallow water

    equation requires that the collision time be given by

    t ¼ ðy þ y eÞ=g�h. The code evaluates the collision time atevery computational node in order to account for the

    variability of eddy viscosity with time and space.

    For free-surface flows, shear stresses are commonly

    modelled by the following quadratic friction law

    (Schlichting and Gersten 2000, Pope 2002),

    �tar¼ cf ~ua

    ffiffiffiffiffiffi~u 2k

    p2

    ; ð7Þ

    where cf is the friction coefficient. For flows over a smooth

    bottom, the following semi-empirical law is often

    implemented for the friction coefficient (Schlichting and

    Gersten 2000, Pope 2002),

    1ffiffiffifficf

    p ¼ 24 log 1:254Re

    ffiffiffifficf

    p� �

    ð8Þ

    3. Numerical model

    Systems (4)–(8) are solved using a conservative finite

    volume method on irregular grids, where the algorithm for

    the mass and momentum fluxes at the control surface of

    the finite volume is obtained from the solution of the

    Bhatnagar–Gross–Krook (BGK) Boltzmann equation.

    Such approach is selected because of its conservation

    properties, its ability to handle complex geometry, and to

    solve waves and turbulence without the need for operator

    splitting (Ghidaoui et al. 2001, Liang et al. 2007). Such

    properties are critical for the authors’ current and future

    research on shallow flows, which entails the study of 2D

    and 3D instabilities in shallow flows in geometries with

    varying degrees of complexities and the interaction

    between turbulence and waves.

    The discretised form of (4) and (5) is (details are

    available in Refs. Ghidaoui et al. (2001), Liang et al.

    (2007)),

    Unþ1i; j ¼Uni; jþSni; jDt

    21

    Vi; j

    XSi; js¼1

    Ls g1þg3 ››t

    � �ðAsiþAsoÞ

    þ Dt2g1þg6 ››t

    � �Esþðg2þg3Þ

    £ ½ns·7ðBsiþBsoÞþts·7ðDsiþDsoÞ�

    þl5ðns·7Gsþts·7IsÞ�: ð9Þ

    Where U¼ b�h; ~ua �hc ; S ¼ [›b/›xa]; (i, j) denotes spatialposition; n indicates time step; V is the area of a grid;Ls is the length of the side s of a grid; ns is the unit vectornormal to the side s; ts is the unit vector normal to the side s;

    g1¼t 2te2Dt=t; g2¼2t2þðt2þtDtÞe2Dt=t; g3 ¼ tg1;g5¼2t Dtþ2t2ð12e2Dt=tÞ2Dtte2Dt=t; g6¼tDt2t2ð12e2 Dt=tÞ, with t¼ ðy þyeÞ=gh. The matrices in (9) aregiven in the Appendix. It is important to note that the collision

    time and node (n,i,j) is given by:

    t ni; j¼yþðyeÞni; jghni; j

    : ð10Þ

    In what follows, the streamwise direction x1 is denoted by x

    and the cross-stream direction x2 by y.

    4. Hydraulic and numerical parameters

    The numerical test rigs correspond to the experiments

    conducted by Uijttewaal and Booij (2000) and van

    Prooijen and Uijttewaal (2002a). Shallow mixing layers

    develop from parallel streams of different velocities

    separated by a splitter at the inflow boundary. Hydraulic

    parameters for the tests are summarised in Table 1.

    A variety of Smagorinsky coefficients were tested. It is

    found that Cs ¼ 0.2 gives the best correlation with theexperimental data. The sensitivity of the results to the

    Smagorinsky coefficient is given in Figure 17.

    Table 1. Hydraulic conditions for shallow mixing layer tests.

    Case U1 (m/s) U2 (m/s) H (m) cf ( £ 1023)1 0.23 0.11 0.042 6.42 0.32 0.14 0.067 5.4

    International Journal of Computational Fluid Dynamics 525

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • At the inflow (i.e. x ¼ 0), zero lateral velocity isimposed and a mean flow profile is prescribed for

    streamwise velocity. In order to minimise the reflection

    from the outflow boundary, a radiation boundary condition

    (Durran 1999) is imposed at x ¼ Lx. At the lateralboundary (i.e. y ¼ ^ (1/2)W), a free-slip condition isprescribed. It is shown by Socolofsky and Jirka (2004) that

    the stability of co-flowing shallow mixing layers are of

    convective nature. This means that the flow behaves like a

    ‘noise-amplifier’ (Huerre and Monkewitz 1990), and

    large-scale vortical structures inside a shallow mixing

    layer are sustained by perturbation prescribed at the inflow.

    In natural and laboratory flows, the flow is perturbed by the

    presence of developed turbulence, as well as environmen-

    tal (natural) factors. In numerical simulations, however,

    artificial forcing is introduced at the inflow boundary to

    mimic the perturbations in the laboratory experiments. In

    the current study, white noise is prescribed at the inflow.

    Sensitivity of flow evolution to inflow perturbations will be

    explored in a later section.

    4.1 Mean inflow profile

    In all laboratory set-ups (see Refs. Chu and Babarutsi

    1988, Uijttewaal and Booij 2000), a splitter plate is located

    at the inflow to separate streams of different velocities.

    Due to the boundary layers attached to the splitter plate,

    wakes develop downstream of the splitter plate. The effect

    of splitter wake is reflected by a velocity deficit at the low

    speed side of the mean velocity profile at the inflow and it

    is shown to be obvious in laboratory measurements (see

    Figures 4 and 5 in Ref. Uijttewaal and Booij 2000). In the

    current study, the effect of splitter wake is modelled in the

    mean inflow profile by adding a secant hyperbolic term

    which is typical to wakes (Monkewitz 1988) to the

    hyperbolic tangent profile typical to mixing layers (van

    Prooijen and Uijttewaal 2002a) as follows,

    UðyÞ ¼ Uc þ DU0 tanh 2yd

    � �2 0:5 1 þ sinh2 y

    d

    � �h i� �:

    ð11ÞFigure 1 compares the measured inflow profile, the

    mean inflow profile with splitter wake effects given by

    Equation (11) (continuous curve in Figure 1) and the

    hyperbolic tangent profile (dashed curve in Figure 1). It

    shows that the measured profiles just after the edge of

    the splitter plate collapse nicely into Equation (11), while

    the conventional hyperbolic profile hardly matches the

    measured data.

    The development of mixing layers from an inflow

    profilewith andwithout the effect of splitterwakes is shown

    in Figure 2. It can be seen that the development of shallow

    mixing layers is generally the same for both mean inflow

    profiles. The mixing layer width for the mean profile with

    the effect of splitter wake is larger near the inflow

    (x , 2 m). This can be ascribed to the fact that the meaninflow profile with the effect of splitter wake actually

    broadens the mixing layer (refer to Figure 3). At the region

    where x . 2 m, the mean flow recovers to the traditional

    –0.4

    0

    0.4

    0.8

    1.2

    –2 –1.5 –1 –0.5 0 0.5 1 1.5 2

    tanh profiletanh with splitter wakedata at 0.05m for 42mm flowsdata at 0.05m for 67mm flows

    y/δ

    (u – u1) / (u2 – u1)

    Figure 1. Comparison of analytical and measure velocityprofile at the inlet of a mixing layer (d tanh denotes mixing layerwidth based on tangent hyperbolic profile; d wake denotesmixing layer width based on mean flow profile with wake effect).

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0 2 4 6 8 10 12 14 16

    x (m)

    With wakeWithout wake

    δ (m)

    Figure 2. Development of mixing layer width from inflowprofiles with and without the effects of splitter wakes (case 1).

    0

    0.2

    0.4

    0.6

    0.8

    0 10 15

    x (m)

    0.1u0.2u0.3u0.4u

    δ (m)

    5

    Figure 3. Development of shallow mixing layers under inflowforcing of different amplitudes (case 1).

    M. Ghidaoui and J.H. Liang526

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • tangent hyperbolic profile in a manner similar to the

    laboratory tests (see Figures 4 and 5 in Ref. Uijttewaal and

    Booij 2000). In addition, the computed mixing layer widths

    from two inflow profiles are comparable in this region.

    4.2 Inflow perturbations

    Co-flowing mixing layers are highly susceptible to inflow

    perturbations due to their convective instability nature.

    Laboratory experiments in deep mixing layers also showed

    that the development of mixing layers is highly sensitive to

    inflow perturbations, in particular, the forcing at the most

    unstable frequency (harmonic) and its subharmonics (for a

    review on experimental findings, see Ref. Ho and Huerre

    1984). In numerical studies, several attempts have been

    made to simulate a natural perturbation both in deepmixing

    layers (e.g. Stanley and Sarkar 1997) and in shallowmixing

    layers (e.g. van Prooijen and Uijttewaal 2002b). A note of

    caution, all this sophisticated forcing can never duplicate

    the real turbulent field since a lot of detailed information,

    like amplitudes and phases of different modes, is uncertain.

    The imposed forcing will undergo an adjustment stage to

    evolve into the real turbulence.

    In the current study, white noise is prescribed at the

    inflow. Stanley and Sarkar (1997) and Yang et al. (2004)

    in their studies of deep mixing layers found that a naturally

    developed mixing layer forms under this inflow forcing. To

    investigate the sensitivity of shallow mixing layers to the

    perturbation, perturbations of amplitudes ranging from

    10% of the inflow velocity difference (DU0) to 40% of theinflow velocity difference are imposed in both tests.

    Figures 3 and 4 show the development of the mixing

    layer width for both cases. As is also found by van Prooijen

    and Uijttewaal (2002b), the growths of the mixing layer

    width forced by disturbances of different amplitudes differ

    largely (see Figures 3 and 4). Near the inflow (x , 1.5 m),growths of mixing layer width are almost the same for

    inflow forcing of different amplitudes. Regardless of the

    amplitude of forcing imposed, no vortices are present in this

    region. Thus, mixing between two streams and sub-

    sequently growths of mixing layers is totally due to

    turbulent diffusion. From the position at around x ¼ 2 monward, mixing layers forced by perturbation of larger

    amplitudes grow faster. For mixing layers forced by 0.3DU0perturbations, vortices roll up between the position x ¼ 2 mand the position x ¼ 3 m, while vortices roll up between theposition x ¼ 3 m and the position x ¼ 4 m for mixing layersforced by 0.1DU0 perturbations. Furthermore, the inten-sities of coherent structures are weaker for the case forced

    by 0.1DU0. This illustrates the importance of vorticalstructures in exchange and mixing between two streams. In

    both tests, perturbation of amplitude 0.3DU0 is found toprovide the closest results to experimental data from

    Uijttewaal and Booij (2000). This forcing amplitude will be

    adopted for later tests.

    5. General flow features

    Tests under both hydraulic conditions shown in Table 1 are

    carried out. All results displayed in this section are forced by

    white noisewith 30%of the velocity difference at the inflow.

    Figures 5–8 display the vorticity field, as well as the passive

    scalar field and illustrate the spatial development of the KH

    instability. The roll up mechanism, which results in the

    formationofKHvortices, occurs in the regionbetween x ¼ 2and 3 m. This indicates that the instability begins at small

    values of x and growth of disturbances reaches the saturation

    state near x ¼ 2–3 m, where vortices roll up. The vorticesexperience growth until around x ¼ 7 m for case 1 andaround x ¼ 13 m for case 2. Vortex pairing is obviousthroughout the growth of the mixing layers. The general

    features of vortex roll-up and pairing are consistent with

    those in Plates 7 and 8 (Lesieur 1997). Comparing the

    contour of the passive scalar (Figures 6 and 8) and the

    contour of vorticity (Figures 5 and 7), it is clear that mixing

    and the subsequent growth of the shallow mixing layer is

    much more prominent in the region after the roll-up of

    vortices, indicating that the vortices are important for mixing

    between two streams.

    The effects of flow confinement and bed friction can be

    understood by invoking the bed friction parameter S (Chen

    and Jirka 1997, Chu and Babarutsi 1988, Ghidaoui and

    Kolyshkin 1999),

    Sh ¼ cfðU1 þ U2Þ=4hðU1 2 U2Þ=d ; ð12Þ

    where the subscript h emphasises the dependence of S on the

    water height. The local value of Sh is a measure of the local

    ratio of energy output from the KH vortices by the bed

    friction to the energy input into the KH vortices by the

    Reynolds stresses (Chu et al. 1991, Uijttewaal and Booij

    2000). The local value of S for the case h ¼ 42 mm (S42) is

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 10 15

    x (m)

    0.1u0.2u0.3u0.4u

    δ (m)

    5

    Figure 4. Development of shallow mixing layers under inflowforcing of different amplitudes (case 2).

    International Journal of Computational Fluid Dynamics 527

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • larger than that for the case h ¼ 67 mm (S67), for all x. Thefact that S42 . S67 explains why the vortical structures aremore evident and experience larger spatial growth for

    h ¼ 67 mm than for h ¼ 42 mm (refer to Figures 5 and 7).Although, in both cases, one finds that the KH vortices roll up

    and merge and that the confinement of the flow by the water

    depth suppresses the vortex stretching mechanism leading to

    the enstrophy cascade (Lesieur 1997), the extent of the

    enstrophy cascade is much smaller for the shallower case

    than for the deeper case. This is a result of the fact that

    S42 . S67 (i.e. the ratio of the energy supply to the largerscales by the pairing of KH vortices to the energy loss from

    the larger scales by the bottom friction is greater for

    h ¼ 67 mm than for h ¼ 42 mm). This conclusion isconsistent with the autocorrelation function and energy

    spectrum found in Uijttewaal and Booij (2000). They show

    that (i) the range of scales for which the slope of energy

    spectrum follows the -3 law is much more evident for the case

    with h ¼ 67 mm and (ii) the spatial growth of the vorticalstructures is much more apparent for the case h ¼ 42 mm.

    Large scale vortical structures in shallow flows are

    often called coherent structures (Jirka 2001, Jirka and

    Uijttewaal 2004). The coherence of large-scale vortical

    structures in the shallow mixing layers is recognised by

    x (m)

    Y (

    m)

    0 2 4 6 8 10 12 14

    –1

    0

    1

    Figure 5. Vorticity contour for mixing layer corresponding to the test with h ¼ 42 mm inflow forcing amplitude ¼ 0.3DU0.

    Figure 6. Scalar contour for mixing layer corresponding to the test with h ¼ 42 mm inflow forcing amplitude ¼ 0.3DU0. Available incolour online.

    x (m)

    Y (

    m)

    0 2 4 6 8 10 12 14

    –1

    0

    1

    Figure 7. Vorticity contour for mixing layer corresponding to the test with h ¼ 67 mm inflow forcing amplitude ¼ 0.3DU0.

    Figure 8. Scalar contour for mixing layer corresponding to the test with h ¼ 67 mm inflow forcing amplitude ¼ 0.3DU0. Available incolour online.

    M. Ghidaoui and J.H. Liang528

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • invoking a spatial temporal correlation coefficient as

    follows (Roshko 1976),

    Corrð~uðx0; t0Þ; ~uðx0 þ x 0; t0 þ tÞÞ

    ¼ Covð~uðx0; t0Þ; ~uðx0 þ x0; t0 þ tÞÞ

    SDð~uðx0; t0ÞÞSDð~uðx0 þ x 0; t0 þ tÞÞ ; ð13Þ

    where Corr stands for correlation, Cov denotes covariance,

    SD indicates standard deviation, x0 and x0 þ x 0 are twospatial sampling, and t can be related to x 0 by theconvective velocity of perturbations (Uconvective) as

    follows, t ¼ x 0/Uconvective.The correlation coefficient defined by Equation (9)

    measures how much change the flow structures experience

    when they travel from x0 to x0 þ x0 at the speed ofUconvective. When the correlation coefficient equals unity,

    the flow structures are unchanged between x0 to x0 þ x0. In amixing layer, growth and decay of vortices and disturbances

    contributes to the decrease of this correlation coefficient.

    Figures 9–11 compare correlation coefficients under

    different convective velocities. It is shown that the

    downstream velocity fluctuations are positively correlated

    when the mean velocity of two streams is chosen to be the

    convective velocity (see Figure 9). When other velocities

    are chosen as convective velocities, negative correlation

    coefficients are prominent (see Figures 10 and 11).

    A negative correlation coefficient between points (x0, t0)

    and (x0 þ x 0, t0 þ t) means that flow features at positionx0 þ x 0 and time t0 þ t are no longer at the same phase asthe flow features at position x0 and time t0.

    Figures 9 and 12 display the correlation coefficients of

    transverse velocities originating from different downstream

    positions in shallowmixing layers for both case 1 and case 2.

    –0.2

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 5 10 15

    x' (m)

    Corrx0 = 0.0m

    x0 = 0.05m

    x0 = 0.25m

    x0 = 0.5m

    x0 = 1.0m

    x0 = 2.0m

    x0 = 3.0m

    x0 = 5.0m

    x0 = 7.0m

    x0 = 9.0m

    x0 = 11.0m

    Figure 9. Correlation coefficients of transverse velocities in a shallow mixing layer with 42mm water depth, convective velocity is Uc.

    –0.8

    –0.6

    –0.4

    –0.2

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 5 10x' (m)

    Corrx0 = 0.0m

    x0 = 0.05m

    x0 = 0.25m

    x0 = 0.5m

    x0 = 1.0m

    x0 = 2.0m

    x0 = 3.0m

    x0 = 5.0m

    x0 = 7.0m

    x0 = 9.0m

    x0 = 11.0m

    15

    Figure 10. Correlation coefficients of transverse velocities in a shallow mixing layer with 42mm water depth, convective velocity is0.9Uc.

    International Journal of Computational Fluid Dynamics 529

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • Velocity fluctuations at a certain spatial position are all, to

    some extent, correlated to downstream velocity fluctuation.

    Correlation coefficients all decrease in the downstream

    direction. In shallow mixing layers, vortices undergo

    different processes including roll-up, growth due to

    entrainment, amalgamation and decay. These vortex

    evolutions all contribute to the decrease in correlation

    coefficients. Comparison of Figures 9 and 12 shows that the

    decrease in correlation for case 2 is more significant than for

    case 1 for x 0 , 11 m. This is ascribed to the fact that thevortices in case 2 continue to change within the range

    x 0 , 11 m by the process of pairing, while the vortices incase 1 exhibit little change in structure beyond x 0 . 7 m, dueto the suppression of pairing by the bottom friction.

    6. Mean flow quantities

    This section investigates how well the depth-averaged

    shallow water equations model mean flow quantities of

    shallow mixing layers. Simplified analytical solutions,

    semi-empirical expressions as well as experimental data

    (Uijttewaal and Booij 2000, van Prooijen and Uijttewaal

    2002a) are adopted for comparison.

    6.1 Free surface water profile

    To achieve steady state in free surface flows, the drag force

    by bottom friction must be balanced by either a pressure

    gradient or gravity force by tilting the channel bed or both.

    In the laboratory experiments, the channel bed is horizontal

    and the flow is driven by the a pressure gradient which is

    given as follows,

    dh

    dx¼ 2cf 1

    gh=U2c�

    2 1: ð14Þ

    Figure 13 shows that the computed results are in good

    agreement with both experimental measurements and

    Equation (14).

    –0.8

    –0.6

    –0.4

    –0.2

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 5 10 15x' (m)

    Corrx0 = 0.0m

    x0 = 0.05m

    x0 = 0.25m

    x0 = 0.5m

    x0 = 1.0m

    x0 = 2.0m

    x0 = 3.0m

    x0 = 5.0m

    x0 = 7.0m

    x0 = 9.0m

    x0 = 11.0m

    Figure 11. Correlation coefficients of transverse velocities in a shallowmixing layer with 42mmwater depth, convective velocity is 1.1Uc.

    –0.2

    0

    0.2

    0.4

    0.6

    0.8

    1

    0 5 10 15

    x' (m)

    Corr

    x0 = 0.0m

    x0 = 0.05m

    x0 = 0.25m

    x0 = 0.5m

    x0 = 1.0m

    x0 = 2.0m

    x0 = 3.0m

    x0 = 5.0m

    x0 = 7.0m

    x0 = 9.0m

    x0 = 11.0m

    Figure 12. Correlation coefficients of transverse velocities in a shallow mixing layer with 67mm water depth, convective velocity is Uc.

    M. Ghidaoui and J.H. Liang530

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • 6.2 Velocity difference between two ambient streamsof the mixing layer

    In deepmixing layers, the velocity differences of the ambient

    streams are the same in the flow direction (Pope 2002). This

    fact does not hold for shallow mixing layers. In a shallow

    mixing layer, friction is bigger for the faster stream and

    smaller for the slower stream. In addition, since transverse

    water depth gradient is negligible (van Prooijen and

    Uijttewaal 2002a), i.e. the streamwise pressure gradient

    (›p/›x) is the same for both the fast and the slow streams.Thus, the fast stream slows down and the slow stream speeds

    up in a shallow mixing layer. The velocity difference along

    the stream is derived by van Prooijen and Uijttewaal (2002a)

    based on a quasi-1D shallow water model. Laboratory

    observations (van Prooijen and Uijttewaal 2002a) suggest

    that the variation of mean velocity in the streamwise

    direction (Uc) is within 3% and thus, can be assumed to be

    constant. Velocity difference is,

    DUðxÞ ¼ U1 2 U2 ¼ DU0 exp 2 cfhx

    � �; ð15Þ

    where the subscript 0 denotes variable at the inflow.

    Figures 14 presents the variation of velocity difference

    across a shallow mixing layer for both 42 and 67mmwater

    depth, respectively. The computed results show good

    agreement with expression (15) and laboratory data.

    6.3 Velocity distribution in lateral direction

    In a series of laboratory tests, van Prooijen and Uijttewaal

    (2002a) noted that the transverse distributions of

    streamwise velocity away from the inflow collapse

    approximately into a curve of tangent hyperbolic profile, i.e.

    Uðx; yÞ ¼ Uc þ DUðxÞ2

    tanhy2 ycðxÞ

    12

    � dðxÞ

    !; ð16Þ

    where d is the width of a mixing layer and the subscript cindicates that the variable is evaluated at the centreline of a

    mixing layer.

    Figures 15 and 16 plot the computed and theoretical

    lateral profiles of the mean streamwise velocity at different

    streamwise locations for the mixing layer with water

    heights of 42 and 67mm, respectively. It is shown that the

    velocity profiles at different spatial locations away from

    the inflow boundary (x . 2.0 m) essentially collapse ontoa single curve (Equation (16)). Near the inflow boundary

    condition (x , 2 m), the splitter wake is obvious. More-over, computed results at the high speed side of the mixing

    layer with 42 mm water depth depart from the empirical

    relation (16). The combined effects of streamwise pressure

    gradient and bottom friction result in acceleration of the

    low-speed flows and deceleration of fast-speed flows. This,

    in turn, results in the widening of a shallow mixing layer at

    the fast speed side. All these characteristics are also

    observed in laboratory tests (see Figures 4 and 5 in Ref.

    Uijttewaal and Booij 2000).

    6.4 Development of shallow mixing layer widths

    The growth of a shallow mixing layer implies that the scale

    of the vortices inside the mixing layer is growing in the

    0.035

    0.045

    0.055

    0.065

    0.075

    0.085

    0 2 4 6 8 10 12 14 16

    x(m)

    h(m)Computed results (67mm)Measurements (67mm)Integral model (67mm)Computed results (42mm)Measurements (42mm)Integral model (42mm)

    Figure 13. Comparison of water depth.

    International Journal of Computational Fluid Dynamics 531

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • flow direction, due to both vortex pairing (Roshko 1976)

    and vorticity induction (Dimotakis 1986). Laboratory

    observations indicate that for deep mixing layers, the

    spatial growth rate of a mixing layer is approximated by

    (see Ref. Balaras et al. 2001),

    dd

    dx¼ aDUðxÞ

    Uc; ð17Þ

    where a is entrainment coefficient and has an empiricalvalue of 0.85 (van Prooijen and Uijttewaal 2002a).

    Integration over the streamwise direction (x) results in

    the expression for width of shallow mixing layers (van

    Prooijen and Uijttewaal 2002a),

    dðxÞ ¼ aDU0Uc

    h

    cf1 2 exp 2

    cf

    hx

    � �h iþ d0; ð18Þ

    where d0 is the mixing layer width at the inflow and isapproximately equal to the water depth.

    Figures 17 and 18 depict the computed mixing layer

    thickness, the empirical solutions given by Equation (18)

    and data measured by van Prooijen and Uijttewaal

    (2002a). The computed mixing layer width shown in

    Figure 17 is performed for Cs ¼ 0.1, 0.2 and 0.3. The valueof 0.2 provides the closest fit with the data. Indeed, this

    value is used throughout the paper.

    0.01

    0.1

    10 2 4 6 8 10 12 14 16

    x (m)

    Computed results (67mm)Measurements (67mm)Integral model (67mm)Computed results (42mm)Measurements (42mm)Integral model (42mm)

    u1 - u2 (m/s)

    Figure 14. Comparison of velocity difference.

    –0.2

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    –2 –1.5 –1 –0.5 0 0.5 1 1.5 2

    0.05m

    0.25m

    0.5m

    1.0m

    1.5m

    2.0m

    2.5m

    4.0m

    5.8m

    7.5m

    Theory

    y/δ

    (u-u1)/(u2-u1)

    Figure 15. Comparison of mean streamwise velocities (case 1).

    M. Ghidaoui and J.H. Liang532

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • The computed shallow mixing layer experiences slow

    growth just downstream of the inflow boundary, causing it

    to deviate from the laboratory data (see Figures 17 and 18).

    This can be attributed to the different flow conditions

    between the model and the experiment. In the simulation,

    the flow undergoes linear KH instability and mixing in this

    region is mostly due to turbulent diffusion, while in

    laboratory tests, vortical structures in splitter wakes play an

    important role in mixing between two streams. The slow

    growth just downstream of the inflow boundary is also

    reported by Stanley and Sarkar (1997), Li and Fu (2003),

    and Yang et al. (2004) for 2D deep mixing layer

    simulations forced by stochastic disturbances. Expression

    (18) seems able to model the mixing layer growth near the

    inflow, which is not surprising if one notices that expression

    (18) is valid when vortices in mixing layers play an

    important role in the growth of mixing layers and it seems

    that mixing due to vortices in splitter wakes can also be

    modelled by this expression. Downstream of the position at

    3 m, where vortices begin to roll up in the simulation, the

    model predicts a slightly larger growth of the mixing layers

    compared to the data and the semi-empirical relation (18).

    Considering that 2D direct numerical simulation can

    reproduce correctly the growth rate of mixing layers

    (Stanley and Sarkar 1997), this slightly larger prediction is

    ascribed to the use of the Smagorinsky parameterisation

    schemes for turbulent diffusion.

    6.5 Centre of shallow mixing layer

    An interesting phenomenon in mixing layer flows is that

    the centre of the mixing layer will be displaced in the

    lateral direction to the low velocity side. In deep mixing

    layers, Dimotakis (1986) proposed that the differences

    between the entrainment fluxes from high-speed stream

    –0.2

    0

    0.2

    0.4

    0.6

    0.8

    1

    1.2

    –2 –1 0 1 2

    0.05m

    0.25m

    0.5m

    1.0m

    2.0m

    5.8 m

    11m

    Theory

    y/δ

    (u-u1)/(u2-u1)

    Figure 16. Comparison of mean streamwise velocities (case 2).

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0 2 4 6 8 10 12

    Computed results (42mm Cs = 0.1)Computed results (42mm Cs = 0.2)Computed results (42mm Cs = 0.3)Measurements (42mm)Integral model (42mm)

    δ (m)

    x(m)

    Figure 17. Comparison of shallow mixing layer width (case 1).

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0 2 4 6 8 10 12

    x(m)

    Computed results (67mm)Measurements (67mm)Integral model (67mm)

    δ(m)

    Figure 18. Comparison of shallow mixing layer width (case 2).

    International Journal of Computational Fluid Dynamics 533

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • and those from the low-speed streams due to the increasing

    spacing between vortices contribute to this shift of

    centreline. In shallow mixing layers, van Prooijen and

    Uijttewaal (2002a) noted that the deceleration of the high

    velocity side and the acceleration of the low velocity side

    is the major factor for the shift of mixing layer centre.

    Assuming that the centreline of the mixing layer is a

    streamline, the lateral shift of the centreline can be

    estimated from the conservation of mass as follows (van

    Prooijen and Uijttewaal 2002a),

    ðycðxÞ2W=2

    HUðx; yÞdy ¼ HW2U2ðx ¼ 0Þ; ð19Þ

    where W is the width of the channel.

    Figures 19 and 20 show acceptable agreements between

    the computed and analytical positions of the centre of the

    mixing layer (yc), where the solution for yc is obtained from

    the mass conservation principle (Equation (19)).

    7. Turbulent intensities

    In this section, the ability of shallow water equations in

    modelling turbulence intensities of shallow mixing layers

    is examined. Figures 21–24 show the spatial development

    of streamwise turbulence intensities and spanwise

    turbulence intensities for both shallow mixing layers of

    42 and 67mm water depth, respectively. In all figures,

    turbulence intensities decay near the inflow (x , 1 m).Disturbances undergo a selective amplification process in

    this region, where only a small portion of modes are

    unstable and grow, while others are stable and decay

    rapidly. In the region from x ¼ 1 to 5.8 m, turbulenceintensities experience substantial growth, indicating that

    energy extracted from the horizontal velocity shear is

    larger than energy consumed by bottom friction and

    horizontal diffusion in this region. Further downstream,

    turbulence intensities decay implying the dominance of

    damping due to bottom friction and horizontal diffusion

    over instability due to horizontal velocity shear. These

    decays and growths of turbulence intensities are also

    observed in laboratory measurements (see Figures 10–13

    in Ref. Uijttewaal and Booij 2000).

    –0.5

    –0.4

    –0.3

    –0.2

    –0.1

    00 2 4 6 8 10 12 14 16

    x(m)

    Computed results (42mm)Measurement (42mm)Integral Model (42mm)

    yc(m)

    Figure 19. Comparison of mixing layer center (case 1).

    –0.5

    –0.4

    –0.3

    –0.2

    –0.1

    00 2 4 6 8 10 12 14 16

    x(m)

    Computed results (67mm)

    Measurements (67mm)Integral model (67mm)

    yc(m)

    Figure 20. Comparison of mixing layer centre (case 2).

    0

    0.001

    0.002

    0.003

    0.004

    0.005

    0.006

    0.007

    0.008

    0.009

    0.01

    –2 –1 0 1 2

    0.05m0.25m0.5m1.0m1.5m2.0m2.5m4.0m5.8m7.5m

    y/δ

    u' (m/s)

    Figure 21. Profile of streamwise turbulence intensities atvarious downstream positions of a shallow mixing layer (case 1).

    0

    0.002

    0.004

    0.006

    0.008

    0.01

    0.012

    0.014

    –2 –1 0 1 2

    0.05m0.25m0.5m1.0m1.5m2.0m2.5m4.0m5.8m7.5m

    y/δ

    v' (m/s)

    Figure 22. Profile of spanwise turbulence intensities at variousdownstream positions of a shallow mixing layer (case 1).

    M. Ghidaoui and J.H. Liang534

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • Table 2 compares the maximum measured turbulence

    intensities by Uijttewaal and Booij (2000) and the

    maximum computed ones in the current study. The

    computed streamwise turbulence intensities are compar-

    able to the measured one, while the computed spanwise

    turbulence intensities are significantly larger than the

    measured ones. The computed distribution of turbulence

    intensities also spread more across the mixing layers than

    the measured ones. These phenomena are also noted in the

    study of the deep mixing layer by Stanley and Sarkar

    (1997), where they compare their 2D numerical results

    with 3D direct numerical simulation results and laboratory

    data. Indeed, the 3D energy cascade processes are totally

    absent in 2D models and energy can only be dissipated by

    bottom friction after they are transferred to large scale.

    8. Conclusions

    Turbulent shallow mixing layers and their associated

    vortical structures are ubiquitous in rivers, estuaries and

    coasts. Examples of these flows can be found in compound

    and/our composite channels, at the confluence of two

    rivers, at harbour entrances, and at groyne fields. A 2D

    BGK-based finite volume model for shallow water is

    applied to shallow mixing layers for which experimental

    results are available. The BGK scheme is explicit, second

    order in time and space and conserves both mass and

    momentum. The BGK relaxation time is locally evaluated

    from the classical turbulence model of Smagorinsky. The

    KH instability and the mean flow quantities are well

    represented by the shallow water model. The following

    remarks in relation to the applicability of the shallow water

    model to shallow mixing layers can be made.

    (1) The model provides the spatial development of the KH

    instability. The instability develops in the region very

    close to the splitter plate. Downstream of this region,

    the disturbances roll up, resulting in the formation of

    KHvortices.This is then followedbya regionwhere the

    vortices undergo pairing and result in the formation of

    coherent vortical structures. While vortex pairing acts

    to increase the length scale of the vortical structures, the

    bed friction tends to suppress these structures. The

    degree of suppression becomes more significant as

    the structures become larger. The scale of the final size

    of these structures at large distances from the splitter

    plate is governed by the bed friction parameter.

    (2) Shallow mixing layers are highly sensitive to

    perturbations at the inflow boundary. These pertur-

    bations are selectively amplified in their linear region

    and cease to grow when they reach saturation. Further

    downstream where bottom friction is dominant, the

    perturbations are dissipated.

    (3) Coherence of large-scale vortical structures in shallow

    mixing layers decreases due to growth, amalgamation

    and decay of these vortices. It is also found that these

    large-scale vortices are advected at the mean speed

    between two streams in the shallow mixing layer, a

    result which is the same as that for deep mixing layers.

    (4) The Smagorinsky scheme for 2D shallowwater flows is

    a bit diffusive since it predicts a slightly larger growth

    rate than observed. However, other mean flow

    quantities, like the velocity difference, centreline and

    water surface, are well predicted by the current model.

    (5) Shallow water model is incapable of accurately

    predicting turbulent intensities and their distribution.

    A 3Dmodel would be required to accurately predict the

    distribution of turbulence intensities.

    Table 2. Comparison of measured and computed turbulenceintensities.

    u 0max (m/s) v 0max (m/s)

    Measured (42mm) 0.095 0.007Computed (42mm) 0.0085 0.12Measured (67mm) 0.028 0.018Computed (42mm) 0.24 0.03

    0

    0.005

    0.01

    0.015

    0.02

    0.025

    0.03

    –2 –1 0 1 2

    0.05m0.25 m0.5m1.0m2.0m5.8m11.0m

    y/δ

    u' (m/s)

    Figure 23. Profile of streamwise turbulence intensities atvarious downstream positions of a shallow mixing layer (case 2).

    0

    0.005

    0.01

    0.015

    0.02

    0.025

    0.03

    0.035

    0.04

    –2 –1 0 1 2

    0.05m0.25m0.5m1.0m2.0m5.8 m11.0m

    y/δ

    v' (m/s)

    Figure 24. Profile of spanwise turbulence intensities at variousdownstream positions of a shallow mixing layer (case 2).

    International Journal of Computational Fluid Dynamics 535

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • (6) The close agreement between the model and exper-

    iments suggests that the processes of flow instability,

    and its associated transversal mixing of mass

    momentum and flow entrainment are well represented

    by the shallow water model.

    Although, the computed and measured turbulent

    intensities do not agree, flow instabilities and mean flow

    quantities are well represented by the shallow water model.

    It is plausible to conclude that the shallowwater models can

    be used to conduct stability of shallow flows and to

    determine mean flow features of shallow flows. Shallow

    water model is not be suitable for investigating small scale

    features such as flow intermittency, instantaneous structure

    of the flow and possible breakdown of vortical motion into

    3D turbulence. Such breakdown is expected to occur in

    shallow flows, where the water depth is of the order of the

    bottomboundary layer. Such very shallowflowcan occur in

    laboratory studies. Indeed, experiments show that the

    vortical motion eventually disintegrates for the case with

    water depth equal to 42mm.

    Acknowledgements

    The authors wish to thank the Research Grants Council of HongKong for financial support under Project HKUST6227/04E andProject 613306.

    References

    Balaras, E., Piomelli, U., and Wallace, J., 2001. Self-similarstates in turbulent mixing layers. Journal of FluidMechanics, 446, 1–24.

    Batchelor, G.K., 1967. An introduction to fluid dynamics.Cambridge: Cambridge University Press.

    Chen, D. and Jirka, G.H., 1995. Experimental study of planeturbulent wakes in shallow water layers. Fluid DynamicsResearch, 16, 11–41.

    Chen, D. and Jirka, G.H., 1997. Absolute and convectiveinstabilities of plane turbulent wakes in a shallow waterlayer. Journal of Fluid Mechanics, 338, 157–172.

    Chu, V.H., Wu, J.H. and Khayat, R.E., 1991. Stability oftransverse shear flows in shallow open channels. Journal ofHydraulic Engineering, ASCE, 117, 1370–1388.

    Chu, V.H. and Babarutsi, S., 1998. Confinement and bed-frictioneffects in shallow turbulent mixing layers. Journal ofHydraulic Engineering, ASCE, 114, 1257–1274.

    Dimotakis, P.E., 1986. Two-dimensional shear-layer entrain-ment. AIAA Journal, 24, 1791–1796.

    Durran, D.R., 1999. Numerical methods for wave equations ingeophysical fluid dynamics. New York: Springer-Verlag.

    Ghidaoui, M.S. and Kolyshkin, A.A., 1999. Linear stabilityanalysis of lateral motions in compound open channel.Journal of Hydraulic Engineering, 125, 871–880.

    Ghidaoui, M.S., Deng, J.Q., Gray, W.G. and Xu, K., 2001.A Boltzmann based model for open channel flows.International Journal for Numerical Methods in Fluids, 35(4), 449–494.

    Ghidaoui, M.S., et al., 2006. Linear and nonlinear analysis ofshallow wakes. Journal of Fluid Mechanics, 548, 309–340.

    Ho, C.M. and Huerre, P., 1984. Perturbed free shear layers.Annual Review of Fluid Mechanics, 16, 365–424.

    Huerre, P. and Monkewitz, P.A., 1990. Local and globalinstabilities in spatially developing flows. Annual Review ofFluid Mechanics, 22, 473–537.

    Jirka, G.H., 2001. Large scale flow structures and mixingprocesses in shallow flows. Journal of Hydraulic Research,39, 567–573.

    Jirka, G.H. and Uijttewaal, W.S.J., 2004. Shallow flows: adefinition. In: G.H. Jirka and W.S.J. Uijttewaal, eds. Shallowflows. Leiden: A.A. Balkema Publishers.

    Kolyshkin, A.A. and Ghidaoui, M.S., 2002. Linear stabilityanalysis of lateral motions in compound open channel.Journal of Hydraulic Engineering, 128, 1076–1086.

    Lesieur, M., 1997. Turbulence in fluids. Boston: KluwerAcademic Publishers.

    Li, Q. and Fu, S., 2003. Numerical simulation of high-speedplanar mixing layer. Computers & Fluids, 32, 1357–1377.

    Liang, J.H., Ghidaoui, M.S., Deng, J.Q. and Gray, W.G., 2007.A Boltzmann-based finite volume algorithm for surfacewater flows on cells of arbitrary shapes. Journal of HydraulicResearch, 45 (2), 147–164.

    Mellor, G.L., 2004. User guide for a 3D, primitive equation,numerical ocean model. http://www.aos.princeton.edu/WWWPUBLIC/htdocs.pom/

    Monkewitz, P.A., 1988. The absolute and convective nature ofinstability in two-dimensional wakes at Low Reynoldsnumbers. Physics of Fluids, 31, 999–1006.

    Oey, L.-Y., Ezer, T., and Lee, H.C., 2005. Loop current, rings andrelated circulation in the Gulf of Mexico: a review ofnumerical models and future challenges. GeophysicalResearch Letters, 32, L1266, doi: 10.1029/2005GL023253.

    Pope, S.A., 2002. Turbulent flows. Cambridge: CambridgeUniversity Press.

    Roshko, A., 1976. Structure of turbulent shear flows: a new look.AIAA Journal, 14, 1349–1357.

    Schlichting, H. and Gersten, K., 2000. Boundary layer theory.Berlin, Heidelberg: Springer-Verlag.

    Socolofsky, S.A. and Jirka, G.H., 2004. Large-scale flowstructures and stability in shallow flows. Journal ofEnvironmental Engineering and Science, 3, 451–462.

    Stanley, S. and Sarkar, S., 1997. Simulations of spatiallydeveloping 2D shear layers and jets. Theoretical andComputational Fluid Dynamics, 9, 121–147.

    Uijttewaal, W.S.J. and Booij, R., 2000. Effects of shallowness onthe development of free-surface mixing layers. Physics ofFluids, 12 (2), 4392–4402.

    Uijttewaal,W.S.J. and Jirka,G.H., 2003.Grid turbulence in shallowflows. Journal of Fluid Mechanics, 489, 325–344.

    van Prooijen, B.C. and Uijttewaal, W.S.J., 2002a. A linearapproach for the evolution of coherent structures in shallowmixing layers. Physics of Fluids, 14 (12), 4105–4114.

    van Prooijen, B.C. and Uijttewaal, W.S.J., 2002b. On theinitiation of large scale turbulence structures in the numericalsimulation of shallow mixing layers. In: Bousamr and Zech,eds. River flow 2002. Lisse: Swets & Zeitlinger.

    Xu, X., Lee, J.S., and Pletcher, R.H., 2005. A compressible finitevolume formulation for large eddy simulation of turbulentpipe flows at low Mach number in Cartesian coordinates.Journal of Computational Physics, 203 (1), 22–48.

    Yang, W.B., Zhang, H.Q., Chan, C.K., and Lin, W.Y., 2004.Large eddy simulation of mixing layer. Journal ofComputational and Applied Mathematics, 163, 311–318.

    Zhou, J.G., 2004. Lattice Boltzmann methods for shallow waterflows. Berlin: Springer-Verlag.

    M. Ghidaoui and J.H. Liang536

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13

  • Appendix: Matrices in Equation (5)

    Asi ¼ hsiffiffiffiffiffiffiffiffighsi

    p2

    Vnerfcð2VnÞ þ 1ffiffiffipp e2V 2nVnvþ n

    ffiffiffiffigh

    p2

    � �erfcð2VnÞ þ v 1ffiffiffipp e2V2n

    2664

    3775si

    Aso ¼ hsoffiffiffiffiffiffiffiffiffighso

    p2

    VnerfcðVnÞ2 1ffiffiffipp e2V2nVnvþ n

    ffiffiffiffigh

    p2

    � �erfcðVnÞ2 v 1ffiffiffipp e2V2n

    2664

    3775so

    Bsi ¼ghesi

    2

    V2n þ 12�

    erfcð2VnÞ þ 1ffiffiffipp Vne2V2nV2nvþ v2 þ vnn�

    erfcð2VnÞ þ 1ffiffiffipp Vnvþ ffiffiffiffiffighp n� e2V2n264

    375si

    Bso ¼gh2so

    2

    V2n þ 12�

    erfcðVnÞ2 1ffiffiffipp Vne2V2nV2nvþ v2 þ vnn�

    erfcðVnÞ2 1ffiffiffipp Vnvþ ffiffiffiffiffighp n� e2V2n264

    375so

    Dsi ¼ hsiffiffiffiffiffiffiffiffighsi

    p2

    vt Vnerfcð2VnÞ þ 1ffiffiffipp e2V2nn ovtffiffiffiffiffigh

    pV2n þ 12�

    erfcð2VnÞ þ 3ffiffiffipp Vne2V2nn ov2t þ gh2�

    Vnerfcð2VnÞ þ 1ffiffiffipp e2V2nn o

    26666664

    37777775si

    Dso ¼ hsoffiffiffiffiffiffiffiffiffighso

    p2

    vt VnerfcðVnÞ2 1ffiffiffipp e2V2nn ovtffiffiffiffiffigh

    pV2n þ 12�

    erfcðVnÞ2 3ffiffiffipp Vne2V2nn ov2t þ gh2� �

    VnerfcðVnÞ2 1ffiffiffipp e2V2nn o

    26666664

    37777775so

    Es ¼ hsvn

    vnvþ gh2 n

    24

    35s

    Gs ¼ hsv2n þ gh2v2nvþ 3gh2 vnn

    24

    35s

    Is ¼ hsvnvt

    vnvtvþ gh2 ðvtnþ vntÞ

    24

    35s

    International Journal of Computational Fluid Dynamics 537

    Dow

    nloa

    ded

    by [

    Uni

    vers

    ity o

    f C

    alif

    orni

    a, L

    os A

    ngel

    es (

    UC

    LA

    )] a

    t 00:

    55 0

    5 Ja

    nuar

    y 20

    13