Reynolds Stress ARSM

download Reynolds Stress ARSM

of 114

Transcript of Reynolds Stress ARSM

  • 8/10/2019 Reynolds Stress ARSM

    1/114

    Helsinki University of Technology. Laboratory of Aerodynamics.

    Series B

    Teknillinen korkeakoulu. Aerodynamiikan laboratorio.

    Sarja B

    Espoo 2001, FINLAND

    IMPLEMENTING AN EXPLICIT ALGEBRAIC REYNOLDS STRESS

    MODEL INTO THE THREE-DIMENSIONAL FINFLO FLOW SOLVER

    Report B-52

    Ville Hmlinen

    TEKNILLINEN KORKEAKOULU

    TEKNISKA HGSKOLAN HELSINKI UNIVERSITY OF TECHNOLOGY

  • 8/10/2019 Reynolds Stress ARSM

    2/114

  • 8/10/2019 Reynolds Stress ARSM

    3/114

    Helsinki University of Technology. Laboratory of Aerodynamics.

    Series B

    Teknillinen korkeakoulu. Aerodynamiikan laboratorio.

    Sarja B

    Espoo 2001, FINLAND

    IMPLEMENTING AN EXPLICIT ALGEBRAIC REYNOLDS STRESS

    MODEL INTO THE THREE-DIMENSIONAL FINFLO FLOW SOLVER

    Report B-52

    Ville Hmlinen

    Approved by Seppo Laine

    Helsinki University of Technology

    Department of Mechanical Engineering

    Laboratory of Aerodynamics

    Teknillinen korkeakoulu

    Konetekniikan osastoAerodynamiikan laboratorio

  • 8/10/2019 Reynolds Stress ARSM

    4/114

    Distribution:Helsinki University of TechnologyLaboratory of AerodynamicsP.O.Box 4400FIN-02015 HUTTel. +358-9-451 3421Fax +358-9-451 3418

    Ville Hmlinen

    ISBN 951-22-5605-3ISBN 951-22-6824-8 (PDF)ISSN 1456-6990

    Printed in OtamediaEspoo 2001, FINLAND

  • 8/10/2019 Reynolds Stress ARSM

    5/114

    iii

    Abstract

    This thesis describes the implementing of an explicit algebraic Reynolds

    stress model (EARSM) into the three-dimensional FINFLO flow solver. The

    EARSM replaces the Boussinesq eddy-viscosity assumption by a more

    general constitutive relation for the second-order correlation in the Rey-

    nolds averaged Navier-Stokes equations.

    The thesis begins by describing the Navier-Stokes equations, in both their

    time-accurate and Reynolds averaged form. Some general aspects of tur-

    bulence are also discussed and a differential Reynolds stress model is pre-

    sented. Then an algebraic Reynolds stress model (ARSM) is formulated

    and simplified to its explicit form. After that some programming aspects of

    the model are discussed. Finally, four test cases are selected, calculated

    and analysed to verify the correct implementation of the model.

    As the first test case a two-dimensional boundary layer over a flat plate is

    calculated to verify the connection between the EARSM and the rest of the

    FINFLO. The second order effects of the model are studied with the secondtest case, a fully developed flow inside a rectangular duct. Thirdly, a fully

    developed flow inside a rotating pipe is calculated to examine the third or-

    der effects of the model. As the last test case an axial flow over a cylinder,

    part of which rotates, is calculated to further study the behaviour of the

    EARSM when a sudden strain component is applied to the flow field. From

    the results of the test cases it is then concluded that the EARSM has been

    correctly implemented. The model gives significant improvements in most

    cases compared to standard eddy-viscosity models without a large in-

    crease in the computational effort.

  • 8/10/2019 Reynolds Stress ARSM

    6/114

  • 8/10/2019 Reynolds Stress ARSM

    7/114

    v

    Contents

    Abstract .................................................................................................... iii

    Contents .................................................................................................... v

    Nomenclature.......................................................................................... vii

    1 Introduction......................................................................................1

    2 Governing equations and eddy-viscosity turbulence models ......3

    2.1 Navier-Stokes equations ........................................................3

    2.2 Reynolds averaging................................................................5

    2.3 General facts of turbulence.....................................................62.4 Eddy-viscosity turbulence models...........................................8

    2.4.1 Algebraic models.........................................................9

    2.4.2 One-equation models................................................10

    2.4.3 Two-equation models................................................11

    3 Differential Reynolds stress models ............................................15

    3.1 Reasons for using the Reynolds stress models ....................15

    3.2 Exact equation for the Reynolds stress tensor......................16

    3.3 Modelling the Reynolds stress tensor ...................................18

    4 Formulation of an explicit algebraic Reynolds stress model .....21

    4.1 Algebraic Reynolds stress model..........................................21

    4.2 Explicit algebraic Reynolds stress model..............................26

    4.3 Simplified EARSM for two-dimensional mean flow................29

    4.4 Simplified EARSM for three-dimensional mean flow.............31

    4.5 Solution of the general quasi-linear ARSM equation.............33

    5 Programming approach and written subroutines........................37

    5.1 Subroutine EARSM1 ............................................................37

    5.2 Subroutine TTIMES..............................................................41

    5.3 Subroutine ANISFLUX..........................................................42

    5.4 Subroutine FLUXP................................................................43

    5.4 Subroutine PEEKOO............................................................43

    5.5 Subroutine VELGRAD ..........................................................44

    5.6 Boundary routines ................................................................44

  • 8/10/2019 Reynolds Stress ARSM

    8/114

    Contents

    vi

    6 Test cases ......................................................................................47

    6.1 Flow over a flat plate.............................................................48

    6.1.1 Grid ...........................................................................48

    6.1.2 Boundary conditions..................................................50

    6.1.3 Results ......................................................................51

    6.2 Flow inside a rectangular duct ..............................................55

    6.2.1 Grid ...........................................................................55

    6.2.2 Boundary conditions..................................................56

    6.2.3 Results ......................................................................57

    6.3 Flow inside a rotating pipe ....................................................63

    6.3.1 Grid ...........................................................................64

    6.3.2 Boundary conditions..................................................656.3.3 Results ......................................................................65

    6.4 Flow over a rotating cylinder .................................................69

    6.4.1 Grid ...........................................................................70

    6.4.2 Boundary conditions..................................................71

    6.4.3 Results ......................................................................71

    7 Summary and conclusions............................................................79

    Bibliography ............................................................................................81

    Appendix A, Subroutine EARSM1..........................................................87

    Appendix B, Subroutine TTIMES ...........................................................91

    Appendix C, Subroutine ANISFLUX.......................................................93

    Appendix D, Subroutine FLUXP .............................................................95

    Appendix E, Subroutine PEEKOO .........................................................97

    Appendix F, Subroutine VELGRAD........................................................99

  • 8/10/2019 Reynolds Stress ARSM

    9/114

    vii

    Nomenclature

    Roman alphabet

    A Model coefficient

    ija Reynolds stress anisotropy tensor

    C Model coefficient

    fC Skin friction coefficient( )aijC Coriolis tensoreffC Effective eddy-viscosity coefficient

    E Total energye Specific internal energy

    if Body force vector

    H Matrix defined by the equation (4.44)

    J Matrix defined by the equation (4.45)

    k Turbulent kinetic energy, Heat transfer coefficient

    l Turbulent length scale

    N Symbol defined bytheequation(4.25), Non-dimensional rotation rate

    in Cell face unit normal vector

    P Production of turbulence

    ijP Production of turbulence tensor

    p Pressure

    iq Heat flux vector

    Re Reynolds number

    S Cell face area

    ijS Mean strain tensor

    T Temperature

    t Time

    iu Velocity vector

    u Friction velocity

    u Fluctuating component of the u

    V Cell volume

    ix Cartesian coordinate vector

    +y Dimensionless wall distance

  • 8/10/2019 Reynolds Stress ARSM

    10/114

    Nomenclature

    viii

    Greek alphabet

    Coefficient for the terms of the ija

    ij Kroneckers delta

    Dissipation rate

    ij Dissipation rate tensor

    ijk Permutation tensor

    ij Redistribution tensor

    Molecular viscosity

    Density

    Turbulent time-scale

    w Wall shear stress

    ij Viscous stress tensorT Kinematic eddy-viscosity

    ij Mean vorticity tensor

    *ij Absolute mean vorticity tensor

    Rij Effective mean vorticity tensor

    Sij System rotation tensor

    kspecific dissipation rateSk Constant system rotation rate vector

    Invariants

    SII Second invariant of the ijS

    II Second invariant of the ij

    III Third invariant of the ijS

    IV Third invariant defined as kijkijS

    V Fourth invariant defined as likljkijSS

  • 8/10/2019 Reynolds Stress ARSM

    11/114

    1

    Chapter 1

    Introduction

    In earlier days fluid dynamics, like other physical sciences, was divided into

    theoretical and experimental branches. The equipment and vehicles in-

    volving fluid flow were designed and analysed by these two methods. With

    the evolution of the digital computer, a third method called ComputationalFluid Dynamics (CFD) has become available. In this computational ap-

    proach the equations that govern a process of interest are solved numeri-

    cally at certain discrete points of space.

    The evolution of numerical methods for solving ordinary and partial differ-

    ential equations began approximately at the beginning of the twentieth

    century. The automatic digital computer was invented in the early 1940s

    and was used from nearly the beginning to solve problems in fluid dynam-

    ics. The explosion in computational activity did not, however, begin until the

    high-speed digital computers began to be generally available in the 1960s

    (Ref. [1]).

    The three key elements of CFD are algorithm development, grid generation

    and turbulence modelling. In this study only the third element is scrutinised.

    Turbulence is inherently three-dimensional and time dependent, and an

    enormous amount of information is thus required to completely describe a

  • 8/10/2019 Reynolds Stress ARSM

    12/114

    Introduction

    2

    turbulent flow. This is beyond the capability of the existing computers for

    virtually all practical flows. Thus, some kind of approximate and statistical

    method, called a turbulence model, is needed.

    Helsinki University of Technology has quite a long history in the field of

    CFD. The development of the parallel multi-block Navier-Stokes flow solver

    called FINFLO was initiated in 1987. The original two-man development

    group has grown along with new applications. Nowadays, in 1995 estab-

    lished CFD group consists of about 15 researchers from the Laboratory of

    Aerodynamics, the Laboratory of Applied Thermodynamics and the Labo-

    ratory of Ship Hydrodynamics, all utilising the FINFLO code. The FINFLO is

    nowadays suitable for compressible, time dependent, laminar and turbulent

    flows and has been applied to dozens of demanding research and devel-

    opment projects. The code is able to handle structured multi-block grids

    and the equations are solved by an implicit pseudo-time integration scheme

    using Roes flux splitting.

    The CFD-group has always tried to keep up with the most recent turbu-

    lence modelling. Prior to this work the FINFLO included the following tur-

    bulence models; Baldwin-Lomax [2], Cebeci-Smith [3], Chiens low Rey-

    nolds number

    k model [4] and different variants of Menters

    k

    model (BSL), (SST) [5] [6] and (RCSST) [7]. The aim of this thesis has

    been to implement into the FINFLO a new turbulence model called the Ex-

    plicit Algebraic Reynolds Stress Model (EARSM) developed by Stefan Wal-

    lin and Arne V. Johansson from Sweden. The formulation of the model is

    presented briefly in this thesis, but the reader is encouraged to find the

    complete description in reference [8]. The EARSM was programmed using

    Fortran 77 and 90 as they are used throughout the main code. After the

    programming was completed, four test cases were selected, calculated and

    analysed to verify the correct implementation of the model.

  • 8/10/2019 Reynolds Stress ARSM

    13/114

    3

    Chapter 2

    Governing equations and eddy-viscosity turbu-lence models

    In this chapter the equations governing the flow field are presented in their

    time-accurate and Reynolds averaged form. Then some general remarks of

    turbulence are made and importance for the turbulence modelling is em-phasised. At the end of the chapter the basic aspects of the so-called eddy-

    viscosity turbulence models are discussed.

    2.1 Navier-Stokes equations

    The equations of viscous flow have been known for more than 100 years.

    The exact number of basic equations depends upon personal preference,

    but some relations are, however, more basic than others. Usually the basicsystem of equations is considered to be the three laws of conservation for

    physical systems:

    - Conservation of mass (i.e. the continuity equation)

    - Conservation of momentum (i.e. the Newtons second law)

    - Conservation of energy (i.e. the first law of thermodynamics)

  • 8/10/2019 Reynolds Stress ARSM

    14/114

    Governing equations and eddy-viscosity turbulence models

    4

    The continuity equation simply states that the mass must be conserved. In

    the Cartesian coordinates ix this equation can be written as

    ( ) 0=

    +

    i

    i

    xu

    t (2.1)

    where is the density of the fluid, ttime and iu the velocity vector. Ten-

    sor notation will be used throughout the text whenever practical. The sec-

    ond equation, conservation of momentum, states that momentum must be

    conserved. It can be written in the Cartesian coordinates as

    ( )

    j

    ij

    ii

    i

    jii

    xx

    p

    fx

    uu

    t

    u

    +

    =

    +

    (2.2)

    where if is a body force, p the pressure. The viscous stress tensor ij is

    +

    =k

    kij

    i

    j

    j

    iij

    x

    u

    x

    u

    x

    u

    3

    2(2.3)

    where is the molecular viscosity and ij the Kroneckers delta. The third

    equation, conservation of energy, states that the energy must be con-served. It can be written as

    ( )[ ] ( ) 0=

    +

    +

    jiji

    j

    j

    j

    qux

    pEuxt

    E (2.4)

    whereEis defined as the total energy and can be written as

    += iiuueE2

    1 (2.5)

    where e is the specific internal energy. Assuming Fouriers heat transfer

    law the heat flux vector can be written as

    i

    ix

    Tkq

    = (2.6)

    where kin the heat transfer coefficient and T the temperature.

  • 8/10/2019 Reynolds Stress ARSM

    15/114

    Governing equations and eddy-viscosity turbulence models

    5

    2.2 Reynolds averaging

    In a turbulent flow the local pressure, density, velocity components and

    temperature vary randomly with time. A reasonable approach is to separate

    the flow quantities into stationary and random parts. The quantities are thus

    usually presented as a sum of the mean flow value and the fluctuating part

    as ii uu + . This formulation is inserted into the equations (2.1), (2.2) and(2.4); a process called Reynolds averaging is performed. There are three

    forms of averaging present in the turbulence-model research, which are a

    time average, a spatial average and an ensemble average. The general

    term used to describe these averaging processes is mean. The most

    commonly used averaging is the first, time averaging. It can be used onlyfor stationary turbulence, i.e. for a turbulent flow that does not vary with

    time on the average. For such a flow the mean flow value is defined as

    ( )+

    =

    Tt

    t

    iT

    i dttuT

    u 1lim (2.7)

    In practice taking the limit to infinity means that the integration time T

    needs to be long enough relative to the maximum period of the assumed

    velocity fluctuations.

    The Reynolds averaged Navier-Stokes equations (RANS) are obtained

    from the continuity and momentum equations, (2.1) and (2.2), by taking the

    time average of all the terms in the equations. The continuity equation does

    not change since it is linear in terms of the velocity. However, the momen-

    tum equation is non-linear, which means that all the fluctuating components

    do not vanish. An extra term, called Reynolds stress jiuu , appears in the

    momentum equation (2.2). The result can be written as

    ( ) ( ) ( )j

    jiij

    i

    i

    i

    jii

    x

    uu

    x

    pf

    x

    uu

    t

    u

    +

    =

    +

    (2.8)

    where the overbar has been dropped from the mean values. This conven-

    tion will be used throughout the text whenever obvious. The equation (2.8)

    presents the fundamental problem of turbulence. In order to compute all the

  • 8/10/2019 Reynolds Stress ARSM

    16/114

    Governing equations and eddy-viscosity turbulence models

    6

    mean-flow properties of the turbulent flow we need a reasonably accurate

    way to compute the Reynolds stress jiuu . This is the fundamental reasonfor the need of the turbulence models.

    The complete set of the Reynolds averaged Navier-Stokes equations are

    not presented in this text, because they can be found from many refer-

    ences, for example [9]. Scalar transport equations are also needed, for ex-

    ample to describe the transport of the concentration of species or the mass

    fraction of species. Their exact formulation can be found from reference [9]

    or [10]. The presentation of the exact formulation of the FINFLO along with

    the solution methods used is beyond the scope of this text. The reader is

    encouraged to consult references [11], [12] and [13].

    2.3 General facts of turbulence

    Many basic definitions of turbulence have been written, but probably the

    most accurate was formed by Bradshaw in 1974:

    "Turbulent fluid motion is irregular condition of flow in which the various

    quantities show a random variation with time and space co-ordinates, so

    that statistically distinct average values can be discerned. Turbulence has a

    wide range of scales."(Ref. [14])

    Analysis of the solutions to the Navier-Stokes equations shows that turbu-

    lence develops as an instability of the laminar flow. At low Reynolds num-

    bers, fluid layers slide smoothly past each other and the molecular viscosity

    dampens the randomly occurring small-scale, high frequency disturbances.

    The flow is laminar and exactly predictable. When Reynolds number in-

    creases, the restraining effects of viscosity are too weak to prevent such

    disturbances from amplifying. The disturbances grow gradually and be-

    come non-linear. They also interact with neighbouring disturbances. At high

    Reynolds number, the flow reaches a chaotic and non-repeating, i.e. tur-

    bulent, state. Turbulent motions are characterised by the following proper-

    ties: three dimensional, unsteady, random, strongly vorticial, strongly dissi-

    pative and strongly diffusive. Turbulence increases friction, heat transfer

  • 8/10/2019 Reynolds Stress ARSM

    17/114

    Governing equations and eddy-viscosity turbulence models

    7

    and mixing and spreading rate. It also reduces the separation tendency of

    the flow by energising the boundary layer (Ref.[15]).

    The non-linearity of the Navier-Stokes equations leads to interactions be-tween turbulent fluctuations of different wavelengths and directions. Ac-

    cording to reference [14] the wavelengths of motion usually extend all the

    way from a maximum comparable to the width of the flow to a minimum

    fixed by viscous dissipation of energy. The main physical process that

    spreads the motion over a wide range of wavelengths is called vortex

    stretching. The turbulence gains energy if the vortex elements are primarily

    oriented in a direction in which the mean velocity gradients can stretch

    them. This is called production of turbulence. The turbulent kinetic energy is

    then convected, diffused and dissipated.

    The larger-scale turbulent motion carries most of the energy and is mainly

    responsible for the enhanced diffusivity and attending stresses. The large

    eddies have memory and their orientation is sensitive to the mean flow.

    The large eddies randomly stretch the smaller eddies, cascading energy to

    them. Small eddies lose their orientation preference and become statisti-

    cally isotropic. Energy is dissipated by viscosity in the shortest wave-

    lengths, although the long-wavelength motion at the start of the cascadesets the rate of dissipation of energy. The shortest wavelengths simply ad-

    just accordingly. This kind of turbulence is called equilibrium turbulence.

    The ratio of largest to smallest scales increases rapidly as the Reynolds

    number increases (Ref. [14]).

    If the unsteady Navier-Stokes equations were calculated, a vast range of

    time and length scales would have to be computed. In other words, a very

    fine grid and very small time steps would be required. This is possible with

    todays computers for only very simple problems with small Reynolds num-

    bers. Thus a turbulence model is needed to account at least some of the

    fluctuating motion by a statistical approach. The model should be, in princi-

    ple, simple, broadly general and applicable, based on physics rather than

    intuition, computationally stable and coordinately invariable. This is, of

    course, an impossible list of requirements. The next sections of this chapter

    describe traditional ways to model the turbulence using a so-called Boussi-

    nesq eddy-viscosity approximation.

  • 8/10/2019 Reynolds Stress ARSM

    18/114

    Governing equations and eddy-viscosity turbulence models

    8

    2.4 Eddy-viscosity turbulence models

    In this section the so-called eddy-viscosity turbulence models are briefly

    discussed. The models can be divided in three categories; zero-, one- and

    two-equation models. These models use the Boussinesq eddy-viscosity

    approximation. The word approximation is a little misleading here because

    it is really more of a hypothesis than an approximation. It can be written for

    one- and two-equation models as

    kx

    u

    x

    u

    x

    uuu ij

    k

    kij

    j

    i

    i

    j

    Tji 3

    2

    3

    2

    +

    = (2.9)

    For the zero-equation models, which do not make use of the turbulent ki-

    netic energy 2iiuuk , the last term of the equation is zero. The Boussi-nesq approximation is used to compute the Reynolds stress tensor as the

    product of the kinematic eddy-viscosity T and the mean strain-rate tensor.

    The kinematic eddy-viscosity will be referred hereafter as the eddy-viscosity

    for simplicity. The Reynolds stress components are calculated from the

    eddy-viscosity for an incompressible flow as

    kSuu ijijTji 3

    22 = (2.10)

    where the mean strain-rate tensor is defined as

    +

    =i

    j

    j

    iij

    x

    u

    x

    uS

    2

    1(2.11)

    The eddy-viscosity is a scalar field variable and the different Reynolds

    stress components of a certain point of the flow field thus scale with the

    mean strain-rate tensor, i.e. the Reynolds stress components are linearly

    proportional to the mean strain-rate tensor. Different ways to calculate the

    eddy-viscosity are described in the next subsections.

    The equations (2.10) and (2.11) as well as all the following equations in this

    text are presented for an incompressible flow. The FINFLO calculation rou-

    tine calculates the flow as compressible, of course, while all of the turbu-

  • 8/10/2019 Reynolds Stress ARSM

    19/114

    Governing equations and eddy-viscosity turbulence models

    9

    lence models approximate the flow as incompressible. This approximation

    is a so-called Morkovins hypothesis meaning that the compressibility af-

    fects the turbulence quantities only at hypersonic speeds.

    2.4.1 Algebraic models

    Models where the eddy-viscosity is completely determined in terms of the

    local mean flow variables are referred to as zero-equation or algebraic

    models. Algebraic models are the simplest of all turbulence models. They

    often use some form of Prandtls mixing length hypothesis in order to com-

    pute the eddy-viscosity using terms of mixing length in analogy to the mo-

    lecular mixing phenomenon. Whereas the molecular viscosity is an intrinsic

    property of the fluid, the eddy-viscosity (and hence the mixing length) de-

    pends upon the flow. Because of this, the eddy-viscosity and mixing length

    must be specified in advance by an algebraic relation between eddy-

    viscosity and length scales of the mean flow. Thus, algebraic models are,

    by definition, incomplete models of turbulence. (Ref. [14]).

    The eddy-viscosity is usually calculated as

    j

    i

    mixT

    dy

    dul 2= (2.12)

    where mixl is the mixing length analogous with the molecular mean free

    path in molecular mixing.

    Well-known and much used algebraic models are the Baldwin-Lomax [2]

    and the Cebeci-Smith [3] models, both of which are two-layer models with

    separate expressions for the computation of the eddy-viscosity in eachlayer. The Cebeci-Smith model is simple and easy to implement. Most

    computational effort goes to the calculation of the boundary layer velocity

    thickness. The Baldwin-Lomax model was formulated for use in computa-

    tions where boundary layer properties such as the boundary layer and the

    boundary layer velocity thickness are difficult to determine.

    Algebraic models are conceptually very simple, economical in terms of

    computer resources and rarely cause any unexpected numerical difficulties.

  • 8/10/2019 Reynolds Stress ARSM

    20/114

    Governing equations and eddy-viscosity turbulence models

    10

    The user of the algebraic models must always be aware of the issue of in-

    completeness. Both the Cebeci-Smith and Baldwin-Lomax models work

    well only for the boundary layer flows for which they have been fine-tuned.

    They reproduce skin friction and velocity profiles well for incompressible

    turbulent boundary layers provided the pressure gradient is not too strong.

    However, they can not be applied to other flow cases, such as wake flows

    or free jets.

    2.4.2 One-equation models

    In one-equation models, one of the two turbulence scales or a combination

    of both is determined from a transport equation. The transport equation for

    the turbulence kinetic energy can be written as

    +

    = jjiijjj

    iij upuuu

    x

    k

    xx

    u

    Dt

    Dk

    1

    2

    1 (2.13)

    where the tensor notation jj xutDtD + is used to denote the rateof change following the mean flow. The term jiij xu is the production

    of the turbulence, P , the term jxk is the molecular diffusion, the term2jii uuu is the turbulent flux of the turbulent kinetic energy, k, and the last

    term jup is the so-called pressure diffusion term. The last term is usu-ally neglected because its contribution is very small.

    The is the dissipation rate per unit mass, defined as

    k

    i

    k

    i

    x

    u

    x

    u

    = (2.14)

    The eddy-viscosity is usually calculated as

    klC mixT = (2.15)

    where the C is a model coefficient. As for algebraic models, additional

    information from the mean flow field or geometrical measures is needed.

    Most commonly used one-equation models are the Spalart-Allmaras [16]

  • 8/10/2019 Reynolds Stress ARSM

    21/114

    Governing equations and eddy-viscosity turbulence models

    11

    and Baldwin-Barth [17] models. Actually, in the models the transported

    quantity is 2k , which follows from the definition of the turbulent Reynolds

    number TTT luRe , since 21~kuT and 23~klT .

    According to reference [15], the one-equation models give usually more

    accurate results than algebraic models because more realism is obtained

    as a result of deducing the turbulent velocity scale from a turbulence quan-

    tity k rather than from gradients of mean velocity. This also implies that,

    by obtaining the solution for the turbulent velocity scale from a differential

    transport equation, some account is taken of the history effects, which be-

    come important in non-equilibrium flows. The differential transport equation

    also removes the problems associated to the flow field points where the

    local mean velocity gradient is zero.

    The main drawbacks with one-equation models are largely the same as for

    the mixing-length models. The history effects are not accounted for the

    length scale l, which is still prescribed as an algebraic function of local

    quantities. According to reference [18] the Spalart-Allmaras model performs

    better than the Chiens k model in a decelerating flow with adversepressure gradient because it is specially designed for aerodynamical pur-

    poses. It is one of the most popular turbulence models in the field of aero-nautical applications, especially in the U.S.A.

    2.4.3 Two-equation models

    Two-equation models have served as the foundation for much of the tur-

    bulence model research until very recent years. The most significant differ-

    ence between the other eddy-viscosity models and the two-equation mod-

    els is that the latter are complete, i.e. can be used to predict properties of agiven turbulent flow with no prior knowledge of the turbulence structure or

    flow geometry. These models provide an equation not only for computation

    of k, but also for the turbulence length scale or equivalent.

    The starting point for virtually all linear two-equation eddy-viscosity models

    is the Boussinesq approximation (2.9) and the turbulence kinetic energy

    equation (2.13). The choice of the second variable is arbitrary and many

  • 8/10/2019 Reynolds Stress ARSM

    22/114

    Governing equations and eddy-viscosity turbulence models

    12

    proposals have been presented. By far two of the most popular dependent

    variables for the second variable have been the dissipation rate and the

    k-specific dissipation rate . Nowadays there is also increasing interest in

    the use of the turbulent time-scale . The most common proposals are

    presented in the table 2.1

    Table 2.1. Proposals for the second variable (Ref. [15]).

    Proposer Year Variable Symbol Physical meaning

    Kolmogorov 1942 121

    lk Frequency

    Rotta 1951 l l Length scale

    Rotta 1968 kl kl k times the length scale

    Harlow & Nakayama 1968 123

    lk Energy dissipation rate

    Spalding 1969 2

    kl Vorticity fluctuations squared

    Nee & Kovasznay 1969 21

    lkT Eddy viscosity

    Speziale 1992

    21

    lk Time-scale

    The k model is the best-known two-equation turbulence model be-cause it is simple to understand and use and relatively easy to program.

    The model dates back to the late 1960s but the most used formulation, re-

    ferred as the standard k model, was presented in 1972 by Jones andLaunder [19]. The eddy-viscosity used with the Boussinesq approximation

    is calculated as

    2k

    CT= (2.16)

    where C is a model coefficient. The k model is used widely in practi-cal engineering calculations even though the standard model can be used

    only with attached flows with thin shear layers. The model fails to reproduce

    the correct flow behaviour in many important flow situations, such as: ad-

  • 8/10/2019 Reynolds Stress ARSM

    23/114

    Governing equations and eddy-viscosity turbulence models

    13

    verse pressure gradients, bluff-body flows, separation, streamline curva-

    ture, swirl, buoyancy, turbulence driven secondary motion, compressibility

    and unsteadiness.

    The k model is based on the choice of the specific dissipation rate asthe second variable. The eddy viscosity is calculated as

    kT= (2.17)

    Even though the wall boundary condition for the is more difficult to for-

    mulate and program than for the , the k model has found its way to

    many engineering, especially aerodynamic, applications. This is mainly dueto the fact that it can also reproduce the flow field behaviour with adverse

    pressure gradient. It usually also predicts separation better than the other

    linear two-equation eddy-viscosity models. Actually, the k SST model[7] predicts the separation fairly well, but it is no longer really a linear

    model.

  • 8/10/2019 Reynolds Stress ARSM

    24/114

  • 8/10/2019 Reynolds Stress ARSM

    25/114

    15

    Chapter 3

    Differential Reynolds stress models

    In this chapter a different approach to model the turbulence is reviewed.

    The Boussinesq approximation is abandoned and the transport equation for

    the Reynolds stresses is presented instead. In the end of the chapter mod-

    elling propositions of the different terms of the transport equation are dis-cussed.

    3.1 Reasons for using the Reynolds stress models

    Eddy-viscosity models perform reasonably well in attached boundary layer

    flows as long as only one component of the Reynolds stress tensor is of

    significant importance. In these cases the eddy-viscosity can be thought as

    a representative for the significant Reynolds stress component. However, ifthe flow becomes more complicated the eddy-viscosity assumption fails.

    There is thus not much hope for a more general validity of the eddy-

    viscosity approach (Ref. [8]).

    The development of the so-called differential Reynolds stress models (also

    called Reynolds stress transport, second-order closure, second-

    moment closure and second-order modelling) started officially in 1968,

  • 8/10/2019 Reynolds Stress ARSM

    26/114

    Differential Reynolds stress models

    16

    when C. Donaldson lectured in Stanford of the need to model the Reynolds

    stress terms. The Boussinesq eddy-viscosity hypothesis is abandoned and

    the unknown Reynolds stress components are obtained directly from the

    solution of differential transport equations in which they are the dependent

    variables. The Reynolds stress tensor is symmetric meaning that four

    equations need to be solved in two-dimensional and six equations in three-

    dimensional flows. An additional equation for the length scale or equivalent

    is also solved in each case.

    The Reynolds stress models are thus more complicated than the eddy-

    viscosity models. Still, they are attractive because they provide a more ac-

    curate representation of the turbulence and are valid over a wider range of

    flows. They can capture many of the complex effects encountered in nature

    and in engineering practice without recourse to the ad-hoc modifications

    found necessary in lower-order models. According to reference [15], exam-

    ples of flow situations where Reynolds stress models were found to give

    accurate predictions include flows with streamline curvature, rotation, swirl

    and buoyancy.

    3.2 Exact equation for the Reynolds stress tensor

    The exact transport equation for the Reynolds-stress tensor jiuu is ob-tained from the momentum equation (2.2) by multiplying the instantaneous

    iu component equation by ju , the ju component equation by iu , addingthe two and then time-averaging the result. Following reference [20], the

    result can be written for constant-density flows with no body forces after

    some rearrangement as

    =

    t

    uuD ji

    +

    k

    ikj

    k

    j

    kix

    uuu

    x

    uuu Terms I and II

    ( )

    ++

    k

    ji

    ikjjkikji

    k x

    uuupupuuu

    x

    1Term III

    +

    +i

    j

    j

    i

    x

    u

    x

    up

    k

    j

    k

    i

    x

    u

    x

    u2 Terms IV and V

    (3.1)

  • 8/10/2019 Reynolds Stress ARSM

    27/114

    Differential Reynolds stress models

    17

    Term I is the convection term. It represents the rate of change of jiuu

    along a streamline. In steady flows this is equal to the rate at which Rey-

    nolds stresses are convected by the mean fluid motion.

    Term IIis the production term. It represents the rate of production of jiuu

    by mean shear, which is the reason for turbulence as discussed earlier.

    The shear stress is generated by interaction of transverse normal stress

    and shear strain. The production term is both large and exact, which is a

    partial reason for the success of the Reynolds stress models. It will be

    hereafter referred to as ijP .

    Term III is the diffusion term. It represents the rate of spatial transport of

    jiuu by the action of turbulent fluctuations, pressure fluctuations and mo-lecular diffusion.

    Term IV is the redistribution term, which is also called the pressure-strain

    term. It represents the redistribution of the available turbulent kinetic energy

    amongst the fluctuating velocity components. The mean flow direction be-

    ing 1x , only 11uu is generated in the shear layer, meaning that it is usuallymuch larger than the other Reynolds stress components. The redistribution

    term shares 11uu s energy out to 22uu and 33uu . It has no effect on theoverall level of turbulent kinetic energy since it has zero trace. The redistri-

    bution term just drives turbulence towards isotropy by redistributing energy.

    This term will be hereafter referred to as ij .

    Usually 21uu has the opposite sign to the boundary layer shear strain. Theredistribution term 12 reduces also 21uu since isotropic turbulence mustbe shear-free. The same applies to the possible shear strains to other di-

    rections.

    Term V is the dissipation term. It represents the dissipation rate of jiuu

    due to molecular viscous action. The dissipation term will be hereafter re-

    ferred to as ij .

  • 8/10/2019 Reynolds Stress ARSM

    28/114

    Differential Reynolds stress models

    18

    3.3 Modelling the Reynolds stress tensor

    With the exception of the convection and production terms of the equation

    for the Reynolds stress tensor, all the other terms introduce new unknown

    correlations that must be modelled in terms of known or knowable quanti-

    ties in order to close the equations.

    The diffusion term is a sum of three parts, which are called the viscous,

    pressure and turbulent diffusion. The contribution of the viscous diffusion

    term to the total rate of transport of jiuu is small at high values of the tur-bulence Reynolds number. This term is thus often neglected in high Rey-

    nolds number applications even though it could be included in the numeri-

    cal simulations without any difficulty.

    Little is known about the pressure diffusion term since direct measurements

    can not be conducted. Estimates of its magnitude are done by indirect

    methods, which contain all the errors made in the measurements of the

    other terms. The consensus of several experiments, however, suggests

    that this term is relatively unimportant and may therefore be neglected (Ref.

    [15]).

    Daly and Harlow [21] were the first to model the turbulent diffusion term

    using the gradient transport hypothesis, meaning that the diffusion of a

    quantity is assumed to be proportional to the spatial gradient of the same

    quantity. According to reference [15] the turbulent diffusion can be written

    as

    l

    ji

    lkSkjix

    uuuu

    kCuuu

    =

    (3.2)

    where SC is an empirical model coefficient typically set equal to 0.22,

    which is obtained by computer optimisation. The ratio k represents a

    characteristic time scale of the energy containing eddies.

    Alternative models for the turbulent diffusion exist, but they all are consid-

    erably more complex than Daly & Harlows model without necessarily pro-

  • 8/10/2019 Reynolds Stress ARSM

    29/114

    Differential Reynolds stress models

    19

    ducing better overall performance. Examples of such models are Hanjalic &

    Launder [22] and Lumley & Khajeh-Nouri [23].

    The redistribution term modelling is based on the Poisson equation forthe instantaneous pressure. It is obtained by differentiating the Navier-

    Stokes equations in a manner that is beyond the scope of this text. Then,

    by subtracting the mean, the following equation for the fluctuating pressure

    is obtained

    ( )

    +

    =

    j

    i

    j

    i

    ij

    jiji

    i x

    u

    x

    u

    xx

    uuuu

    x

    p2

    1 2

    2

    2

    (3.3)

    where a sum of two very distinct terms is identified in the parentheses. The

    first term of the sum contains only turbulence quantities and second term

    contains mean-velocity gradients. When the equation is solved for homo-

    geneous turbulence, the so-called Chous integral for the pressure-strain

    correlation is obtained

    +

    +

    =

    wij

    jl

    im

    m

    l

    jml

    iml

    j

    i

    r

    dvol

    rr

    uu

    x

    u

    rrr

    uuu

    x

    up,

    *2**3

    24

    1

    (3.4)

    where * denotes quantities evaluated at position *x and xxr = * . wij ,is a surface integral, which is important only when the typical size of the

    energy containing eddies is of the same order as distance from a wall (Ref.

    [15]).

    Launder, Reece and Rodi (LRR) proposed that the first and the second

    term in the parentheses of the equation (3.4) are modelled separately [24].

    This is the traditional approach to modelling Chous integral and the great

    majority of Reynolds-stress model developers have followed it.

    Speziale, Sarkar and Gatski (SSG) proposed that the Chous integral could

    be modelled as a whole [25]. This approach is rapidly becoming more

    popular than (LLR) as it does not appear to require a specific model for the

    surface integral wij , .

  • 8/10/2019 Reynolds Stress ARSM

    30/114

  • 8/10/2019 Reynolds Stress ARSM

    31/114

  • 8/10/2019 Reynolds Stress ARSM

    32/114

    Formulation of an explicit algebraic Reynolds stress model

    22

    Also a transport equation for the Reynolds stress anisotropy tensor can be

    construed in a similar was as the transport equation (3.1) for the Reynolds

    stress component tensor. According to reference [8] it can be presented in

    a rotating Cartesian coordinate system as

    ( )aij

    ijijijji

    l

    lji

    l

    ljiijC

    PP

    k

    uu

    x

    ku

    k

    uu

    x

    uuu

    Dt

    Dak+++

    =

    1

    1

    (4.2)

    where the dissipation rate tensor ij and the redistribution tensor ij need

    to be modelled. On the other hand, the production terms ijP and 2iiPP=and the Coriolis term ( )aijC do not need any modelling since they can be

    calculated directly from the Reynolds stress tensor.

    In flows where the Reynolds stress anisotropy varies slowly in time and

    space, the transport equation for the Reynolds stress anisotropy tensor is

    reduced to an implicit algebraic relation. Many inhomogeneous flows of en-

    gineering interest consist of a steady flow and equilibrium turbulence as-

    sumption can thus be made. Therefore the convection and diffusion of the

    Reynolds stress anisotropy may be neglected. This is equivalent to the as-

    sumption made in reference [27] that the convection and diffusion of each

    Reynolds stress component scale with the convection and diffusion of the

    turbulent kinetic energy. This is indeed the traditional ARSM idea, to ne-

    glect convection and diffusion terms in the exact transport equation for the

    Reynolds stress anisotropy. This means that the left-side terms in the

    equation (4.2) can be neglected and set to zero

    01

    =

    l

    lji

    l

    ljiij

    x

    ku

    k

    uu

    x

    uuu

    Dt

    Dak

    (4.3)

    The convection term DtDaij is exactly zero for all stationary parallel mean

    flows, such as fully developed channel and pipe flows. For inhomogeneous

    flows the assumption of negligible diffusion effects can cause problems,

    particularly in regions where the production term is small or where the in-

    homogeneity is strong. However, the ARSM assumption incorporates in a

    natural way the effects of rotation, effects of streamline curvature and

    three-dimensionality of the flow. The assumption results in an implicit alge-

  • 8/10/2019 Reynolds Stress ARSM

    33/114

    Formulation of an explicit algebraic Reynolds stress model

    23

    braic equation for the Reynolds stress anisotropy tensor, which can be

    written as

    ( )aij

    ijijijjiCPP

    kuu ++=

    1 (4.4)

    The production terms ijP and 2iiPP= and the Coriolis term ( )aijC do notneed any modelling since they can be calculated directly from the Reynolds

    stress tensor. However, the dissipation rate tensor ij and the redistribution

    ij need to be modelled. In a non-rotating coordinate system, the produc-

    tion term is normally written as implied in the equation (3.1). It is rewritten

    here for the sake of completeness as

    k

    ikj

    k

    j

    kiijx

    uuu

    x

    uuuP

    = (4.5)

    To illustrate the natural way in which rotational effects enter in this type of

    formalism it is convenient to split the mean velocity gradient tensor into a

    mean strain and a mean vorticity tensor. Symbols ijS and ij are usedhere to represent these tensors, normalised with the turbulent time-scale as

    =

    +

    =

    i

    j

    j

    iij

    i

    j

    j

    iij

    x

    u

    x

    u

    x

    u

    x

    uS

    2

    2

    (4.6)

    where the turbulent time-scale is defined as

    k

    (4.7)

    A consistent formulation of the equation (4.4) can then be obtained by re-

    placing the mean vorticity tensor by the absolute vorticity tensor. According

    to reference [8], this can formally be done introducing the absolute mean

    vorticity tensor

    Sijijij +=* (4.8)

  • 8/10/2019 Reynolds Stress ARSM

    34/114

    Formulation of an explicit algebraic Reynolds stress model

    24

    where the system rotation tensor Sij is defined as

    Skjik

    Sij = (4.9)

    where ijk is the permutation tensor and Sk is the constant rotation rate

    vector of the system. The permutation tensor is unity when the indexes are

    123, 231 or 312. It is minus unity when the indexes are 213, 321 or 132. In

    other cases it is zero. Now, it is possible to rewrite the production term (4.5)

    using the equations (4.1), (4.6) and (4.8). It is presented here normalised

    by the dissipation rate as

    ( ) ( )kjikkjikkjikkjikij

    ij

    aaaSSaS

    P**

    3

    4

    +++= (4.10)

    The Coriolis term ( )aijC arises from the transformation of the convection

    term. It is presented in this text for completeness, even though it was not

    programmed into the FINFLO, as the EARSM was only implemented into a

    non-rotating coordinate system. For rotational purposes, FINFLO uses a

    semi-rotational formulation, which is presented in reference [28]. The

    Coriolis term can be written according to reference [8] as

    ( )kj

    S

    ik

    S

    kjik

    a

    ij aaC = (4.11)

    For the present modelling purpose the dissipation rate tensor is assumed to

    be isotropic. The equation is written here normalised by the dissipation rate

    as

    ijij

    3

    2= (4.12)

    The redistribution term is modelled in two subparts, slow and fast redistri-

    bution, as proposed in reference [8]. According to reference [29], the slow

    redistribution rate can be though to be linear in terms of the anisotropy ten-

    sor as

    ( )

    ij

    sij

    aC1=

    (4.13)

  • 8/10/2019 Reynolds Stress ARSM

    35/114

    Formulation of an explicit algebraic Reynolds stress model

    25

    where 1C is a model constant. For the rapid redistribution rate the general

    linear Launder, Reece and Rodi (LRR) model [24] is chosen. It is normally

    written for a non-rotating system as

    ( )

    +

    +=

    PDC

    x

    u

    x

    uk

    CPP

    C

    ijij

    i

    j

    j

    iijij

    rij

    3

    2

    11

    28

    55

    230

    3

    2

    11

    8

    2

    22

    (4.14)

    where the 2C is a model constant and ijD is calculated as

    i

    kkj

    j

    kkiij

    x

    uuu

    x

    uuuD

    = (4.15)

    A simple way of obtaining a consistent, frame-independent formulation of

    the rapid pressure-strain rate model is to apply the same methodology as

    for the production term, i.e. to use the mean strain and vorticity tensors de-

    fined in equation (4.4.) and to normalise by . The result can be written as

    ( )

    ( )kjikkjik

    ijmkkmkjikkjikij

    rij

    aaC

    SaaSSaC

    S

    **2

    2

    11

    107

    3

    2

    11

    69

    5

    4

    +

    ++

    +=

    (4.16)

    When the equations for the production term (4.5), the Coriolis term (4.11),

    the dissipation term (4.12), the slow redistribution term (4.13) and the fast

    redistribution terms (4.16) are inserted to the equation (4.4), the implicit al-

    gebraic equation for the Reynolds stress anisotropy tensor is then obtained

    as

    ( )

    +

    +

    +=

    +

    kiikijkjikkjik

    kjRik

    Rkjikijij

    SaaSSaC

    aaC

    SaP

    C

    3

    2

    11

    95

    11

    17

    15

    81

    2

    2

    1

    (4.17)

    Here the effective mean vorticity tensor Rij is dependent on the modelconstant 2C and therefore on the choice of the redistribution model.

  • 8/10/2019 Reynolds Stress ARSM

    36/114

    Formulation of an explicit algebraic Reynolds stress model

    26

    The effective mean vorticity tensor Rij of the equation (4.17) can be writtenas

    Sijij

    Sijij

    Rij

    CC

    C

    +++=

    ++=

    17127

    1711

    2

    2

    2

    * (4.18)

    It should be noted, however, that in the present FINFLO formulation the

    effective mean rotation rate tensor is the same as the vorticity tensor

    ijRij = , for the reasons discussed earlier. The dissipation rate must be

    modelled by a transport equation similar to the one used by the two-

    equation eddy-viscosity models. It should be noted that equation (4.17) rep-

    resents a non-linear relation, since the production to dissipation ratio is de-

    fined as

    kiikSaP

    (4.19)

    The algebraic Reynolds stress model has been formulated in the form of

    equation (4.17). In the next section this is further simplified.

    4.2 Explicit algebraic Reynolds stress model

    The implicit ARSM relation for the Reynolds stress anisotropy tensor has

    been found to be numerically and computationally cumbersome since there

    is no diffusion or damping present in the equations. This usually means

    convergence problems. In many applications the computational effort has

    been found to be excessively large and the benefits of using ARSM instead

    of the full Reynolds stress transport form are then successively lost, as

    Wallin discusses [8]. Therefore a more simple and straightforward way to

    calculate the Reynolds stress anisotropy is needed. A model, where the

    Reynolds stresses are explicitly related to the mean flow field, is called an

    explicit algebraic Reynolds stress model. It is much more numerically ro-

    bust and has been found to have almost a negligible effect on the compu-

    tational effort as compared to the linear k or the linear k model.

    The procedure used here to obtain an explicit form is to avoid the non-

    linearity by considering the production to dissipation ratio P as an extra

  • 8/10/2019 Reynolds Stress ARSM

    37/114

    Formulation of an explicit algebraic Reynolds stress model

    27

    unknown. The resulting linear equation system can then be formally written

    as

    ( ) 0,,, = PSa RklklklijL (4.20)

    and may, in principal, be solved directly. However, by the inspection of the

    equation (4.17) it is realised that the anisotropy tensor is dependent on only

    two other tensors, ijS andRij , which can be used to form a complete base

    for the anisotropy. The most general form for ija in terms of ijS andRij

    consists of ten tensorially independent groups to which all higher-order ten-

    sor combinations can be reduced with the aid of the Caley-Hamilton theo-

    rem. Various techniques have been developed with different level of simpli-

    fications in references [26], [30], [31], [32] and [33], for example. The Rey-

    nolds stress anisotropy tensor is, however, written here following reference

    [8] as

    ( )SSIISSa +

    +

    += 423221

    3

    1

    3

    1 IIIIS

    ( )

    ++

    +++ ISSISSSS VIV

    3

    2

    3

    2 22227

    22

    6

    22

    5 (4.21)

    ( ) ( ) ( )+++ 222210

    22

    9

    22

    8 SSSSSSSS

    For simplicity, boldface has been used to denote second-rank tensors as

    ija=a , ijS=S and Rij= . The inner product of two matrices is definedas ( ) ( ) kjikijij SS

    2SSS and I represents the identity matrix. The

    coefficients may be functions of the five independent invariants of S

    and , which can be written as

    ( )( )

    ( )

    ( )

    ( ) likljkij

    kijkij

    kijkij

    jiij

    jiijS

    SSV

    SIV

    SSSIII

    II

    SSII

    ====

    ====

    ==

    22

    2

    3

    2

    2

    trace

    trace

    trace

    trace

    trace

    S

    S

    S

    S

    (4.22)

    Other scalar parameters may also be involved.

  • 8/10/2019 Reynolds Stress ARSM

    38/114

  • 8/10/2019 Reynolds Stress ARSM

    39/114

    Formulation of an explicit algebraic Reynolds stress model

    29

    The procedure to solve this equation system is the following:

    1) The general form of the anisotropy, the tensor polynomial (4.21), is in-

    serted into the simplified ARSM equation (4.24) where N is not yet

    determined. Now N is considered to be a known parameter. This re-

    sults in a linear equation system for the coefficients.

    2) The linear equation system can be solved using the fact that higher-

    order tensor groups can be reduced with the aid of the Cayley-

    Hamilton theorem. The solution is unequivocal because the ten groups

    in the general form (4.21) form a complete basis for the system. The

    solution consist of many pages of complicated tensor algebra and it is

    not essential to reproduce it here. The -coefficients are now functions

    of N , or the production to dissipation ratio.

    3) The next step is to formulate and solve the non-linear scalar equation

    for N . This is done by inserting the solution of the coefficients into

    the tensor polynomial (4.21) for the Reynolds stress anisotropy tensor.

    4) The final step is to insert the resulting equation to the simplified ARSM

    equation (4.24). Then a non-linear scalar equation for N is obtained,

    which for a general three-dimensional flow field is of sixth order.

    The solution for the simplified ARSM equation (4.24) is presented in the

    following section 4.3 for two- and 4.4 for three-dimensional mean flows. In

    section 4.5, the solution for the model without the simplification made from

    the equation (4.17) to (4.23) is also presented.

    4.3 Simplified EARSM for two-dimensional mean flow

    For two-dimensional mean flows the simplified solution of the ARSM equa-

    tion is reduced to only two non-zero coefficients, which can be expressed

    according to reference [8] as

    =

    =

    IIN

    IIN

    N

    2

    1

    5

    6

    25

    6

    24

    21

    (4.27)

  • 8/10/2019 Reynolds Stress ARSM

    40/114

    Formulation of an explicit algebraic Reynolds stress model

    30

    In the equation (4.27) it is clearly seen that the denominator cannot become

    singular since II is always negative. The non-linear equation for Nin two-

    dimensional mean flow can be derived by inserting the tensor polynomial

    (4.21) for the Reynolds stress anisotropy tensor with the coefficients from

    the equation (4.27), into the definition of N, equation (4.24). The resulting

    equation is cubic and can be written following reference [8] as

    02210

    271

    21

    3 =+

    + IICNIIIINCN S (4.28)

    The equation can be solved in a closed form with the solution for the posi-

    tive root being

    ( ) ( )

    ( )