sdarticle(33)

Upload: pramod-badyankal

Post on 08-Apr-2018

220 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/6/2019 sdarticle(33)

    1/13

    An efficient solver for the RANS equations and a one-equation turbulence model

    R.C. Swanson ,1, C.-C. Rossow

    DLR, Deutsches Zentrum fr Luft- und Raumfahrt, Lilienthalplatz 7, D-38108 Braunschweig, Germany

    a r t i c l e i n f o

    Article history:

    Received 23 October 2009

    Received in revised form 13 July 2010Accepted 19 October 2010

    Available online 28 October 2010

    Keywords:

    Runge-Kutta

    Implicit preconditioner

    Multigrid

    Reynolds-averaged NavierStokes equations

    SA turbulence model

    a b s t r a c t

    A three-stage Runge-Kutta (RK) scheme with multigrid and an implicit preconditioner has been shown to

    be an effective solver for the fluid dynamic equations. Using the algebraic turbulence model of Baldwin

    and Lomax, this scheme has been used to solve the compressible Reynolds-averaged NavierStokes(RANS) equations for transonic and low-speed flows. In this paper we focus on the convergence of the

    RK/Implicit scheme when the effects of turbulence are represented by the one-equation model of Spalart

    and Allmaras. With the present scheme the RANS equations and the partial differential equation of the

    turbulence model are solved in a loosely coupled manner. This approach allows the convergence behavior

    of each system to be examined. Point symmetric Gauss-Seidel supplemented with local line relaxation is

    used to approximatethe inverse of theimplicit operator of the RANS solver. To solve theturbulence equa-

    tion we consider three alternative methods: diagonally dominant alternating direction implicit (DDADI),

    symmetric line Gauss-Seidel (SLGS), and a two-stage RK scheme with implicit preconditioning. Compu-

    tational results are presented for airfoil flows, and comparisons are made with experimental data. We

    demonstrate that the two-dimensional RANS equations and a transport-type equation for turbulence

    modeling can be efficiently solved with an indirectly coupled algorithm that uses RK/Implicit schemes.

    2010 Elsevier Ltd. All rights reserved.

    1. Introduction

    Reliable and sufficient convergence for steady-state computa-

    tions of turbulent flows continues to be a challenge in computa-

    tional fluid dynamics. Here sufficient convergence means that the

    residuals of the fluid dynamic equations and the equation set of

    a turbulence model are reduced to the level of the truncation error

    of the numerical scheme. In many applications a turbulence model

    has one or more partial differential equations (PDEs) which have a

    transport form and represent the effects of turbulence on the flow.

    When solving the transport-type equations of turbulence models,

    either directly or indirectly coupled to the flow equations, the

    residuals are frequently reduced only two orders of magnitude.

    In addition, the poor convergence of these transport-type equa-

    tions adversely affects the convergence of the flow equations. Of

    course, when adequate convergence is not achieved, there is no

    assurance that the results obtained represent an acceptable

    approximation of the solution even from an engineering perspec-

    tive. Thus, there is a strong need for improved numerical methods

    for not only obtaining steady-state solutions but also unsteady

    solutions when using a dual time-stepping scheme.

    When developing an improved numerical method for solving

    the Reynolds-averaged NavierStokes (RANS) equations, a neces-

    sary consideration is the coupling of the RANS equations and the

    equation or equations of the turbulence model being applied. If

    both the fluid dynamic and turbulence equations are directly cou-

    pled, then the characterization of the discrete system can change.

    That is, with appropriate discretization the fluid dynamic equa-

    tions are positive definite (sometimes called a vector positive sys-

    tem [1]), making them amenable to relaxation, but the directly

    coupled system may not be, due to the equation set for the turbu-

    lence model [2]. The numerical stiffness of the entire system is also

    much higher due to the source terms of the turbulence model. An

    alternative is to use indirect coupling of the two equation sets.

    Generally, in an iterative solution process with this approach the

    flow variables are updated while the turbulence variables are fro-

    zen; and then, the turbulence variables are updated while the flow

    variables are treated as fixed quantities. Strategies for implement-

    ing indirect coupling depend on the algorithm being used. For

    example, when applying multigrid methods, there are two princi-

    pal strategies for indirect coupling. The first approach [3,4] is to

    solve the mean flow and turbulence equations in sequence on each

    grid level. This can augment the coupling effects, which may be

    beneficial for certain types of problems. The second strategy [5,6]

    is to use the eddy viscosity determined on the finest grid on all

    coarser grids in the multigrid procedure for the mean flow

    equations. Then, the turbulence model equation set is solved with

    multigrid or another algorithm. This method can provide an

    0045-7930/$ - see front matter 2010 Elsevier Ltd. All rights reserved.doi:10.1016/j.compfluid.2010.10.010

    Corresponding author.

    E-mail addresses: [email protected](R.C. Swanson), [email protected]

    (C.-C. Rossow).1 Corresponding author was visiting scientist at the Center for Computer Applica-

    tions in AeroSpace Science and Engineering (C2A2S2E), DLR, Braunschweig, Germany.

    Computers & Fluids 42 (2011) 1325

    Contents lists available at ScienceDirect

    Computers & Fluids

    j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / c o m p fl u i d

    http://dx.doi.org/10.1016/j.compfluid.2010.10.010mailto:[email protected]:[email protected]://dx.doi.org/10.1016/j.compfluid.2010.10.010http://www.sciencedirect.com/science/journal/00457930http://www.elsevier.com/locate/compfluidhttp://www.elsevier.com/locate/compfluidhttp://www.sciencedirect.com/science/journal/00457930http://dx.doi.org/10.1016/j.compfluid.2010.10.010mailto:[email protected]:[email protected]://dx.doi.org/10.1016/j.compfluid.2010.10.010
  • 8/6/2019 sdarticle(33)

    2/13

    advantage in flexibility for independent evaluations of the schemes

    for solving the two equation sets. In addition, since there is latitude

    in choosing the iterative scheme and update procedure for the tur-

    bulence quantities, one can also investigate methods to enhance

    the coupling effects.

    By indirectly coupling the equations one can focus on the

    specific properties of each equation set to obtain the best possible

    convergence of the two systems of equations. Furthermore, the

    essential properties of an algorithm for efficiently solving the di-

    rectly coupled system can be identified. There are common design

    criteria for the algorithms of both equation sets. These require-

    ments are as follows: (1) high CourantFriedrichsLewy (CFL) lim-

    it, (2) convergence with weak dependency on mesh density, (3)

    suitable for stiff discrete systems. In addition, for the equation

    set of the turbulence model there must also be appropriate treat-

    ment of any source terms so that convergence is not adversely

    affected.

    A candidate for the flow solver of the loosely coupled system is

    an RK/Implicit scheme with three stages and three evaluations of

    the numerical dissipation. Subsequently, this scheme is designated

    as the RK3/Implicit scheme. Previously, Rossow [7] and Swanson

    et al. [8] demonstrated that fast convergence can be obtained for

    both the two-dimensional (2-D) and three dimensional (3-D) RANS

    equations with the RK3/Implicit scheme with multigrid when

    using the Baldwin-Lomax (BL) algebraic eddy viscosity model [9].

    Although there is some slowdown in the convergence rate of the

    3-D scheme relative to the 2-D scheme, Swanson et al. demon-

    strate that this scheme is more than 10 times faster than a well-

    tuned standard RK scheme with scalar implicit residual smoothing

    and multigrid. The underlying three-stage RK scheme of this algo-

    rithm is important for clustering of the eigenvalues associated with

    the error components of the iterative process. Preconditioning with

    a fully implicit operator, which allows a CFL number of 1000, treats

    the discrete stiffness problemassociated with viscous-layer resolu-

    tion. The operator of the discrete implicit system can be approxi-

    mately inverted with symmetric Gauss-Seidel (SGS).

    The main purpose of this work was to initiate an effort to satisfythe need to significantly augment the effectiveness (as measured

    by reliability and efficiency) of algorithms for solving the RANS

    equations and the PDEs of turbulence models. Since establishing

    a highly effective scheme in two dimensions is a prerequisite for

    constructing an efficient scheme in three dimensions, the focus

    of the present effort is on a 2-D scheme. Moreover, we assess the

    performance of an efficient RANS solver (i.e., RK3/Implicit scheme

    with multigrid) when the turbulent viscosity field is generated by

    solving a transport-type equation.

    To represent the effects of turbulence we use the SpalartAllm-

    aras (SA) model, which is a transport-type equation model that is

    frequently used in solving a variety of fluid dynamics problems.

    In the first section of this paper this turbulence model is described,

    and the specifics of its implementation are given. Next the numer-ical schemes for solving the mean flow and turbulence equations

    are presented and discussed. Modifications of the RK3/Implicit

    scheme that have produced improved efficiency and robustness

    are emphasized. Then three approaches for solving the SA equation

    are considered. These methods are as follows: diagonally dominant

    alternating direction implicit (DDADI), symmetric line Gauss-Sei-

    del (SLGS), and a two-stage RK scheme (RK2/Implicit) with implicit

    preconditioning. In the results section the convergence behavior of

    the methods for solving the RANS equations and the SA equation is

    examined. The effectiveness of the loosely coupled algorithm at

    high Reynolds numbers and low Mach number is presented. Rapid

    evolution of global quantities such as lift and drag coefficients is

    demonstrated. Furthermore, we show that the convergence of

    the RK3/Implicit scheme with the SA model is similar to that ob-tained with the BL model.

    2. SpalartAllmaras turbulence model

    Here we provide a sufficient description of the SA model to al-

    low implementation. A detail discussion explaining the modeling

    of the physical terms in the single transport-type equation is given

    in the paper by Spalart and Allmaras [10]. Let mt be the eddy viscos-ity, which is defined by

    mt ~mfv1; fv1 v3

    v3 C3v1

    ; v ~mm; 2:1

    where m is the kinematic viscosity. The transport-type equation forem given in Ref. [10] is written as

    @~m@t

    uj@~m@xj

    Cb11 ft2eS~m

    1

    r@

    @xjm ~m

    @~m@xj

    ! Cb2

    @~m@xj

    @~m@xj

    & ' Cw1fw

    Cb1j2

    ft2

    ~md

    2 S; 2:2

    where t is time, xj and uj are Cartesian coordinates and velocitycomponents, respectively, and

    eS S ~mj2d2

    fv2; fv2 1 v

    1 vfv1; 2:3

    with S being the magnitude of the vorticity (jXj), d delineating the

    distance to the closest wall boundary, and j denoting the vonKrmn constant. The first, second, and third terms on the right-

    hand side of Eq. (2.2) represent the production, diffusion, and

    destruction terms, respectively. The last term is a source term,

    which is defined by

    S ft1DU2; 2:4

    where DU is the norm of the difference between the velocity at thetransition location and that at a field point being considered. The

    function fw in Eq. (2.2) is given by

    fw g1 C6w3

    g6 C6w3

    1=6; 2:5

    where g and r are defined by

    g r Cw2r6 r; r

    ~meSj2d2 : 2:6For large values ofrthe functionfw goes to a constant, and a value of

    10 is appropriate. The function ft2 is defined as

    ft2 Ct3 expCt4v2: 2:7

    Spalart includes the transition function given by

    ft1 Ct1gtexp Ct2x2tDU2

    d2

    g2td2t

    !; 2:8

    where dt is the distance from the field point to the boundary-layer

    trip (where trip refers to a known location for transition), xt is thewall vorticity at the trip,

    gt min 0:1;DU=xtDxt : 2:9

    andDxt is the grid spacing along the wall at the trip.

    In the present implementation of the model we do not include

    the trip function, which is usually neglected when applying the

    model (e.g., Ref. [11]). In addition, for the purposes of grouping

    terms similar in form and numerical implementation, we rewrite(after some algebra and rearranging of terms), Eq. (2.2) as

    14 R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325

  • 8/6/2019 sdarticle(33)

    3/13

    @~m@t

    uj@~m@xj

    SP1 ~m SP2 ~m SD~m

    D~m; 2:10

    where SP1 and SP2 are the two contributions to the production term,

    SD is the destruction term, and D is the diffusion term, and they are

    given by

    SP1 ~m Cb11 ft2jXj~m;

    SP2 ~m SD~m Cb11 ft2fv2 ft2j2 Cw1fw

    ~md

    2;

    D~m 1

    r@

    @xjm 1 Cb2~m

    @~m@xj

    !

    Cb2r

    ~m@2~m@x2j

    :

    The constants of the model are as follows:

    Cb1 0:1355; r 2

    3; Cb2 0:622; j 0:41; 2:11

    Cw1 Cb1j

    1 Cb2

    r;

    Cw2 0:3; Cw3 2; Cv1 7:1; Ct1 1; Ct2 2;

    Ct3 1:2; Ct4 0:5:

    On a solid boundary ~m 0. Originally, the free-stream ~m was set to1.342m1, where m1 is the free-stream kinematic viscosity. In orderto avoid the possibility of a delayed transition, the free-stream va-

    lue of ~m is set to 3m1, as suggested by Rumsey [12].

    3. Numerical schemes

    To solve the two-dimensional RANS equations we use the RK3/

    Implicit scheme. Complete details of the scheme are presented in

    the papers of Rossow [7] and Swanson et al. [8]. The SA turbulence

    model requires the solution of one transport-type equation. In the

    present work we do not directly couple the solutionof the fluid dy-

    namic equations with the additional equation of the turbulencemodel. To solve the transport-type equation of the SA turbulence

    model we consider the DDADI, SLGS, and RK2/Implicit schemes.

    In the first part of this section the essential elements of the RK3/

    Implicit scheme are presented. Recent enhancements of the origi-

    nal RK3/Implicit scheme are also introduced. Then the three meth-

    ods considered for solving the SA equation are described.

    3.1. RK/implicit scheme

    We apply a finite-volume approach to discretize the fluid

    dynamic equations and use the approximate Riemann solver of

    Roe [13] to obtain a second-order discretization of the convective

    terms. The viscous terms are discretized with a second-order cen-

    tral difference approximation. To obtain an explicit update to the

    solution vector for the flow equations we use a three-stage RK

    scheme. The update for the qth stage of the RK scheme is given by

    Wq W0 dWq; 3:1

    where the change in the solution vector W is

    dWq Wq W0 aq

    Dt

    VLW

    q1; 3:2

    and L is the complete difference operator for the system of equa-

    tions. Here aq is the RK coefficient of the qth stage, Dt is the timestep, and V is the volume of the mesh cell being considered. For

    the three-stage scheme we use the coefficients

    a1; a2; a3 0:15; 0:4; 1:0

    from Ref. [14].

    To extend the support of the difference scheme we consider

    implicit residual smoothing. Applying the smoothing technique

    of Ref. [15] we have the following:

    LidWq dWq; 3:3

    where Li is an implicit operator. By approximately inverting the

    operator Li we obtain

    dWq aqDt

    VPL; Wq1 aq

    Dt

    VP

    Xall faces

    Fq1n S; 3:4

    where P is a preconditioner defined by the approximate inverseeL1i , Fn is the normal flux density vector at the cell face, and S isthe area of the cell face. The change dWq replaces the explicit

    update appearing in Eq. (3.1). Thus, each stage in the RK scheme

    is preconditioned by an implicit operator.

    A first-order upwind approximation based on the Roe scheme is

    used for the convective derivatives in the implicit operator. To

    derive this operator one treats the spatial discretization terms in

    the flow equations implicitly and applies linearization. For a

    detailed derivation see Rossow [7]. Substituting for the implicit

    operator in Eq. (3.3), we obtain for the qth stage of the RK scheme

    I eDt

    V

    Xall faces

    AnS" #

    dWq aqDt

    V

    Xall faces

    Fq1n S

    bRq1; 3:5where the matrix An is the flux Jacobian associated with Fn at a cell

    face, bRq1 represents the residual function for the (q1)th stage,and e is an implicit parameter, which will be defined later in thissection.

    The matrix An can be decomposed into An and A

    n , which are

    associated with the positive and negative eigenvalues of An and

    defined by

    An

    1

    2An jAnj; A

    n

    1

    2An jAnj: 3:6

    If we substitute for An in Eq. (3.5) using the definitions of Eq. (3.6),

    then the implicit scheme can be written as

    I eDt

    V

    Xall faces

    An S

    " #dW

    qi;j

    bRq1i;j eDtV Xall faces

    An dW

    qNBS; 3:7

    where the indices (i,j) indicate the cell of interest, and NB refers to

    all the direct neighbors of the cell being considered.

    To solve the implicit system of Eq. (3.7) for the changes in con-

    servative variables dWq

    i;j , the implicit operator must be inverted. It

    is sufficient to approximate the inverse of the implicit operator.

    Based on analysis and numerical testing, an adequate approximate

    inverse is obtained with two pointwise symmetric Gauss-Seidel

    (SGS) sweeps, with each complete sweep followed by one local

    (boundary layer and near wake) symmetric line sweep. Previously

    [16], we have demonstrated that line relaxation can be used to effi-ciently approximate the inverse. By using local line relaxation, we

    make the scheme amenable to application on unstructured grids.

    To initialize the iterative process the unknowns are set to zero.

    The choice of the implicit parameter e in Eq. (3.5) affects thehigh-frequency damping of the scheme, and thus, its effectiveness

    as a smoother for multigrid. Fig. 1a shows the amplification factorg

    of the original RK3/Implicit scheme [7,8] with variation ofe. The gwas determined with the one-dimensional Fourier analysis of

    Swanson et al. [8] and is a function of the phase angle hx, which

    is proportional to frequency. At e = 0.5 there is a significant in-crease ingat the highest frequencies, and when e = 0.4, the schemeis unstable. Thus, for the original scheme we chosee = 0.6. By intro-ducing non-standard weighting of the explicit numerical dissipa-

    tion on stages of the RK scheme, the lower bound on e can bedecreased, which results in faster convergence. Here, weighting

    R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325 15

  • 8/6/2019 sdarticle(33)

    4/13

    of the numerical dissipation means that the dissipation in the

    residual function on a given RK stage q (q > 1) is weighted with that

    from a previous stage. In general, there is weighting of both the

    numerical dissipative and physical diffusive terms, and the opera-

    tor L in Eq. (3.2) is defined by

    LWq

    1

    VLcW

    q Xqr0

    cqrLvWr

    Xqr0

    cqrLdWr

    " #; 3:8

    withPcqr 1 for consistency. The operatorsLc;Lv, andLd relate to

    the convective, viscous, and numerical dissipative terms. The coef-

    ficients cqr are the weights of the viscous and dissipative terms oneach stage, and for the 3-stage scheme,

    c00 c1; c10 1 c2; c11 c2; c20 1 c3;

    c21 0; c22 c3: 3:9

    When the weights c1; c2; c3 are [1,1, 1], this is called standardweighting. Based upon analysis and numerical testing we have

    determined that the modified weights [1,0.5,0.5] lead to improved

    robustness of the smoother. With this weighting there is a shift in

    the intersection of the locus of the residual eigenvalues to the left

    along the negative real axis of the complex plane, which increases

    the parabolic stability limit of the basic RK scheme by more than

    a factor of two. A detailed discussion of dissipation weighting is

    given in Jameson [17] and Swanson and Turkel [18]. Fig. 1b shows

    the effect on g due to weighting of the dissipation on the second

    and third stages by a factor of 0.5. The modified scheme is now

    stable when e = 0.4. However, in practice, for stability, good damp-ing of the highest frequencies, and best convergence the parameter

    e is taken to be 0.5.

    In the application of the RK3/Implicit scheme as the smoother ofa full approximation storage (FAS) multigrid method, the CFL num-

    ber is increased to 1000 after 10 multigrid cycles. The hyperbolic

    tangent function defined in

    CFL 1000 tanh0:0052N1

    n o; 3:10

    where N is the number of cycles, is used to smoothly increase the

    CFL number. A W-type cycle (see Ref. [19]) is used to execute the

    multigrid. Coarse meshes are created by eliminating every other

    mesh line in each coordinate direction (i.e., full coarsening). Details

    of the multigrid method are given in the paper by Swanson et al. [8].

    Although we have used lexicographic ordering in applying

    Gauss-Seidel, alternative ordering strategies can also be used. For

    example, the lexicographic ordering of the solution points can bereplaced with red-black ordering. Roberts and Swanson [20] have

    applied eigensystem analysis to show that the RK/ Implicit scheme

    with a Gauss-Seidel preconditioner and red-black ordering is an

    effective smoother for multigrid. With a different data structure

    such as that employed for unstructured grids, an alternative order-

    ing can be more convenient to implement and even lead to a more

    robust iterative method. Due to the larger stencils that are pro-

    duced on unstructured grids when approximating the spatial

    derivatives of the governing equations, the red-black ordering

    must be replaced by a multicolor ordering to ensure that each

    point of a particular color only directly connects to points of a dif-

    ferent color, and thus, obtain a Gauss-Seidel type scheme.

    By using multicoloring the algorithm can be highly parallelized

    (see Refs. [21,22]), and the convergence rate is the same as for

    sequential processing. We can define each color in the multicolor-

    ing ordering as a member of an independent set. Then, each solu-

    tion point is only directly connected to points with a different

    color. First, all points of a particular color on all subdomains areupdated in parallel. Next, another set of points of a different color

    is updated, using the latest information available. This procedure is

    continued until all points have been updated. In updating the solu-

    tion points of the initial color, Jacobi relaxation is used since no up-

    dated points are available. However, this would also be true for

    red-black (odd-even) Gauss-Seidel on a structured mesh. The num-

    ber of colors required to build independent sets is a function of the

    number of points in the stencil.

    3.2. Schemes for SA equation

    After discretizing Eq. (2.10), we consider the implicit form

    I Lx Ly SJD~m R~m; 3:11

    where Lx and Ly are the linear discrete operators for the terms of the

    transport-type equation, SJ is a Jacobian of the source term contain-

    ing the production and destruction of turbulence contributions, and

    R~m is the residual function. The operators for the two coordinatedirections are as follows:

    Lx Dt

    Vh dux dxb1 b2dxdx

    ;

    Ly Dt

    Vh duy dyb1 b2dydy

    h i; 3:12

    where du is a first-order upwind operator for the convective term, d

    is a standard central difference operator, and the coefficients b1, b2

    are defined by the diffusion term of the turbulence model. Theparameter h indicates temporal accuracy. If h = 1/2, then the time

    x

    g

    0 0.5 1 1.5 2 2.5 30

    0.2

    0.4

    0.6

    0.8

    1

    = 1.0

    = 0.8

    = 0.6

    = 0.5

    = 0.4

    x

    g

    0 0.5 1 1.5 2 2.5 30

    0.2

    0.4

    0.6

    0.8

    1

    = 1.0

    = 0.8

    = 0.6

    = 0.5

    = 0.4

    (a) (b)

    Fig. 1. Effecton amplification factorof RK3/Implicit scheme (applied to 1-D Euler equations) due to variation of implicit parametere (3-stage scheme). (a) Standard weightingof numerical dissipation, (b) modified weighting of dissipation.

    16 R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325

  • 8/6/2019 sdarticle(33)

    5/13

    derivative is approximated by a central difference, which is second-

    order accurate (i.e., CrankNicolson type scheme). When h = 1 the

    approximation is a first-order backward difference, and we have a

    fully (an Euler) implicit scheme. The parameter h may also be

    viewed as a measure of implicitness withh > 1 a n d 0 < h < 1 indicat-

    ing under-relaxation and over-relaxation, respectively (to see this

    consider the effect of h for large Dt). The source term Jacobian

    and the residual function are defined by

    SJ Dt

    Vh

    @

    @~mSP2 ~m SD~m

    ; R~m Dt

    VR~m: 3:13

    For the convective and diffusive terms of the residual function we

    use first-order upwind difference and central difference approxima-

    tions, respectively. A first-order approximation of convective terms

    is frequently applied in the implementation of turbulence models to

    promote positivity of the turbulence variables. In general, this is not

    sufficient to ensure positivity, so usually there is also limiting (clip-

    ping) of the turbulence quantities and/or certain terms (e.g., pro-

    duction term) in the set of turbulence field equations.

    An appropriate linearization of the source term is extremely

    important to allow the use of large CFL numbers. One approach

    for solving Eq. (3.11) is to factor the implicit (left-hand side) oper-

    ator and apply the DDADI scheme. Define the diagonal contribu-tion in Eq. (3.11) as

    D I Dx Dy SJ; 3:14

    where Dx and Dy are the diagonal parts of Lx and Ly, respectively.

    Then, after factoring out D, we factor the resulting operator,

    obtaining

    I Ly Dx SJ

    D1 I Lx Dy SJ

    D~m R~m: 3:15

    To invert this implicit operator, we solve the sequence of one-

    dimensional systems corresponding to the two coordinate direc-

    tions. To prevent deterioration in the allowable CFL number and

    damping behavior of the DDADI scheme due to the factorization

    error and possible boundary condition lagging error, we use the

    subiterative procedure described by Klopfer et al. [23]. Using Fou-rier analysis and some applications of the iterative DDADI scheme

    (also called the modified approximate factorization (MAF) scheme)

    MacCormack and Pulliam [24] and Walsh and Pulliam [25] have

    demonstrated that a few subiterations (e.g., two to four) makes

    the DDADI scheme unconditionally stable and improves the damp-

    ing properties. In Pulliam et al. [26] best performance for a diago-

    nalized DDADI was obtained with three to six iterations, and

    three iterations were recommended. Certainly, the effectiveness

    and reliability of the DDADI scheme depends on the convergence

    behavior of the subiterative process and on the magnitude of the

    implicit parameter h. By numerical testing we have found that four

    subiterations produces reliable convergence and best performance

    when DDADI is used by itself or as a smoother for multigrid. Cur-

    rently we use four subiterations when performing one outer itera-tion. Convergence with iteration and subiteration can be

    enhanced by choosing an appropriate implicit parameter. Over a

    range of mesh densities a h between 1.2 and 2.0 works well. Addi-

    tional discussion of the present implementation is given in Swanson

    and Rossow [16].

    With the SLGS scheme the implicit operator of Eq. (3.11) is

    approximately inverted in each iteration with two symmetric

    Gauss-Seidel line relaxation sweeps (line solves performed in ra-

    dial direction only). The RK2/Implicit scheme involves two RK

    stages and an implicit preconditioner, and it is the same type of

    scheme used to solve the mean flow equations. One point SGS

    sweep and one local (boundary layer + near wake) symmetric line

    relaxation sweep are applied twice to obtain an approximate inver-

    sion of the implicit preconditioner. The coefficients for the two-stage scheme are

    a1; a2 0:25; 1:0 :

    Subsequently this method is designated as the RKI-SGS scheme.

    Due to the strong nonlinearities of the source terms, we have

    employed numerical evaluation to determine an appropriate num-

    ber of relaxation sweeps for the SLGS and RKI-SGS schemes. In the

    evaluation we also considered the effect of mesh density on the

    number of relaxation sweeps for solving the turbulence equation.

    To achieve favorable convergence rates the turbulence equationis solved on each stage of the RK/Implicit smoothing scheme for

    the mean flow equations. The CFL number for all solvers of the

    SA equation is 1000. When solving the mean flow equations, solu-

    tion of the turbulence equation is performed on the fine mesh only,

    and the eddy viscosity is frozen on the coarser meshes. For addi-

    tional enhancement of efficiency and robustness when solving

    Eq. (2.2) the three different solution strategies, namely DDADI,

    SLGS, and RKI-SGS, are supported by a V-cycle multigrid algorithm.

    The multigrid algorithm is called at each stage of the fine mesh RK/

    Implicit scheme when solving the mean flow equations.

    One issue that can arise in solving the turbulence equation is

    the unbounded growth of the solution. Exponential growth of ~mcan occur when there is a sufficiently large imbalance of the pro-

    duction and destruction terms so as to produce an instability. Aspointed out by Allmaras [27], this is a consequence of the Jacobian

    of the production and destruction terms becoming positive. Such a

    behavior can become a significant problem especially when using

    multistage (e.g., RK) relaxation and only updating the precondi-

    tioner on the zeroth stage. In the current formulationwe do not ob-

    serve this type of problem. There are several possible reasons for

    this. One is the form of the weak coupling used, where the multi-

    grid scheme for the turbulence equation is separate from that of

    the mean flow equations. Another reason is that the preconditioner

    is updated on each stage of the RK scheme. There is also the poten-

    tial benefit from reducing the positive contribution to the Jacobian

    of the production and destruction terms (see Eq. (3.13)). These fac-

    tors have contributed to the increased reliability of the solver.

    During the course of this work we have made the followingconvergence behavior observations, which are similar to the ones

    reported by Walsh and Pulliam [25]. The rate of development of

    the turbulence field can signficantly affect the convergence of the

    flow solver. Conversely, how well the flow solver converges can

    have an impact on the effectiveness of the scheme for solving the

    equation set of the turbulence model. Moreover, when the RANS

    and turbulence equations are being solved in a loosely coupled

    manner, an essential requirement for an effective total algorithm

    is that the numerical solution vector of each equation set exhibits

    a similar evolution rate.

    4. Computational results

    Computations for turbulent, viscous flow over the RAE 2822 air-foil were performed to evaluate the convergence behavior of the

    RK3/Implicit scheme when applying the SA turbulence model.

    The airfoil solutions were primarily calculated with the Cases 1

    and 9 flow conditions given in Table 1 from the experimental

    investigation of Cook et al. [28]. In the table M1 is the free-stream

    Mach number, a denotes the angle of attack, Rec represents theReynolds number based on chord length, and xtr/cis the transition

    Table 1

    Flow conditions for RAE 2822 airfoil.

    Cases M1 a (deg.) Rec xtr/c

    Case 1 0.676 1.93 5.7 106 0.11

    Case 9 0.730 2.79 6.5 106 0.03

    R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325 17

  • 8/6/2019 sdarticle(33)

    6/13

    location divided by the chord length. For Case 1 the flow is

    primarily subsonic with a relatively small region of supersonic

    flow. For Case 9 the flow is transonic, with a shock wave occurring

    on the upper surface at approximately the 55% chord location. In

    addition, an incompressible (M1 = 0.001) airfoil flow calculation

    was made with a = 2.79 and Rec = 6.5 106.

    In solving the flow equations structured meshes with a C-type

    topology were used. We primarily considered three mesh densi-

    ties, with the finest having 1280 cells around the airfoil and 256

    cells in the radial direction. Successively coarser grids (640 128

    and 320 64) were generated with half as many cells as the next

    finer grid in each coordinate direction. There is clustering of the

    grids at the leading and trailing edges of the airfoil and also at

    the surface in the radial direction. The finest mesh has 1024 cells

    on the airfoil and a minimum normal mesh spacing of 3 106.

    On the airfoil surface the maximum cell aspect ratio is 2032. The

    outer boundary is located at 20 chords away from the airfoil. To

    investigate the RANS solver for a range of Reynolds (Re) numbers

    we used a set of meshes (adapted to the Re of the flow [29]) con-

    taining 368 88 cells.

    In all the applications the same boundary conditions were im-

    posed for the fluid dynamic equations. On the surface the no-slip

    condition was applied. At the outer boundary Riemann invariants

    were used. A far-field vortex effect was included to specify the

    velocity for an inflow condition at the outer boundary. A detailed

    discussion of the boundary conditions is given in Ref. [18]. In the

    computations two types of initial conditions were considered.

    One type uses the free-stream values of the dependent variables.

    The other one uses an initial solution determined by applying

    full multigrid (FMG). With FMG a grid sequencing process is

    used to generate an initial solution on successively finer meshes.

    Multigrid is used to solve the discrete problem on each grid in

    the sequence. All computations were performed on a Fujitsu

    computer with an Intel core two duo CPU 6750 processor at

    2.66 GHz.

    When comparing the computational performance of the RK3/

    Implicit scheme for different turbulence model solvers and for dif-ferent mesh densities, the computational time is included. These

    computing times provide a reasonable estimate of performance

    since all solvers were programmed in Fortran 77 by the same per-

    son using the same coding practices. This also applies to compari-

    sons that are made with a frequently used scheme for solving the

    RANS equations. Furthermore, by providing a description of the

    processor used, the computational times required on other com-

    puters can be determined.

    Fig. 2 shows convergence histories for Case 9 of the schemes for

    the RANS and turbulence equations. For these three results on the

    320 64 grid the SA equation was solved with DDADI, SLGS, and

    RKI-SGS schemes. The L2 norm of the residual of the continuity

    equation is used as a measure of convergence for the flow equa-

    tions. With each scheme the residual of the mean flow equations

    is reduced 13 orders of magnitude in less than 75 multigrid cycles

    (for an average reduction rate of about 0.65). In fact, the residual

    histories for the mean flow equations essentially coincide. This is

    not surprising since for all schemes the residuals of the turbulence

    equation are reduced between seven and eight orders. The DDADI

    scheme requires less CPU time than the other two schemes, as seen

    in Table 2. However, the residual is reduced about an order of mag-

    nitude more with the RKI-SGS scheme than the DDADI scheme. The

    RKI-SGS scheme has the advantage of being compatible with the

    solver of the mean flow equations, allowing the possibility of con-

    structing a fully coupled solver. In addition, it is amenable to appli-

    cation in an unstructured flow solver, since it does not require lines

    across the entire domain for the solution algorithm. For these rea-

    sons we use the RKI-SGS scheme to solve the SA equation.

    4.1. Essentially subsonic flow

    For Case 1 we first consider the effect of the approximation

    order for the convective terms of the mean flow equations on the

    coarse grids in the multigrid method. Usually, only first-order

    accurate spatial discretization is used for these terms. In Fig. 3

    the effect on convergence behavior when using a second-order

    approximation is shown. Clearly there is a significant improvement

    in convergence with the second order, as the number of multigrid

    cycles to reduce the residual of the flow equations 13 orders is de-

    creased from 82 cycles to 65 cycles. The residual of the SA equation

    (using the RKI-SGS scheme) is reduced to almost the same level

    (exceeding nine orders) in both calculations. In all subsequent re-

    sults for subcritical flows the second-order approximation is used.

    Furthermore, to elimnate the possibility of convergence effects due

    to limiting, no limiter is applied.The effect of mesh refinement on convergence for Case 1 is

    shown Fig. 4. The finest mesh (1280 256) contains over

    300,000 cells. Similar convergence behavior is obtained on all grids

    for both the mean flow and turbulence equations. As revealed in

    Table 3 the convergence rate in solving the mean flow equations

    is approximately 0.6, and the CPU time is increased by about a fac-

    tor of four as the number of mesh points is doubled in each coor-

    dinate direction, indicating convergence without mesh

    Cycles

    Log(||Res||

    2)

    CL

    0 20 40 60 80 100 1 20-14

    -12

    -10

    -8

    -6

    -4

    -2

    0

    0.2

    0.4

    0.6

    0.8

    1

    DDADI

    CL

    SLGS

    CL

    RKI-SGS

    CL

    RK3/Implicit, SA Model, RAE 2822: Case 9

    M

    = 0.73, = 2.79o, Re = 6.5 x 10

    6

    Grid: 320 x 64

    Cycles

    Log(||Restur|

    |2)

    0 20 40 60 80 100 120-12

    -10

    -8

    -6

    -4

    -2

    0

    2

    4

    DDADI

    SLGS

    RKI-SGS

    SA Model, RAE 2822: Case 9, 320 x 64

    M

    = 0.73, = 2.79o, Re - 6.5 x 10

    6(a) (b)

    Fig. 2. Convergence histories for solvers of flow and turbulence equations (Case 9, grid: 320 64). SA equation solved with three different methods: DDADI, SLGS, RKI-SGS.(a) Flow equations, (b) SA equation.

    18 R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325

  • 8/6/2019 sdarticle(33)

    7/13

    dependency. In Table 4 a comparison is made of the computational

    efficiency of the current RK3/Implicit scheme (denoted by RK3/I)

    with the SA and BL turbulence models. In addition, the computa-

    tional effort required by a highly tuned standard five stage RK

    scheme (RK5/S) with three evaluations of numerical dissipative

    and physical diffusive terms, scalar implicit residual smoothing,

    and multigrid is given. The BL model was used when applying

    the RK5/S algorithm. With the BL model the RK3/Implicit scheme

    is about four times faster than the RK5/S scheme. Convergence his-

    tories with the BL model are given in Fig. 5. Even with the addi-

    tional computing time required by the SA model, the RK3/

    Implicit scheme is still about two times faster than the RK5/S

    scheme. The computing time with the SA model is increased by

    roughly a factor of 1.6 relative to that with the BL model.

    To provide an additional perspective on the efficiency of the

    RK/Implicit algorithm Table 4 also includes a comparison with

    the SLGS scheme when used to solve both the mean flow and SA

    Table 2

    Comparison for Case 9 of solution strategies for solving the turbulence equation of the

    SA model (grid: 320 64).

    Method CPU time (s) MG cycles

    DDADI 63 69

    SLGS 65 70

    RKI-SGS 75 69

    Cycles

    L

    og(||Res||

    2)

    CL

    0 20 40 60 80 100 120-14

    -12

    -10

    -8

    -6

    -4

    -2

    0

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    1st order

    CL

    2nd order

    CL

    RK3/Implicit, SA Model, RAE 2822: Case 1

    M

    = 0.676, = 1.93o, Re = 5.7 x 10

    6

    Grid:3 20 x 64

    Cycles

    Log(||Restur|

    |2)

    0 20 40 60 80 100 1 20-12

    -10

    -8

    -6

    -4

    -2

    0

    2

    4

    1st order

    2nd order

    RKI-SGS, SA Model, RAE 2822: Case 1

    M

    = 0.676, = 1.93o, Re - 5.7 x 10

    6

    Grid: 320 x 64

    (a) (b)

    Fig. 3. Effect on convergence of approximation order of convective terms in themean flowequations on coarse grids of themultigrid method (SAmodel, Case 1, grid density:

    320 64). (a) Flow equations, (b) SA equation.

    Cycles

    Log(||Res||

    2)

    CL

    0 20 40 60 80 100 120-14

    -12

    -10

    -8

    -6

    -4

    -2

    0

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    320 x 64

    CL

    640 x 128

    CL

    1280 x 256

    CL

    RK3/Implicit, SA Model, RAE 2822: Case 1

    M

    = 0.676, = 1.93o, Re = 5.7 x 10

    6

    Cycles

    Log(||Restur|

    |2)

    0 20 40 60 80 100 120-12

    -10

    -8

    -6

    -4

    -2

    0

    2

    4

    320 x 64

    640x 128

    1280 x 256

    RKI-SGS, SA Model, RAE 2822:Case 1

    M

    = 0.676, = 1.93o, Re - 5.7 x 10

    6(a) (b)

    Fig. 4. Convergence histories for solvers of flow and turbulence equations for Case 1 on three grids. SA equation solved with RKI-SGS. (a) Flow equations, (b) SA equation.

    Table 3

    Effect of mesh density on convergence of RK3/Implicit scheme (Case 1). SA model

    solved with RKI-SGS scheme.

    Mesh size CPU time (s) MG cycles Convergence rate

    320 64 71 65 0.629

    640 128 299 63 0.619

    1280 256 1242 60 0.604

    Table 4

    Comparison of computational efficiency of RK3/Implicit scheme with that of SLGS and

    tuned RK5/S schemes. Case 1 on the 320 64 grid.

    Scheme Turb. model CPU time (s) MG cycles Convergence rate

    RK3/I SA 71 65 0.629

    SLGS SA 152 344 0.917

    RK3/I BL 44 64 0.624

    SLGS BL 128 351 0.918

    RK5/S BL 181 1792 0.983

    R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325 19

  • 8/6/2019 sdarticle(33)

    8/13

    equations. As a multigrid smoother for the mean flow, the SLGSscheme, due to stability, required a reduced CFL of 100 and under-

    relaxation on the coarse grids. When solving the SA equation, the

    RK3/Implicit scheme requires less than half the computer time of

    the SLGS scheme.

    In Fig. 6 convergence plots with FMG, starting on a 160 32

    grid, are displayed. Four grids were used on all levels of grid refine-

    ment, except for the 1280 256 level, which used five grids. The

    benefit of the FMG in accelerating the convergence of a global

    quantity such as lift coefficient (CL) is evident. In Table 5 the com-

    puted CL and drag coefficient (CD) are presented for each grid level.

    The two contributions to the total drag coefficient, pressure drag

    (CD)p and skin-friction drag (CD)f coefficients, are also given. The

    development of these coefficients after three, five, and 10 multigrid

    cycles on each level are included in the table as well. In 10 cyclesthe CL and CD are obtained to at least four significant digits. Fur-

    thermore, with just three cycles on each level of the FMG the error

    in these quantities is less than 0.1%.

    Fig. 7 shows a comparison of the computed surface pressure

    and skin-friction distributions on the 1280 256 grid with exper-

    imental data. In general, there is very good agreement with the

    data. The computational pressure distribution does exhibit a weak

    shock on the upper surface of the airfoil in the transition region

    (11% chord location). There is insufficient data in the region to ver-

    ify this behavior.

    So far we have presented results for grids with moderately high

    aspect ratio cells. Fig. 8 shows the residual histories for Case 1

    when the Re number is varied by more than an order of magnitude

    (from 5.7 106 to 100 106). The SA equation was solved with the

    RKI-SGS scheme. Even at a Re = 100 106 a good convergence rate

    (0.751) is still obtained for the RK3/Implicit scheme. Despite a Rey-

    nolds number increase exceeding an order of magnitude, there is

    only a factor of about two increase in computational effort. A com-

    parison of the RK3/Implicit and RK5/S schemes (with the SA and BL

    models, respectively) reveals that the RK3/Implicit method is more

    than five times faster when Re = 100 106.

    4.2. Incompressible flow

    Since the numerical dissipation matrix of the present scheme is

    written as a function of Mach number (see Refs. [7,8]), the dissipa-

    tion can be scaled appropriately for low-speed flows. To demon-

    strate the effectiveness of the present algorithm at a low Mach

    number we consider an incompressible airfoil flow. Except for

    the free-stream Mach number of M1 = 0.001, the flow conditions

    are the same as for Case 9. Fig. 9 exhibits the residual histories.

    Here the density residual is decreased by only nine orders of mag-

    nitude to avoid round-off errors [7]. Removal of round-off errors at

    lowMach number canbe achieved by introducing a gauge pressure

    [30].

    4.3. Transonic flow

    In Fig. 10 the convergence histories on three grids is presented

    for Case 9. For these results the limiter was activated, and first-

    order differencing was used for coarse-grid convective terms. As

    for Case 1, similar convergence behavior is obtained on all grids.

    Cycles

    Log(||Res||

    2)

    CL

    0 20 40 60 80 100 1 20-14

    -12

    -10

    -8

    -6

    -4

    -2

    0

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    320 x 64

    CL

    640 x 128

    CL

    RK3/Implicit, BL Model, RAE 2822: Case 1

    M

    = 0.676, = 1.93o, Re = 5.7 x 10

    6

    Fig. 5. Convergence histories with BL model for Case 1. Second-order approxima-

    tion of convective terms on coarse grids.

    Cycles

    Log(||Res||

    2)

    CL

    0 50 100 150 200

    -14

    -12

    -10

    -8

    -6

    -4

    -2

    0

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    Residual

    CL

    RK3/Implicit, SA Model, RAE 2822: Case 1

    160 x 32 320 x 64 640 x 128 1280 x 256

    Cycles

    Log(||Restur|

    |2)

    0 50 100 150 200-12

    -10

    -8

    -6

    -4

    -2

    0

    2

    4RKI-SGS, SA Model, RAE 2822:Case 1

    160 x 32 320 x 64 640 x 128 1280 x 256

    (a) (b)

    Fig. 6. Convergence histories with FMG for solvers of flow and turbulence equations (Case 1). SA equation solved with RKI-SGS. (a) Flow equations, (b) SA equation.

    Table 5

    Effect of mesh density on computed lift and drag coefficients for Case 1.

    Mesh size Cycles (FMG) CL CD (CD)p (CD)f

    160 32 50 0.5840 0.010030 0.004012 0.006023

    320 64 50 0.5915 0.008541 0.002669 0.005872

    640 128 50 0.5903 0.008298 0.002497 0.005802

    1280 256 50 0.5884 0.008261 0.002484 0.005777

    640 128 3 0.5900 0.008300 0.002501 0.005798

    640 128 5 0.5901 0.008303 0.002503 0.005800640 128 10 0.5903 0.008298 0.002497 0.005802

    1280 256 3 0.5887 0.008254 0.002484 0.005770

    1280 256 5 0.5885 0.008254 0.002480 0.005775

    1280 256 10 0.5884 0.008261 0.002484 0.005777

    20 R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325

  • 8/6/2019 sdarticle(33)

    9/13

    The rate of convergence on the three grids is between 0.63 and0.65. From Table 6 we see that the increase in CPU time in going

    from the 640 128 grid to the 1280 256grid is slightly greater thana factor of four, which suggests a weak dependence of convergence

    x/c

    Cp

    0 0.2 0.4 0.6 0.8 1

    -1.5

    -1

    -0.5

    0

    0.5

    1

    Exp.

    1280 x 256

    RK3/Implicit, SA Model, Case 1

    x/c

    Cf

    0 0.2 0.4 0.6 0.8 1-0.002

    0

    0.002

    0.004

    0.006

    0.008

    Exp.

    1280 x 256

    RK3/Implicit, SA Model, Case 1(a) (b)

    Fig. 7. Comparison of computed surface pressures and skin friction with experimental data (Case 1, grid: 1280 256). (a) Surface pressures, (b) surface skin friction.

    Cycles

    Log(||Res||

    2)

    0 20 40 60 80 100 1 20-14

    -12

    -10

    -8

    -6

    -4

    -2

    0RK3/Implicit, SA Model, RAE 2822: Case 1

    M

    = 0.676, = 1.93 o, Re = 5.7 x 10 6

    5 .7 m 2 0 m 5 7 m 1 00 m

    Grid:368 x 88

    Cycles

    Log(||Restur|

    |2)

    0 20 40 60 80 100 1 20-12

    -10

    -8

    -6

    -4

    -2

    0

    2

    4

    5.7 x 106

    20 x 106

    57 x 106

    100 x 106

    RKI-SGS, SA Model, RAE 2822: Case 1

    M = 0.676, = 1.93o, Re = 5.7 x 106(a) (b)

    Fig. 8. Effect of Reynolds number variation on convergence of solvers for RANS and turbulence equations (Case 1, grid: 368 88). RKI-SGS scheme used to solve SA equation.

    (a) Flow equations, (b) SA equation.

    Cycles

    Log(||Res||

    2)

    CL

    0 20 40 60 80-10

    -8

    -6

    -4

    -2

    0

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    320 x 64

    CL

    640 x 128C

    L

    RK3/Implicit, SA Model, Incomp.

    M

    = 0.001, = 2.79o, Re = 6.5 x 10

    6

    Cycles

    Log(||

    Restur|

    |2)

    0 20 40 60 80-10

    -8

    -6

    -4

    -2

    0

    2

    4

    320 x 64

    640 x 128

    RKI-SGS, SA Model, Incomp.

    M

    = 0.001, = 2.79o, Re - 6.5 x 10

    6(a) (b)

    Fig. 9. Convergence histories for solvers of flow and turbulence equations for incompressible case on two grids. SA equation solved with RKI-SGS. (a) Flow equations, (b) SA

    equation.

    R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325 21

  • 8/6/2019 sdarticle(33)

    10/13

    on mesh density. The computational efficiency of the RK3/Implicit

    scheme with both the SA and BL models is given in Table 7.

    For this case the computer time of the RK3/Implicit scheme is

    approximately a factor of 3.5 smaller than that of the SLGS scheme.

    Again, the RK3/Implicit scheme with the SA model is about two

    times faster than the standard scheme RK5/S with the BL model.

    The convergence of the RK3/Implicit scheme with the BL model

    is similar to that obtained with the SA model, as revealed in the

    convergence plots of Fig. 11.

    The convergence behavior with FMG for Case 9 is displayed in

    Fig. 12. As in Case 1, we observe a rapid evolution of the CL. Table

    8 gives the CL and CD after three, five, 10, and 50 multigrid cycles.

    Even for this transonic case these coefficients are obtained to four

    Cycles

    Log(||Res||

    2)

    CL

    0 20 40 60 80 100 120-14

    -12

    -10

    -8

    -6

    -4

    -2

    0

    0.2

    0.4

    0.6

    0.8

    1

    320 x 64

    CL

    640 x 128

    CL

    1280 x 256

    CL

    RK3/Implicit, SA Model, RAE 2822: Case 9

    M

    = 0.730, = 2.79o, Re = 6.5 x 10

    6

    Cycles

    Log(||Restur|

    |2)

    0 20 40 60 80 100 120-12

    -10

    -8

    -6

    -4

    -2

    0

    2

    4

    320 x 64

    640 x 128

    1280 x 256

    RKI-SGS, SA Model RAE 2822: Case 9

    M

    = 0.730, = 2.79o, Re - 6.5 x 10

    6(a) (b)

    Fig. 10. Convergence histories for solvers of flow and turbulence equations for Case 9 on three grids. SA equation solved with RKI-SGS. (a) Flow equations, (b) SA equation.

    Table 6

    Effect of mesh density on convergence of RK3/Implicit scheme (Case 9). SA model

    solved with RKI-SGS scheme.

    Mesh size CPU time (s) MG cycles Convergence rate

    320 64 75 69 0.648

    640 128 308 64 0.626

    1280 256 1307 66 0.632

    Table 7

    Comparison of computational efficiency of RK3/Implicit scheme with that of SLGS and

    tuned RK5/S schemes. Case 9 on 320 64 grid.

    Scheme Turb. model CPU time (s) MG cycles Convergence rate

    RK3/I SA 75 69 0.648

    SLGS SA 268 632 0.954

    RK3/I BL 44 62 0.616

    SLGS BL 211 599 0.951

    RK5/S BL 191 1891 0.984

    Cycles

    Log(||Res||

    2)

    CL

    0 20 40 60 80 100 1 20-14

    -12

    -10

    -8

    -6

    -4

    -2

    0

    0.2

    0.4

    0.6

    0.8

    1

    320 x 64

    CL

    640 x 128

    CL

    RK3/Implicit, BL Model, RAE 2822: Case 9

    M

    = 0.73, = 2.79o, Re = 6.5 x 10

    6

    Fig. 11. Convergence history for Case 9 using the BL model.

    Cycles

    Log(||Res||

    2)

    CL

    0 50 100 150 200

    -14

    -12

    -10

    -8

    -6

    -4

    -2

    0

    0.2

    0.4

    0.6

    0.8

    1

    Residual

    CL

    RK3/Implicit, SA Model, RAE 2822: Case 9

    160 x 32 320 x 64 640 x 128 1280 x 256

    Cycles

    Log(||Restur|

    |2)

    0 50 100 150 200-12

    -10

    -8

    -6

    -4

    -2

    0

    2

    4RKI-SGS, SA Model, RAE 2822: Case 9

    160 x 32 320 x 64 640 x 128 1280 x 256

    (a) (b)

    Fig. 12. Convergence histories with FMG for solvers of flow and turbulence equations (Case 9). SA equation solved with RKI-SGS. (a) Flow equations, (b) SA equation.

    22 R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325

  • 8/6/2019 sdarticle(33)

    11/13

    significant digits in just 10 cycles. After five cycles the computed CLand CD, on both the 640 128 and 1280 256 grids, are obtained

    to within about 0.25% of their final values. On the finest grid with

    three cycles on each refinement level the coefficients have an error

    of less than 0.2%. In Fig. 13 a comparison is made of the calculated

    surface pressure and skin-friction variations on the finest grid at

    three, five, and 10 cycles. With just three cycles on each level there

    are bearly discernible differences on the upper airfoil surface and

    at the shock. The distributions on the finest grid are compared with

    the experimental data in Fig. 14. There is fairly good agreement

    with the data.

    For all the computations the local line solves of the RKI-SGS

    scheme were terminated at the jl/4 location, where jl is the number

    of cells in the normal direction to the airfoil. In Fig. 15 the effect of

    varying the number of points in the line solves is shown. The con-

    vergence is only slightly faster by doubling the number of points in

    the normal direction.

    5. Concluding remarks

    In this work the fluid dynamic (RANS) equations and the trans-

    port-type equation of the SA turbulence model have been solved in

    a loosely coupled manner. The RANS equations have been solved

    with a RK3/Implicit scheme (RK3/I) and multigrid. This scheme

    has been enhanced by weighting the numerical dissipative and

    physical diffusive terms, a smooth initial increase of the CFL num-

    ber, and local line implicit relaxation. Three different methods have

    been considered for solving the SA equation: diagonally dominant

    alternating direction implicit (DDADI), symmetric line Gauss-Seidel

    (SLGS), and a RK2/Implicit with local line solves (RKI-SGS). To

    enhance efficiency and robustness of these schemes multigrid

    acceleration has also been applied. For both the fluid dynamic

    and turbulence equations a CFL of 1000 has been used.

    Similar convergence behavior has been observed for the three

    schemes evaluated for solving the turbulence model equation.

    Although the computational effort required with the RKI-SGS

    scheme is somewhat larger than that needed with the other two

    schemes, it provides important advantages. With appropriate

    ordering for Gauss-Seidel, this method can be implemented in an

    unstructured grid algorithm. In addition, it allows for the possibil-

    ity to solve the mean flow and turbulence equations in a fully cou-

    pled manner. Thus, the present loosely coupled algorithm is based

    on RK/Implicit schemes. This algorithm also has the advantage that

    it can be readily incorporated into many existing codes that

    employ RK smoothers for multigrid methods.

    The performance of the loosely coupled algorithm (RK3/I +

    RKI-SGS schemes) has been investigated by computing solutions

    to subsonic and transonic airfoil flows. We have demonstrated that

    there is no significant slowdown in convergence of the RK/Implicit

    scheme when the SA model is used instead of the algebraic model

    of Baldwin and Lomax. This is quite important since it suggests

    that for at least similar 3-D problems, such as wing flows, the

    performance of the 3-D scheme for the SA model will be similar

    to that observed for the BL model. In addition, even with the SA

    model, the loosely coupled algorithm is approximately two to five

    times faster, depending on the Reynolds number, than the highly

    tuned standard RK scheme (RK5/S) with the BL model. It should

    be emphasized that the RK5/S scheme includes three evaluations

    of the dissipative and diffusive terms, multigrid, and scalar implicit

    residual smoothing. The RK/Implicit algorithm has also been com-

    pared to the SLGS scheme when applied to both the mean flow and

    SA equations. It is between two and 3.5 times faster than the SLGS

    scheme, depending on the flow conditions.

    Although the indirectly coupled algorithm uses local line solves

    (in boundary layer and wake) rather than line solves extending

    across the entire domain, there is no significant deterioration in

    convergence. The RK/Implicit schemes applied to the mean flowand turbulence equations have exhibited a low sensitivity to dis-

    crete stiffness associated with large aspect ratio mesh cells. Fur-

    thermore, it has been shown that the algorithm can also

    effectively solve a low-speed flow; and thus, the analytical stiffness

    due to disparity in wave speeds has been removed.

    By using FMG to generate the initial conditions on the solution

    grid, we have observed rapid development of the aerodynamic

    Table 8

    Effect of mesh density on computed lift and drag coefficients for Case 9.

    Mesh size Cycles (FMG) CL CD (CD)p (CD)f

    160 32 50 0.7955 0.01748 0.01192 0.005563

    320 64 50 0.8185 0.01678 0.01120 0.005576

    640 128 50 0.8227 0.01665 0.01113 0.005510

    1280 256 50 0.8238 0.01655 0.01108 0.005474

    640 128 3 0.8152 0.01650 0.01103 0.005471

    640 128 5 0.8213 0.01661 0.01111 0.005500640 128 10 0.8227 0.01665 0.01113 0.005510

    1280 256 3 0.8222 0.01654 0.01107 0.005471

    1280 256 5 0.8235 0.01655 0.01108 0.005474

    1280 256 10 0.8238 0.01655 0.01108 0.005474

    x/c

    Cp

    0 0.2 0.4 0.6 0.8 1

    -1.5

    -1

    -0.5

    0

    0.5

    1

    3 cycles

    5 cycles

    50 cycles

    RK3/Implicit, SA Model, Case 9, 1280 x 256

    x/c

    Cf

    0 0.2 0.4 0.6 0.8 1-0.002

    0

    0.002

    0.004

    0.006

    0.008

    3 cycles5 cycles

    50 cycles

    RK3/Implicit, SA Model, Case 9, 1280 x 256(a) (b)

    Fig. 13. Effect of number of cycles in each level of FMG on computed surface pressures and skin friction (Case 9, grid: 1280 256). (a) Surface pressures, (b) Surface skinfriction.

    R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325 23

  • 8/6/2019 sdarticle(33)

    12/13

    coefficients for turbulent viscous flows at both subsonic and tran-

    sonic speeds. Moreover, in just three cycles on each refinement le-

    vel of the FMG, the lift and drag coefficents for both the subsonic

    and transonic cases have an error less than 1.0%.

    References

    [1] Dick E. A flux-vector splitting method for steady NavierStokes equations. Int J

    Numer Methods Fluids 1988;8:31726.

    [2] Wackers J, Koren B. Multigrid solution method for thesteady RANS equations. J

    Comput Phys 2007;226:1784807.

    [3] Mavriplis DJ, Venkatakrishnan V. Agglomeration multigrid for twodimensional viscous flows. Comput Fluids 1995;24(5):55370.

    [4] Pierce NA, Giles MB. Preconditioned multigrid methods for compressible flow

    calculations on stretched meshes. J Comput Phys 1997;136:42545.

    [5] Jespersen DC, Pulliam TH, Buning PG. Recent enhancements to overflow. AIAA

    paper 97-0644; January 1997.

    [6] Venkatakrishnan V. Improved convergence of compressible NavierStokes

    solvers. AIAA paper 98-2967; June 1998.

    [7] Rossow C-C. Efficient computation of compressible and incompressible flows. J

    Comput Phys 2007;220:87999.

    [8] Swanson RC, Turkel E, Rossow C-C. Convergence acceleration of Runge-Kutta

    schemes for solving the NavierStokes equations. J Comput Phys

    2007;224:36588.

    [9] Baldwin BS, Lomax H. Thin layer approximation and algebraic model for

    separated flows. AIAA paper 78-257; January 1978.

    [10] Spalart PR, Allmaras SR. A one-equation turbulence model for aerodynamic

    flows. La Recherche Aerospatiale 1994;1:521.

    [11] Krist SL, Biedron RT, Rumsey CL.CFL3D users manual. NASA TM 1998-208444;

    June 1998.

    [12] Rumsey CL. Apparent transition behavior of widely-used turbulence models.Int J Heat Fluid Flow 2007;28:146071.

    [13] Roe PL. Approximate Riemann solvers, parameter vectors and difference

    schemes. J Comput Phys 1981;43:35772.

    [14] Van Leer B, Tai C-H, Powell KG. Design of optimally smoothing multi-stage

    schemes for the Euler equations. AIAA paper 89-1933; June 1989.

    [15] Jameson A. The evolution of computational methods in aerodynamics. J Appl

    Mech 1983;50(4b):105270.

    [16] Swanson RC, Rossow C-C. An initial investigation of the effects of turbulence

    models on the convergence of the RK/Implicit scheme. NASATM-2008-

    215342; 2008.

    [17] Jameson A. Numerical solution of the Euler equation for compressible inviscid

    fluids. MAE report 1643, Princeton University, Princeton, New Jersey;

    December 1983.

    [18] Swanson RC, Turkel E. Multistage schemes with multigrid for Euler and

    NavierStokes equations. NASA TP 3631; 1997.

    [19] Trottenberg U, Oosterlee CW, Schller A. Multigrid. Academic Press; 2001.[20] Roberts TW, Swanson RC. A study of multigrid preconditioners using

    eigensystem analysis. AIAA paper 2005-5229; June 2005.

    [21] Adams M, Brezina M, Hu J, Tuminaro R. Parallel multigrid smoothing:

    polynomial versus Gauss-Seidel. J Comput Phys 2003;188(2):593610.

    [22] Chow E, Falgout RD, Hu JJ, Tuminaro RS, Yang UM. A survey of parallelization

    techniques for multigrid solvers. In: Heroux MA, Raghaven P, Simon HD,

    editors. Parallel Processing for Scientific Computing. SIAM; 2006. p. 179201

    [chapter 10].

    [23] Klopfer GH, Van der Wijngaart RF, Hung CM, Onufer JT. A diagonalized

    diagonal dominant alternating direction implicit (D3ADI) scheme and

    subiteration correction. AIAA paper 98-2824; June 1998.

    [24] MacCormack RW, Pulliam TH. Assessment of a new numerical procedure for

    fluid dynamics. AIAA paper 98-2821; June 1998.

    [25] Walsh PC, Pulliam T. The effect of turbulence model solution on viscous flow

    problems. AIAA paper 2001-1018; January 2001.

    [26] Pulliam TH, MacCormack RW, Venkateswaran S. Convergence characteristics

    of approximate factorization methods. In: Sixteenth international conference

    on numerical methods in fluid dynamics. Lecture notes in physics, vol.515. Berlin/Heidelberg: Springer; 1998. p. 40914.

    x/c

    Cp

    0 0.2 0.4 0.6 0.8 1

    -1.5

    -1

    -0.5

    0

    0.5

    1

    Exp.

    1280 x 256

    RK3/Implicit, SA Model, Case 9

    x/c

    Cf

    0 0.2 0.4 0.6 0.8 1-0.002

    0

    0.002

    0.004

    0.006

    0.008

    Exp.

    1280 x 256

    RK3/Implicit, SA Model, Case 9(a) (b)

    Fig. 14. Comparison of computed surface pressures and skin friction with experimental data. (Case 9, grid: 1280 256). (a) Surface pressures, (b) Surface skin friction.

    Cycles

    Log(||Res||

    2)

    CL

    0 20 40 60 80 100 1 20-14

    -12

    -10

    -8

    -6

    -4

    -2

    0

    0.2

    0.4

    0.6

    0.8

    1

    jl/4

    CL

    jl/2

    CL

    jl

    CL

    RK3/Implicit, SA Model, RAE 2822: Case 9

    M

    = 0.73, = 2.79o, Re = 6.5 x 10

    6

    Cycles

    Log(||Restur

    ||2

    )

    0 20 40 60 80 100 1 20-12

    -10

    -8

    -6

    -4

    -2

    0

    2

    4

    jl/4

    jl/2

    jl

    RKI-SGS, SA Model, RAE 2822:Case 9

    M

    = 0.73, = 2.79o, Re - 6.5 x 10

    6(a) (b)

    Fig. 15. Effect of number of points in line solves on convergence (Case 9, grid: 640 128). (a) Flow equations, (b) SA equation.

    24 R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325

  • 8/6/2019 sdarticle(33)

    13/13

    [27] Allmaras SR. Multigrid for the 2-D compressible NavierStokes equations.

    AIAA paper 99-3336; July 1999.

    [28] Cook PH, McDonald MA, Firmin MCP. Aerofoil RAE 2822 pressure

    distributions and boundary layer and wake measurements. AGARD-AR-

    138; 1979.

    [29] Fabender J. Improved robustness of numerical simulation of turbulent flows

    around civil transport aircraft at flight Reynolds numbers. Ph.D. thesis,

    Technical University of Braunschweig, Braunschweig, Germany; 2004.

    [30] Feng J, Merkle CL. Evaluation of preconditioning methods for time-marching

    systems. AIAA paper 90-0016; January 1990.

    R.C. Swanson, C.-C. Rossow / Computers & Fluids 42 (2011) 1325 25