Comparison Les Rans Bluff-body

download Comparison Les Rans Bluff-body

of 15

Transcript of Comparison Les Rans Bluff-body

  • 8/3/2019 Comparison Les Rans Bluff-body

    1/15

    Journal of Wind Engineering

    and Industrial Aerodynamics 89 (2001) 14711485

    Comparison of LES and RANS in

    bluff-body flows

    H. L .ubcke*, St. Schmidt, T. Rung, F. Thiele

    Hermann-F.ottinger-Institut f.ur Str .omungsmechanik, TU Berlin, M.uller-Breslau-Strasse 8,

    D-10623 Berlin 10589, Germany

    Abstract

    The turbulent flow around bluff-bodies features a variety of complex phenomena, e.g.

    streamline curvature, separation and the formation of large unsteady vortical structures. In

    particular, an accurate representation of the interaction between mean transient motion and

    residual turbulence poses a challenge to numerical simulation procedures. In virtually all

    commercial simulation packages, the representation of turbulence relies on Reynolds-averagedNavierStokes (RANS) equations in conjunction with Boussinesq-viscosity models (BVM).

    Various studies have demonstrated the inability of established RANS methodologies to render

    the fundamental physics when applied to transient flows. In contrast to this, the

    computationally more demanding large eddy simulation (LES) is known to be a viable

    approach to simulate unsteady turbulent flows. The present study aims to assess the predictive

    prospects of advanced recent RANS practices, i.e. explicit algebraic stress models (EASM), in

    unsteady bluff-body flows. Results are reported in comparison to conventional BVM, LES

    and measurements. Examples included refer to three different cylinder flows, which indicate

    that the predictive accuracy obtained from an EASM is in close proximity to LES

    results, whereas the computational surplus remains moderate in comparison with a linear

    BVM. r 2001 Published by Elsevier Science Ltd.

    1. Introduction

    Due to the progress in computer technology, computational fluid dynamics (CFD)

    is now able to tackle industrial flow problems at moderate costs and time-to-

    solution. The prospect and success of CFD will therefore depend on the accuracy of

    the approach, in particular the predictive realms of the employed physical models. A

    *Corresponding author. Tel.: +49-30-314-26283; fax: +49-30-314-21101.

    E-mail address: [email protected] (H. L .ubcke).

    0167-6105/01/$ - see front matter r 2001 Published by Elsevier Science Ltd.

    P I I : S 0 1 6 7 - 6 1 0 5 ( 0 1 ) 0 0 1 3 4 - 9

  • 8/3/2019 Comparison Les Rans Bluff-body

    2/15

    specifically challenging example of industrial relevance is the simulation of

    unsteady turbulent flows around bluff-bodies, e.g. vehicles, buildings or

    flame stabilizers. The principal phenomenon of bluff-body flows, i.e. the unsteady

    shedding of vortical structures from the obstacle, hinges on the interactionbetween mean-transient motion and residual turbulence. Hence, an appropriate

    treatment of turbulence is crucial for the simulation quality. When attention is

    drawn to industrial applications, a direct resolution of the turbulent motion is,

    however, unfeasible as the associated grid resolution would cause prohibitive

    computational expenses. The simulation of engineering turbulent flows is thus

    restricted to turbulence-modeling practices based on Reynolds-averaged Navier

    Stokes (RANS) equations and large eddy simulation (LES). In the framework of

    RANS, all aspects of turbulence are modeled, which enhances the numerical

    efficiency at the expense of a strong model dependency. As opposed to the RANS

    approach, a major portion of the turbulent scales is numerically resolved within

    LES. The primary advantage of LES is the reduced influence of the turbulence

    model, which significantly increases the computational effort in comparison to

    RANS. Recently developed explicit algebraic stress models revealed remarkable

    improvements of the RANS methodology for several steady flows at low

    computational costs, which motivates their investigation in unsteady turbulent

    bluff-body flows.

    2. Mathematical model

    The present study is confined to incompressible and isothermal flows, thus the flow

    field is fully determined by the conservation of mass and momentum. The influence

    of turbulence is either entirely modeled by means of a statistical turbulence model

    (RANS), or partially resolved (LES) with only the small, dissipative scales being

    closed by a subgrid-scale model. Since only parts of the energy spectrum are

    numerically resolved, it is necessary to ensemble average (RANS) or filter (LES) the

    governing equations. In both cases the resulting equation system can be cast into a

    common form based on a Cartesian coordinates xi

    q %ui

    qxi 0; 1

    D %ui

    Dt

    1

    r

    q %p

    qxiq %tij %Tij qxj

    ; 2

    where D=Dt; %ui; %p; r denote the material derivative and the mean-average or filteredvelocity, pressure and density. The viscous stresses are given by %tij 2nSij; where

    Sij q %ui=qxj q %uj=qxi =2 represents the strain-rate tensor. All unresolved effectsof turbulence are absorbed into %Tij: The closure models %Tij are described in theremainder of this section.

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 147114851472

  • 8/3/2019 Comparison Les Rans Bluff-body

    3/15

    2.1. Rans

    Due to the ensemble averaging process, further unknowns are introduced to the

    momentum equations by means of the Reynolds stresses %Tij u0ju0j: In industrialsimulations, two-equation Boussinesq-viscosity models (BVM) for u0ju

    0j have gained

    the most popularity. Employing the Boussinesq-viscosity hypothesis, the Reynolds

    stresses are modeled in analogy to the viscous stresses

    u0iu0

    j 23kdij 2ntSij: 3

    In Eq. (3), k u0iu0i=2 denotes the turbulent kinetic energy, nt marks the turbulent

    viscosity and dij is the Kronecker delta. The present study is confined to the standard

    k2o model of Wilcox [1], which employs two additional transport equations for k

    and the specific dissipation rate o; viz.

    nt k

    o; with o e= cmk

    ;

    Dk

    Dt P cmok

    q n nt=2

    qk=qxj

    qxj; 4

    Do

    Dt a

    o

    kP bo2

    q n nt=2

    qo=qxj

    qxj:

    Here, P %Tijq %ui=qxj denotes the production of turbulent energy and e is the energydissipation rate. The coefficients a; b and cm employ heir standard values 0.55, 0.075and 0.09, respectively. Despite their numerical advantages, BVM fail to represent the

    complex interaction mechanisms between Reynolds-stresses and mean velocity field.

    As a well-known example, the conventional BVM (3) fails to mimic effects related to

    streamline curvature or secondary motion. Although a general approach would,

    arguably, be based on a full differential-stress model for u0iu0

    j; such models are hardlyused in industry due to their increased numerical effort and reduced robustness. To

    remedy the shortcomings of the linear BVM, Gatski and Speziale [2] derived an

    explicit solution for the algebraic stress model of Rodi [3], that features most benefitsof the differential stress model while retaining the numerical advantages

    of the BVM. The derivation of the explicit algebraic stress model (EASM) stems

    from the transport equation for the Reynolds stress-anisotropy tensor bij

    u0iu0

    j=2k dij=3 assuming an isotropic representation of the diffusion process, viz.

    Dbij

    Dt bij P e =k

    2

    3Sij

    1

    2kFij bjlSli bilSlj

    2

    3bmlSmldij

    bjlWli bilWlj

    : 5

    Here, Fij is the pressurestrain correlation and Wij q %ui=qxj q %uj=qxi =2 therotation-rate tensor. Employing the structural equilibrium assumption Dbij=Dt 0;Rung et al. [4] devised an EASM along the route suggested by Gatski and

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 14711485 1473

  • 8/3/2019 Comparison Les Rans Bluff-body

    4/15

    Speziale, viz.

    u0iu0

    j 2

    3

    kdij 2cnmk

    cmo

    Sij b2k

    e

    SikWkj SjkWki b3k

    e

    SikSkj S2kkdij=3 !

    6

    with cnm b1= 1 2Z2=3 2x2

    and

    Z2 b3S 2

    =8; x2 b2O 2

    =2;

    S k=e ffiffiffiffiffiffiffiffi

    2S2jj

    q; O k=e

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffi2W2jj

    q;

    b1 2=3 C2=2

    =g; b2 1 C4=2

    =g;

    b3 2 C3 =g; g fgC1 1 C ;

    C 0:421 0:526S2= 4 1:83ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

    0:2S2 0:8O2p

    fg 1 0:95 1 tanh 0:465S 2

    The coefficients of the employed pressurestrain correlation model are listed in

    Table 1. In contrast to the Gatski and Speziale approach, the present model does not

    require a regularization, due to an improved representation of the production-to-

    dissipation ratio CEP=e within the stressstrain relation. The unregularized EASM

    satisfies the positivity requirement for the eddyviscosity and obeys the realizabilityprinciple.

    The structural equilibrium assumption is at the root of almost any EASM. It is

    reasonably accurate in flows which involve only weak streamline curvature. Girimaji

    [5] pointed out, that the predictive response of the EASM to strong streamline

    curvature can be improved when the structural-equilibrium hypothesis is applied to

    the coordinates of the stress-anisotropy tensor in a mean-acceleration system ba

    ij

    rather than a Cartesian system (cf. Eq. (5)), viz.

    Dbij

    Dt

    D ba

    ij%

    ea

    i%

    ea

    j

    h iDt

    E ba

    ij

    D%

    ea

    i%

    ea

    j

    Dt

    : 7

    Restricting our interest to the convective part of Eq. (7), all curvature related

    modifications of the EASM can be absorbed into the definition of an effective

    vorticity tensor [5]

    #Wij Wij 2

    2 C4Tri%ul

    qTrj

    qxl

    !; 8

    Table 1

    Coefficients of the employed EASM

    C1 C2 C3 C4

    2.5 0.39 1.25 0.45

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 147114851474

  • 8/3/2019 Comparison Les Rans Bluff-body

    5/15

    which replaces the rotation-rate tensor in Eq. (6). In Eq. (8) Tij denotes the

    transformation matrix between the base vectors%

    ea

    i of the acceleration system

    and the Cartesian base%

    ei; that is traditionally used for the fully conservative

    numerical formulation, i.e.%e a i

    %e a j TipTjp

    %ep

    %eq: The determination of the accelera-

    tion system and its derivatives with respect to Cartesian coordinates is a non-trivial

    task, particularly when focusing upon general three-dimensional flows. In the

    considered two-dimensional mean flows, the acceleration vector is always

    perpendicular to the homogeneous direction, which significantly simplifies the

    procedure.

    2.2. LES

    In contrast to the RANS methodology, the spatial filtering of the instantaneousequations employed in LES explicitly considers the three-dimensional,

    unsteady character of turbulent motion. Turbulent scales which exceed the

    utilized filter width %D are resolved by the numerical procedure. Smaller

    scales, which are assumed to behave in a universal-isotropic way, are attributed

    to a subgrid-scale (sgs) model. The present study employs the dynamic one-

    equation model of Davidson [6] to approximate the unknown subgrid stress

    tensor

    %Tij tsgs

    ij

    2C%D ffiffiffiffiffiffiffiffiksgsp %Sij 2nt %Sij: 9Quantities of influence to the modeled stresses are the local grid-resolution quality%D DxDyDz 1=3 and the subgrid turbulent kinetic energy ksgs; which is determinedby its transport equation [6]

    Dksgs

    Dt Psgs C

    nksgs%D

    q n nt qksgs=qxj

    qxj; 10

    where Psgs tsgsij q %ui=qxj denotes the production of subgrid energy. The model

    parameters C LijMij=2MijMij and Cn Psts gPsgsPsgs Cngk1:5sgsk1:5sgs= %Dh i *%DK1:5sts are

    evaluated in a dynamic procedure, based on an explicit test-filter (sts) which is set

    twice the grid size *%D 2 %D: The filtered Leonard-stresses are evaluated via Lij g%ui%uj%ui%uj *%ui*%uj: The kinetic energy of the test-filter level Ksts is computed via Ksts gksgsksgs Lii=2: A scale similarity tensor Mij *%D gffiffiffiffiffiffiffiffiKstspffiffiffiffiffiffiffiffiKstsp *%Sij %D gffiffiffiffiffiffiffiffiffiffiffiffiffiksgsSijpffiffiffiffiffiffiffiffiffiffiffiffiffiksgsSijp supplementsthe basis of the dynamic model.

    The model is able to capture different flow phenomena, such as backscattering and

    relaminarization, by the adaption of the model parameters Cand C to the local flow

    structure. The benefit of David-sons dynamic one-equation approach is itsinsensitivity to the existence of homogeneous directions, which recommends the

    model for complex flow geometries.

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 14711485 1475

  • 8/3/2019 Comparison Les Rans Bluff-body

    6/15

    3. Numerical method

    To preclude the effects of different numerical algorithms, all simulations were

    performed with the same implicit finite-volume NavierStokes procedure of Xue [7].The general accuracy of the procedure is of second-order in time and space. The

    algorithm is based on the strong conservation form within general body-fitted

    coordinates, and employs a fully co-located storage arrangement for all transport

    properties. Diffusion terms are approximated using either second or fourth-order

    accurate central differences, whereas advective fluxes are approximated by central

    differencing or third-order bounded (monotonic) upwind-biased schemes. The odd

    even decoupling problem of the cell-centered formulation is suppressed with a

    fourth-order artificial pressure term in the continuity equation as introduced by Rhie

    and Chow [8]. The solution is iterated to convergence using a SIMPLE pressure-

    correction scheme. The procedure is parallelized by means of domain decomposition

    technique. All RANS simulations were performed on a PC-cluster, whereas the LES

    were carried out on the CRAY T3E of Berlins supercomputing center.

    4. Results

    To assess the influence of the turbulence-representation practice, three different

    cylinder flows have been investigated in the present study. The selected test cases

    refer to the flow around a square (Re=22 000) and a circular cylinder (Re=3900,Re=140 000), which feature geometry-fixed and pressure-induced separation,

    impinging flow and streamline curvature.

    4.1. Square cylinder at Re=22 000

    The experimental study of the flow around a square cylinder at Re=22 000

    conducted by Lyn et al. [9], is a well-known example for the validation of numerical

    simulation techniques (see e.g. [10] or [11]). A comprehensive overview of this case

    related to LES is given by Voke [12]. The computational grid employed in the LES

    mode consists of 32 grid nodes in the homogeneous (z-) direction and 32 304 gridnodes in the main-flow (xy-) plane. The physical domain covers 20 diameters in

    streamwise direction (5ox/Do15), 14 diameters in normal direction (7oy/

    Do7) and four diameters in the homogeneous direction. RANS computations were

    restricted to two-dimensional flow conditions, which were carried out on the xy-

    portion of grid. The resolution of the cylinder edges in the xy-plane was based on 30

    grid nodes. Approximately 125 70 grid nodes were used to resolve the wake regime

    aft of the cylinder. A uniform velocity profile with a turbulence level of 2% and an

    eddyviscosity ratio of nt=n or L=D{1 was prescribed at the inlet, which is in linewith recommendations of former studies [11]. Fig. 1 depicts the phase-averaged

    streamlines for four selected phases, where each period was split into 20 phases. Thestrong influence of the turbulence-modeling practice is evident at first glance. Both,

    the EASM and LES simulations return approximately the same topological features.

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 147114851476

  • 8/3/2019 Comparison Les Rans Bluff-body

    7/15

    The overprediction of the recirculation length Lr=D; as indicated by Table 2, revealsthat the EASM slightly underpredicts the intensity of the transient motion in

    comparison to the experiments (cf. CL; rms). In contrast to this, the LES, returns anaugmented intensity of the transient motion, which is confirmed by an under-

    estimation of the recirculation length. When attention is drawn to the BVM,

    Fig. 1. Phase-averaged streamlines.

    Table 2

    Global parameters for square cylinder

    CD Lr=D CL;rms St

    Exp. 2.1 1.38 F 0.13

    LES 2.178 1.06 1.47 0.13

    BVM 1.68 2.12 0.03 0.12

    EASM 2.206 1.64 0.95 0.15

    LESa 2.3 1.46 1.15 0.13

    BVMb 2.11 F F 0.15

    aLES is taken from Breuer [10].bBVM from [11].

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 14711485 1477

  • 8/3/2019 Comparison Les Rans Bluff-body

    8/15

    however, a completely different picture emerges. The recirculation zone becomes

    larger and the wake is much more stretched, which is in line with a significant

    damping of the transient motion. The BVM fails to mimic the dynamics of the flow,

    which is indicated by the reduced root-mean-square value of the lift coefficient

    (CL; rms).

    In conclusion, the EASM offers a reasonable predictive response to unsteady flow

    conditions, which is in close proximity to LES results and the traditional BVM fails

    to capture the transient phenomenon. Fig. 2 provides an explanation for the poorperformance of the BVM. Distinct from the EASM and LES simulations, the BVM

    model predicts almost no resolved kinetic energy kres 0:5 uiui uiui : In contrastto this, the level of total kinetic energy predicted by the EASM is much higher, still,

    the LES is closer to the measured data. The recirculation length is resolved with a

    comparable error by both EASM and LES; however, the EASM shows a faster

    recovery and consequently overpredicts the downstream velocity.

    The misrepresentation of the velocity profile attached to the wall (x 0:0), inparticular the boundary-layer thickness, suggests that the defect of the BVM is

    related to an insufficient modeling of the quasi-steady flow regime upstream of the

    separation. The failure is attributed to the models tendency towards a prematuretransition to turbulence when the flow is exposed to irrotational strain in the vicinity

    of the impingement zone [11].

    -0.5 0 0.5 1 1.5

    U

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    y

    x=0.0

    0 0.5 1

    U

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    y

    EXP

    LES

    WILCOX

    EASM

    x=2.0

    0 0.5 1

    U

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    y

    x=4.0

    0 0.5 1

    U

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    y

    x=6.0

    0 0.5 1

    U

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    y

    x=8.0

    0 0.2

    k

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    y

    x=0.0

    0 0.2 0.4

    k

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    y

    EXP

    LES

    WILCOX

    EASM

    x=2.0

    0 0.2 0.4

    k

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    y

    x=4.0

    0 0.2 0.4

    k

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    y

    x=6.0

    0 0.2 0.4

    k

    0

    0.5

    1

    1.5

    2

    2.5

    3

    3.5

    4

    y

    x=8.0

    EXP

    LES

    EASM

    BVM

    EXP

    LES

    BVM

    EASM

    EXP

    EXP

    LES

    BVM

    EASM

    LES

    EASM

    BVM

    Fig. 2. Time-averaged resolved kinetic energy and mean streamwise velocity in the wake.

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 147114851478

  • 8/3/2019 Comparison Les Rans Bluff-body

    9/15

    4.2. Circular cylinder at Re=3900

    In cases where the flow exhibits sharp corners, such as the square-cylinder

    example, the separation is fixed and the solution is less sensitive to the impact ofcurvature and near-wall modeling. In contrast to this, the separation of the pressure

    driven boundary layer from the continuous surface of a circular cylinder is a much

    more delicate task. The location of the separation depends on the details of the

    attached boundary layer. Moreover, impingement issues are less severe than in the

    square-cylinder case, due to the streamlined shape of the obstacle.

    The first round-cylinder example refers to subcritical conditions at Re=3900,

    where experiments are reported by Lourenco and Shih [13] and Ong and Wallace

    [14]. For comparison, the LES results of Beaudan and Moin [15] are included in the

    figures. In the present LES, the velocity at the inflow is constant with no turbulence

    superimposed, whereas in the RANS a uniform profile with a low turbulence level of

    0.12% and nt 0:02n was prescribed. The boundary layer along the cylinder remainslaminar, and transition to turbulence occurs in the separated shear-layer,

    approximately one diameter aft of the cylinder. The computational mesh employed

    in the (2D) RANS mode consists of approximately 15 000 grid points, covering 32

    diameters in streamwise direction (8ox/Do24) and 16 diameters in normal

    direction (8oy/Do8). The LES grid extends in spanwise direction to pD and 19D

    in streamwise direction and covers 14 diameters in normal direction (-7oy/Do7)

    with approximately 900 000 cells.

    Fig. 3 displays the time-averaged streamlines obtained from the LES. Thetopology involves two small counter-rotating vortices attached to the rear of the

    obstacle, followed by two larger vortices in the near-wake region. In conjunction

    with the linear BVM, the counter-rotating vortices are completely suppressed. The

    X

    Y

    _1 0 1 2 3

    _1.5

    _1

    _0.5

    0

    0.5

    1

    Fig. 3. Time-averaged streamlines (LES).

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 14711485 1479

  • 8/3/2019 Comparison Les Rans Bluff-body

    10/15

    EASM, however, is able to capture the flow topology due to the application of thecurvature correction described in Eq. (8).

    Again, the BVM returns an excessively overestimated recirculation, which is

    obvious from the evolution of the wake centerline velocity shown in Fig. 4 and the

    flow global parameters listed in Table 3. As opposed to this, both the LES and

    EASM return a fair agreement with the experimental data. However, the necessity of

    the curvature correction in conjunction with the EASM is quite evident from the

    global parameters in Table 3. A standard (std.) EASM without the modification,

    Eq. (8), returns almost the same result as the BVM.

    The EASM exhibits the most pronounced recovery due to the augmented intensity

    of the flow reversal. A more complete picture of the simulation quality emerges fromthe streamwise evolution of the wake-velocity profiles displayed in Fig. 5. All

    simulations give v-shaped profiles inline with the experiments, whereas Kravchenko

    U

    _0.4

    _0.3

    _0.2

    _0.1

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    EXP [LS93]EXP [OW96]LESBVMEASMLES [BM94]Std EASM

    X

    _1 0 1 2 3 4 5 6 7 8

    Fig. 4. Mean centerline velocity (Re=3900).

    Table 3

    Global parameters for circular cylinder (Re=3900)

    CD Lr=D St

    Exp. 0.99 1.181.33 0.215

    LES 1.314 1.00 0.216

    BVM 0.88 2.1 0.198

    EASM 0.98 1.2 0.203

    Std. EASM 0.89 1.9 0.200

    LES [14] 1.00 1.3 0.203

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 147114851480

  • 8/3/2019 Comparison Les Rans Bluff-body

    11/15

    [16] reported an alteration to u-shaped profiles when increasing the spanwise

    resolution. The application of the EASM clearly improves the predictive

    performance. The curvature correction facilitates an enhanced sensitivity to a

    curvature-induced attenuation of turbulence, which is obvious from the reducedshear-stress levels returned by the EASM (cf. Fig. 6).

    4.3. Circular cylinder at Re=140 000

    The final example refers to the flow around a circular cylinder at Re=140 000,

    where experiments were reported by Cantwell and Coles [17]. Both grids cover the

    same area as in the former case, but feature an increased resolution of the boundary

    layer region (3D-LES: 2 600 000 points; 2D-RANS: 18 000 points). The resolution of

    the LES grid in spanwise direction is 64 nodes and the wall distance of the first grid

    point is Dn=D=0.0001. Results obtained from the three modeling approachesconsidered in the present study confirm the above-mentioned findings, thus reported

    details are confined to the global flow parameters listed in Table 4 and the evolution

    U

    y

    0 0.5 1

    _2

    _1

    0

    1

    2

    x=1.06

    U

    y

    0 0.5 1

    _2

    _1

    0

    1

    2

    EXP [LS93]

    LESBVMEASMLES [BM94]

    StdEASM

    x=1.54

    U

    y

    0 0.5 1

    _2

    _1

    0

    1

    2

    x=2.02

    U

    y

    0.4 0.6 0.8 1

    _2

    _1

    0

    1

    2

    x=3.00

    U

    y

    0.6 0.8 1

    _2

    _1

    0

    1

    2EXP [OW96]EXP [LS93]LESBVMEASMStdEASM

    x=4.00

    U

    y

    0.6 0.8 1

    _2

    _1

    0

    1

    2

    x=5.00

    Fig. 5. Evolution of the %u-velocity in the wake (Re=3900).

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 14711485 1481

  • 8/3/2019 Comparison Les Rans Bluff-body

    12/15

    of the wake centerline depicted in Fig. 7. With the LES the drag coefficient is

    underpredicted by E50% compared to the experiment and the reference LES of

    Breuer [18]. This great discrepancy is caused by an earlier separation at Y 921

    against Breuers LES with Y 961: This leads to a wrong pressure distributionaround the cylinder, wherein the pressure coefficient on the back cpb=0.649 is only

    uv

    y

    _0.4 _0.2 0 0.2 0.4

    _2

    _1

    0

    1

    2

    x=1.06

    uv

    y

    _0.8 _0.4 0 0.4 0.8

    _2

    _1

    0

    1

    EXP [LS93]

    LESBVM

    EASM

    Std EASM

    x=1.54

    uv

    y

    _0.4 _0.2 0 0.2 0.4

    _2

    _1

    0

    1

    2

    x=2.02

    uv

    y

    _0.1 0 0.1

    _2

    _1

    0

    1

    2

    x=3.00

    uv

    y

    _0.1 _0.05 0 0.05 0.1

    _2

    _1

    0

    1

    2

    EXP [OW96]

    EXP [LS93]LES

    BVMEASM

    Std EASM

    x=4.00

    uv

    y

    _0.1 _0.05 0 0.05 0.1

    _2

    _1

    0

    1

    2

    x=5.00

    Fig. 6. Evolution of the shear stress in the wake (Re=3900).

    Table 4

    Global parameters

    CD Lr=D St

    Exp. 1.237 0.50 0.2

    LES 0.68 0.40 0.2

    BVM 0.41 1.19 0.3

    EASM 1.16 0.59 0.22

    LES [14] 1.239 0.572 0.204

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 147114851482

  • 8/3/2019 Comparison Les Rans Bluff-body

    13/15

    half the value of Breuers LES (cpb=1.398). Moreover, the experiment is

    subcritical, i.e. the boundary stays laminar and the transition take place in the

    separated shear layer. In the present LES, small laminar separation zones cause a

    reattachment of a turbulent boundary layer premature to the experimental observed

    transition. Despite this, the LES give a fair representation of the recirculation length

    and the centerline velocity.

    It should be noted, thatFalthough not elaborated in great detail hereFthe merits

    of the EASM are, in this case, to a large extent related to a non-equilibrium

    modification of the production term in the o-equation as outlined by Menter [19],

    who introduced the reciprocal value of the turbulent viscosity in the equation. In an

    analogy to Menter, the ratio cm=cn

    m appears in the present approach.

    DoDt

    a cmo

    cnmkP bo2 q n n

    t=2 qo=qxj qxj

    : 11

    The modification, which was exclusively used in this testcase, aims to suppress the

    EASM related linearization ofP in the generation term ofo and thus adds to the

    models predictive benefits in non-equilibrium flows cnmocm

    :

    5. Conclusion

    Generally, unsteady statistical modeling approaches (URANS) rely on the formalexistence of a spectral gap between the time scales of the mean transient flow and the

    residual turbulence. This gap does not almost exist in bluff-body flows. Particularly

    X

    U

    _1 0

    _0.4

    _0.3

    _0.2

    _0.1

    0

    0.1

    0.2

    0.3

    0.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    EXP [CC83]

    LES

    BVM

    EASM

    LES [Br99]

    1 2 3 4 5 6 7 8

    Fig. 7. Mean centerline velocity (Re=140 000).

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 14711485 1483

  • 8/3/2019 Comparison Les Rans Bluff-body

    14/15

    in the wake regime, the duration of the flow distortion due to unsteady effects is

    much smaller than the intrinsic time scale of the turbulence and the assumption of a

    spectral gap is excessively violated. This, however, does not undermine the

    defensibility of the URANS approach in engineering computational fluid dynamics.It is important to note, that under such circumstances there is insufficient time for the

    turbulence to effect the mean flow, which significantly reduces the impact of an

    inaccurate modeling approach. In contrast to this, the attached shear layers close to

    the wall are governed by turbulent time scales which are much smaller than the time

    scale of the transient distortion. In such situations, the turbulence model operates in

    quasi-steady mode, and there are no more reasons to distrust the model than in

    steady flow conditions. In conclusion, the predictive failures of a turbulence closure

    often pertain to an inaccurate representation of steady phenomena, e.g. flow-

    impingement, streamline curvature or massive straining. The RANS approaches

    adopted in the present research effort confirm, that deficits inherent to the baseline

    turbulence model tend to foster a deterioration of the overall performance in

    unsteady flows.

    The investigated explicit algebraic stress model, aims to return improved modeling

    capabilities for complex engineering shear flows by means of an efficient non-linear

    stressstrain relations and related modifications to the background model. The

    methodology is seen to enhance the predictive accuracy in unsteady flow conditions

    and significantly outperforms the linear BVM. Superficially, results obtained from

    the present EASM show many features of the LES. The EASM, however, does not

    reproduce the same level of agreement with experimental data as achieved by theLES. Taking into account that the EASM needs only 5% of the LES computing

    time, the capability of the EASM to capture the important flow feature makes

    EASM attractive in engineering calculations.

    Acknowledgements

    The authors greatly acknowledge the financial support by the Deutsche

    Forschungsgemeinschaft under the umbrella of the SFB 557 Beeinflussung

    komplexer turbulenter Scherstr.omungen. Supercomputer access on the CRAY

    T3E was provided by the Zentrum f.ur Informationstechnik Berlin.

    References

    [1] D.C. Wilcox, Reassessment of the scaledetermining equation for advanced turbulence models,

    AIAA J. 26 (11) (1988) 1299.

    [2] T.B. Gatski, C.G. Speziale, On explicit algebraic stress models for complex turbulent flows, J. Fluid

    Mech. 254 (1993) 59.

    [3] W. Rodi, A new algebraic relation for calculating the Reynolds stresses, Z. Angew. Math. Mech. 54

    (1976) T219.

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 147114851484

  • 8/3/2019 Comparison Les Rans Bluff-body

    15/15

    [4] T. Rung, H. L .ubcke, M. Franke, L. Xue, F. Thiele, S. Fu, Assessment of explicit algebraic stress

    models in transonic flows, Engineering Turbulence Modelling and Experiments 4, Ajaccio, 1999,

    p. 659.

    [5] S.S. Girimaji, A Galilean invariant explicit algebraic Reynolds stress model for turbulent curvedflows, Phys. Fluids A 9 (4) (1997) 1067.

    [6] L. Davidson, A dynamic one-equation subgrid model for three dimensional flow, 11th International

    Symposium on Turbulent Shear Flow, Vol. 3, Grenoble, 1997, p. 26.1.

    [7] L. Xue, Entwicklung eines effizienten parallelen L.osungsalgorithmus zur 3d Simulation komplexer

    turbulenter Str.omungen, Dissertation, TU Berlin, 1998.

    [8] C. Rhie, W. Chow, Numerical study of the turbulent flow past an airfoil with trailing edge separation,

    AIAA J. 21 (11) (1983) 1525.

    [9] D.A. Lyn, S. Einav, W. Rodi, J.H. Park, A laser-Doppler velocimetry study of ensemble-averaged

    characteristics of the turbulent near wake of a square cylinder, J. Fluid. Mech. 304 (1995) 285.

    [10] W. Rodi, J.H. Ferziger, M. Breuer, M. Pourqui!e, Status of large eddy simulation: results of a

    workshop, J. Fluids Eng. 119 (1997) 248.

    [11] G. Bosch, W. Rodi, Simulation of vortex shedding past a square cylinder, Proceedings of the tenthSymposium on Turbulent Shear Flows, Pennsylvania State University, USA, 1995, p. 4.13.

    [12] P.R. Voke, Flow past a square cylinder: test case LES2, in: J.P. Chollet, P.R. Voke, L. Kleiser (Eds.),

    Direct and Large-Eddy Simulation II, ERCOFTAC Series, Vol. 5, Kluwer Academic Press,

    Dordrecht, 1997, p. 355.

    [13] L.M. Lourenco, C. Shih, Characteristics of the plane wake of a circular cylinder, A particle image

    velocimetry. From Kravchenko & Moin Report TF-73, 1998, Stanford University, California, 1993.

    [14] L. Ong, J. Wallace, The velocity field of the turbulent very near wake of a circular cylinder, Exp.

    Fluids 20 (1996) 441.

    [15] B. Beaudan, P. Moin, Numerical experiments on the flow past a circular cylinder at sub-critical

    Reynolds-number, Report TF-62, Center for Trubulence Research, Stanford University, 1994.

    [16] A.G. Kravchenko, P. Moin, Numerical studies of flow over circular cylinder at Re=3900, Phys.

    Fluids A 12 (2000) 403.[17] B. Cantwell, D. Coles, An experimental study of entrainment and transport in the turbulent wake of a

    circular cylinder, J. Fluid Mech. 136 (1983) 321.

    [18] M. Breuer, A challenging test case for large eddy simulation: high Reynolds number circular cylinder,

    in: Banerjee, et al., (Eds.), Proceedings of the first Turbulence and Shear Flow Phenomena,

    Begelhouse Inc., New York, 1999, p. 735.

    [19] F.R. Menter, Two-equation eddyviscosity turbulence models for engineering applications, AIAA J.

    32 (8) (1994) 1598.

    H. L .ubcke et al. / J. Wind Eng. Ind. Aerodyn. 89 (2001) 14711485 1485