tr_cfd_08_36.pdf

Post on 04-Apr-2018

212 Views

Category:

Documents

0 Downloads

Preview:

Click to see full reader

TRANSCRIPT

  • 7/29/2019 TR_CFD_08_36.pdf

    1/60

    Effects of mesh resolution on Large Eddy Simulation

    of reacting flows in complex geometry combustors

    G. Boudier a, L.Y.M. Gicquel a, and T.J. Poinsot b

    aCERFACS, 42 Avenue G. Coriolis, 31057 Toulouse cedex, France

    bInstitut de M ecanique des Fluides de Toulouse, Avenue C. Soula, 31400 Toulouse, France

    Abstract

    The power of present parallel computers is becoming sufficient to apply Large Eddy Simu-

    lation (LES) tools to reacting flows not only in academic configurations but also in real gas

    turbine chambers. The most limiting factor to perform LES of real systems is the mesh size

    which directly controls the overall cost of the simulation, so that the effects of mesh resolu-

    tion on LES results become a key issue. In the present work, an unstructured compressible

    LES solver is used to compute the reacting flow in a domain corresponding to a sector of a

    realistic helicopter chamber. Three grids ranging from 1.2 to 44 million elements are used

    for LES and results are compared in terms of mean and fluctuating fields as well as of pres-

    sure spectra. Results show that the mean temperature, species and velocity fields are almost

    insensitive to the grid size. The RMS field of the resolved velocity is also reasonably in-

    dependent of the mesh while the RMS fields of temperature exhibit more sensitivity to the

    grid as expected by the fact that most of the combustion process proceeds at small scales.

    The acoustic field exhibits a limited sensitivity to the mesh, suggesting that LES is adapted

    to the computation of combustion instabilities in complex systems.

    Key words: Mesh resolution, Industrial GT, Large Eddy Simulations

    Corresponding author.Email address: lgicquel@cerfacs.fr (L.Y.M. Gicquel).

    Preprint submitted to Elsevier 2 April 2008

  • 7/29/2019 TR_CFD_08_36.pdf

    2/60

    1 Introduction

    Ongoing developments for the next generation of gas turbines focus on lean pre-

    mixed operating regimes to satisfy emission regulations. The design of these com-

    bustion chambers is complex because combustion concepts leading to minimum

    emissions are also sensitive to combustion instabilities [13]. These instabilities

    are due to a combination of the natural unstable modes of swirling flows (Precess-

    ing Vortex Cores or PVCs [48]) with acoustics and unsteady heat release [912].

    To study combustion instabilities but also to provide more accurate results for sta-

    ble reacting flows, the best numerical technique developed today is Large Eddy

    Simulation (LES) [1315]. LES has been used successfully for many academic

    flames in simple geometries [1622] but still very rarely for complex realistic

    combustion chambers. Multiple issues remain to be investigated before LES can

    be used efficiently for design of combustion chambers: high-order schemes, Sub-

    Grid Scale (SGS) tensor and flux vectors, flame/turbulence interaction, chemical

    schemes, boundary conditions, parallel efficiency, etc.

    In this framework, a fundamental and unresolved question is often neglected: the

    effects of mesh resolution on LES results. Multiple authors have underlined the im-

    portance of this point for LES [23,24]. Although LES results depend on mesh res-

    olution (unlike Reynolds-Averaged Navier-Stokes simulation (RANS) which must

    produce grid independent results as soon as the mesh is sufficiently refined), they

    must satisfy multiple properties: time-averaged values must converge, Root Mean

    Square (RMS) resolved values must increase when the mesh cell size decreases

    and the SGS turbulence level diminishes, the resolved velocity spectra must fill

    towards larger wave-numbers. In practice, these behaviors are expected to be con-

    trolled by the LES models, the flow Reynolds number, the grid resolution as well as

    the accuracy of the numerical solver (in the context of implicit filtering [23,25,26]).

    Mesh dependency analysis of non-reacting LES predictions has recently been ad-

    dressed [2729] and quality criteria have been proposed for a posteriori evaluation

    2

  • 7/29/2019 TR_CFD_08_36.pdf

    3/60

    of the LES flow predictions [24,3033].

    For reacting flows, the computer power needed to simulate realistic geometries is

    so large that the grids used for LES are usually as large as possible while being still

    too coarse to resolve all flow zones: in these circumstances, multiplying the num-

    ber of grid points by a significant factor to verify the effects of grid resolution on

    the LES results was impossible until very recent times. Very few LES of reacting

    flows have been devoted to mesh dependency in simple configurations [2729] and

    none of them has addressed this issue in complex geometry combustors. The situ-

    ation has changed in the last two years: porting LES codes on massively parallel

    machines in the Top 20 list has allowed a sudden increase of power for combus-

    tion computations. For example, Fig. 1 shows the speed-up obtained on such a

    parallel architecture for a 40 million cell configuration corresponding to a full gas

    turbine combustion chamber with the parallel solver used in the present paper [34].

    Speed-ups of nearly 95 percent as obtained here on 4096 processors allow to ad-

    dress the problem of mesh dependency by performing one coarse grid simulation

    with a reasonable mesh (typically 1.5 million cells) and then comparing it with an

    intermediate grid simulation (8 times more cells) and finally with a fine grid

    simulation (32 times more cells). In the present work, the mesh dependency of the

    LES predictions is studied for a helicopter combustion chamber [35]. Results on the

    coarse, intermediate and fine grids for the same regime allow a direct investigation

    in terms of mean flow, RMS values, unsteady activity and acoustic mode excitation.

    Although this grid dependency exercise must also be performed on simple aca-

    demic geometries, using a real Rich-burn, Quick-mix, Lean-burn (RQL) com-

    bustor case is an interesting test case because this configuration puts constraints on

    meshes which are not found in most academic chambers where simple structured

    meshes can be used. Using a real helicopter chamber guaranties that issues relevant

    to industrial cases will be taken into account. However a drawback of this choice

    is that very limited experimental information is available. Therefore the present pa-

    per must be viewed only as a partial response to the problem of LES resolutions

    3

  • 7/29/2019 TR_CFD_08_36.pdf

    4/60

    in combustion since no experimental result will be used for validation. For exten-

    sive comparaisons of the present solver with experimental data, readers are referred

    to previous studies [8,3638] where velocity and/or temperature fields have been

    compared in various configurations.

    The LES solver used for this study and the SGS models required for such a com-

    parison are described in Section 2. Details on the combustion model are given in

    Section 3, while the target configuration (a sector of a helicopter combustion cham-

    ber) is presented in Section 4. Section 5 then discusses results obtained on the three

    grids.

    2 Massively parallel Large Eddy Simulations of reacting flows

    LES for reactive multi-species mixtures involves the spatial filtering operation which

    reduces for spatially, temporally invariant and localised filter functions [39,40], to:

    f(x, t) =1

    (x, t)

    +

    Z

    (x,t) f(x, t) G(x

    x) dx, (1)

    where G denotes the filter function andf(x, t) is the Favre filtered value of the

    variable f(x, t) [41].

    In the mathematical description of compressible turbulent flows with chemical re-

    actions and species transport, the primary variables are the species volumic mass

    fractions(x, t), the velocity vector ui(x, t), the total energyE(x, t) es +1/2 uiui,and the density (x, t) = N=1 (x, t). Note that (x, t) is linked to the species

    mass fractions Y(x, t) and mass conservation imposes for a mixture of N species:

    N=1 Y(x, t) = 1.

    The fluid follows the ideal gas law, p = r T and es =RT

    0 Cp dTp/, where esis the mixture sensible energy, T the temperature, Cp =

    N=1 Cp,Y the fluid heat

    4

  • 7/29/2019 TR_CFD_08_36.pdf

    5/60

    capacity at constant pressure and r is the mixture gas constant, which varies with

    composition and is obtained by r= RW

    = RN=1YW

    , where R = 8.314 kg m2/(s2K)

    and W is the molecular weight of the species . The viscous stress tensor, the

    heat diffusion vector and the species molecular transport use classical gradient ap-proaches. The fluid viscosity follows Sutherlands law, the heat diffusion coefficient

    follows Fouriers law, and the species diffusion coefficients are obtained using a

    constant species Schmidt number and diffusion velocity corrections for mass con-

    servation [15].

    The application of the filtering operation to the instantaneous set of compressible

    Navier-Stokes transport equations with chemical reactions yields the LES equa-

    tions, Eqs. (2)-(4), which need modelling for the system to be closed [23,42]:

    uit

    +

    xj( ui uj) =

    xj[Pi j i j i jt], (2)

    Et

    +

    xj( E uj) =

    xj[ui (Pi j i j) + qj + qjt] + T, (3)

    Yk

    t+

    xj(

    Yk

    uj) =

    xj[Jj,k+Jj,k

    t] + k. (4)

    In Eq. (2), the unresolved SGS stress tensor i jt = (uiuj ui uj) is usually ad-

    dressed through the concept of SGS turbulent viscosity model and the Boussinesq

    assumption [26]. The model henceforth reads (Smagorinsky [43]):

    i jt 1

    3kk

    ti j = 2t Si j, (5)with,

    Si j = 12

    uixj

    +ujxi

    1

    3

    ukxk

    i j. (6)

    In Eq. (5) and (6), Si j is the resolved strain tensor and t is the SGS turbulentviscosity. The Smagorinsky model [43] is used here. It expresses t as:

    t = (CS)2 S. (7)

    5

  • 7/29/2019 TR_CFD_08_36.pdf

    6/60

    In Eq. (7), denotes the filter characteristic length and is approximated by thecubic-root of the cell volume, CS is the model constant (CS = 0.18) and S =(2

    Si j

    Si j)

    1/2.

    In Eqs. (3) and (4), the SGS species flux Jit = (uiY ui Y), and the SGS energy

    flux qit = (uiE ui E), are respectively modelled by use of the species SGS turbu-

    lent diffusivity Dt =t/Sct , where Sc

    t is the turbulent Schmidt number (Sc

    t = 0.9

    for all ). The eddy diffusivity is also used along with a turbulent Prandtl number

    Prt = 0.6, so that t = tCp/Prt:

    Jit

    =

    DtW

    W

    X

    xi

    YVc

    i and qi

    t = t T

    xi+

    N

    =1

    Jit

    hs . (8)

    In Eq. (8) the mixture molecular weight W and the species molecular weight W can

    be combined with the species mass fraction to yield the expression for the molar

    fraction of species : X = YW/W. In expression (8), Vc

    i is the diffusion correc-

    tion velocity resulting from the Hirschfelder Curtis approximation [15] and T isthe modified filtered temperature which satisfies the modified filtered state equa-

    tion [4446], p = rT. Finally, hs stands for the enthalpy of species . Althoughthe performance of the models could be improved through the use of a dynamic

    formulation [25,44,4749], they are considered sufficient to address the present

    investigation. Note that throughout the work, the variations of the molecular coef-

    ficients resulting from the unresolved fluctuations are neglected so that the various

    expressions for the molecular coefficients become only functions of the filtered

    field [15].

    3 Combustion modelling

    This section presents the models necessary to take into account chemical kinet-

    ics and flame/turbulence interactions. For simplicity, a one-step chemistry model,

    Eq. (9), is derived for C10H16 based on a detailed model ofC10H16/Air combustion

    6

  • 7/29/2019 TR_CFD_08_36.pdf

    7/60

    with 43 species and 174 steps (Turbomeca private communication).

    C10H16 + 14 O2 10 CO2 + 8 H2O (9)

    The reduced one-step scheme guaranties proper flame speed predictions only in

    the lean regime (i.e. with equivalence ratios, < 1). For the target configuration,

    such a chemical scheme is not sufficient to predict proper flame positionning since

    the local equivalence ratio reaches a wide range of values. To circumvent such a

    shortcoming the pre-exponential constant of the one-step scheme is adjusted versus

    local equivalence ratio to reproduce the proper flame speed dependency on the rich

    side [50]. The final expression for the rate of reaction of Eq. (9), is:

    Q= A()

    YC10H16WC10H16

    n1 YO2WO2

    n2exp(

    TaT

    ) (mol.m3.s1), (10)

    where n1 = 1.5,n2 = 0.55,Ta = 3608.4 K and the A() function is:

    A() =3.841014

    2(1 + tanh(

    1.390.26

    ))

    +0.33

    4(1 + tanh(

    1.60.8

    ))(1 + tanh(1.85

    0.8)). (11)

    Figure 2 shows that the adjusted one-step scheme matches the detailed scheme

    reasonably well in terms of flame speed and adiabatic temperature for premixed

    laminar flames at 8 bar, which is the target pressure for the full combustor.

    The flame/turbulence interaction is modelled using the Dynamic Thickened Flame

    (DTF) model. Following the theory of laminar premixed flames [51] the flame

    speed S0

    L and the flame thickness 0

    L of a premixed front may be expressed as:

    S0L A and 0L

    S0L=

    A, (12)

    where is the thermal diffusivity and A the pre-exponential constant. Increasing

    the thermal diffusivity by a factor F, the flame speed is kept unchanged if the pre-

    exponential factor is decreased by the same factor. This operation leads to a flame

    7

  • 7/29/2019 TR_CFD_08_36.pdf

    8/60

    thickness which is multiplied by F and more easily resolved on a coarser mesh.

    While in reacting zones, diffusion and source terms issued from the thickened re-

    action are well resolved and turbulence is solely represented by the efficiency func-

    tion E [52], molecular and thermal diffusion cannot be over-estimated by a factor Fin mixing zones where no combustion occurs (it would yield over-estimated mixing

    and wrong flame positions). Dynamic thickening is thus introduced to account for

    these points [15,36,53,54]. The baseline idea of the Dynamically Thickened Flame

    (DTF) model is to detect reaction zones using a sensor S and to thicken only these

    reaction zones, leaving the rest of the flow unmodified. Thickening depends on the

    local grid resolution and it locally adapts the combustion process to reach a numer-

    ically resolved flame front. The flame sub-grid scale wrinkling and interactions at

    the SGS level are supplied by the efficiency function [15,36,53,54].

    4 LES solver and target configuration

    The LES code used here solves the fully compressible, multi-species (variable heat

    capacities) Navier-Stokes equations using a finite-volume discretization, on struc-tured, unstructured and hybrid meshes. Second and third-order temporal and spa-

    tial numerical schemes [55,56] offer reliable unsteady solutions for complex ge-

    ometries as encountered in the field of aeronautical gas turbines. In this study the

    Lax-Wendroff second-order numerical scheme is chosen for the three meshes.

    The configuration (Fig. 3) corresponds to a helicopter combustion chamber where

    fuel is injected using an inverted cane injection system, also called pre-vaporizer.

    The computational domain focuses on a 36 degree section of a full annular reverse-

    flow combustion chamber designed by Turbomeca (Safran group). A premixed

    gaseous mixture of C10H16 enters the chamber through the pre-vaporizer, Fig. 3

    (a) & (b). Fresh gases are consumed in the primary zone, delimited by the chamber

    dilution holes and the liner dome of the combustion chamber, Fig. 3 (a). To ensure

    full combustion, this region of the chamber is fed with air by primary jets located

    8

  • 7/29/2019 TR_CFD_08_36.pdf

    9/60

    on the inner liner, Fig. 3 (b). Burnt gases are then cooled by dilution jets or cooling

    films located on the inner and outer liners as well as on the return bend of the com-

    bustion chamber. Multi-perforated plates also ensure local wall cooling in areas of

    the chamber shown on Fig. 3 (b).

    The combustion regime expected in such burners mixes rich partially premixed

    flames in the chamber primary zone (the gases injected in the canes can be con-

    sidered as premixed gases at an equivalence ratio of 3.17) and diffusion flames

    in the dilution region: i.e. RQL concept. A turbulent combustion model able to

    handle both regimes is therefore needed and the DTF model offers this capac-

    ity [36,54,57,58]. The three meshes used to assess the impact of the grid resolution

    on reacting LES are shown on Fig. 4. Grid characteristics are summarized in Ta-

    ble 2 in terms of number of cells, points and minimum/maximum cell volumes. The

    meshes are refined in the primary zone, particularly in the lower part where combus-

    tion occurs, and in the regions of cooling films. Due to the restrictive size of cooling

    films, manipulations of the fluxes through the faces are needed to specify the cor-

    rect Boundary Conditions. This particular approach is applied for the cooling air

    effusing from the multi-perforated plates. These plates are also homogeneized:

    the flow issuing from the perforations is redistributed over the entire surface and

    corrections are performed for momentum fluxes and turbulent law-of-the-walls to

    account for jet penetrations and wall interactions [59]. The Navier-Stokes Charac-

    teristic Boundary Conditions (NSCBC) [56,60] are applied on the other inlet and

    outlet boundaries to control the acoustic behavior of the system. Walls are adiabatic

    and are treated with a turbulent law-of-the-wall to take into account boundary layer

    effects, Table 1. Side boundaries of the computational domain are axi-periodic.

    The operating point corresponds to cruising conditions and is the same for the three

    grids. The initial condition for the three computations is built as follows: a statisti-

    cally stabilized instantaneous solution, obtained with the coarse grid, is interpolated

    on the two other meshes. These interpolated solutions are then computed until the

    same physical time t0 to ensure independence on the initial guesses. Finally tem-

    9

  • 7/29/2019 TR_CFD_08_36.pdf

    10/60

    poral integration is started for all three grids at t0 until statistical convergence of

    the various mean quantities. The averaging time, the time steps and the local CPU

    effort are summarized in Table 2. Note that the cost of the three computations for

    the same physical time goes from 315 hours on the coarse mesh to 30 ,200 on thefine one (96 times more) so that knowing whether fine grids are really needed and

    which additional information they provide is indeed a critical issue to choose a

    mesh resolution.

    5 Results and discussion

    To evaluate the impact of the grid resolution on the LES predictions, the arguments

    developed in Pope [24] are addressed. Indeed, it is now recognized by the LES com-

    munity that two LES instantaneous fields may differ for various reasons [23,25,61].

    Figures 5 and 6 illustrate the statement as three instantaneous views obtained from

    the three meshes depict degrees of similarity and local differences. The questions of

    interest are in this context whether these LES converge towards the same statistics

    and whether these statistics are independent of the grid resolution, filter size, LES

    model... In this work, grid resolution is investigated as the filter to grid size ratio is

    kept fixed and equal to one. The LES models are the same for the three simulations:

    the Smagorinsky closure for the SGS stress tensor and the DTF model (Section 2)

    for the combustion model.

    Instantaneous flow field visualizations are first presented to illustrate the main flow

    features and the general contribution of grid refinement when performing reactingLES of a complex geometry. In the remaining part of the section, temporally av-

    eraged fields are presented and analyzed. For that study, the following notation is

    adopted to refer to the Reynolds statistical operator:

    f(x,t)

    T

    =1

    T

    t0+TZ

    t0

    f(x, t) dt, (13)

    10

  • 7/29/2019 TR_CFD_08_36.pdf

    11/60

    where T is the time interval used for temporal integration (i.e.: five flow-through

    times, based on the mass flow rate and the primary zone volume). In Eq. (13), the

    spatial dependency of

    T

    is omitted but inferred 1 . The Favre notationf(x, t) is

    introduced to emphasize the fact that statistics constructed here are obtained fromFavre filtered quantities and might differ from the true turbulent statistics obtained

    in raw experimental measurements for example [15,62]. Furthermore, because of

    the filtering procedure introduced in LES, SGS contributions can only be retrieved

    based on modeling criteria as underlined by Veynante [63] and Pope [24]. Assess-

    ment of the LES resolved first moment is investigated first (Section 5.2). Then,

    resolved standard deviations are probed in Section 5.3 followed by estimates of the

    SGS contributions [24,32,63]. Finally acoustic fields are presented in Section 5.4

    using pressure resolved standard deviations and spectra at various locations within

    the computational domain.

    5.1 Instantaneous flow topology and flame structure

    Figure 5 compares instantaneous fields of the axial component of the velocity vec-

    tor scaled by the mean velocity at the inlet of the pre-vaporizer for the three com-

    putational domains. Although instantaneous fields on the different meshes are dif-

    ferent, the main flow structures appear to be very similar: the spreading of the jets

    exiting from the pre-vaporizer is well predicted and the high-speed zones induced

    by multi-perforated plates are present. Likewise, the impact of the primary jets (bot-

    tom liner wall of the main chamber) is equally well captured on all three meshes.

    Nevertheless, and as expected, the fine grid computation exhibits more structures

    perturbating the primary jet as well as an overall greater amount of small turbulent

    structures throughout the computational domain.

    A crucial requirement for the LES method when applied in such complex config-

    1

    Tdepends on x and Tsince the studied flow is not truly stationary. However for long

    enough T,

    Tcan be considered as mainly a function of space.

    11

  • 7/29/2019 TR_CFD_08_36.pdf

    12/60

    urations is the right prediction of the combustion phenomenon. Modelling, which

    is needed to supply proper combustion enhancement due to lack of interactions at

    the unresolved scales, is paramount in that context. If improperly parameterized, a

    turbulent combustion model can yield different flame positions for LES computedwith different mesh resolutions.

    Figure 6 compares instantaneous fields of temperature for the three resolutions.The cutting plane goes through one of the pre-vaporizer outlets (Plane 1, Fig. 3)

    and is colored by the instantaneous field of temperature scaled by the inlet mean

    temperature. The observations drawn from Fig. 5 for the axial component of the

    velocity field also apply to Fig. 6: the temperature fields are clearly enriched

    with increasing grid resolution and the impact on the temperature levels seems

    reduced.

    Figure 7 shows the instantaneous fields of the reaction rate (left column) andthe equivalence ratio in the reaction zone (right column) for the three grids. This

    last quantity is the equivalence ratio conditioned by an index equal to zero in

    the gases where no combustion takes place and to one in the reaction zones. The

    flame position is almost independent of the mesh used and, as expected, the re-

    solved wrinkling of the flame front increases with the mesh resolution. These

    observations confirm the proper behavior of the turbulence/combustion and the

    underlying mechanisms built in the DTF model. Combustion starts close to the

    lips of the canes at an equivalence ratio below the value of the one in the canes

    (which approaches 3). Indeed, because of the large jet velocities in the cane outlet

    regions, combustion can not be sustained immidiately at the cane exits. It allows

    partial mixing of the rich premixed fuel mixture with the surrounding fresh air

    coming from the upper wall multi-perforated plate (cf. Fig. 3 (b)). Combustion

    then occurs for equivalence ratio values approaching 1.7. It then proceeds along

    the walls of the chamber (left column). All three LES, even the coarse one, pre-

    dict very similar fields of reaction rate.

    Figure 8 displays the thickening factor fields (left column) along with the corre-

    12

  • 7/29/2019 TR_CFD_08_36.pdf

    13/60

    sponding efficiency function fields (right column) for the three grids. These re-

    sults complement the predictions of the reaction rate fields, Fig. 7 (left column).

    An overall agreement is observed between the three simulations. The coarse

    mesh solution exhibits less wrinkling and wider flame fronts, compensated bylarger values of the efficiency function. Contours of reaction rates coincide with

    contours of thickening factor and efficiency function on the three meshes. The

    grid dependency of the thickening factor accompanied by a proportional adap-

    tation of the efficiency function leads to similar overall positions of the flame

    front. From a statistical point of view, all simulations yield similar mean flame

    locations within the combustion chamber and differences may be expected for

    higher moments which needs to be determined and quantified for proper assess-

    ment. These points are addressed in section 5.2.

    As underlined by the right column results of Fig. 7, increasing the filtered field

    content with smaller turbulent structures might also impact the mixture field to yield

    different local filtered compositions. Figure 9 investigates that issue by showing the

    Probability Density Function (PDF) of the instantaneous resolved equivalence ratio

    in reacting zones, i.e. zones where the local reaction rate is higher than a tenth of

    the mean volumetric reaction rate. Combustion mainly occurs in lean premixed

    zones, and the fitted one-step chemical scheme is seen to operate suitably as very

    few flame elements burn in rich conditions ( > 1). The fine mesh results allow

    to identify the existence of three peaks which can be identified as follows: (1) the

    peak around = 1.7 corresponds to the rich premixed flames created at the cane

    outlet, (2) the peak at = 1 is due to the presence of the diffusion flamelets and

    (3) the lean peak coincides with the mean chamber equivalence ratio ( = 0.33).

    The differences in shapes and more specifically the presence of a secondary peak

    at stoechiometry ( = 1) that is more pronounced as the grid is refined, underline

    potential shift in combustion regimes. Further investigations are however needed

    as these observations only stem from instantaneous flow studies which will differ

    for different grid resolutions as underlined previously. Only temporal statistics are

    13

  • 7/29/2019 TR_CFD_08_36.pdf

    14/60

    generally of interest to LES predictions.

    The preliminary investigation of the instantaneous flow field confirms the good

    quality and behavior of the predictions obtained with all three resolutions. Overall

    agreement is clear and differences are only local as expected by the working con-

    text of LES. Further analyses are presented below to quantify the impact of mesh

    resolution on LES statistics.

    5.2 Mean flow results

    Spatial fields of the mean temporal statistics, Eq. (13), are good indicators of the

    mesh influence on LES results. In this sub-section, mean fields are investigated for

    the three grids for Plane 1, Fig. 3. Figures 10 and 11 respectively show the mean

    axial component of the velocity vector and the mean temperature field scaled by

    corresponding mean values at the inlet of the pre-vaporizer.

    A good agreement is observed between the three predictions for both quantities,

    Figs 10 and 11. The jet spreading at the cane outlet observed on the mean field of

    velocity of Fig. 10 is recovered on the three meshes. High temperature pockets

    are located in the same zones and fill similar volumes, assessing the similar mean

    behavior of combustion on the three grids (Fig. 11).

    Grid resolution has an impact in various highly localized regions and some dis-crepancies are detected, especially in the mean temperature fields of Fig. 11. The

    improved mesh quality of the intermediate and fine grids makes the flow behave

    differently in near-wall regions where the chamber flow interacts with the flow

    issuing from cooling devices. For example, the thermal boundary layer created

    by the multi-perforated plates (Zone 1 and 3 on Fig. 10 (a)) on the intermediate

    mesh is thinner than the one on the coarse grid. Likewise, the penetration of the

    cooling film located in the upper part of the liner dome (Zone 2 on Fig. 10 (a)) is

    different on the coarse grid. Despite these discrepancies, the agreement between

    14

  • 7/29/2019 TR_CFD_08_36.pdf

    15/60

    these sets of predictions underlines mesh independence of the first moments for

    these calculations. Convergence in terms of mesh resolution seems to be reached

    for the two finer grids at least in terms of first moments.

    More quantitative studies are presented on Fig. 12 where profiles in transverse cuts

    going through the cane jets and primary jets at different axial positions and for the

    same quantities are displayed. Comparisons between the three sets of LES predic-

    tions confirm the previous observations on the mean field maps (Figs. 10 and 11).

    Axial velocity profiles, Fig. 12 (a), illustrate a slight over-estimation of the axial

    velocity magnitude along the center of the cane for the coarse grid predictions.

    Likewise, upper boundary layer and recirculating zones located behind the primary

    jets seem over-estimated on the coarse grid. The intermediate and fine mesh pro-

    files collapse onto the same curves at all stations which indicates that convergence

    is reached on the intermediate mesh for that quantity. Furthermore, on the finer

    meshes, velocity curves prior to the chamber dome are more diffused on the coarse

    grid. Figure 12 presents temperature profiles at the same stations: all grid predic-

    tions are in very good agreement and discrepancies appear only in the upper part

    of the primary zone (Zones 2 and 3) where the stream coming from the top cooling

    film does not penetrate the chamber as much on the finer meshes as on the coarse

    grid. Results on the finer meshes predict more local mixing between the chamber

    flow and fresh gases coming from the cooling devices. Monotoneous convergence

    is achieved for the mean temperature profiles: the intermediate mesh predictions

    are positioned in-between the coarse and fine grid profiles. At other locations, such

    a statement is not always true and is explained by the complexity of the flow field

    which is thus very sensitive to various flow and grid parameters.

    Spatial distributions of the mean reaction rates in Plane 1 are displayed in Fig. 13.

    As underlined from the instantaneous snapshots and the mean field of temperature,

    combustion occurs in the region delimited by the cane outlet and chamber dome

    for all meshes. Mean flame anchoring is similar on all three meshes and differences

    only appear when localizing the peak reaction rate distributions. Indeed, as the grid

    15

  • 7/29/2019 TR_CFD_08_36.pdf

    16/60

    resolution increases, the peak value of the mean reaction rate propagates towards

    the lips of the cane injector. Convergence for this purely reacting quantity is more

    difficult to assess. Such results corroborate the importance of the interaction be-

    tween the various models and the local available grid resolution when conductingLES. They also underline the potential differences in composition as illustrated in

    Fig. 9. Such local changes, which may impact mean statistical fields, directly im-

    pact Arrhenius-like combustion models such as the DTF model.

    5.3 Standard deviations and SGS statistics

    One criterion for true LES stems from the observation that at a given value of the

    Reynolds number (preferably very large) the total variances, that is to say the sec-

    ond central moment based on the unfiltered variables, should remain independent of

    the mesh resolution. The evolution with the grid resolution of the resolved standard

    deviation as obtained from the LES field is henceforth of great interest. Similarly

    the SGS contributions need to be evaluated for a proper assessment of the asymp-

    totic behavior (if existing) of the total RMS values. The following sub-section dis-

    cusses such aspects and presents results obtained from the three grid resolutions for

    the resolved RMS values and turbulent kinetic energy for the resolved and residual

    motions [24]. The analysis is based on the spatial evolution of the mean temporal

    operator, Eq. (13).

    Resolved RMS fields for the filtered axial velocity component and temperature are

    shown in Figs. 14 and 15 for all meshes.

    Figure 14 shows no significant difference between the three predictions: the RMSaxial velocity fields are similar and values reach approximately the same levels.

    This global agreement highlights the proper action of the classical Smagorin-

    sky model which results in consistent predictions regardless of the mesh reso-

    lution. The SGS velocity model does not infer major changes in the resolved

    16

  • 7/29/2019 TR_CFD_08_36.pdf

    17/60

    RMS velocity fields. Some local discrepancies can be observed in the wake of

    the primary jet and the cane jet, especially on the coarse mesh. These differences

    more likely arise from the over-estimated mean velocity field at these locations

    resulting from an under-resolved mesh to allow proper use of the SGS model.Despite that shortcoming on the coarse mesh, convergence is almost reached on

    the intermediate grid and for most of the computational domain.

    The resolved RMS temperature, Fig. 15, exhibits larger differences than the re-solved RMS velocity when the mesh resolution evolves. Higher RMS tempera-

    tures are observed on the fine grid, even if levels reached in the primary zone are

    very similar between the intermediate and the fine grids. A direct observation of

    the flow (Figs. 5 and 6) reveals that these RMS levels are directly due to strong

    flame motions which can be captured on the finer grids and not on the coarse

    one. Moreover the finer grids are able to capture more flame wrinkling, implying

    higher RMS temperature levels. As mentioned before, the local mean reaction

    rates are similar on all meshes (Fig. 13) because the DTF model compensates

    the reduced resolved flame wrinkling on the coarse mesh by a larger sub-grid

    efficiency. However, the model cannot explicitely reconstruct the resolved wrin-

    kling and RMS temperatures observed in Fig. 15.

    Results of Fig. 14 and 15 can be quantitatively confirmed by plotting cuts extracted

    from Plane 1, as done on Fig. 16. The profiles show that the resolved RMS velocity

    depends weakly on the mesh while the resolved RMS temperature grows signifi-

    cantly when the mesh resolution increases. At this point, it is worth discussing why

    these standard deviations behave in such different ways. The main reason has ac-

    tually been mentioned many times before in the literature: the velocity field (and

    its standard deviation) is mainly controlled by the large structures and even the

    coarse grid is sufficient to capture the RMS levels because the smallest scales are

    not modelled by the classical Smagorinsky model (it only models vortex dissipa-

    tion beyond a cutting length). On the other hand, combustion is mainly a sub-grid

    scale phenomenon, progressing at the smallest scales so that each increase of the

    17

  • 7/29/2019 TR_CFD_08_36.pdf

    18/60

    mesh resolution will reveal more flame wrinkling and produce higher RMS temper-

    ature. The best which can be expected from any flame/turbulence interaction model

    is to conserve the zeroth-order moment which is the mean local reaction rate be-

    cause it controls the flame position. Predicting the true RMS temperature remainsa challenge for LES of reacting flows 2 which can be tackled only with more so-

    phisticated models providing a precise description of the temperature and species

    PDF [21,6466] (which is not done with the DTF model).

    The quality of a LES computation mostly relies on the scale separation between

    the length scales of turbulence captured and the one modeled by the afforded grid

    resolution. A quantitative measurement of the turbulence resolution is introduced

    in Pope [24] by comparing SGS turbulent kinetic energy kSGS with the resolved

    turbulent kinetic energy kresolved, which reads:

    kresolved =1

    2

    UiUiT

    UiUiT

    T

    (14)

    The estimated quantity kSGS is derived from the following expression [23]: SGS =

    CM

    kSGS, where SGS = t is the SGS turbulent viscosity given by the classical

    Smagorinsky model [43], CM is a constant value and is the filter cutting charac-

    teristic length (equal to the cell characteristic length in this study). The resolution

    criterion M is thus defined as M = kSGS/(kresolved + kSGS). M varies between 0

    (equivalent to DNS where no model is needed) and 1 (corresponding to RANS

    where the entire turbulent spectrum is modeled). Figure 17 depicts the spatial evo-

    lution ofMas obtained from the temporally averaged LES and shows no significant

    discrepancy between the three grids. Regions where M> 0.2 (corresponding to re-

    gions where less than 80 % of the turbulent kinetic energy is resolved) are reduced

    to the near-wall regions and the potential core of jets on both grids. This criterion

    2 Note that eventhough all models for flame/turbulence interactions will exhibit the same

    difficulties, some models may require less points to resolve the flame front for a given mesh

    resolution. For example, the G-equation model [14] will exhibit more wrinkling than the

    DTF model for the same resolution.

    18

  • 7/29/2019 TR_CFD_08_36.pdf

    19/60

    thus guaranties the quality of the results obtained on these three grids, even if care

    must be taken when dealing with near-wall regions on the coarse grid. Note that the

    Pope criterion only quantifies the applicability of SGS modeling for the velocity

    field and its extension to the conserved scalar, Z, is needed to assess the modeledto resolved contributions for mixing, which is of great interest from a combustion

    point of view. In that case, MZ = ZZSGS/(ZZresolved +ZZSGS) is proposed and pre-

    sented on Fig. 18. Similarly to Popes criterion, the SGS term is estimated following

    the expression of [63]. For mixing, the impact of the grid resolution is observed

    in the same regions. Figure 18 also underlines the reduced effect of the grid reso-

    lution on MZ at the flame front, which is expected since the DTF model provides

    a resolved field with no SGS contribution at that specific location, the interaction

    between turbulence and combustion being represented through the efficiency func-

    tion. Note in that respect that other LES combustion models are expected to behave

    differently at the flame front.

    5.4 Acoustic fields

    The prediction of thermo-acoustic phenomena is of growing interest for industrial

    purposes as gas turbine manufacturers are constrained, mostly for environmental

    concerns, to design combustion devices able to operate at lean regimes, which are

    known to be sometimes unstable. The contribution of LES in this framework is

    essential and has already proved its capability to predict flame/acoustics interac-

    tions [15,36,54,58,67]. A central question to use LES for combustion instability

    studies is to know whether unsteady LES results (pressure fields for example) are

    grid-independent. Since the previous section has indicated that the RMS temper-

    ature is dependent on grid resolution, checking whether the pressure field is also

    dependent on the mesh resolution is a logical next step.

    RMS pressure fields exhibit very similar patterns on all meshes (Fig. 19) even

    though the finer meshes lead to slightly higher p levels. The three simulations iden-

    19

  • 7/29/2019 TR_CFD_08_36.pdf

    20/60

    tify the same acoustic activity in the combustor as shown by the pressure spectra

    displayed in Fig. 20 for two probes: Probe 1 is located near one of the pre-vaporizer

    outlets and Probe 2 in the primary zone (Fig. 3). LES reveals two modes around

    8700 and 9300Hz which are recovered on the different meshes with reasonableaccuracy. In such complex geometries, these modes can not be associated to any

    simple mode (i.e. quarter wave, etc). The only statement which can be made is that

    they are clearly associated to transverse activity between the upper and lower walls

    of the chamber.

    The fact that all LES provide similar results for the p fields must be interpreted

    with care:

    The unsteady pressure field is the convolution of the source terms (due to un-steady reaction rate) and of acoustic propagation in the chamber. The acoustic

    propagation takes place on long wave lengths (typically tens of centimeters at

    10kH z in the burnt gases) and is probably not very sensitive to mesh changes.

    The mean value of the source term has also been shown to be insensitive to the

    mesh but the dependence of its unsteady component on mesh resolution, on the

    other hand, is not studied here. This is obviouly the next step to take.

    Predicting unsteady pressure in LES of reacting flows remains a significant chal-lenge and the present results only suggest that predicting unsteady pressure in a

    real combustor with LES is actually not more difficult than predicting the mean

    flame position or the velocity statistics. Other studies [54,6870] have indicated

    the same trend and this optimistic outcome is probably justified only for the pre-

    diction of the acoustic mode frequencies (which depend mainly on the geometry

    and not much on the source terms) but not for the amplitudes of the modes.

    In other words, predicting whether a combustor will oscillate and at which fre-

    quency is a task which is within the possibilities of LES. However predicting the

    exact amplitude of the oscillation might remain more difficult for a long time.

    Indeed, this amplitude will depend not only on the mesh and on the prediction of

    the unsteady source terms as well as on physical dissipation mechanisms but also

    20

  • 7/29/2019 TR_CFD_08_36.pdf

    21/60

    on the dissipation of the numerical method, on the acoustic boundary conditions,

    etc.

    6 Conclusions

    Although LES is the most promising tool to study unsteady phenomena in complex

    geometry combustion chambers, few studies have addressed the question of the de-

    pendence of LES results to mesh resolution in realistic configurations. This paper

    considers this problem by performing three LES of the same helicopter combustion

    chamber, the first one with 1,242,086 elements, the second one with 10,620,245

    elements and the third one with 43,949,682 elements. Multiple comparisons be-

    tween the three sets of LES results are performed. The flame position, its unsteady

    behaviour and the mean flow fields (velocity, temperature, reaction rate) are shown

    to be reasonably insensitive to mesh resolution. The standard deviation of the re-

    solved velocity field is also shown to depend weakly on resolution, as confirmed by

    the Pope criterion, showing that most of the kinetic energy spectrum is resolved on

    the different grids. However, the RMS temperature is shown to increase when the

    mesh resolution increases, consistently with the fact that combustion is a sub-grid

    phenomenon and that its largest part is in the unresolved component. Finally, even

    though the RMS temperature depends strongly on mesh resolution, the acoustic ac-

    tivity is shown to be less dependent, exhibiting similar acoustic modes (frequency

    and amplitudes) on all the three meshes, suggesting that LES is reasonably adapted

    to the study of combustion instability in complex geometry combustors.

    Acknowledgments

    Numerical simulations and visualizations have been conducted on the computers of

    the French National Computing Center (CINES) in Montpellier, at the Barcelona

    21

  • 7/29/2019 TR_CFD_08_36.pdf

    22/60

    Supercomputing Center (BSC) located in Spain and CERFACS in-house computing

    facility.

    22

  • 7/29/2019 TR_CFD_08_36.pdf

    23/60

    References

    [1] W. Polifke, A. Fischer and T. Sattelmayer, ASME journal of engineering for gas

    turbines and power 125 (2003) 20.

    [2] V. Sankaran and S. Menon, J. Turb. 3 (2002) 011.

    [3] D. Smith and E. Zukoski, 21st Joint Propulsion Conference, pp. AIAA paper 851248,

    Monterey, 1985.

    [4] P. Billant, J.M. Chomaz and P. Huerre, J. Fluid Mech. 376 (1998) 183.

    [5] A. Gupta, D. Lilley and N. Syred, Swirl flows (Abacus Press, 1984).

    [6] M. Hall, Ann. Rev. Fluid Mech. 4 (1972) 195.

    [7] O. Lucca-Negro and T. ODoherty, Prog. Energy Comb. Sci. 27 (2001) 431.

    [8] S. Roux et al., Combust. Flame 141 (2005) 40.

    [9] Y. Mizobuchi et al., Proc. of the Combustion Institute 30 (2004) 611.

    [10] L. Muniz and M. Mungal, Combust. Flame 111 (1997) 16.

    [11] H. Pitsch and H. Steiner, Proc. of the Combustion Institute 28 (2000) 41.

    [12] L. Vervisch et al., J. Turb. 5 (2004) 004.

    [13] N. Peters, Turbulent combustion (Cambridge University Press, 2000).

    [14] H. Pitsch, Ann. Rev. Fluid Mech. 38 (2006) 453.

    [15] T. Poinsot and D. Veynante, Theoretical and numerical combustion (R.T. Edwards,

    2nd edition., 2005).

    [16] H. Pitsch and H. Steiner, Phys. Fluids 12 (2000) 2541.

    [17] H. Pitsch and L. Duchamp de la Geneste, Proc of the Comb. Institute 29 (2002) 2001.

    [18] C. Duwig et al., AIAA Journal 45 (2007) 624.

    [19] J. Janicka and A. Sadiki, Proc. of the Combustion Institute 30 (2004) 537.

    23

  • 7/29/2019 TR_CFD_08_36.pdf

    24/60

    [20] H. El-Asrag and S. Menon, Proc. of the Combustion Institute 31 (2007) 1747.

    [21] H. El-Asrag and S. Menon, AIAA/SAE/ASME/ASEE 41st Joint Propulsion

    Conference, Tucson, USA, 2005.

    [22] O. Stein and A. Kempf, Proc. of the Combustion Institute 31 (2007) 1755.

    [23] P. Sagaut, Large Eddy Simulation for incompressible flows (Springer-Verlag, 2000).

    [24] S.B. Pope, New Journal of Physics 6 (2004) 35.

    [25] S. Ghosal and P. Moin, J. Comput. Phys. 118 (1995) 24 .

    [26] S.B. Pope, Turbulent flows (Cambridge University Press, 2000).

    [27] B. Vreman, B. Geurts and H. Kuerten, International Journal for Numerical Methods

    in Fluids 22 (1996) 297.

    [28] J. Meyers, B. Geurts and M. Baelmans, Phys. Fluids 15 (2003) 2740.

    [29] C. Priere et al., AIAA Journal 43 (2005) 1753.

    [30] A. Vreman, B. Geurts and J. Kuerten, J. Eng. Math. 29 (1995) 299.

    [31] B. Geurts and J. Frohlich, Physics of Fluids (2002) L41.

    [32] M. Klein, Flow Turb. and Combustion 75 (2005) 131.

    [33] I. Celik, Z. Cehreli and I. Yavuz, Journal of Fluids Engineering 127 (2005) 949.

    [34] N. Allsopp et al., Unfolding the IBM eServer Blue Gene Solution, First ed. (IBM

    Redbooks, 2005) pp. 277285.

    [35] G. Boudier et al., Proc. of the Combustion Institute 31 (2007) 3075.

    [36] P. Schmitt et al., J. Fluid Mech. 570 (2007) 17.

    [37] L. Selle et al., Combust. Flame 137 (2004) 489.

    [38] C. Priere et al., J. Turb. 5 (2004) 1.

    [39] A.A. Aldama, Lecture Notes in Engineering Vol. 49 (Springer-Verlag, New York,

    1990).

    24

  • 7/29/2019 TR_CFD_08_36.pdf

    25/60

    [40] B. Vreman, B. Geurts and H. Kuerten, Phys. Fluids 6 (1994) 4057.

    [41] A. Favre, Problems of hydrodynamics and continuum mechanics, pp. 231266, SIAM,

    Philadelphia, 1969.

    [42] J. Ferziger and M. Peric, Computational Methods for Fluid Dynamics (Springer

    Verlag, Berlin, Heidelberg, New York, 1997).

    [43] J. Smagorinsky, Monthly Weather Review 91 (1963) 99.

    [44] P. Moin et al., Phys. Fluids A 3 (1991) 2746.

    [45] G. Erlebacher et al., J. Fluid Mech. 238 (1992) 155.

    [46] F. Ducros, P. Comte and M. Lesieur, J. Fluid Mech. 326 (1996) 1.

    [47] D.K. Lilly, Phys. Fluids A4 (1992) 633.

    [48] M. Germano, J. Fluid Mech. 238 (1992) 325.

    [49] C. Meneveau, T. Lund and W. Cabot, J. Fluid Mech. 319 (1996) 353.

    [50] A. Linan and F. Williams, Fundamental aspects of combustion (Oxford University

    Press, 1993).

    [51] F. Williams, Combustion theory (Benjamin Cummings, Menlo Park, CA, 1985).

    [52] O. Colin et al., Phys. Fluids 12 (2000) 1843.

    [53] J.P. Legier, Simulations numeriques des instabilites de combustion dans les foyers

    aeronautiques, Phd thesis, INP Toulouse, 2001.

    [54] C. Martin et al., AIAA Journal 44 (2006) 741.

    [55] O. Colin and M. Rudgyard, J. Comput. Phys. 162 (2000) 338.

    [56] V. Moureau et al., J. Comput. Phys. 202 (2005) 710.

    [57] Y. Sommerer et al., J. of Turbulence 5 (2004).

    [58] A. Roux et al., Combust. Flame 152 (2007) 154.

    [59] S. Mendez et al., Summer Program, pp. 5772, Center for Turbulence Research,

    NASA AMES, Stanford University, USA, 2006.

    25

  • 7/29/2019 TR_CFD_08_36.pdf

    26/60

    [60] T. Poinsot, T. Echekki and M. Mungal, Combust. Sci. Tech. 81 (1992) 45.

    [61] M. Lesieur, New tools in turbulence modelling, edited by O. Metais and J. Ferziger,

    pp. 1 28, Les Editions de Physique - Springer Verlag, 1997.

    [62] J. Panda and R. Seasholtz, AIAA Journal 44 (2006) 1952.

    [63] D. Veynante and R. Knikker, J. Turb. 7 (2006) 1.

    [64] S. Pope et al., Proc. of the Summer Program, pp. 283294, Center for Turbulence

    Research, Stanford, USA, 2004.

    [65] A. Klimenko and R. Bilger, Prog. Energy Comb. Sci. 25 (1999) 595 .

    [66] S. James, J. Zhu and M. Anand, Proc. of the Combustion Institute 31 (2007) 1737.

    [67] A. Giauque et al., J. Turb. 6 (2005) 1.

    [68] A. Sengissen, Simulation aux grandes echelles des instabilites de combustion : vers le

    couplage fluide/structure - TH/CFD/06/12, PhD thesis, Universite de Montpellier II,

    2006.

    [69] R.A. Huls et al., J. Sound Vib. 304 (2007).

    [70] A. Sengissen et al., Combust. Flame 150 (2007) 40.

    26

  • 7/29/2019 TR_CFD_08_36.pdf

    27/60

    List of Tables

    1 BC types and specificities for LES. 29

    2 Mesh characteristics used for the coarse, the intermediate and

    fine LES meshes. 30

    27

  • 7/29/2019 TR_CFD_08_36.pdf

    28/60

    TABLES

    28

  • 7/29/2019 TR_CFD_08_36.pdf

    29/60

    Boundary type Acoustic properties Specified quantities Relax values

    Cane inlet, dilution

    and primary holesNSCBC inlet un, T, P, Y 50

    Outlet NSCBC outlet P 200

    Inlet films Purely reflective inlet un, T, P, Y -

    Multi-perforated

    plates

    Purely reflective inlet

    with momentum

    correction

    un, T, P, Y -

    Adiabatic wallsPurely reflective wall

    with law-of-the-wall

    - -

    Table 1

    BC types and specificities for LES.

    29

  • 7/29/2019 TR_CFD_08_36.pdf

    30/60

    Coarse mesh Intermediate mesh Fine mesh

    Total number of points 230,118 1,875,835 7,661,005

    Total number of cells 1,242,086 10,620,245 43,949,682

    Max. cell volume [m3] 3.12671108 8.97802109 4.05748 109

    Min. cell volume [m3] 1.81795 1011 8.29479 1012 1.1828 1012

    Time step [s] 1.52107 0.88107 0.49107

    Averaging time [ms] 10 10 10

    Number of iterations during averaging 65,790 108,695 204,082

    Total CPU effort for a 10ms(hours) LES 315 4,550 30,200

    Table 2

    Mesh characteristics used for the coarse, the intermediate and fine LES meshes.

    30

  • 7/29/2019 TR_CFD_08_36.pdf

    31/60

    List of Figures

    1 Typical speed-up curve as obtained for the configuration studied in

    this work. 34

    2 Chemical scheme validation: (a) is flame speed prediction, (b)

    is adiabatic flame temperature as functions of the equivalence

    ratio. The detailed scheme for C10H16 contains 43 species and 174

    chemical steps. 35

    3 Combustion chamber: (a) 3D view and (b) side view of the

    computational domain. 36

    4 Mesh resolution for (a) the coarse mesh, (b) the intermediate mesh

    and (c) the fine mesh, as obtained in a transversal cut passing

    through one of the canes (Plane 1). 37

    5 Instantaneous axial velocity field scaled by the mean inlet velocity

    and obtained on (a) the coarse grid, (b) the intermediate one and

    (c) the fine one. 38

    6 Instantaneous temperature field scaled by the mean inlet

    temperature and obtained on (a) the coarse grid, (b) the

    intermediate one and (c) the fine one. 39

    7 Instantaneous fields of reaction rate (a, c and e) and equivalence

    ratio conditioned by the local reaction (b, d and f) on Plane 1

    obtained on the coarse grid, (a and b), the intermediate grid (c and

    d) and the fine one (e and f). 40

    31

  • 7/29/2019 TR_CFD_08_36.pdf

    32/60

    8 Instantaneous fields of the thickening factor (a, c and e) and the

    efficiency function (b, d and f) on Plane 1 obtained on the coarse

    grid (a and b), the intermediate grid (c and d) and the fine one (e

    and f). 41

    9 Probability density function of local equivalence ratio in reacting

    zones (zones where local reaction rate is higher than a tenth of

    the mean volumetric reaction rate) on (a) the coarse grid, (b) the

    intermediate one and (c) the fine one. 42

    10 Mean axial velocity field (scaled by the mean velocity at the

    pre-vaporizer inlet) obtained on (a) the coarse grid, on (b) the

    intermediate grid and on (c) the fine grid. 43

    11 Mean temperature field (scaled by the mean temperature at the

    pre-vaporizer inlet) obtained on (a) the coarse grid, on (b) the

    intermediate grid and on (c) the fine grid. 44

    12 Transverse profiles in Plane 1, Fig. 3, passing through one cane

    outlet: (a) mean axial velocity profiles scaled by the pre-vaporizer

    inlet velocity and (b) mean temperature profiles scaled by by the

    pre-vaporizer inlet temperature. 45

    13 Mean reaction rate on Plane 1 obtained on (a) the coarse grid, (b)

    the intermediate one and (c) the fine one. 46

    14 Fields of RMS resolved axial velocity on (a) the coarse grid, (b)

    the intermediate grid and (c) the fine grid. The quantities are

    non-dimensionalized by pre-vaporizer inlet conditions. 47

    15 Fields of RMS resolved temperature on (a) the coarse grid, (b)

    the intermediate grid and (c) the fine grid. The quantities are

    non-dimensionalized by pre-vaporizer inlet conditions. 48

    32

  • 7/29/2019 TR_CFD_08_36.pdf

    33/60

    16 Transverse profiles in Plane 1 (cfFig. 3), passing through one cane

    outlet: (a) RMS axial velocity profiles scaled by squared inlet

    velocity and (b) RMS temperature profiles scaled by squared inlet

    temperature. 49

    17 Spatial evolution of the temporal average of the resolution criterion

    M: (a) is the coarse grid, (b) is the intermediate grid and (c) is the

    fine grid. Black isoline corresponds to M= 0.2. 50

    18 Spatial evolution of the temporal average of the resolution criterion

    MZ: (a) is the coarse grid, (b) is the intermediate grid and (c) is the

    fine grid. Black isoline corresponds to M= 0.2. 51

    19 RMS pressure field obtained on (a) the coarse grid, (b) the

    intermediate grid and (c) the fine grid. The quantities are

    non-dimensionalized by pre-vaporizer inlet conditions. 52

    20 Spectra of the fluctuating pressure temporal evolution at (a) Probe

    1 and (b) Probe 2 (cf Fig. 3). 53

    33

  • 7/29/2019 TR_CFD_08_36.pdf

    34/60

    FIGURES

    4000

    3000

    2000

    1000Speedup

    4000300020001000

    Number of processors

    Ideal Speedup

    40.106

    cells mesh

    Fig. 1. Typical speed-up curve as obtained for the configuration studied in this work.

    34

  • 7/29/2019 TR_CFD_08_36.pdf

    35/60

    (a)

    2.0

    1.5

    1.0

    0.5Flamespeed[m.s

    -1]

    3.02.52.01.51.00.5

    Equivalence ratio [-]

    Detailed scheme

    1-step schemeFitted 1-step scheme

    (b)

    2600

    2400

    22002000

    1800Tempera

    ture[K]

    3.02.52.01.51.00.5

    Equivalence ratio [-]

    Detailed scheme

    1-step schemeFitted 1-step scheme

    Fig. 2. Chemical scheme validation: (a) is flame speed prediction, (b) is adiabatic flame

    temperature as functions of the equivalence ratio. The detailed scheme for C10H16 contains

    43 species and 174 chemical steps.

    35

  • 7/29/2019 TR_CFD_08_36.pdf

    36/60

    (a)

    Primary holes

    Fuel

    inlet

    Dilution holes

    Probe 2

    Liner dome

    Probe 1

    Plane 1

    (b)

    Multi-perforated plates Cooling

    film

    Multi-perforated plates

    Cooling films

    Cooling film for

    Nozzle Guide Vane

    Primary holes

    Dilution holes

    Exit

    Fuel

    inlet

    Fig. 3. Combustion chamber: (a) 3D view and (b) side view of the computational domain.

    36

  • 7/29/2019 TR_CFD_08_36.pdf

    37/60

    (a)

    (b)

    (c)

    Fig. 4. Mesh resolution for (a) the coarse mesh, (b) the intermediate mesh and (c) the fine

    mesh, as obtained in a transversal cut passing through one of the canes (Plane 1).

    37

  • 7/29/2019 TR_CFD_08_36.pdf

    38/60

    (a)

    1.422

    0.711

    0

    -0.711

    -1.422

    U/Uinlet

    "[ ]

    (b)

    1.422

    0.711

    0-0.711

    -1.422

    U/Uinlet

    "[ ]

    (c)

    1.422

    0.711

    0

    -0.711

    -1.422

    U/Uinlet

    "[ ]

    Fig. 5. Instantaneous axial velocity field scaled by the mean inlet velocity and obtained on

    (a) the coarse grid, (b) the intermediate one and (c) the fine one.

    38

  • 7/29/2019 TR_CFD_08_36.pdf

    39/60

    (a)

    4.613

    3.708

    2.803

    1.898

    1

    T/Tinlet

    "[ ]

    (b)

    4.613

    3.708

    2.8031.898

    1

    T/Tinlet

    "[ ]

    (c)

    4.613

    3.708

    2.803

    1.898

    1

    T/Tinlet

    "[ ]

    Fig. 6. Instantaneous temperature field scaled by the mean inlet temperature and obtained

    on (a) the coarse grid, (b) the intermediate one and (c) the fine one.

    39

  • 7/29/2019 TR_CFD_08_36.pdf

    40/60

    (a)

    5000

    3750

    2500

    1250

    0

    Reaction rate[mol.m-3.s-1]

    (b)

    3

    2.25

    1.5

    0.75

    0

    Equivalence ratio []

    (c)

    5000

    3750

    2500

    1250

    0

    Reaction rate [mol.m-3.s-1]

    (d)

    3

    2.25

    1.5

    0.75

    0

    Equivalence ratio []

    (e)

    5000

    3750

    2500

    1250

    0

    Reaction rate [mol.m-3.s-1]

    (f)

    3

    2.25

    1.5

    0.75

    0

    Equivalence ratio []

    Fig. 7. Instantaneous fields of reaction rate (a, c and e) and equivalence ratio conditioned

    by the local reaction (b, d and f) on Plane 1 obtained on the coarse grid, (a and b), the

    intermediate grid (c and d) and the fine one (e and f).

    40

  • 7/29/2019 TR_CFD_08_36.pdf

    41/60

    (a)

    35

    26.5

    18

    9.5

    1

    Thickening

    (b)

    8

    6.25

    4.5

    2.75

    1

    Efficiency

    (c)

    35

    26.5

    18

    9.5

    1

    Thickening

    (d)

    8

    6.25

    4.5

    2.75

    1

    Efficiency

    (e)

    35

    26.5

    18

    9.51

    Thickening

    (f)

    8

    6.25

    4.5

    2.751

    Efficiency

    Fig. 8. Instantaneous fields of the thickening factor (a, c and e) and the efficiency function

    (b, d and f) on Plane 1 obtained on the coarse grid (a and b), the intermediate grid (c and d)

    and the fine one (e and f).

    41

  • 7/29/2019 TR_CFD_08_36.pdf

    42/60

    (a)

    1.6

    1.2

    0.8

    0.4

    0.0Probability

    density

    []

    3.02.52.01.51.00.50.0

    Equivalence ratio []

    (b)

    1.6

    1.2

    0.8

    0.4

    0.0

    Probability

    density

    []

    3.02.52.01.51.00.50.0

    Equivalence ratio []

    (c)

    1.6

    1.2

    0.8

    0.4

    0.0

    Probability

    density

    []

    3.02.52.01.51.00.50.0

    Equivalence ratio []

    Fig. 9. Probability density function of local equivalence ratio in reacting zones (zones where

    local reaction rate is higher than a tenth of the mean volumetric reaction rate) on (a) the

    coarse grid, (b) the intermediate one and (c) the fine one.

    42

  • 7/29/2019 TR_CFD_08_36.pdf

    43/60

    (a)

    1.422

    0.711

    0

    -0.711

    -1.422

    Zone 1 Zone 2

    Zone 3

    U /Uinlet

    "[ ]

    (b)

    1.422

    0.711

    0

    -0.711

    -1.422

    U /Uinlet

    "[ ]

    (c)

    1.422

    0.711

    0

    -0.711

    -1.422

    U /Uinlet

    "[ ]

    Fig. 10. Mean axial velocity field (scaled by the mean velocity at the pre-vaporizer inlet)

    obtained on (a) the coarse grid, on (b) the intermediate grid and on (c) the fine grid.

    43

  • 7/29/2019 TR_CFD_08_36.pdf

    44/60

    (a)

    4.464

    3.598

    2.732

    1.866

    1

    Zone 1 Zone 2

    Zone 3

    T /Tinlet

    "[ ]

    (b)

    4.464

    3.598

    2.7321.866

    1

    T /Tinlet

    "[ ]

    (c)

    4.464

    3.598

    2.732

    1.866

    1

    T /Tinlet

    "[ ]

    Fig. 11. Mean temperature field (scaled by the mean temperature at the pre-vaporizer inlet)

    obtained on (a) the coarse grid, on (b) the intermediate grid and on (c) the fine grid.

    44

  • 7/29/2019 TR_CFD_08_36.pdf

    45/60

    (a)

    -5

    0

    5

    Verticaldistancey/R

    c

    ane

    [-]

    -2.50.0-0.40.4 -20

    Fine mesh

    Intermediate mesh

    Coarse mesh

    (b)

    4202.50.0 420

    Fine mesh

    Intermediate mesh

    Coarse mesh

    -5

    0

    5

    Verticaldistan

    cey/R

    cane

    [-]

    Fig. 12. Transverse profiles in Plane 1, Fig. 3, passing through one cane outlet: (a) mean

    axial velocity profiles scaled by the pre-vaporizer inlet velocity and (b) mean temperature

    profiles scaled by by the pre-vaporizer inlet temperature.

    45

  • 7/29/2019 TR_CFD_08_36.pdf

    46/60

    (a)

    5000

    3750

    2500

    1250

    0

    Reaction rate[mol.m-3.s-1]

    (b)

    5000

    3750

    2500

    1250

    0

    Reaction rate [mol.m-3.s-1]

    (c)

    5000

    3750

    2500

    1250

    0

    Reaction rate [mol.m-3.s-1]

    Fig. 13. Mean reaction rate on Plane 1 obtained on (a) the coarse grid, (b) the intermediate

    one and (c) the fine one.

    46

  • 7/29/2019 TR_CFD_08_36.pdf

    47/60

    (a)

    1.193

    0.925

    0.656

    0.388

    0.119

    URMS

    /Uinlet

    "[ ]

    (b)

    1.193

    0.925

    0.6560.388

    0.119

    URMS

    /Uinlet

    "[ ]

    (c)

    1.193

    0.925

    0.656

    0.388

    0.119

    URMS

    /Uinlet

    "[ ]

    Fig. 14. Fields of RMS resolved axial velocity on (a) the coarse grid, (b) the intermediate

    grid and (c) the fine grid. The quantities are non-dimensionalized by pre-vaporizer inlet

    conditions.47

  • 7/29/2019 TR_CFD_08_36.pdf

    48/60

    (a)

    1

    0.775

    0.55

    0.225

    0.1

    TRMS

    /Tinlet

    "[ ]

    (b)

    1

    0.775

    0.550.225

    0.1

    TRMS

    /Tinlet

    "[ ]

    (c)

    1

    0.775

    0.55

    0.325

    0.1

    TRMS

    /Tinlet

    "[ ]

    Fig. 15. Fields of RMS resolved temperature on (a) the coarse grid, (b) the intermediate

    grid and (c) the fine grid. The quantities are non-dimensionalized by pre-vaporizer inlet

    conditions.48

  • 7/29/2019 TR_CFD_08_36.pdf

    49/60

    (a)

    -5

    0

    5

    Verticaldistancey/R

    ca

    ne

    [-]

    0.60.01.00.00.50.0

    Fine mesh

    Intermediate mesh

    Coarse mesh

    (b)

    1.00.00.80.40.00.80.40.0

    Fine mesh

    Intermediate mesh

    Coarse mesh

    -5

    0

    5

    Verticaldistan

    cey/R

    cane

    [-]

    Fig. 16. Transverse profiles in Plane 1 (cfFig. 3), passing through one cane outlet: (a) RMS

    axial velocity profiles scaled by squared inlet velocity and (b) RMS temperature profiles

    scaled by squared inlet temperature.

    49

  • 7/29/2019 TR_CFD_08_36.pdf

    50/60

    (a)

    1

    0.75

    0.5

    0.25

    0

    M criterion

    (b)

    1

    0.75

    0.50.25

    0

    M criterion

    (c)

    1

    0.75

    0.5

    0.25

    0

    M criterion

    Fig. 17. Spatial evolution of the temporal average of the resolution criterion M: (a) is the

    coarse grid, (b) is the intermediate grid and (c) is the fine grid. Black isoline corresponds

    to M= 0.2.50

  • 7/29/2019 TR_CFD_08_36.pdf

    51/60

    (a)

    1

    0.75

    0.5

    0.25

    0

    Mzcriterion

    (b)

    1

    0.75

    0.50.25

    0

    Mzcriterion

    (c)

    1

    0.75

    0.5

    0.25

    0

    Mzcriterion

    Fig. 18. Spatial evolution of the temporal average of the resolution criterion MZ: (a) is the

    coarse grid, (b) is the intermediate grid and (c) is the fine grid. Black isoline corresponds

    to M= 0.2.51

  • 7/29/2019 TR_CFD_08_36.pdf

    52/60

    (a)

    0.0658

    0.051

    0.0362

    0.0214

    0.0066

    PRMS

    /Pinlet

    "[ ]

    (b)

    0.0658

    0.051

    0.03620.0214

    0.0066

    PRMS

    /Pinlet

    "[ ]

    (c)

    0.0658

    0.051

    0.0362

    0.0214

    0.0066

    PRMS

    /Pinlet

    "[ ]

    Fig. 19. RMS pressure field obtained on (a) the coarse grid, (b) the intermediate grid and

    (c) the fine grid. The quantities are non-dimensionalized by pre-vaporizer inlet conditions.

    52

  • 7/29/2019 TR_CFD_08_36.pdf

    53/60

    (a)

    20000

    15000

    10000

    5000

    0

    Amplitude[Pa.s

    -1]

    1600012000800040000

    Frequency [Hz]

    Coarse meshIntermediate meshFine mesh

    (b)

    30000

    20000

    10000

    0A

    mplitude[Pa.s

    -1]

    1600012000800040000

    Frequency [Hz]

    Coarse meshIntermediate meshFine mesh

    Fig. 20. Spectra of the fluctuating pressure temporal evolution at (a) Probe 1 and (b) Probe

    2 (cf Fig. 3).

    53

  • 7/29/2019 TR_CFD_08_36.pdf

    54/60

    ADDITIONALS

    1 Color version of Fig. 5: i.e. Instantaneous axial velocity field

    scaled by the mean inlet velocity and obtained on (a) the coarse

    grid, (b) the intermediate one and (c) the fine one. 56

    2 Color version of Fig. 6: i.e. Instantaneous temperature field scaled

    by the mean inlet temperature and obtained on (a) the coarse grid,

    (b) the intermediate one and (c) the fine one. 57

    3 Color version of Fig. 7: i.e. Instantaneous fields of reaction rate (a,

    c and e) and equivalence ratio conditioned by the local reaction

    (b, d and f) on Plane 1 obtained on the coarse grid, (a and b), the

    intermediate grid (c and d) and the fine one (e and f). 58

    4 Color version of Fig. 8: i.e. Instantaneous fields of the thickening

    factor (a, c and e) and the efficiency function (b, d and f) on Plane

    1 obtained on the coarse grid (a and b), the intermediate grid (c

    and d) and the fine one (e and f). 59

    5 Color views of the instantaneous fields of temperature on Plane 1

    along with an isosurface of reaction rate (in blue) obtained on the

    coarse grid (a), the intermediate grid (b) and the fine one (c). 60

    54

  • 7/29/2019 TR_CFD_08_36.pdf

    55/60

    ADDITIONALS

    55

  • 7/29/2019 TR_CFD_08_36.pdf

    56/60

    (a)

    (b)

    (c)

    Additional 1: Color version of Fig. 5: i.e. Instantaneous axial velocity field scaled

    by the mean inlet velocity and obtained on (a) the coarse grid, (b) the intermediate

    one and (c) the fine one.

    56

  • 7/29/2019 TR_CFD_08_36.pdf

    57/60

    (a)

    (b)

    (c)

    Additional 2: Color version of Fig. 6: i.e. Instantaneous temperature field scaled by

    the mean inlet temperature and obtained on (a) the coarse grid, (b) the intermediate

    one and (c) the fine one.

    57

  • 7/29/2019 TR_CFD_08_36.pdf

    58/60

    (a) (b)

    (c) (d)

    (e) (f)

    Additional 3: Color version of Fig. 7: i.e. Instantaneous fields of reaction rate (a, c

    and e) and equivalence ratio conditioned by the local reaction (b, d and f) on Plane

    1 obtained on the coarse grid, (a and b), the intermediate grid (c and d) and the fine

    one (e and f).

    58

  • 7/29/2019 TR_CFD_08_36.pdf

    59/60

    (a) (b)

    (c) (d)

    (e) (f)

    Additional 4: Color version of Fig. 8: i.e. Instantaneous fields of the thickening

    factor (a, c and e) and the efficiency function (b, d and f) on Plane 1 obtained on

    the coarse grid (a and b), the intermediate grid (c and d) and the fine one (e and f).

    59

  • 7/29/2019 TR_CFD_08_36.pdf

    60/60

    (a)

    (b)

    (c)

top related