17.01.2014 Views

NUMERICAL SIMULATION OF LOW-SPEED STALL AND ... - IAG

NUMERICAL SIMULATION OF LOW-SPEED STALL AND ... - IAG

NUMERICAL SIMULATION OF LOW-SPEED STALL AND ... - IAG

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Conference on Modelling Fluid Flow (CMFF’12)<br />

The 15 th International Conference on Fluid Flow Technologies<br />

Budapest, Hungary, September 4-7, 2012<br />

<strong>NUMERICAL</strong> <strong>SIMULATION</strong> <strong>OF</strong> <strong>LOW</strong>-<strong>SPEED</strong> <strong>STALL</strong> <strong>AND</strong> ANALYSIS <strong>OF</strong><br />

TURBULENT WAKE SPECTRA<br />

Philipp Peter GANSEL 1 , Sebastian ILLI 2 , Thorsten LUTZ 3 , Ewald KRÄMER 4<br />

1 Corresponding Author. Institute of Aerodynamics and Gas Dynamics, University of Stuttgart. Pfaffenwaldring 21, 70569 Stuttgart, Germany.<br />

Tel.: +49 711 685 63416, Fax: +49 711 685 63438, E-mail: gansel@iag.uni-stuttgart.de<br />

2 E-mail: illi@iag.uni-stuttgart.de<br />

3 Head of branch “Aircraft aerodynamics“. E-mail: lutz@iag.uni-stuttgart.de<br />

4 Head of Institute. E-mail: kraemer@iag.uni-stuttgart.de<br />

ABSTRACT<br />

At the Institute of Aerodynamics and Gas<br />

Dynamics (<strong>IAG</strong>) numerical studies on the low-speed<br />

stall of aircraft wings, the development of the wake<br />

and its interaction with the empennage are performed.<br />

In the present paper unsteady Reynolds-averaged<br />

Navier-Stokes (URANS) calculations as well as highly<br />

resolved Delayed Detached-Eddy Simulations (DDES)<br />

of the turbulent wake downstream of a boundary<br />

layer separation on an airfoil section are described.<br />

For these simulations a NACA 0012 airfoil test case<br />

with a free stream Mach number of 0.3 and a<br />

Reynolds number of 6 million at a high angle of<br />

attack was chosen. URANS calculations on different<br />

grids and with different turbulence models show that<br />

a fine structured discretisation of the wake region<br />

facilitates the transport of turbulent viscosity with low<br />

dissipation. The DDES approach allows propagation<br />

of resolved turbulent structures far downstream of<br />

the flow separation. The spectral analysis shows a<br />

good agreement of measured and simulated pressure<br />

fluctuations in the junction of energy-containing range<br />

and inertial sub-range of the spectrum. Furthermore<br />

semi-empirical turbulence spectrum models were<br />

used to obtain spectral turbulence information from<br />

URANS solutions. The resulting turbulent velocity<br />

fluctuation spectra are compared to the resolved<br />

velocity fluctuations of the DDES.<br />

Keywords: CFD, DDES, stall, spectra, wake<br />

NOMENCLATURE<br />

E 11 [m 3 /s 2 ] one-dimensional longitudinal<br />

turbulent kinetic energy spectrum<br />

F+ [−] reduced frequency<br />

L 11 [m] longitudinal integral length scale<br />

Ma [−] Mach number<br />

R 11 [m 2 /s 2 ] spatial velocity autocorrelation<br />

Re [−] chord Reynolds number<br />

U ∞ [m/s] free stream velocity<br />

c [m] airfoil chord length<br />

c l [−] lift coefficient<br />

c p [−] static pressure coefficient<br />

k [m 2 /s 2 ] specific turbulent kinetic energy<br />

r [m] distance<br />

t [s] time<br />

u [m/s] velocity inx-direction<br />

x, y,z [m] space coordinates<br />

y 1 + [-] dimensionless wall coordinate<br />

α [ ◦ ] angle of attack<br />

∆ [m] maximal cell edge length<br />

in the structured wake mesh<br />

ε [m 2 /s 3 ] dissipation rate ofk<br />

κ 1 [1/m] longitudinal wave number<br />

κ e [1/m] wave number of the most<br />

energy containing eddies<br />

ω [1/s] turbulent dissipation frequency<br />

Subscripts and Superscripts<br />

¯ temporal mean<br />

′<br />

fluctuation<br />

1. INTRODUCTION<br />

At the boarders of the flight envelope flow effects<br />

occur which still are not completely understood and far<br />

from being reproduced correctly by today’s numerical<br />

methods in industry relevant flow cases. At low Mach<br />

numbers and sufficiently high angles of attack large<br />

boundary layer separations on the wing cause a rapid<br />

lift drop and form a highly unsteady turbulent wake.<br />

This phenomenon called low-speed stall is the subject<br />

of the presented studies, with a focus on the spectral<br />

information in the wake downstream of the separation.<br />

The fluctuations can affect the inflow conditions and<br />

the flow state of the safety critical horizontal tail plane<br />

(htp) [1] or even excite its structural modes.<br />

However, only few published studies concerning<br />

transport aircraft configurations deal with turbulent<br />

fluctuations in the wake. Whitney et al. [2] developed<br />

199


an URANS based model to predict htp yaw and roll<br />

moment fluctuations measured near low-speed stall.<br />

In the German HINVA research project in-flight stall<br />

tests with the DLR ATRA aircraft are planned to<br />

collect amongst others unsteady pressure signatures<br />

on wing and htp as a validation basis for unsteady<br />

CFD methods [3]. A couple of investigations of<br />

turbulent wake fluctuations have been published in<br />

the context of active flow control devices. Seifert<br />

and Pack [4] evaluated the spectral content of the<br />

turbulent wake of a stalling airfoil with and without<br />

periodic blowing near the leading edge by means<br />

of unsteady pressure transducers. Wokoeck et<br />

al. [5] used PIV measurements of a stalling tail<br />

plane airfoil to validate RANS results with different<br />

transition prediction methods. Lopez Mejia et al.<br />

[6] explored time-averaged and instantaneous vorticity<br />

fields around an airfoil with tangentially aligned<br />

synthetic jet actuators using an incompressible DDES<br />

code (EDDES) and PIV measurements. Since hybrid<br />

RANS-LES methods as DES were developed to model<br />

massive separation, at first applications concentrated<br />

on flows around blunt bodies or airfoils in deep stall<br />

[7–9]. Beyond that, Onera performed zonal and<br />

delayed DES on three-element airfoils [10–12] and<br />

on a civil aircraft type configuration [13]. Nagy<br />

et al. [14] conducted zonal RANS-LES and LDA<br />

measurements of a fan airfoil with intent to calculate<br />

the flow generated noise. In their numerical studies<br />

Durrani and Qin [15] compared Reynolds stresses in<br />

the boundary layer and the near wake region of an<br />

airfoil from DDES, DES, URANS and experiments.<br />

Probst and Radespiel [16] applied different DES<br />

approaches and showed deficiencies in the prediction<br />

of mild trailing edge separation. Stall of a NACA<br />

0012 airfoil was investigated by Lehmkuhl et al. [17]<br />

by comparing Direct Numerical Simulation and Large<br />

Eddy Simulation (LES) results.<br />

In the presented research three different ways to<br />

assess the spectrum of turbulent motions in separated<br />

airfoil flow are tested and compared: unsteady RANS<br />

calculations with evaluation of the resolved large<br />

scale fluctuations of the flow variables, DDES with<br />

fine resolution of turbulent fluctuations down to the<br />

dissipation range and a tubulence spectrum model<br />

based on URANS turbulence model entities.<br />

2. FUNDAMENTALS<br />

2.1. Turbulent Spectrum Model<br />

Proccesses in turbulent flows at high Reynolds<br />

numbers can be described by the concept of the energy<br />

cascade [18]. It states that turbulence consists of eddies<br />

of different sizes. Energy is entered into turbulence at<br />

large scales driven by the boundary conditions given as<br />

the adjacent flow field, then transferred to successively<br />

smaller eddies and finally dissipated to heat at the<br />

smallest scale, called the Kolmogorov scale. The so<br />

called Kolmogorov hypotheses state that at high Re<br />

i) small scale turbulence is locally isotropic, ii) the<br />

statistics of turbulent motions are of universal form<br />

and uniquely determined by turbulent dissipation rate<br />

ε and viscosity and iii) in a certain range of scale<br />

the statistics of turbulent motions are independent<br />

of viscosity as well and thus solely defined by ε.<br />

Accordingly the turbulent spectrum can be devided<br />

into the energy-containing range, where big eddies are<br />

formed and depend on the flow conditions, the inertial<br />

subrange, where energy is passed on to smaller scales<br />

by inviscid processes and the dissipation range where<br />

viscous effects dominate. The universal functional<br />

form of the turbulent spetrum allows to construct<br />

models independent of the particular class of flow.<br />

Von Kármán developed a turbulent energy spectrum<br />

model, wherefrom Lysack and Brungart [19] derived<br />

the one-dimensional longitudinal spectrum<br />

E 11 (κ 1 ) = 8.73<br />

55<br />

[<br />

k<br />

κ e<br />

(<br />

κ1<br />

1+<br />

κ e<br />

) 2<br />

] 5/6<br />

(1)<br />

with the most energy containing eddies’ wave number<br />

κ e = 1.9 ε<br />

k 3/2 (2)<br />

and the turbulent kinetic energy k. When derived<br />

from a time and space dependent data sample,E 11 (κ 1 )<br />

equals twice the Fourier transform of the velocity<br />

fluctuation two-point auto-correlation<br />

R 11 (r,t) = u ′ (x+r,t)·u ′ (x,t) . (3)<br />

2.2. Numerical Methods<br />

The calculation of the fluid flow describing<br />

incompressible Navier-Stokes equations for industry<br />

relevant cases with complex geometries and high<br />

Reynolds numbers today is still not affordable, since<br />

the number of grid points required for sufficient<br />

resolution and the CPU-time needed for computation<br />

are proportional to Re 9/4 and Re 3 respectively [20].<br />

One approach is to decompose the flow variables in the<br />

Navier-Stokes equations into a temporal mean value<br />

and a time-dependent fluctuation. This leads to the<br />

RANS equations which are solved for the mean values.<br />

The remaining fluctuation-dependent terms (called<br />

Reynolds stresses) are modelled. To determinate the<br />

Reynolds stresses eddy viscosity turbulence models<br />

are widely used. For the URANS simulations<br />

conducted in the presented studies mainly the Menter<br />

SST two-equation turbulence [21] model was used.<br />

It provides transport equations for k and either its<br />

dissipation rate ε (in the free flow) or the dissipation<br />

frequency ω = ε/k (in near wall flow).<br />

If turbulent fluctuations shall be resolved in the<br />

simulations RANS methods are not applicable by<br />

definition. DES became popular for flows with large<br />

scale boundary layer separation in the recent years.<br />

200


It combines the advantage of LES to resolve the big<br />

turbulent structures with relatively small grid sizes.<br />

The outer part of the separated flow is calculated with<br />

LES, where the total grid points scale with Re 0.4<br />

[20]. In the boundary layer, where LES would require<br />

much more cells, RANS equations are applied. In<br />

contrast to RANS, in LES the Navier-Stokes equations<br />

are not averaged in time but spatially filtered, where<br />

the filter width is proportional to the biggest cell<br />

extension. In the equations terms depending on the<br />

residual fluctuations filtered out remain according to<br />

the Reynolds stresses in RANS and are named residual<br />

or subgrid scale (SGS) stresses. Again they can be<br />

modelled using eddy viscosity models, of which the<br />

Smagorinsky model is used in the present study.<br />

Spalart et al. [22] presented an implementation of<br />

DDES where the SA turbulence model is used in the<br />

RANS domain and the SGS model is formulated in<br />

such a way that the transport equation can be transfered<br />

into the SA model equation by replacing the local<br />

grid size by the distance to the nearest wall. A flow<br />

dependent blending function controls the application<br />

of the different approaches. Nevertheless the adequate<br />

selection of the required spatial grid resolution is<br />

crucial for correct operation of the hybrid scheme.<br />

3. CONDUCTED <strong>SIMULATION</strong>S<br />

3.1. Test Case Description<br />

As a test case wind tunnel tests were chosen<br />

from literature with available unsteady measurements<br />

in the wake [4]. The analysed NACA 0012 airfoil<br />

section with a chord length c of 0.165m was tested<br />

in a transonic cryogenic wind tunnel under free stream<br />

conditions of Ma = 0.3 and Re = 6·10 6 . The data<br />

chosen for the presented simulations were recorded at<br />

an angle of attack of α = 16 ◦ , where a big turbulent<br />

separation area occurs on the airfoil.<br />

Due to the high memory consumption of full flow<br />

field solutions, the spectral analysis, which all time<br />

steps are required for, was conducted only in three<br />

defined points listed in Table 1. Position A is very close<br />

to the trailing edge and position B lies approximately<br />

in the middle of the wake at a downstream position.<br />

In the experiments unsteady pressure transducers were<br />

placed in the wake at position C. The origin of the used<br />

coordinate system is placed at the airfoil leading edge,<br />

the x-axis points towards the trailing edge, the y-axis<br />

in span wise direction and thez-axis upwards.<br />

Table 1. Analysed measurement positions<br />

Position A B C<br />

x/c 1.2121 3.3750 3.3750<br />

z/c 0.12121 0.60606 1.2198<br />

3.2. Numerical Setup<br />

The presented simulations have been performed<br />

using the unstructured finite volume code TAU (version<br />

2011.1.0) from DLR (German Aerospace Center) [23].<br />

It can handle the mixed-type element hybrid meshes<br />

created with the meshing software Gridgen V15.17.<br />

For the URANS calculations the computational<br />

domain is a two-dimensional (2d) hybrid mesh with<br />

a C-type boundary layer mesh, a big structured wake<br />

area and an unstructured farfield region. The boundary<br />

layer grids of the three resolution levels fine, medium<br />

and coarse have 60 cells in wall normal direction with<br />

a first cell size of y 1 + ≤ 1 independently of the grid<br />

resolution. The wake mesh is directly attached to the<br />

boundary layer mesh with a smooth transition of cell<br />

size. Table 2 shows cell sizes in the homogeneously<br />

structured wake region, overall and surface point<br />

counts for the different grid resolutions. To save<br />

computational costs, the wake mesh is restricted to the<br />

part of the flow field, where the measurement positions<br />

are defined and the turbulent wake is assumed to pass<br />

through (Figure 1). The farfield boundary distance<br />

from the airfoil geometry is 100c.<br />

Table 2. Point counts and wake region cell size ∆ of<br />

the used RANS and DES meshes<br />

RANS<br />

DES<br />

fine medium coarse<br />

points 219000 94000 47000 17.8·10 6<br />

surf. pts. 445 243 139 445×81<br />

∆ 0.5%c 1.0%c 2.0%c 0.5%c<br />

U ∞<br />

unstructured<br />

tetrahedral mesh<br />

A<br />

structured<br />

hexahedral mesh<br />

Figure 1. Computational domain with the analysed<br />

measurement positions A, B and C<br />

In the URANS simulations the second order<br />

Jameson central discretisation scheme was used with<br />

matrix dissipation for stabilisation. The dual time<br />

stepping method was applied with a physical time<br />

step size of 25µs which is approximately 1/100 times<br />

the convective time scale c/U ∞ . Amongst others<br />

the one equation turbulence model from Spalart and<br />

Allmaras [24] in its original formulation (SAO), the<br />

shear stress transport model (SST) from Menter [21]<br />

and the Reynolds stress model (RSM) implemented in<br />

TAU [23] were used as turbulence models. For DDES<br />

calculations the structured parts of the fine 2d URANS<br />

mesh were extruded span wise by 40%c with the same<br />

cell size ∆ in y-direction yielding three-dimensional<br />

(3d) isotropic structured cells in the wake region. The<br />

farfield mesh is filled up with tetrahedra. On the<br />

span wise boundary planes periodic conditions were<br />

C<br />

B<br />

201


applied to simulate an infinite wing without any tip or<br />

wind tunnel side wall effects. The distance between<br />

the periodic planes was chosen as big as possible<br />

under consideration of limited computational costs to<br />

minimise the error introduced through the forced span<br />

wise periodicity of the flow. The numerical schemes<br />

were the same as applied to URANS. During one time<br />

step of 2.5µs the flow passes the shortest cells in<br />

the wake which, in combination with the isotropically<br />

structured mesh, ensures a low dissipation in the<br />

propagation of the turbulent wake structures.<br />

4. RESULTS<br />

4.1. URANS Simulations<br />

The analysis of the 2d URANS simulations<br />

concentrates on the SST turbulence model because<br />

it is used for the semi-empirical turbulent fluctuation<br />

spectrum model later. The results with the other<br />

eddy-viscosity turbulence models are comparable. In<br />

the simulations the boundary layer flow separates from<br />

the suction side of the airfoil, as expected from the<br />

compared experimental data. The unsteady flow is<br />

exactly periodic with a very long main cycle of about<br />

11.5 convective time scales (see Figure 2). Over a<br />

big part of this period the separated flow reattaches to<br />

the upper surface of the airfoil and forms a separation<br />

bubble (see Figure 3 a)). Gradually the separation point<br />

is moving upstream and the vortex inside the bubble is<br />

getting stronger. This becomes significant in the rising<br />

lift coefficient during this phase. There is virtually no<br />

large-scale turbulent movement in the wake.<br />

1.6<br />

1.4<br />

c l<br />

1.2<br />

1<br />

0.8<br />

0 5 10 15 20 25 30 35 40<br />

t U c ∞<br />

Figure 2. c l time series of a NACA 0012 airfoil at<br />

16 ◦ angle of attack (URANS SST)<br />

At one moment a counter rotating vortex develops<br />

behind the separation bubble and a shedding process<br />

begins which means a drastic decrease of lift (see<br />

Fig. 2). The streamlines in Fig. 3 b) show the flow<br />

state shortly after this event. Finally several alternately<br />

rotating vortices are shed from the airfoil, the flow<br />

reattaches in the upstream part due to the reduced<br />

circulation and the remaining small separation bubble<br />

starts again to grow and increase the lift (see Fig. 3 c)).<br />

The time-averaged pressure distribution on the<br />

airfoil simulated with the three different turbulence<br />

models is compared to experimental data in Figure 4.<br />

The numerical results lie very close to each other.<br />

However, the measurements of Seifert and Pack [4] are<br />

qualitatively different. In contrast to the simulations<br />

a)<br />

b)<br />

c)<br />

Figure 3. Instantaneous streamlines at a) t U∞ c<br />

=<br />

14.5, b) 17.5 and c) 19 as highlighted with◦in Fig. 2<br />

the airfoil flow is separated over the whole chord<br />

length which is expressed in the lower pressure in<br />

the rear part of the airfoil. The drop of circulation<br />

causes a decrease of the leading edge suction peak<br />

and overall lift level. Yet there are older wind tunnel<br />

tests conducted by Ladson et al. [25] in the same<br />

facility that fit the simulations very good. This leads<br />

to the assumption that the test case is exactly at the<br />

boarder between those two flow patterns and little<br />

disturbances can determine it. Since it is known<br />

that eddy-viscosity turbulence models tend to predict<br />

boundary layer separation too late in the sense of<br />

higher incidence, some calculations at increased α<br />

were accomplished. The surface pressure distribution<br />

at α = 18.5 ◦ was found to be in very good agreement<br />

with the experiment of Seifert and Pack (see Fig. 4).<br />

-6<br />

-5<br />

-4<br />

c<br />

-3<br />

p<br />

-2<br />

-1<br />

0<br />

1<br />

16° Exp. Seifert<br />

16° Exp. Ladson<br />

16° C-grid URANS SAO<br />

16° C-grid URANS SST<br />

16° C-grid URANS RSM<br />

18.5° C-grid URANS SST<br />

0 0.2 0.4 0.6 0.8 1<br />

x/c<br />

Figure 4. Fine grid URANS mean pressure<br />

distributions compared to experimental data<br />

In the comparison of the pressure distribution<br />

simulated on the different grids (Figure 5) it can be<br />

202


seen that with increasing spatial resolution the pressure<br />

on the suction side in front of about 25%c is increasing<br />

a bit, but decreasing behind that point. On the pressure<br />

side there are no noticeable differences.<br />

-6<br />

-5<br />

-4<br />

-3<br />

c p<br />

-2<br />

-1<br />

0<br />

1<br />

Exp. Ladson<br />

SST ∆ = 0.5% c<br />

SST ∆ = 1.0% c<br />

SST ∆ = 2.0% c<br />

0 0.2 0.4 0.6 0.8 1<br />

x/c<br />

Figure 5. URANS meanc p with diff. resolutions<br />

The resolved pressure fluctuation spectra at<br />

position C in the simulation on the fine and medium<br />

grids are dominated by the main flow oscillation seen<br />

above in the lift characteristics and its higher harmonics<br />

(see Figure 6). The bad results on the coarse mesh<br />

can be explained with high numerical dissipation of the<br />

fluctuations in the insufficiently resolved wake mesh,<br />

since the pressure distributions on the airfoil surface<br />

are almost equal for all URANS grids. The medium<br />

grid simulation seems to match the experimental data<br />

in its peaks, but in terms of grid convergence the<br />

peaks of the fine mesh simulations overestimate the<br />

amplitudes. Compared to the URANS calculations<br />

the experiment shows a smoother spectrum without<br />

harmonic peaks. This fact suggests a more continuous<br />

vortex shedding progress in the experiment than the<br />

URANS simulations show.<br />

c p<br />

’<br />

Position C<br />

Exp. Seifert<br />

10 -1 x = 3.375c<br />

SST ∆ = 0.5%c<br />

z = 1.220c<br />

SST ∆ = 1.0%c<br />

SST ∆ = 2.0%c<br />

10 -3<br />

10 -5<br />

10 -7<br />

10 -9<br />

10 -11<br />

10 -1 10<br />

F+<br />

0 10 1<br />

Figure 6. Pressure fluctuations resolved by URANS<br />

on fine, medium and coarse meshes<br />

4.2. DDES Calculations<br />

The DDES calculation produces a physically more<br />

reasonable flow pattern with continuously shed vortices<br />

of different scales. The 3d vortices can be visualised<br />

using the λ 2 criterion as illustrated in Figure 7. It is<br />

evident that the fine turbulent structures dissipate as<br />

soon as they leave the structured wake mesh and enter<br />

the coarser areas of the tetrahedra.<br />

Figure 7. λ 2 iso-surface of turbulenct wake (DDES)<br />

The time-averagedc p distribution from the DDES<br />

shows an earlier and more clearly visible separation<br />

point (see Figure 8) than the URANS. It is<br />

remarkable that even the RANS simulation with the<br />

SAO turbulence model produces a different pressure<br />

distribution than DDES, where in the boundary layer<br />

enclosing RANS area the same turbulence model is<br />

used. This shows the feedback of the separated flow<br />

on the boundary layer state.<br />

-6<br />

-5<br />

-4<br />

-3<br />

c p<br />

-2<br />

-1<br />

0<br />

1<br />

Exp. Ladson<br />

SAO DDES<br />

URANS SST<br />

URANS SAO<br />

0 0.2 0.4 0.6 0.8 1<br />

x/c<br />

Figure 8. DDES span wise averaged mean c p<br />

compared to URANS and experiment<br />

To give a better insight in the differences of<br />

URANS and DDES simulated flow, the instantaneous<br />

vorticity in a x-z plane is plotted in Figure 9 (for the<br />

URANS plot the same time step was chosen as shown<br />

in Fig. 3 c)). While the URANS results show few<br />

big discrete vortices the DDES generates structures of<br />

different sizes. Furthermore small vortex cores of high<br />

vorticity are preserved longer.<br />

The pressure fluctuation spectrum derived from<br />

DDES time series at position C is in good agreement<br />

with the measurement data (see Figure 10). The<br />

spectrum has no outstanding peaks like the URANS<br />

spectrum in Fig. 6 which confirms the assumption<br />

about the character of the DDES calculated flow.<br />

The spectra derived at positions A and B are plotted<br />

additionally in Fig. 10. The spectrum at position A<br />

shows the highest amplitudes due to its proximity to<br />

the airfoil trailing edge where the shedding turbulent<br />

203


medium mesh results in Fig. 6) the same characteristics<br />

are expected as with the fine DDES mesh. According<br />

studies also discussing different DES formulations are<br />

to be published by Illi et al. [26].<br />

Another approach to gain information about the<br />

turbulent spectrum is to perform the relatively cheap<br />

URANS and model a turbulence spectrum from the<br />

turbulent quantities calculated by the turbulence model.<br />

This is done for the presented test case in the following<br />

section.<br />

Figure 9. Vorticity in x-z plane from URANS SST<br />

(upper picture) and DDES (lower picture)<br />

vortices are formed driven by the boundary conditions.<br />

In position B the fluctuation level is already about one<br />

order of magnitude smaller because of the decaying<br />

processes in the energy cascade and the spreading of<br />

the wake while propagated downstream. In point C the<br />

amplitudes are even smaller, although in this position<br />

the flow has passed roughly the same distance from<br />

the airfoil as in point B. However, point C lies close to<br />

the boarder of the wake to the outer free stream region<br />

where the turbulence intensity is significant lower.<br />

10 -3<br />

4.3. Modelled Fluctuation Spectra<br />

For an applicability assessment of the velocity<br />

spectrum model presented by Lysack and Brungart<br />

[19] for separated wake flow first the turbulent kinetic<br />

energy k and its dissipation range ε were extracted<br />

from URANS solutions. For comparison k = 0.5 ·<br />

(u ′2 + v ′2 + w ′2 ) also has been calculated from the<br />

DDES results and listed in Table 3. Here the amount of<br />

turbulence modelled by the SGS model is neglected.<br />

Table 3. Time-averaged turbulent kinetic energy k<br />

from URANS SST and DDES<br />

k [m 2 /s 2 ] pos. A pos. B pos. C<br />

DDES 678.28 43.670 0.041669<br />

URANS fine 264.90 49.921 6.3069<br />

URANS medium 301.34 44.718 0.33582<br />

URANS coarse 398.00 46.848 0.00047421<br />

10 -2 Exp. Seifert Pos. C<br />

c’ p<br />

10 -4<br />

10 -5<br />

10 -6<br />

10 -7<br />

DDES ∆ = 0.5%c Pos. A<br />

DDES ∆ = 0.5%c Pos. B<br />

DDES ∆ = 0.5%c Pos. C<br />

10 -1 10 0 10<br />

F+<br />

1 10 2<br />

Figure 10. Pressure fluctuations resolved by DDES<br />

in wake positions A, B and C<br />

It should be noted in this place that the simulated<br />

physical time covered by the DDES calculation is<br />

much shorter than in the URANS cases. A total of<br />

6000 physical time steps with 200 inner iterations each<br />

were computed within the DDES on 500 computing<br />

cores in parallel mode. The last 2000 time steps<br />

were considered for spectral analysis after the time<br />

averaged flow quantities have reached convergence.<br />

Due to the big grid, the small time step size<br />

and limited computational resources the spectrum is<br />

limited to relatively high frequencies, since the lowest<br />

representable frequency is reciprocally proportional to<br />

the total time length of the evaluated signal.<br />

One possibility to overcome this problem without<br />

adding computational time is to perform the DDES<br />

on a coarser mesh with less points and thus bigger<br />

time steps. Like shown for URANS (compare fine and<br />

The values show that close to the trailing edge<br />

(position A) the URANS simulations give a much<br />

lower turbulence level than the DDES. Additionally<br />

k is increasing with decreasing spatial resolution. At<br />

position B in the middle of the wake downstream<br />

of the airfoil k is very comparable for all presented<br />

simulations. At the boarder of the turbulent wake<br />

in point C k is very low and the relative differences<br />

between the methods and grid resolutions are in the<br />

range of several orders of magnitude.<br />

Utilising Taylor’s hypothesis of frozen turbulence<br />

the time dependent velocity data logged at A, B and C<br />

during the DDES calculation can be transferred into a<br />

space domain [18]. The one-dimensional longitudinal<br />

velocity spectra E 11 (κ 1 ) from DDES are derived by<br />

duplicating and Fourier transformingR 11 (Eq. (3)). For<br />

URANS E 11 (κ 1 ) is calculated via Eq. (1).<br />

In Figure 11 E 11 (κ 1 ) at the different evaluation<br />

positions is plotted. As seen from the turbulent kinetic<br />

energy in Table 3 the modelled spectra in point A<br />

underpredict the DDES calculated amplitudes. At<br />

position B in the middle of the wake it is the other way<br />

around. In both cases the URANS modelled spectra are<br />

almost the same for all grid resolutions. Also the wave<br />

number range of the transition region between large<br />

scale range and inertial subrange which corresponds to<br />

κ e is in good agreement to the DDES spectra.<br />

In position C the large differences of k in the<br />

URANS simulations give a very dissimilar description<br />

204


a)<br />

b)<br />

c)<br />

10 0<br />

E 11<br />

10 -1<br />

DDES ∆ = 0.5% c<br />

10 2 SST ∆ = 0.5% c<br />

10 1<br />

SST ∆ = 1.0% c<br />

SST ∆ = 2.0% c<br />

10 -2<br />

10 -3<br />

10 -4<br />

10 -5<br />

10 -1<br />

10<br />

E -2<br />

11<br />

10 -3<br />

10 -4<br />

10 -5<br />

10 -6<br />

E<br />

10 -4<br />

11<br />

10 -5<br />

Position A<br />

x = 1.212c<br />

z = 0.121c<br />

10 0 10 1 10 2 κ<br />

10 3 10 4 10 5<br />

1<br />

DDES ∆ = 0.5% c<br />

10 1 SST ∆ = 0.5% c<br />

10 0<br />

SST ∆ = 1.0% c<br />

SST ∆ = 2.0% c<br />

10 -3<br />

10 -6<br />

10 -7<br />

10 -8<br />

10 -9<br />

Position B<br />

x = 3.375c<br />

z = 0.606c<br />

10 0 10 1 10 2 κ<br />

10 3 10 4 10 5<br />

1<br />

-1 DDES ∆ = 0.5% c<br />

10<br />

SST ∆ = 0.5% c<br />

10 -2<br />

SST ∆ = 1.0% c<br />

SST ∆ = 2.0% c<br />

Position C<br />

x = 3.375c<br />

z = 1.220c<br />

10 0 10 1 10 2 10 3 κ<br />

10 4 10 5 10 6<br />

1<br />

Figure 11. E 11 (κ 1 ) from DDES and turbulent<br />

fluctuation spectrum model at positions A, B and C<br />

of the flow and cannot be used for detailed analysis.<br />

Also the very low velocity fluctuations in the DDES<br />

are not useful for spectral evaluation. Nevertheless the<br />

pressure fluctuation spectrum from DDES matches the<br />

experimental measurement as shown in the previous<br />

section. This fact leads to the assumption that even<br />

if the turbulent wake passes by a certain point in such<br />

a way that the turbulent vortices are not recognisable<br />

in the local velocity, still you can find a significant<br />

footprint of the turbulent structures in the pressure.<br />

5. CONCLUSIONS<br />

The presented studies revealed the ability of<br />

URANS and DDES calculations to propagate turbulent<br />

vortical structures on structured meshes in the wake<br />

of the flow past an airfoil in low-speed stall. The<br />

dissipation of coherent turbulent structures strongly<br />

depends on the local grid cell size in the wake region.<br />

While URANS simulates big discrete periodically<br />

shed vortices, DDES shows a continuously forming<br />

of turbulent 3d structures of different scales in the<br />

wake. Hence evaluated pressure spectra from URANS<br />

simulations base on the fundamental frequency of the<br />

periodic vortex shedding pattern and its harmonics,<br />

whereas DDES calculations give a smooth spectrum<br />

which fits the experimental measurements better.<br />

The turbulent spectrum model based on URANS<br />

turbulence model entities is able to give a rough<br />

estimate of the velocity fluctuation spectrum, at least<br />

the wave number of the most energy containing eddies.<br />

Furthermore it could be observed that on the<br />

boarder of the wake virtually no meaningful velocity<br />

spectrum could be derived from the DDES results,<br />

while the measurable pressure fluctuations fit the<br />

experimental data. This can be of relevance when<br />

simulating the wake of an aircraft with boundary layer<br />

separation occurring at the inner wing sections. Even<br />

when the wake of the separation passes underneath<br />

the htp it can have an effect on the surface pressure<br />

distribution and cause unsteady fluctuations in the<br />

integral forces.<br />

ACKNOWLEDGEMENTS<br />

The presented studies are part of the HINVA<br />

and the ComFliTe project funded by the “German<br />

Federal Ministry of Economics and Technology”.<br />

Computational resources were kindly provided by the<br />

“Leibniz Supercomputing Center” Munich during the<br />

DEISA 2 call and the “High Performance Computing<br />

Center Stuttgart” within the bwGRiD project [27].<br />

REFERENCES<br />

[1] Havas, J., and Jenaro Rabadan, G., 2009,<br />

“Prediction of Horizontal Tail Plane Buffeting<br />

Loads”, Technical Report IFASD-2009-128.<br />

[2] Whitney, M. J., Seitz, T. J., and Blades,<br />

E. L., 2009, “Low-Speed-Stall Tail Buffet Loads<br />

Estimation Using Unsteady CFD”, Technical<br />

Report IFASD-2009-127.<br />

[3] Rudnik, R., Reckzeh, D., and Quest, J.,<br />

2012, “HINVA - High lift INflight VAlidation<br />

- Project Overview and Status”, Proc. 50th<br />

AIAA Aerospace Sciences Meeting including the<br />

New Horizons Forum and Aerospace Exposition,<br />

Nashville, TN, AIAA-2012-0106.<br />

[4] Seifert, A., and Pack, L. G., 1999, “Oscillatory<br />

Excitation of Unsteady Compressible Flows over<br />

Airfoils at Flight Reynolds Numbers”, Technical<br />

Report AIAA-99-0925.<br />

[5] Wokoeck, R., Grote, A., Krimmelbein, N.,<br />

Ortmanns, J., Radespiel, R., and Krumbein, A.,<br />

2006, “RANS Simulation and Experiments on the<br />

Stall Behaviour of a Tailplane Airfoil”, in New<br />

Results in Numerical and Experimental Fluid<br />

205


Mechanics V (Rath, H.-J., Holze, C., Heinemann,<br />

H.-J., Henke, R., and Hönlinger, H., Eds.), Notes<br />

Num Fluid Mech and Multidisciplinary Design,<br />

Vol. 92, pp. 208–216.<br />

[6] Lopez Mejia, O. D., Moser, R. D., Brzozowski,<br />

D. P., and Glezer, A., 2011, “Effects of<br />

Trailing-Edge Synthetic Jet Actuation on an<br />

Airfoil”, AIAA J, Vol. 49(8), pp. 1763–1777.<br />

[7] Weinman, K., 2006, “NACA0012 beyond Stall”,<br />

in FLOMANIA – A European Initiative on Flow<br />

Physics Modelling (Haase, W., Aupoix, B.,<br />

Burge, U., and Schwammborn, D., Eds.), pp.<br />

233–246.<br />

[8] Garbaruk, A., Shur, M., Strelets, M., and<br />

Travin, A., 2009, “NACA0021 at 60 Incidence”,<br />

in DESider – A European Effort on Hybrid<br />

RANS-LES Modelling (Haase, W., Braza, M., and<br />

Revell, A., Eds.), pp. 127–139.<br />

[9] Illi, S., Lutz, T., and Krämer, E., 2010,<br />

“Simulation of Pressure and Shock Induced<br />

Separation using DES Implementations in the<br />

DLR-TAU Code”, Proc. Second Symposium<br />

“Simulation of Wing and Nacelle Stall”,<br />

Braunschweig, Germany, pp. 127–139.<br />

[10] Deck, S., 2005, “Zonal Detached Eddy<br />

Simulation of the Flow around a High-Lift<br />

Configuration”, AIAA J, Vol. 43(11), pp.<br />

2372–2384.<br />

[11] Deck, S., 2009, “Three-Element Airfoil”, in<br />

DESider - A European Effort on Hybrid<br />

RANS-LES Modelling (Haase, W., Braza, M., and<br />

Revell, A., Eds.), pp. 156–163.<br />

[12] Deck, S., 2011, “Recent Improvements in<br />

the Zonal Detached Eddy Simulation (ZDES)<br />

Formulation”, Theor Comp Fluid Dyn, pp. 1–28.<br />

[13] Brunet, V., and Deck, S., 2008, “Zonal-Detached<br />

Eddy Simulation of Transonic Buffet on a Civil<br />

Aircraft Type Configuration”, in Advances in<br />

Hybrid RANS-LES Modelling (Peng, S.-H., and<br />

Haase, W., Eds.), Notes Num Fluid Mech and<br />

Multidisciplinary Design, Vol. 97, pp. 182–191.<br />

[14] Nagy, L., Lohász, M. M., Régert, T., and Vad, J.,<br />

2008, “Hybrid/Zonal RANS/LES Computation of<br />

an Airfoil”, Proc. 6th International Conference<br />

on Mechanical Engineering (Gépészet 2008),<br />

Budapest, Hungary.<br />

[15] Durrani, N., and Qin, N., 2011, “Behavior<br />

of Detached-Eddy Simulations for Mild Airfoil<br />

Trailing-Edge Separation”, J Aircraft, Vol. 48(1),<br />

pp. 193–202.<br />

[16] Probst, A., and Radespiel, R., 2010, “A<br />

Comparison of Detached-Eddy Simulation<br />

and Reynolds-Stress Modelling Applied to the<br />

Flow over a Backward-Facing Step and an<br />

Airfoil at Stall”, Proc. 48th AIAA Aerospace<br />

Sciences Meeting including the New Horizons<br />

Forum and Aerospace Exposition, Orlando, FL,<br />

AIAA-2010-920.<br />

[17] Lehmkuhl, O., Baez, A., Rodríguez, I., and<br />

Pérez-Segarra, C. D., 2011, “Direct Numerical<br />

Simulation and Large-Eddy Simulations of the<br />

Turbulent Flow around a NACA0012 Airfoil”,<br />

Proc. 7th International Conference on Comp.<br />

Heat and Mass Transfer, Istanbul, Turkey.<br />

[18] Pope, S. B., 2011, Turbulent Flows, Cambridge<br />

University Press.<br />

[19] Lysack, P. D., and Brungart, T. A., 2003,<br />

“Velocity Spectrum Model for Turbulence<br />

Ingestion Noise from Computational-Fluid-<br />

Dynamics Calculations”, AIAA J, Vol. 41(9), pp.<br />

1827–1829.<br />

[20] Blazek, J., 2005, Computational Fluid Dynamics:<br />

Principles and Applications, Elsevier, 2nd edn.<br />

[21] Menter, F. R., 1994, “Two-Equation Eddy-<br />

Viscosity Turbulence Models for Engineering<br />

Applications”, AIAA J, Vol. 32(8), pp.<br />

1598–1605.<br />

[22] Spalart, P. R., Deck, S., Shur, M. L., Squires,<br />

K. D., Strelets, M. K., and Travin, A., 2006,<br />

“A New Version of Detached-Eddy Simulation,<br />

Resistant to Ambiguous Grid Densities”, Theor<br />

Comp Fluid Dyn, Vol. 20(3), pp. 181–195.<br />

[23] German Aerospace Center (DLR), 2011,<br />

TAU-Code User Guide, Release 2011.1.0.<br />

[24] Spalart, P. R., and Allmaras, S. R., 1992,<br />

“A One-Equation Turbulence Model for<br />

Aerodynamic Flows”, Technical Report<br />

AIAA-92-0439.<br />

[25] Ladson, C. L., Hill, A. S., and Johnson,<br />

W. G., Jr., 1987, “Pressure Distributions from<br />

High Reynolds Number Transonic Tests of an<br />

NACA 0012 Airfoil in the Langley 0.3-Meter<br />

Transonic Cryogenic Tunnel”, Technical Report<br />

NASA-100526.<br />

[26] Illi, S., Gansel, P. P., Lutz, T., and Krämer, E.,<br />

2012, “Hybrid RANS-LES Wake Studies of an<br />

Airfoil in Stall”, submitted to CEAS Aeronaut J.<br />

[27] http://www.bw-grid.de, member of the German<br />

D-Grid initiative, funded by the “Federal<br />

Ministry for Education and Research” and<br />

the “Ministry for Science, Research and Arts<br />

Baden-Württemberg”.<br />

206

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!