Vous êtes sur la page 1sur 13

14th AIAA/CEAS Aeroacoustics Conference (29th AIAA Aeroacoustics Conference) 5 - 7 May 2008, Vancouver, British Columbia Canada

AIAA 2008-2802

Investigation of noise radiation from a turbulent boundary layer


J. Berland and X. Gloerfelt
Arts et M etiers ParisTech, Paris, France

A numerical investigation using compressible large-eddy simulation of the noise generated by a spatially developing turbulent boundary layer is reported in the present paper. The results are compared to reference data to assess the consistency of the computation. It turns out that the present large-eddy simulation approach seems to be able to capture the features of both aerodynamic and acoustic elds.

I.

Introduction

This paper is concerned with the investigation of noise production by turbulent boundary layers (TBL). The inuence of solid surfaces on aerodynamic noise has claimed attention over the past fty years. Investigation of TBL noise using experiments as well as numerical simulations is however especially tedious because of the small amplitude of the generated noise. As shown for instance by Powell,26 the acoustic conversion eciency is expected to vary as M 3 for the boundary layer and as M 5 for free turbulence, where M is the free-stream Mach number. Few experiments are available so far2, 12 and a large scatter exists in the measured data. Several theoretical studies have nevertheless been carried out with the aim of predicting the features of the pressure eld radiated by a plane TBL. This topic was rst investigated by Curle8 and Powell26 using Lighthills analogy.22 Lighthills theory and its extensions have then been extensively used in the study of the noise generated by TBL. Landahl20 for instance developed a two-scale model for the turbulent velocity and pressure elds in order to estimate the noise radiated by a the ow over a solid surface. Tam29 calculated the intensity, the directivity and the spectrum of the sound eld produced a TBL based on an empirical model for the pressure cross-correlation function. In a similar manner, the noise generated by specic features of the TBL has been investigated. In particular, Lauchle21 considered sound wave production during the boundary-layer transition. In an attempt to shed a new light on sound sources of TBL, Hardin13 analyzed the noise generated by various vortical phenomena such as the formation of horseshoe vortices and the emergence of viscous sublayer bursts. The works depicted above use an empirical expression or an assumed shape for the Fourier transform of the pressure cross-correlation function. To cope with the uncertainties associated with the choice of a cross-correlation model, Lighthills analogy22 has also been applied to aerodynamic data issued from numerical simulations. Hu et al.1517 focused on noise generated by turbulent channel ows. They applied Lighthills acoustic analogy to Direct Numerical Simulation (DNS) data in order to obtain an expression for the power spectrum of the far-eld pressure uctuations. Wang et al.33 evaluated the far-eld sound of an unstable wavepacket over a solid surface, the features of the transition being provided by solving the incompressible Navier-Stokes equations. In spite of the numerous works already achieved in the eld of TBL noise, a clear identication of sound sources remains elusive. An investigation approach relying on Direct Noise Calculation (DNC) might therefore be valuable. Such simulations indeed aim at obtaining both aerodynamic and acoustic elds within a same run.3, 7, 10 The method is particularly reliable since the compressible Navier-Stokes equations are directly solved and no additional modelling of the acoustic sources is required. Performing DNC however remains a challenging issue: the full range of length and amplitude scales associated with aerodynamic and acoustic uctuations needs to be resolved and silent non-reecting boundary conditions must be implemented. These requirements are especially dicult to meet for turbulent boundary layers. As mentioned, sound
PhD, Assitant

julien.nerland@paris.ensam.fr. Professor, xavier.gloerfelt@paris.ensam.fr

1 of 13 American Institute of Aeronautics and Astronautics

Copyright 2008 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.

sources in wall bounded ows are likely to be inecient and can therefore be easily overwhelmed by the inherent background noise due to numerical methods and boundary conditions. Discretization schemes based on large stencils and ensuring a wide range of accuracy are hence employed and specic boundary conditions are implemented.30 Surface noise has already been studied in more favorable ow congurations for instance by Gloerfelt et al.11 for the cavity tones or by Marsden et al.24 for the airfoil noise. Direct calculation of the sound produced by a turbulent boundary layer itself is nonetheless still to be performed. The present work aims in particular at demonstrating that the DNC of a spatially developing turbulent layer is feasible. To keep computational cost at a reasonable level while dealing with realistic turbulence conguration, a Large-Eddy Simulation (LES) approach is retained. Only the large scales are thus resolved and the eects of the smaller ones are taken into account thanks to a subgrid-scale model. The simulation of the spatial development of a turbulent boundary layer furthermore raises the diculty of the ignition of the transition to turbulence at the upstream frontier of the domain. Several paths to laminar breakdown have been identied thanks to controlled experiments. To name a few, the most popular boundary layer breakdown scenarios are the fundamental breakdown (K-type) observed in the pioneering works of Klebano et al.,19 and the subharmonic regime (N-type) highlighted later on in the experiments of Kachanov & Levchenko.18 In order to trigger the breakdown to turbulence of the boundary layer, additional disturbances need to be added to the integration domain. Rist & Fasel27 for instance implemented a timedependent blowing and suction within a localized surface at the wall. The method, aiming at reproducing the vibrating-ribbon technique, yielded aerodynamic results in good agreement with those obtained in the wind-tunnel experiments of Kachanov & Levchenko.18 Similarly, Ducros et al.9 added at the inow a randomly chosen white noise to a two-dimensional Tollmien-Schlichting wave in order to ignite the transition to turbulence in their LES calculation of a turbulent boundary layer. However, when dealing with aeroacoustic calculations, the enforcement of aerodynamic perturbations at the inow conditions is a tedious problem. Consider for instance some highly accurate numerical boundary conditions with several order of magnitudes between the incoming aerodynamic uctuations and the spurious acoustic waves generated by the numerics. The resulting sound waves may still mask the physical sound eld radiated by turbulence. An alternative to controlled laminar breakdown is the use of bypass transition. Nagarajan et al.25 for instance included in their computational domain a blunt leading-edge in order to ignite laminar breakdown. The occurrence of Tollmien-Schlichting waves, spanwise vorticity and three dimensionalization of the ow is bypassed. For the present simulation, bypass transition is achieved by introducing a small step located slightly downstream the inow conditions and lying in the spanwise direction. A ne control of the scenario of the laminar breakdown cannot be not achieved, but at least, three-dimensional turbulent motions can be introduced inside the ow without generating unwanted acoustic radiation due to numerical methods. In the present study, the noise radiated by a spatially developing TBL is investigated by means of compressible LES. The free-stream Mach number is equal to M = 0.5 and the Reynolds number Re = U / based on the momentum thickness ranges from 300 to 2000. Bypass transition is ensured by introducing a small step upstream of the computational domain. The LES approach is based on low-dispersive and low-dissipative algorithms.5 In addition, explicit selective ltering of the ow variables is implemented in order to take into account the eects of subgrid scales.6 The main objective of this work is to demonstrate the feasibility of the DNC of TBL noise. The outline of the paper is the following. The numerical procedure as well as the simulation parameters are detailed in section II. The turbulent development and the acoustic eld provided by the LES calculation are then investigated in section III. Concluding remarks are nally drawn in section IV.

II.

Numerical procedure

The spatial development of a three-dimensional turbulent boundary layer is calculated by means of compressible LES. A sketch of the computational domain and of the coordinate system is provided in gure 1: x is the streamwise direction and the ow goes from left to right; the wall is located in the plane y = 0. II.A. LES strategy and numerical methods

The LES equations are discretized on a collocated Cartesian grid. Explicit low-dispersive and low-dissipative 11-point numerical algorithms are implemented in order to perform the space discretization.5 Finite differences are of order four whereas selective ltering is carried out by a scheme of order two but with a

2 of 13 American Institute of Aeronautics and Astronautics

y U x z

Figure 1. Sketch of the computational domain and of the coordinate system (gure not to scale). A small step lying in the spanwise direction is introduced downstream of the inow boundary conditions in order to ignite the transition to turbulence of the boundary layer ow.

spectral-like cut-o. Time integration is nally performed thanks to a second-order six-step Runge-Kutta scheme optimized in the Fourier space.5 II.B. II.B.1. Simulation parameters Inow conditions

At the upstream frontier of the computational domain the boundary layer is assumed to be laminar. A Blasius boundary layer velocity prole, given by u = U (2 2 2 + 3 ) 1 if < 1 if 1 (1)

is therefore enforced at this location. The parameter is the distance to the wall normalized by the boundary layer thickness , and U = M c is the free-stream velocity. The free-stream Mach number is taken to be M = 0.5. The speed of sound far from the boundary layer is evaluated using the relationship c = rT , with T = 298.15 K and where r = 287.06 J.kg1 .K1 is the perfect gas constant. The ambient pressure is initially uniform and equal to p = 101300 Pa. In a similar manner, the density is initialized with a constant value = p /(rT ). The kinematic viscosity is nally taken to be = 1.5 105 m2 .s1 so that the initial Reynolds number Re0 based on the local momentum thickness at the inow is about 300. The initial boundary layer thickness 0 is then equal to 2.5 104 m. Note that the friction velocity u is calculated when the ow is fully turbulent. This reference velocity permits in particular to express simulation parameters in wall units, e.g. u+ = u/u or y + = u y/ . In addition, in what follows, variables with an overbar such as f denote time average (or mean) quantities whereas primed variables f correspond to uctuations with respect to the mean. II.B.2. Boundary conditions and grid design

A three-dimensional planar periodic domain is considered. The domain is hence made periodic in the spanwise direction and a no-slip wall condition is enforced on the lower y -plane. The remaining frontiers are treated using the non-reecting boundary conditions of Tam & Dong.31 A sponge zone4 is furthermore added at the downstream end of the domain so that unhindered passage of aerodynamic perturbations is possible without the generation of spurious acoustic waves. The mesh size is taken to be uniform in the spanwise direction. In a similar manner, the mesh in the streamwise direction is uniform but grid stretching, with a ratio of 1.02, is used in the sponge zone. In the wall normal coordinate, the grid size is stretched using a geometric progression of 2%. In the fully turbulent regime the mesh size corresponds to x+ = 58, y + = 3 and z + = 17.5. The domain of interest, excluding the sponge zone, has dimensions 3000 570 140 and is discretized by 760 200 101 15 106 nodes. The time step t 1.7 108 s corresponds to a Courant-Friedrichs-Levy number CFL = c t/y equal to 1. The parameters of the simulation are compiled in table 1.

3 of 13 American Institute of Aeronautics and Astronautics

Table 1. Simulation parameters for the LES of a three-dimensional spatially developing turbulent boundary layer.

M p T 0 Re0 c U x+ y + z + t Lx Ly Lz Nx Ny Nz

0.5 101300 Pa 298.15 K 2.5 104 m 1.5 105 m2 .s1 300 346 m.s1 173 m.s1 1.18 kg.m3 58 3 17.5 1.7 108 s 300 57 14 15 106

II.B.3.

Laminar breakdown ignition - Bypass transition

Turbulence seeding is achieved by introducing a small step slightly downstream the inow boundary condition. The step lies in the spanwise direction and is located between x/0 = 11 and x/0 = 15, and 0 y/0 0.14. This step has a large width/height aspect ratio and is hence expected to have few impact on the global mean ow while providing a discontinuity igniting the transition to turbulence. As mentioned, controlled laminar breakdown is not considered in the present paper since the enforcement of aerodynamic perturbations at the inow conditions, without generating spurious sound waves, remains a dicult task. This issue is nonetheless briey illustrated with a two-dimensional test case in appendix. It should be noted that the step itself is not enough to ensure ow three-dimensionality. At the very beginning of the calculation, random uctuations are added to the initial ow eld upstream of the step.

III.
III.A. Instantaneous ow eld

Results and validation

A snapshot of the streamwise vorticity eld x in the whole computational domain is rst presented in gure 2. The boundary layer clearly appears to be turbulent with random motions visible in the neighborhood of the solid surface. Flow randomization is furthermore seen to occur at the beginning of the computational domain. Close to the inow frontier the ow is laminar. Further downstream, the small step seems to be responsible for the emergence of vortical structures, which eventually lead to the laminar breakdown. These observations are further conrmed by the instantaneous snapshots of the vorticity modulus | | in the central plane (x, y, z = 0) and in the horizontal plane (x, y = 0.30 , z ) which are provided in gures 3.a and 3.b, respectively. Note that only the upstream part of the calculation domain (x/0 < 100) has been represented. One may observe that the laminar breakdown indeed occurs in the neighborhood of the step (x/0 13). Upstream x/0 = 10 the ow eld appears to be laminar whereas after the step, especially downstream x/0 = 30, a large range of turbulence scales are visible both in gure 3.a and 3.b. III.B. Mean ow

The mean ow provided by the present LES calculation is now compared to the reference DNS data of Spalart28 obtained at Re = 1410.
4 of 13 American Institute of Aeronautics and Astronautics

American Institute of Aeronautics and Astronautics

5 of 13
y x z Figure 2. Instantaneous snapshot of the streamwise vorticity x in the whole computational domain. The ow is from left to right. Color scale represents the magnitude of x 0 /U and ranges from 0 (yellow) to 0.73 (red).

(a)
10

y/0

5 0 0 10 20 30 40 50 60 70 80 90 100

x/0

(b)
5

z/0

0 5 0 10 20 30 40 50 60 70 80 90 100

x/0
Figure 3. Instantaneous snapshot of the vorticity modulus | |0 /U . (a), Vorticity modulus in the central plane (x, y, z = 0) and (b), vorticity modulus in the horizontal plane (x, y = 0.30 , z ). Color scale represents the magnitude of | |0 /U and ranges from 0 (white) to 1.46 (black).

In order to evaluate the streamwise position where data comparison has to be made, the Reynolds number Re based on the local momentum thickness of the boundary layer is given in gure 4 as a function of the streamwise position x/0 . The Reynolds number is observed to linearly increases with the streamwise position, even if the linear slope is perturbed in the neighborhood of the step (x/0 10). As concern the overall evolution of the Reynolds number Re , it is close to Re = 300 at the upstream frontier of the domain and reaches Re = 2000 at x/0 = 300. A linear regression of the curve Re = f (x/0 ) yields a slope equal to 5.5. As a comparison, in the LES results of Lund et al.,23 a linear relationship given by /0 0.0018 x/0 can be observed between the momentum thickness and the streamwise position. The Reynolds number hence varied as Re 5.2 x/0 , which is in agreement with the present results. Finally, it is seen that the present LES calculation reaches a Reynolds number Re of 1410 for x/0 = 190. Data comparison with the DNS data of Spalart28 will consequently be made at this streamwise location. The prole in the wall-normal direction of the mean streamwise velocity is rst presented in wall units in gure 5, where u+ is plotted as a function of y + . The linear and logarithmic laws, as well as the DNS data of Spalart28 are also represented. Overall, a fair agreement is found with the reference data. In the viscous sublayer, the present LES calculation slightly underestimates the magnitude of the mean velocity u, but a linear law is still observed. In the overlap region (30 < y + < 200), the DNS data,28 the log law and the present prole perfectly collapse. Further away from the wall, for y + > 200, the classical departure from the exact log relation is visible. The amplitude of the upwards concavity in the prole is in addition close to the one obtained by Spalart.28
+ + = v v are nally plotted in gure 6.a and 6.b as The turbulent intensities uu = u u and vv + 28 functions of y . The DNS data of Spalart are also represented to provide reference results. The turbulent + intensities uu in gure 6.a exhibit some discrepancies. Between y + = 100 and y + = 500 a gap is especially visible between the two curves. Nonetheless, it should be remarked that the location and the amplitude of the turbulent peak are correctly predicted, as well as the residual turbulent motions further away from the + wall, for approximately y + > 500. As for the turbulent intensity vv in gure 6.b, its amplitude is 10% 28 smaller than the value proposed by the DNS of Spalart. An overall fair agreement is however observed and the location of the maximum of turbulence activity is in particular well reproduced. + +

6 of 13 American Institute of Aeronautics and Astronautics

2000 1500

Re

1000 500 0

50

100

150

200

250

300

x/0
Figure 4. Streamline evolution of the Reynolds number Re based on the local momentum thickness of the boundary layer. The location where mean ow validation is carried out is indicated by the dotted line.

25 20 15 10 5 1 0 10 10
1

u+

10

10

10

y+
Figure 5. Prole in the wall-normal direction of the mean streamwise velocity u+ as a function of y + , at x/0 = 190. , u+ y + ; , u+ log y + ; , DNS data of Spalart;28 , present LES calculation.

III.C.

Radiated acoustic eld

An overview of the radiated eld is given in gure 7 where the pressure uctuation p is plotted in the central plane (x, y, z = 0) as a colormap. Acoustic wavefronts are observed in the near-eld. They are mainly oriented towards the upstream direction and have a relatively large wavelength. The time evolution of the pressure perturbation measured in the near-eld, at (x, y, z ) = (1350 , 380 , 0) is plotted in gure 8 as a function of time tU /0 . The peak amplitude of the signal is about 5 Pa and the sound pressure level based on the root-mean-square of the signal is equal to 103 dB. The uctuations furthermore seem to comprise a broad range of frequency components. In particular, high-frequency oscillations, with a period of about 5U /0 , are visible over the whole time window. In addition, events occur at a rather low frequency rate, as for instance at tU /0 = 200 and tU /0 = 250, or at tU /0 = 650 and tU /0 = 750. These trends are supported by the power spectral densities of the pressure uctuations obtained in the near-eld, which are provided in gure 9 as functions of the Strouhal number ref /U , as dened for instance 29 by Tam. The reference displacement thickness ref is taken to be the average value of with respect to the streamwise direction, so that ref = 1.5 104 m. The far-eld dimensionless pressure spectrum predicted 29 by Tam based on theoretical considerations is also presented for comparison. Tams formulation including mean ow eects has been retained and its amplitude has been rescaled to t with the present results. It turns out that a good collapse between the spectra obtained with the LES computation at the dierent streamwise locations is observed. A broadband low frequency component is clearly visible at Strouhal number St = 0.041. It is interesting to remark that the pressure spectrum provided by Tam29 has similar features, the broadband peak is centered on a slightly higher Strouhal number, St = 0.043, but this value is still very close to the one observed for the LES calculation. The near-eld spectra provided by the LES computation unfortunately exhibit an additional narrow highfrequency peak at Strouhal number St = 0.5. Its rst harmonic is also visible at St = 1. The emergence of such tones in the turbulent boundary layer noise has not been reported so far and is likely to be a consequence

7 of 13 American Institute of Aeronautics and Astronautics

(a)
3 2.5 2
+ uu + vv

(b)
1.5 1.25 1 0.75 0.5 0.25

1.5 1 0.5 0 0 250 500 750 1000

250

500

750

1000

y+

y+

+ Figure 6. Prole in the wall-normal direction of the turbulent intensities as functions of y + , at x/0 = 190. (a), uu and + , DNS data of Spalart;28 , present LES calculation. (b), vv .

50

y/0

25

0 0 50 100 150 200 250 300

x/0
Figure 7. Instantaneous snapshot of the pressure uctuations p in the central plane (x, y, z = 0). Color scale ranges from -10 Pa (blue) to +10 Pa (red).

of the small step introduced to force the laminar breakdown. This assumption is conrmed by gure 10, where a snapshot of the pressure eld downstream the step in the central plane (x, y, z = 0) is presented. Alternating low and high pressure regions are situated downstream the step and correspond to developing aerodynamic perturbations. The emergence of these vortical structures can be driven by a Kelvin-Helmholtz instability process. Just downstream of the step, shear-layers can be observed in the wall-normal velocity proles. For instance, in gure 11, the mean streamwise velocity u/U is reported for various streamwise positions after the step. A good collapse between the present curves and an hyperbolic-tangent prole is indeed visible. In addition, the wavelength of the phenomenon, as shown in gure 10, is closed to 3.70 . The corresponding Strouhal number is St = 0.505; this value is in agreement with the one deduced from the near-eld pressure spectrum provided in gure 9. One may therefore conclude that the high-frequency tones are related to the presence of the small step inside the ow.

IV.

Conclusion

In the present work, the direct noise computation of a spatially developing turbulent boundary layer has been performed by means of compressible LES based on spectral-like discretization algorithms. A small step has been introduced inside the integration domain in order to force the laminar breakdown with a bypass transition. The Reynolds number Re = U / based on the momentum thickness ranges from 300 to 2000. The turbulent development of the present LES turns out to be consistent. The mean ow data and the turbulent intensities compare favorably to the DNS reference data of Spalart.28 In addition, the noise generated by the turbulent ow has also been obtained. The radiated acoustic eld exhibits similar features as those formally predicted by Tam29 thanks to theoretical considerations. In particular, a broadband frequency content in visible in the pressure spectrum. This noise component occurs at a Strouhal number close to the one deduced from Tams model29 and is hence associated with the turbulent boundary layer
8 of 13 American Institute of Aeronautics and Astronautics

10 5

p (Pa)

0 5 10

100

200

300

400

500

600

700

800

900

tU /0
Figure 8. Pressure history in the near-eld. Time evolution of the pressure uctuations at (x, y, z ) = (1350 , 380 , 0) as a function of time tU /0 .

10

10

10

PSD(p )

10

10

10

10

10

10

10

10

ref /U Figure 9. Power spectral density of the pressure perturbations in the near-eld (y/0 = 38) as a function of the Strouhal number ref /U and for various streamwise locations. , x/0 = 113; , x/0 = 135; , x/0 = 158; + , x/0 = 180; , x/0 = 203. , far-eld spectrum predicted by Tam.29

noise. Unfortunately, further investigations have been carried out and have demonstrated that including a small step to ignite laminar breakdown is responsible for an additional high-frequency noise component in the pressure spectrum.

Appendix - Unsteady inow boundary conditions


The present test case aims at illustrating the diculty to enforce an incoming vorticity waves at the upstream frontier of a computational domain without generating spurious acoustic waves. Unsteady inow boundary condition The spatial development of a two-dimensional boundary layer is solved using similar discretization techniques as those employed for the three-dimensional LES calculation presented in this paper. The computational domain is discretized by 201 201 40000 nodes. Non-reecting boundary conditions are implemented so that the radiated noise can leave the integration domain without generating spurious sound waves.31 Following the works of Tam,32 a non-homogeneous set of radiation boundary equations are resolved in order

9 of 13 American Institute of Aeronautics and Astronautics

3.70

y/0
2 0 14

16

18

20

22

24

26

28

30

x/0
Figure 10. Snapshot of the pressure eld downstream the small step in the central plane (x, y, z = 0). Color scale ranges from -100 Pa (blue) to +100 Pa (red).

2 1.5

y/0

0.5 0

0.2

0.4

0.6

0.8

1.2

u/U
Figure 11. Prole in the wall-normal direction of the mean streamwise velocity u/U as a function of y/0 , for various , streamwise locations just downstream the small step. , x/0 = 15; , x/0 = 16; , x/0 = 17; +, x/0 = 18. hyperbolic-tangent prole.

to impose some inow conditions 1 1 1 1 + + + + U= Uin V () t r 2r V () t r 2r (2)

where U = ( u v w p)t is the primitive variable vector, and Uin = (in uin vin win pin )t is the vector containing the inow conditions. Tollmien-Schlichting waves The present test case aims at introducing a Tollmien-Schlichting (TS) wave into the integration domain. TS waves are of special interest since they are not supposed to radiate sound waves.1 The present benchmark hence permits to unambiguously assess whether the inow boundary conditions generate spurious noise or not. For TS waves, the incoming velocity perturbation is given by Uin (x, y, t) = A(r) (y ) cos(kr x t) A(i) (y ) sin(kr x t) eki x (3)

where A(r) (y ) and A(i) (y ) are the real and imaginary parts of the amplitude distribution in the wall normal direction y of the incoming perturbation. The complex quantity A(y ) is given by
2 Qp (y ) A(y ) = Q (y ) U Qu (y ) U Qv (y ) U t

(4)

The amplitude of the incoming TS wave is given by . The pulsation is , and the real and imaginary parts of the wavenumber are denoted by kr and ki . The normalized amplitude distribution Q(y ) of each ow variable is deduced from linear stability theory.
10 of 13 American Institute of Aeronautics and Astronautics

(a)
6

(b)
6

0.2

0.4

0.6

0.8

0.2

0.4

0.6

0.8

u/U

v (Rex /U )

Figure 12. Mean velocity proles in the wall normal direction for the two-dimensional boundary layer test case p as functions of the normalized distance to the wall = y U /(x). (a), streamwise mean velocity u/U and (b), , Blasius analytical solution. Symbols , + and are the results of the present transverse mean velocity v/U . calculation at x/0 = 167, x/0 = 188 and x/0 = 211, respectively.

The initial boundary layer thickness is set to 0 = 2.5 104 m so that the Reynolds number Re = U / based on the displacement thickness at the upstream limit of the calculation domain is equal to 1000. Under these operating conditions, the most unstable mode oscillates with a pulsation /U = 0.1. A linear stability analysis then yields the amplitude distributions Q(y ) and the corresponding wavenumbers: (kr , ki ) = (0.28, 0.73 102 ) if the ow is assumed to be incompressible, and (kr , ki ) = (0.27, 0.67 102 ) when compressibility eects are taken into account. Results Before the unsteady inow conditions are enforced, the mean ow is rst calculated. The amplitude of the excitation is set to zero and the computation is continued as long as the mean ow is not converged. The mean streamwise and transverse velocities u and v obtained with this procedure are plotted in gures 12.a and 12.b, as functions of the distance to the wall and for various streamwise locations. The Blasius solution is also represented for comparison. A very good collapse between the various curves is observed, thus ensuring that the mean ow is consistent. Once the mean ow has been calculated, an incoming TS wave is introduced at the upstream boundary condition. The amplitude of the perturbation is set to = 102 , so that the corresponding pressure uctu2 ations have a magnitude of approximately U 350 Pa. The pressure eld obtained when the inow forcing is turned on is plotted in gures 13.a13.d for an incompressible TS wave, and in gures 13.e13.h for a compressible TS wave. One may rst observe that a TS wave is indeed introduced into the domain. The TS wave corresponds to the alternating dark and white regions visible on the pressure eld in the neighborhood of the solid surface. However, for both incompressible and incompressible TS waves, sound waves are seen to be generated by the inow boundary conditions. Their amplitude is about 0.5 Pa for the incompressible case and about 0.1 Pa for the compressible TS wave. In terms of relative amplitude these spurious waves are rather small since there are about three orders of magnitude between their amplitude and the amplitude of the pseudopressure associated with the incoming vorticity wave. Unfortunately, a spurious acoustic source with an amplitude around 0.1 Pa or 0.5 Pa can easily contaminate the whole acoustic eld and may interfere with the waves emitted by the true sound sources.

Acknowledgments
The authors gratefully acknowledge the Institut du D eveloppement et des Ressources en Informatique Scientique (IDRIS - CNRS) and the Centre de Calcul Recherche et Technologie (CCRT) of the Commissariat a lEnergie ` Atomique (CEA) for providing computing time. The authors also thank X. Merle and J.-C. Robinet for providing linear stability analysis data.

11 of 13 American Institute of Aeronautics and Astronautics

References
D.R., Toplosky N., The sound eld of a Tollmien-Schlichting wave, Phys. Fluids, 29, 685 (1986). S.J., Radiated noise from turbulent boundary layers in dilute polymer solutions, Phys. Fluids, 16, 1387 (1973). 3 Berland J., Bogey C., Bailly C., Numerical study of screech generation in a planar supersonic jet, Phys. Fluids, 19, 075105 (2007). 4 Bogey C., Bailly C., Three-dimensional non-reective boundary conditions for acoustic simulations: far eld formulation and validation test cases, Acta Acustica, 88, 463 (2002). 5 Bogey C., Bailly C., A family of low dispersive and low dissipative explicit schemes for ow and noise computations, J. Comput. Phys., 194, 194 (2004). 6 Bogey C., Bailly C., Large eddy simulations of transitional round jets: inuence of the Reynolds number on ow development and energy dissipation, Phys. Fluids, 18, 065101 (2006). 7 Bogey C., Bailly C., An analysis of the correlations between the turbulent ow and the sound pressure elds of subsonic jets, J. Fluid Mech., 583, 71 (2007). 8 Curle N., The inuence of solid boundaries upon aerodynamic sound, Proc. Roy. Soc. London, Series A, 231, 505 (1955). 9 Ducros F., Comte P. & Lesieur M., Large-eddy simulation of transition to turbulence in a boundary layer developing spatially over a at plate, J. Fluid Mech., 326, 1 (1996). 10 Freund J., Noise sources in a low-Reynolds-number turbulent jet at Mach 0.9, J. Fluid Mech., 438, 277 (2001). 11 Gloerfelt X., Bogey C., Bailly C., Numerical evidence of mode switching in the ow-induced oscillations by a cavity, Int. J. Aeroacoust., 2, 193 (2003). 12 Greshilov E.M., Mironov M.A., Experimental evaluation of sound generated by turbulent ow in a hydrodynamic duct, Sov. Phys. Acoust., 29, 275 (1983). 13 Hardin J.C., Acoustic sources in low Mach number turbulent boundary layer, J. Acoust. Soc. Am., 90, 1020 (1991). 14 Howe M.S., Surface pressures and sound produced by turbulent ow over smooth and rough walls, J. Acoust. Soc. Am., 90, 1041 (1910). 15 Hu Z.W., Morfey C.L., Sandham N.D., Aeroacoustics of wall-bounded turbulent ows, AIAA J., 40, 465 (2002). 16 Hu Z.W., Morfey C.L., Sandham N.D., Sound radiation in turbulent channel ows, J. Fluid Mech., 475, 269 (2003). 17 Hu Z.W., Morfey C.L., Sandham N.D., Sound radiation from a turbulent boundary layer, Phys. Fluids, 18, 098101 (2006). 18 Kachanov Y., Levchenko V., The resonant interaction of disturbances at laminar-turbulent transition in a boundary layer, J. Fluid Mech., 138, 209 (1984). 19 Klebano P., Tidstrom K., Sargent L., The three-dimensional nature of boundary-layer instability, J. Fluid Mech., 12, 1 (1962). 20 Landahl T., Wave mechanics of boundary layer turbulence and noise, J. Acoust. Soc. Am., 57, 824 (1975). 21 Lauchle G., On the radiated noise due to boundary-layer transition, J. Acoust. Soc. Am., 67, 158 (1980). 22 Lighthill M., On sound generated aerodynamically, I, general theory, Proc. R. Soc. London, Series A, 211, 564 (1952). 23 Lund T., Wu X., Squires K., Generation of turbulent inow data for spatially-developing boundary layer simulations, J. Comput. Phys., 140, 233 (1998). 24 Marsden O., Bogey C., Bailly C., Direct noise computation of the turbulent ow around a zero-incidence airfoil, AIAA J., 46, 874 (2008). 25 Nagarajan S., Lele S.K., Ferziger J.H., Leading-edge eects in bypass transition, J. Fluid Mech., 572, 471 (2007). 26 Powell A., Aerodynamic noise and the plane boundary, J. Acoust. Soc. Am., 32, 982 (1960). 27 Rist U., Fasel H., Direct numerical simulation of controlled transition in a at-plate boundary layer, J. Fluid Mech., 298, 211 (1995). 28 Spalart P.R., Direct simulation of a turbulent boundary layer up to Re = 1410, J. Fluid Mech., 197, 61 (1988). 29 Tam C.K.W., Intensity, spectrum, and directivity of turbulent boundary layer noise, J. Acoust. Soc. Am., 57, 25 (1975). 30 Tam C.K.W., Computational aeroacoustics: issues and methods, AIAA J., 33, 1788 (1995). 31 Tam C., Dong Z., Radiation and outow boundary conditions for direct computation of acoustic and ow disturbances in a nonuniform mean ow, J. Comput. Ac., 4, 175 (1996). 32 Tam C., Advances in numerical boundary conditions for computational aeroacoustics, J. Comput. Ac., 6, 377 (1998). 33 Wang M., Lele S.K., Moin P., Sound radiation during local laminar breakdown in a low-Mach-number boundary layer, J. Fluid Mech., 319, 197 (1996).
2 Barker 1 Akylas

12 of 13 American Institute of Aeronautics and Astronautics

p p (Pa)
0
60

0.2

0.4

0.6

0.8

1
60

(a)
40 40

(e)

y/0

20

y/0
20 0 20 40 60 80 0 0

20

40

60

80

60

(x x0 )/0 (b)

60

(x x0 )/0 (f )

40

40

y/0

20

y/0
20 0 20 40 60 80 0 0

20

40

60

80

60

(x x0 )/0 (c)

60

(x x0 )/0 (g)

40

40

y/0

20

y/0
20 0 20 40 60 80 0 0

20

40

60

80

60

(x x0 )/0 (d)

60

(x x0 )/0 (h)

40

40

y/0

20

y/0
20 0 20 40 60 80 0 0

20

40

60

80

(x x0 )/0

(x x0 )/0

Figure 13. Snapshots of the pressure perturbations p p for the two-dimensional boundary layer test case for various time location using, (a-d) the incompressible TS amplitude distributions and, (e-f ) using the compressible TS amplitude proles. (a) or (e), t = 1.2TTS ; (b) or (f ), t = 2.4TTS ; (c) or (g), t = 3.5TTS ; (d) or (h), t = 4.7TTS . TTS is the period of the Tollmien-Schlichting wave.

13 of 13 American Institute of Aeronautics and Astronautics

Vous aimerez peut-être aussi