Sei sulla pagina 1di 15

12th AIAA/CEAS Aeroacoustics Conference (27th AIAA Aeroacoustics Conference) AIAA 2006-2513

8 - 10 May 2006, Cambridge, Massachusetts

Prediction of Flat Plate Self-Noise

K. W. Chang*, J. H. Seo†, and Y. J. Moon‡


Department of Mechanical Engineering, Korea University, Seoul, 136-701, Korea

M. Roger§
Laboratoire de Mecanique des Fluides et Acoustique (LMFA), Ecole Centrale de Lyon, 69134, Ecully, France

Abstract
In this study, the accuracy of the computational methodology developed for prediction of turbulent flow noise at low
Mach numbers is assessed for the flat plate self-noise. The far-field self-noise and the wall-pressure field over the
flat plate (chord=10cm, thickness=3mm, and span=30cm) are measured at Ecole Centrale de Lyon for a flow speed
UO=20m/s at zero angle of attack. The three-dimensional turbulent flow over the plate is computed by
incompressible large eddy simulation (LES), while the near- and far-field acoustics are calculated by the linearized
perturbed compressible equations (LPCE), coupled with the LES solutions. Comparisons are made for the wall
pressure PSD spectra near the trailing-edge, the spanwise coherence function of the surface pressure, and the far-
field sound pressure level spectrum. The computations agree well with the experiment. The tonal and broadband
noise characteristics of the flat plate are also discussed.

I. Introduction

T urbulence trailing-edge noise is one of the important concerns in airframe noise. It is often associated with
broadband noise generated by three-dimensional, spatio-temporal interactions of eddies in the boundary layer
with the cut-off edge and a tone associated with von Kármán vortex shedding for the edge-bluntness.
Comprehensive experimental investigations and analytical modeling works on the trailing-edge noise have been
conducted for years1,2,3. Meanwhile, prediction of turbulence trailing-edge noise also poses a computational
challenge because its occurrences are often concerned with flows at low Mach number. When flow Mach number is
less than 0.1, for example, scale disparity between hydrodynamics and acoustics becomes large and thereby direct
numerical simulation (DNS) becomes a difficult choice.
In the present study, flat plate self-noise at M=0.06 is predicted by a computational methodology developed for
prediction of turbulent flow noise at low Mach numbers4,5. A three-dimensional turbulent flow is computed by
incompressible large eddy simulation (LES), while the near- and far-field acoustics are calculated by the linearized
perturbed compressible equations (LPCE)4. The present method is found effective at low Mach numbers because it
is based on an incompressible flow solver and the grid systems for flow and acoustics can be treated separately. For
predicting the far-field sound pressure level (SPL) for long span, approximate but efficient methods are also
revisited: (i) solving 2D LPCE for acoustics at zero spanwise wavenumber, kz=0, (ii) 2D Kirchhoff method6 for
extrapolation to the far field, (iii) Oberai’s 3D-correction7 for spectral acoustic pressure in the mid-span plane, and
(iv) spanwise coherence function of the surface pressure.
The accuracy of the present methodology is assessed by comparing with the experimental data measured at
Ecole Centrale de Lyon for the self-noise generated by a low-speed flow (UO=20m/s) over a flat plate (chord=10cm,
thickness=3mm, and span=30cm) at zero angle of attack. Comparisons are made for the wall pressure PSD spectra
near the trailing-edge, the spanwise coherence function of the surface pressure, and the far-field SPL spectrum for
the long span used in the experiment. Also, the tonal and broadband noise characteristics of the flat plate are
discussed. In section II, the experimental work is described for the flat plate self-noise measurements, and the
computational methodologies used in the present study are introduced in section III. Computational results are
discussed in section IV, including comparisons between experiment and computation.

*
Graduate student, AIAA student member.

Ph.D. Candidate, AIAA student member.

Professor, AIAA senior member, corresponding author (yjmoon@korea.ac.kr)
§
Professor, AIAA member.

1
American Institute of Aeronautics and Astronautics

Copyright © 2006 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
II. Experiments

A. Experimental Set-up
The experiment has been performed in the low-speed, open-jet anechoic wind tunnel of the École Centrale de
Lyon (ECL). The investigated rectangular model plate is held vertically between horizontal side-plates fixed to the
nozzle. It has a thickness of 3 mm, a chord length of 10 cm and a spanwise extent of 30 cm. It is rounded at the edge
corners by a bevel of radius 0.5 mm. Remote-microphone probes (RMP) are installed on the plate, around mid-span.
Each probe is made of a thin capillary tube inside the plate, progressively enlarged outside and connected to a flush-
mounted small-size microphone8. The effective measuring point has a diameter of 0.5 mm, ensuring that wall-
pressure measurements do not suffer from a significant integration effect at the frequencies of interest. A subset of
RMPs includes five spanwise distributed probes, 2 mm upstream of the trailing edge and one facing backwards in
the middle of the blunt trailing-edge section. These locations are used as checking points when comparing the data
to the numerical results. The acoustic measurements are made at a distance 2 m from the trailing edge in the mid-
span plane by a single B&K ½’’ microphone on a rotating support.
All results given below correspond to a pressure PSD in decibels per Hz, with a reference pressure of 0.00002 Pa,
and the data has been averaged on 200 samples, over the frequency range 0-to-12800 Hz, with a frequency-band 4
Hz and Hanning windowing.

B. Zero Angle of Attack


The measured wall-pressure spectra are shown in Fig. 1(a) for a flow speed of 20 m/s at zero angle of attack.
Due to their locations, the probes are true aerodynamic sensors and measure both the aerodynamic pressure
associated with the boundary layer turbulence convected from upstream, and the one associated with the von
Kármán vortex shedding in the wake. The latter is responsible for the large peak at the Strouhal number 0.2 on the
spectra from the upstream points (black curves in the figure). The remaining, broadband part of the spectrum, at a
typical level of 60 dB, is similar to a classical turbulent boundary-layer behaviour. It can be attributed to the
turbulence in the flow, apart from the vortex shedding. The spectrum measured on the trailing-edge strip is quite
different, with higher-level fluctuations at low frequencies, a faster decay at large frequencies and a secondary peak
at twice the Strouhal frequency. This is explained as follows. Due to the anti-symmetrical structure of the von
Kármán vortex street, the local motion normal to the free stream is reversed as vortices of opposite circulation are
shed in the wake, and repeats itself essentially at the Strouhal frequency. The same holds for the fluctuating lift force
induced on the plate. Conversely, the motion parallel to the flow, and the corresponding drag force, are the same
whatever the circulation of the vortices is, and rather occur at twice the Strouhal frequency.
For the zero angle of attack, the measured coherence function between two spanwise pressure probes at 2 mm
from the trailing edge is found a decreasing function, which is best fitted by a Gaussian model. This is clearly shown

(a) (b) (c)


Fig. 1 (a) Measured wall-pressure spectra close to the trailing-edge of a flat plate at zero angle of attack. Flow
speed 20 m/s. The black curves should collapse as corresponding to spanwise distributed sensors on one side
(2 mm upstream of the trailing edge). The red curve stands for a point located on the trailing-edge strip
surface normal to the flow direction, (b) Coherence plots. o: present computations; (o, o): wind-tunnel
measurements, as deduced from different sets of probes. Data interpolated by a Gaussian law, and (c) Sound
pressure level spectrum measured 2m away from the mid-chord of the plate.

2
American Institute of Aeronautics and Astronautics
in Fig. 1(b), where the computed values, for small spanwise separations, are plotted together with the measurements,
only available for larger separations due to the probe technology. Both sets of data are in the continuation of each
other, especially at the Strouhal frequency. In view of the results, the size of the computational domain normal to the
mean stream is smaller than the spanwise correlation length π / 2 Λ , with Λ=7h, related to the model coherence by
the formula γ 2(z) =exp(-(z/LC)2). As mentioned in an accompanying paper by Roger et al. 9, Λ is also an
exponentially decreasing function of the difference between the actual frequency and the Strouhal frequency. Finally,
Fig. 1(c) shows the sound pressure level spectrum measured 2m away from the mid-chord of the plate. Three runs
are superimposed, showing the reproducibility of the measurements.

III. Computational Methodologies


In the present study, a turbulent flow field is computed by incompressible large eddy simulation (LES), while its
acoustic field at zero spanwise wavenumber, kz=0 is predicted by the linearized perturbed compressible equations
(LPCE), with acoustic source, DP/Dt acquired from the LES solutions. This is based on hydrodynamic/acoustic
splitting method and Fig. 2 schematically represents the methodology. The far-field sound pressure level (SPL) for
an extended span is estimated by a 2D Kirchhoff method in frequency domain for the extrapolation of the spectral
acoustic pressure to the far field, Oberai’s 3D correction for the acoustic pressure in the mid-span plane, and a
spanwise coherence function of the surface pressure (see Fig. 3). Details of the present computational methods are
described below.

A. Incompressible Large Eddy Simulation (LES)


The filtered incompressible Navier-Stokes equations are written as

∂U j
=0 (1)
∂x j

∂U i ∂ ∂P ∂ ⎛ ∂U i ∂U j ⎞ ∂
ρ0 + ρ0 (U iU j ) = − + µ0 ⎜⎜ + ⎟⎟ − ρ0 M ij , (2)
∂t ∂x j ∂xi ∂x j ⎝ ∂x j ∂xi ⎠ ∂x j

u =U + u'
Compressible Incompressible Perturbed
λ a / λ h ∼ 1/ M ∞
Flow & Incompressible LPCE
Acoustics LES

λa

λh
Acoustic Sources

Fig. 2 Computational methodology for prediction of turbulent flow noise at low Mach numbers.

3
American Institute of Aeronautics and Astronautics
where the resolved quantities are denoted by ( ˜ ) and the unknown sub-grid tensor Mij is modeled as

k
M ij = U   2  
iU j − U iU j = −(Cs ∆) S Sij
. (3)

Here, ∆ is the mean radius of a grid cell (computed as the cubic root of its volume).
For solving the filtered incompressible Navier-Stokes equations, an iterative fractional-step method is used. The
governing equations are spatially discretized by a sixth-order compact finite difference scheme to avoid excessive
numerical dissipations and dispersions errors and integrated in time by a four-stage Runge-Kutta method.

B. Linearized Perturbed Compressible Equations (LPCE)


The linearized perturbed compressible equations (LPCE) are based on a hydrodynamic/acoustic splitting method,
where the total flow variables are decomposed into the incompressible and perturbed compressible variables as,

ρ ( x , t ) = ρ0 + ρ '( x , t )
u ( x , t ) = U ( x , t ) + u '( x , t ) . (4)
p( x , t ) = P( x , t ) + p '( x , t )

The incompressible variables represent the turbulent flow field, while the acoustic fluctuations and other
compressibility effects are resolved by perturbed quantities denoted by (‘) .
The LPCE are written in a vector form as,

∂ρ '
+ (U ⋅∇) ρ '+ ρ0 (∇ ⋅ u ') = 0 (5)
∂t

∂u ' 1
+ ∇(u '⋅U ) + ∇p ' = 0 (6)
∂t ρ0

∂p ' DP
+ (U ⋅∇) p '+ γ P(∇ ⋅ u ') + (u '⋅∇) P = − , (7)
∂t Dt

where the total change of the hydrodynamic pressure, DP/Dt is the only explicit noise source term in the formulation.
This term represents the noise sources generated by turbulences resolved with the grids used in the study. Details of
LPCE derivation can be found in reference4.
The LPCE is explicitly integrated in time by a four-stage Runge-Kutta method and spatially discretized by a six-
order compact finite difference scheme. Practically, the application of a high-order compact scheme to the stretched
meshes generates numerical instability arising from the numerical truncation or failure of capturing high wave
number phenomena. Therefore, a tenth-order spatial filtering proposed by Gaitonde et al.10 is used for enhancing the
numerical stability with high frequency cutoff. A buffer-zone type ETA boundary condition proposed by Edgar and
Visbal11 is used at the far field boundaries and a slip boundary condition is used at the solid surface.

C. Computation of Far-Field Sound for Long Span


In LES, a spanwise periodic boundary condition allows us to capture the spanwise-correlated flow structures at a
reduced computational cost. In acoustic calculation, however, applying a periodic boundary condition for the same
span yields unphysical correlated acoustic results12,13,14, because the acoustic length scale is usually much larger than
the hydrodynamic scales at low Mach numbers. To avoid this, one must either use a very long span that fully covers
the acoustic coherence length or apply an absorbing boundary condition at the side boundaries and then correct the
sound pressure level obtained from the simulated span, LS for the actual span, L(>> LS) afterwards. The latter
approach can be a practical choice but one still needs to conduct a full 3D acoustic calculation with the LPCE,
coupled with 3D LES solutions.
In the present study, an approximated but efficient approach is considered (see Fig. 3). First, a 2D LPCE
calculation is conducted for a computational domain of finite boundary with the spanwise-averaged acoustic sources
(DP/Dt) and flow solutions. If an observer’s position is far from the noise source, the acoustic pressure calculated

4
American Institute of Aeronautics and Astronautics
pˆ S' 2 D ( r,θ ,ω ) pˆ S' 3 D ( r,θ ,0,ω )

γ 2 ( z)

pˆ L' (r,θ ,0,ω)

Fig. 3 Computational methodology for prediction of far-field sound from a long-span.

by 2D LPCE is extrapolated by a Kirchhoff method. The 2D Kirchhoff method in frequency domain6 is given by

⎡ ∂pˆ ' (2) ω ⎤


4 i pˆ ' = − ∫
S

⎣ ∂n
H 0 (ω r / c0 ) − ( nˆ ⋅ rˆ) pˆ ' H1(2) (ω r / c0 ) ⎥ dS ,
c0 ⎦
(8)

where cO is the ambient speed of sound, r is the distance from the source to the observer’s position, n̂ is a unit
vector normal to the Kirchhoff surface, and H is the Hankel function.
According to Oberai et al.7, the three-dimensionally radiated, far-field acoustic pressure at the z=0 plane, pˆ '3 D is
related to a two-dimensionally predicted acoustic pressure at the zero spanwise wave number, kz=0, pˆ '2 D by

1+ i ω .
pˆ '3 D (r ,θ , 0, ω ) ≈ pˆ '2 D (r ,θ , ω ) (9)
2 c0π r

Using Eq. (9), the 2D spectral acoustic pressure at observer’s position is corrected to 3D spectral acoustic pressure at
the mid-span plane. This approach is approximate but computationally more efficient than solving the full 3D LPCE
for obtaining the far-field acoustics in the mid-span plane . It can be used also for many long-span bodies.
Finally, to estimate the sound for a long span, the sound pressure level (SPL) for the simulated span (LS) must be
corrected for the actual span (L). According to Kato et al.15, SPL for the long span can be obtained by adding 10log(L/LS),
if a coherence length of the surface pressure functuation, LC is determined as LC ≤ LS. For LC>L, 20log(L/LS) must be
added. This simple correction method has been used for various studies13,14,16,17. Although two asymptotic values are
correct, Kato’s approach is rather ad hoc for Ls≤ LC ≤ L. Recently, Perot et al.12 proposed a correction method,
considering an acoustic spanwise coherence with Curle’s analogy solution. This method gives more realistic results in the
near-field because a retarded time is taken into account but it has not been completed for practical use.

5
American Institute of Aeronautics and Astronautics
pˆ 'L

pˆ 'i pˆ ' j

r

N

3
2
1 Sj
Si

Ls
L
(a) (b)

Fig. 4 (a) schematic of long-span body divided into N subsections, (b) SPL to be added for long span.

In this study, we propose a method, revisiting the previous works of Kato et al. and Perot et al. First, consider a
long-span body, which is divided into N subsections by LS (i.e. L = N∙LS). Let the spectral acoustic pressure radiated
from the i-th subsection be l p 'i (see Fig. 4(a)), then the power spectral density (PSD) of acoustic pressure for the
entire span ( lp ' ) can be written as:
L

* N * N * N N
l
p 'L l
p 'L = ∑ l
p 'i ⋅ ∑ l
p ' j = ∑∑ Re( l
p 'i l
p'j ) , (10)
i=1 j =1 i=1 j =1

where * denotes a conjugate. Now, we need an assumption of ‘statistical homogeneity in the spanwise direction’
which satisfies the following properties. For LS,
1. The power spectral density of the acoustic pressure radiated from each subsection is the same, i.e.

2 2 2 2
l
p '1 = l
p '2 = ..... = l
p 'N = l
p 'S , (11)

where pˆ 'S is the spectral acoustic pressure radiated from the simulated span, LS.
2. The acoustic pressure radiated from each subsection is only lagged by a phase difference which can be
characterized by the following coherence function,

*
Re( l
p 'i l
p'j ) . (12)
γ 'ij =
2 2
l
p' l
p'
i j

3. The coherence function (phase lagging) is a function of ∆zij , the spanwise separation between two subsections, i.e.

γ 'ij = γ '(∆ zij ), ∆ zij = zi − z j = i − j ⋅ LS . (13)

In most turbulent flows around the long-span bodies, these assumptions are not so crude, if the observer’s position is
very far. Applying Eqs. (11)-(13), Eq. (10) can be re-written as

6
American Institute of Aeronautics and Astronautics
2 N N 2
l )⋅ l
p 'L = ∑∑ γ '(∆z
i=1 j =1
ij p 'S , (14)

where lp 'S is our known solution, i.e. the acoustic pressure radiated from the simulated span, LS.
So, we can now estimate the acoustic pressure for the long span by determining γ ’(∆zij). Since the phase lagging in
the spanwise direction tends to follow a Gaussian distribution18,19, γ ’(∆zij) can be expressed as

⎛ ∆zij 2 ⎞
γ '(∆zij ) = exp ⎜⎜ − 2 ⎟

, (15)
⎝ L 'c (ω ) ⎠

where L 'C is the spanwise coherence length.


From Eqs. (14) and (15), the SPL to be added for the long span, L is given by

⎛ pˆ 'L (ω ) 2 ⎞ ⎛ N N ⎛ Ls ⎞ ⎞ ⎞⎟ .
2
2⎛
SPLc (ω ) = 10 log ⎜ ⎟ = 10 log ∑∑ exp −(i − j ) ⎜
⎜ ⎜ ⎟ ⎟⎟ (16)
⎜ pˆ ' (ω ) 2 ⎟ ⎜ ⎜ ⎝ L ' (ω ) ⎠ ⎠ ⎟⎠
⎝ s ⎠ ⎝ i =1 j =1
⎝ c

The SPLC is plotted in Fig. 4(b) against L 'C / LS for various N ( = L / LS ) . The asymptotic behavior of Eq. (16) agrees
with Kato’s formula. If L 'c / Ls < 1/ π (≈ 0.56) , SPLC converges to 10 ⋅ log( N ) , while SPLC → 20 ⋅ log( N ) for
L 'c > L . It is interesting to note that in the range of LS < L 'C < L , SPLC is proportional to log( L 'C / LS ) , when N is
sufficiently large. Actually, γ ’(∆zij) is the coherence function between acoustic pressures emitted from each subsection
and experimentally measuring that is hardly possible. So, a spanwise coherence function of the surface pressure can be
used instead of γ ’(∆zij), with no significant difference.

IV. Computational Results and Discussion

A. Large Eddy Simulation


We consider a flow (UO=20m/s) over the flat plate at an angle of attack, α = 0. The plate has a chord length
c=10cm, thickness h=0.03c, span L=3c, and the plate edges have four corners cut with a radius of 0.5 mm in order
to smoothen the geometrical singularities. The Reynolds number based on the chord length, Rec is 1.3×105 and the
flow Mach number, Mo is 0.06. An incompressible large eddy simulation is performed for the computational domain
shown in Fig. 5. Flow periodicity is assumed in the spanwise direction for the simulated span, LS =0.03c. An o-type

r = 10c

U∞ Lz = h

h = 0.03c

Fig. 5 Computational domain and grid details near the trailing-edge of the flat plate.

7
American Institute of Aeronautics and Astronautics
(a) (b)
Fig. 6 Instantaneous hydrodynamic flow field at Rec=1.3×105 (Uo=20 m/s); (a) spanwise vorticity, (b) iso-
surfaces of Q(=200), non-dimensionalized by (Uo/h)2.

grid is used with 657×201×21 (about 2.8 millions) points in x, y, and z, and the computational domain is divided into
32 blocks for parallel computations. The minimum grid spacing for x and y is 0.0005c (or ∆x+min = ∆y+min =3) and a
uniform grid spacing of 0.0015c (or ∆z+ =15) is used in the spanwise direction. The grid details near the trailing-
edge of the plate are shown in Fig. 5.
Figure 6 shows (a) the instantaneous spanwise vorticity and (b) the iso-surfaces of the second invariant property
of the velocity gradients, Q(=200) over the plate. A separation occurs at the leading-edge and the shear layer
approximately reattaches at x= -0.8c . Then it breaks off into eddies that are convected downstream along the plate
surface. Fig. 6(b) also exhibits the expected von Kármán vortex shedding in the wake region. Figure 7 shows the
boundary layer characteristics over the flat plate: (a) the instantaneous u-velocity contours in a x-z plane at ∆y+ =3
and (b) the time-averaged boundary layer profiles. One can see in Fig. 7(a) that the shear layer is breaking in the
streamwise and spanwise directions and streaks are noticeable in the downstream of the boundary layer.

(a)

(b)
Fig. 7 Boundary layer characteristics; (a) instantaneous U-velocity at ∆y+=3, (b) boundary layer profiles
(time-averaged) at x=-0.2c, -0.4c, and -0.6c.

8
American Institute of Aeronautics and Astronautics
Fig. 8 Time history of surface pressure fluctuations at streamwise stations A, B, C, D, E, F, G, H, I; curves are
separated by a vertical off set of 0.1 with the corresponding zero lines located at -0.1, 0, 0.1, 0.2,…,0.7.

Figure 7(b) also shows the boundary layer profiles over the plate, which are expected to be turbulent but with a
pressure gradient effect imposed upon it. A boundary layer thickness, δ at x=-0.2c (upstream from the trailing-edge)
is 1.12h and the turbulent Reynolds number, Reτ is 231.
Now, unsteady characteristics of the boundary layer are described. Temporal variations of the wall pressure
fluctuations at various stations over the plate (from A to I) are exemplified in Fig. 8. First, one can identify that at
station A, pressure fluctuations with a range of scales are monitored because it is close to the reattachment point of
the shear layer (x= -0.8c). Small scale eddies are rapidly decaying but large-scale structures generated at

-5/3

(a) (b)
Fig. 9 (a) wall pressure PSD spectra along the plate and pressure PSD spectrum at the region of shear layer
breaking-off (green), (b) auto-correlation (SVV) spectrum of v-velocity fluctuations at 2h downstream from
the trailing-edge and on the wake centerline.

9
American Institute of Aeronautics and Astronautics
UC=0.83UO

(a) (b)
Fig. 10 (a) X-t plot of Cp (plate wall): exhibiting large-scale pressure fluctuations at St=0.01 with a convection
velocity, UC=0.83UO, (b) x-t plot of v-velocity fluctuations along the wake centerline: von Kármán vortex
shedding at St=0.2 with a convection velocity, UC=0.56UO.

very low frequency (St=fh/Uo=0.01 or 67Hz) seem to survive to the trailing-edge and influence the region close to
the stations at H and I. Of course, station H clearly shows the pressure fluctuations at the von Kármán vortex
shedding frequency (St=0.2 or 1333Hz). At station I, pressure fluctuations at St=0.01 are only visible.
Figure 9 presents the spectral characteristics of the flow fluctuations over the plate. Figure 9(a) shows power
spectral density (PSD) spectra of the wall pressure at various streamwise positions. At x= -0.8c (i.e. reattachment
point of the shear layer), pressure fluctuations are generated between St=0.01 and 0.4 and the PSD level is 20dB
higher than in the region downstream. Even though the magnitude of the pressure fluctuations is much greater, one
may find in the next acoustic calculation that this will not be the major noise source. The main noise source will be
at the region, where edge-scattering effect can occur, namely the plate trailing-edge region. At x= -0.5c (i.e. mid-
chord of the plate), the spectrum looks similar but the PSD level is significantly dropped. It is interesting to note that
the spectrum near the trailing-edge at x= -0.02c looks quite different from those in the upstream locations because of
the influence of the trailing-edge effect. A peak is noticeable at St=0.2 due to the von Kármán vortex shedding and
the aforementioned low frequency feature at St=0.01 is also moderately pronounced due to the edge-effect. This
edge-effect in the wall pressure PSD spectrum is alos found at station I, presented in Fig. 11(right). The spectrum
also shows high frequency characteristics between St=0.3 and 2 (or 2000Hz ∼ 13kHz), which are generated by the
breaking-up of eddies at the trailing-edge (i.e. transitional characteristics of boundary layer eddies into the wake).
Furthermore, the PSD of the pressure fluctuations at the breaking-off region of the leading-edge shear layer exhibits
a broadened peak at St=0.4. This frequency corresponds to the shear layer instability in the most unstably-amplified
mode20.
Meanwhile, Fig. 9(b) shows an auto-correlation spectrum of the v-velocity fluctuations at 2h downstream from
the trailing-edge along the wake centerline. The PSD profile is well-correlated with the trailing-edge PSD spectrum,
except for low frequency characteristics (0.01< St <0.04). This means that flow fluctuations at low frequency are not
related to the characteristics in the wake region. A x-t plot of the wall Cp presented in Fig. 10(a) clearly indicates
that large-scale pressure bursts ‘originated’ from the spanwise instability of the shear layer have a non-dimensional
time scale close to 100 (= tUO/h), which corresponds to the frequency close to St=0.01. It is also indicated that
pressure bubbles are convected downstream with UC=0.83UO. and On the contrary, the von Kármán vortices shed
from the trailing-edge are convected downstream in the wake with UC=0.56UO (see the x-t plot of v-velocity
fluctuations in Fig. 10(b)). Finally, the computed wall pressure PSD spectra are compared in Fig. 11 with the
experimental data measured at ECL: (a) at x= -0.02c from the plate trailing-edge and (b) at station I. The overall

10
American Institute of Aeronautics and Astronautics
Fig. 11 Comparison of wall pressure PSD spectra; at x=-0.02c (left) and station I (right).

agreement is reasonably good. The spectrum at station I shows two small peaks at St=0.2 (lift) and 0.4 (drag), which
are also confirmed in the computation. The computation over-predicts the measured PSD by 5dB but this
discrepancy is expected to be corrected by more refined computations in the future.

B. LPCE computation for acoustics


The flat plate self-noise at ReC=1.3×105 and Mo=0.06 (for α=0 deg.) is predicted by solving the linearized perturbed
compressible equations (LPCE). The LPCE calculation is conducted in a two-dimensional computational domain,
where it is coupled with the ‘spanwise-averaged’ incompressible LES solution. An o-type acoustic grid is used with
347×247 points and a minimal normal spacing at the wall is five times larger than the hydrodynamic one. Figure 12
shows the grid details near the trailing-edge and the noise sources, i.e. the von Kármán vortex shedding and the
convected eddies in the boundary layer. The noise source, DP/Dt field is acquired from the LES solution and
projected onto the coarse acoustic grid.
The instantaneous pressure fluctuation field ( ∆p ' = ( P + p ') − ( P + p ') ) computed by LPCE is presented in Fig.
13(a). First, a dipole tone generated by the von Kármán vortex shedding at the trailing-edge is clearly noticeable.
The acoustic wavelength, λ/c is close to 2.5, which corresponds to the frequency, St=0.2. Besides, one may note that
the other higher frequency components superimposed in the near-field are being generated from the trailing-edge
and the reattachment point of the shear layer. Due to the grid coarseness, however, acoustic waves with short
wavelengths are expected to be numerically dissipated as propagating away from the plate. Figure 13(b) also shows
the time history of the pressure fluctuations monitored at r=70h (or 2.1c) vertically away from the trailing-edge.

Fig. 12 Instantaneous DP/Dt field (spanwise-averaged) projected onto the acoustic grid (347×247: top) from
the hydrodynamic grid (657×201×21: bottom).

11
American Institute of Aeronautics and Astronautics
λ/c=2.5

(a) (b)
Fig. 13 (a) instantaneous pressure fluctuation field ( ∆p ' = ( P + p ') − ( P + p ') ), (b) time history of pressure
fluctuations at r=70h (or 2.1c) vertically away from the trailing-edge; non-dimensionalized by ρoco2.

In order to predict the far-field SPL spectrum, we follow the computational procedure described in section III-c.
Since the microphone is located 20c from the plate, the 2D acoustic field computed by LPCE within the domain of
10c needs to be extrapolated by the Kirchhoff method and also to be corrected to 3D spectral pressure by
Oberai’s method. Finally, the 3D spectral pressure in the far-field, which is radiated from span (h), needs to be
corrected for the total span (100h or 3c) used in the experiment. In this procedure, we need information on the
spanwise coherence function of the surface pressure , γ 2(z) in the most dominant noise source region. Figure 14(a)
shows the γ 2(z) function, which is computed by LES for h as well as measured for 3c (or 100h). The spanwise
coherence function rapidly drops in most frequencies, except for the tone at St=0.2. The spanwise coherence length
of the surface pressure, LC calculated by the Gaussian law, γ 2(z) =exp(-(z/LC)2) is shown in Fig. 14(b). The largest
value of LC is approximately 7h at St=0.2 but LC is mostly under 2h at other frequencies. The far-field SPL predicted
for the actual span is now compared in Fig. 15 with the measured data at ECL. The comparison is very good, except
at low frequency (St < 0.03). The measured SPL spectrum indicates an increase of SPL as the Strouhal number
decreases towards St=0.01 but the prediction fails to show it. The discrepancy comes from the fact that for the tone
at St=0.2, the microphone location (at 20c vertically away from the mid-chord of the plate) is 8 times larger than the
corresponding acoustic wavelength (2.5c) and one can say that the far-field approximation is quite valid. On the
other hand, at St=0.01, the acoustic wavelength is approximately 50c and the far-field assumption is not valid

(a) (b)
Fig. 14 (a) spanwise coherence; o: computations, (o, o): wind-tunnel measurements, (b) spanwise coherence
length; data interpolated by a Gaussian law, γ 2(z) =exp(-(z/LC)2).

12
American Institute of Aeronautics and Astronautics
Fig. 15 Comparison of SPL spectrum at r=20c vertically away from the mid-chord of the plate; computation
(blue), experiment (black).

any more. As a reference, the acoustic wave generated at St=0.025 has a wavelength comparable to the microphone
location (i.e. 20c). Finally, Fig. 16 shows the directivity patterns of ∆p’rms at r=20c for various Strouhal numbers (or
ratios of the plate chord length to the acoustic wavelength). At the von Kármán vortex shedding frequency (St=0.2
or c/λ=0.4), it clearly shows a dipolar pattern. The figure also shows that as Strouhal number increases (or acoustic

St=0.2 (c/λ=0.4) St=0.4 (c/λ=0.8)

St=1.0 (c/λ=2.0) St=2.0 (c/λ=4.0)

Fig. 16 Directivity patterns of ∆p’rms at r=20c.

13
American Institute of Aeronautics and Astronautics
wavelength becomes shorter than the chord length, the directivity pattern changes to a finger-like shape due to wave
diffractions around the leading and trailing edges of the plate. Finally, one can see that the present computational
methodologies for flow and acoustics reasonably capture the physical aspects of flat plate self-noise, i.e. the noise
generation mechanism and its interactions with the flow and structure.

V. Concluding remarks
The computational methodologies used in the present study accurately predict the flat plate self-noise (broadband
and tone) at ReC =1.3×105 and M=0.06. The comparison of the wall pressure PSD with the experiment confirms that
the present LES results show reasonable spectral accuracy of turbulent fluctuations over the plate. The acoustic field
computed by LPCE with the DP/Dt field projected onto the acoustic grid seems to indicate that noise sources
generated in the flow field are well represented in the acoustic computation. The estimated far-field SPL spectrum
predicted by an approximate but efficient methodology described in section III-c also shows the validity of the
method for a long-span body. From the results of experiment and computation, the flat plate self-noise sources are
characterized as:
(i) Broadened Tone (St~0.2) – von Karman vortex shedding near the trailing-edge
(ii) Broadband (0.02<St<0.1) – TBL eddy scattering at the trailing-edge (wall pressure fluctuations with a convection
velocity of Uc=0.83Uo are the foot-prints of the large-scale eddy motions in the upper boundary layer)
(iii) Low frequency (St<0.02) – spanwise instability of the leading-edge separation bubble
(iv) High frequency (St>0.3) – rapidly decaying.
Finally, it is worth to note that a low frequency noise at low Mach number requires a computational domain
sufficiently large enough to take care of all the validity of the methods we have used.

References
1
Brooks, T. F., and Hodgson, T. H., “Trailing Edge Noise Prediction from Measured Surface Pressures,” Journal of Sound
and Vibration, Vol. 78, No. 1, 1981, pp. 69-117.
2
Howe, M. S., “ A Review of the Theory of Trailing Edge Noise,” Journal of Sound and Vibration, vol. 61, No.3, 1978, pp.
437-465
3
Roger, M. and Moreau, S., “Broadband Self-Noise from Loaded Fan Blades,” AIAA Journal, Vol. 42, No. 3, 2004, pp. 536-
544.
4
Seo, J. H., and Moon, Y. J., “Linearized Perturbed Compressible Equations for Low Mach Number Aeroacoustics,” Journal
of Computational Physics (accepted for publication), 2006.
5
Seo, J. H., and Moon, Y. J., “Perturbed Compressible Equations for Aeroacoustic Noise Prediction at Low Mach Numbers,”
AIAA Journal, Vol. 43, No. 8, pp. 1716-1724, 2005.
6
Scott, J. N., Pilon, A. R., Lyrintzis, A. S., and Rozmajzl, T. J., “A Numerical Investigation of Noise From a Rectangular
Jet,” AIAA-Paper 1997-285, 1997.
7
Oberai, A. A., Roknaldin, F., and Hughes, T. J. R., “Trailing-Edge Noise Due to Turbulent Flows,” Technical Report,
Boston University, Report No. 02-002, 2002.
8
Pérennès S. & Roger M. – Aerodynamic Noise of a Two-Dimensional Wing With High-Lift Devices, 4th AIAA/CEAS
Aeroacoustics Conference, paper 1998- 2338, Toulouse, 1998.
9
Roger M., Moreau S. & Guédel A. – Vortex-Shedding Noise and Potential-Interaction Noise Modelling by a Reversed
Sears’ Problem, 12th AIAA/CEAS Aeroacoustics Conference, paper 2006- 2607, Cambridge Massasuchetts, 2006.
10
Gaitoned, D., Shang, J. S., and Young, J. L., “Practical Aspects of High-Order Accurate Finite-Volume Schemes for
Electromagnetics,” AIAA-paper 97-0363, 1997.
11
Edgar, N. B., and Visbal, M. R., “A General Buffer Zone-type Non-reflecting Boundary Condition for Computational
Aeroacoustics,” AIAA-paper 2003-3300, 2003.
12
Perot, F., Auger, J. M., Giardi, H., Gloerfelt, X., and Baily, C., “Numerical Prediction of the Noise Radiated by a Cylinder,”
AIAA-Paper 2003-3240, 2003.
13
Ewert, R., and Schroder, W., “On the Simulation of Trailing Edge Noise with a Hybrid LES/APE Method,” Journal of
Sound and Vibration, Vol. 270, 2004, pp. 509-524.
14
Manoha, E., Delahay, C., Sagaut, P., Mary, I., Khelil, S., and Guillen, P., “Numerical Prediction of Unsteady Flow and
Radiated Noise from a 3D Lifting Airfoil,” AIAA-Paper 2001-2133, 2001.
15
Kato, C., Iida, A., Takano, Y., Fujita, H., and Ikegawa, M., “Numerical Prediction of Aerodynamic Noise Radiated from
Low Mach Number Turbulent Wake,” AIAA-Paper 93-145, 1993.
16
Boudet, J., Casalino, D., Jacob, M. C., and Ferrand, P., “Prediction of Sound Radiated by A Rod Using Large Eddy
Simulation,” AIAA-Paper 2003-3217, 2003.

14
American Institute of Aeronautics and Astronautics
17
Ewert, R., Meinke, M., and Schroder, W., “Computation of Trailing Edge Noise of a 3D Lifting Airfoil in Turbulent
Subsonic Flow,” AIAA-Paper 2003-3114, 2003.
18
Casalino, D., and Jacob, M., “Prediction of Aerodynamic Sound from Circular Rods via Spanwise Statistical Modelling,”
Journal of Sound and Vibration, Vol. 262, No. 4, 2003, pp. 815-844.
19
Jacob, M. C., Boudet, J., Casalino, D., and Michard, M., “A Rod-Airfoil Experiments as a Benchmark for Broadband Noise
Modeling,” Theoret. Comput. Fluid Dynamics, Vol. 19, No. 3, 2005, pp. 171-196.
20
Blake, W. K., “Mechanics of Flow-Induced Sound and Vibration,” Vol. 1, Academic Press, Inc., 1986, pp. 131-135.

15
American Institute of Aeronautics and Astronautics

Potrebbero piacerti anche