ArticlePDF Available

Self-sustained oscillations with acoustic feedback in flows over a backward-facing step with a small upstream step

AIP Publishing
Physics of Fluids
Authors:

Abstract and Figures

Self-sustained oscillations with acoustic feedback take place in a flow over a two-dimensional two-step configuration: a small forward-backward facing step, which we hereafter call a bump, and a relatively large backward-facing step (backstep). These oscillations can radiate intense tonal sound and fatigue nearby components of industrial products. We clarify the mechanism of these oscillations by directly solving the compressible Navier-Stokes equations. The results show that vortices are shed from the leading edge of the bump and acoustic waves are radiated when these vortices pass the trailing edge of the backstep. The radiated compression waves shed new vortices by stretching the vortex formed by the flow separation at the leading edge of the bump, thereby forming a feedback loop. We propose a formula based on a detailed investigation of the phase relationship between the vortices and the acoustic waves for predicting the frequencies of the tonal sound. The frequencies predicted by this formula are in good agreement with those measured in the experiments we performed.
Content may be subject to copyright.
Self-sustained oscillations with acoustic feedback in flows
over a backward-facing step with a small upstream step
Hiroshi Yokoyama
a
Graduate School of Engineering, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo, 113-8656 Japan
Yuichi Tsukamoto
Hitachi Plant Technologies, Ltd., 4-5-2 Higashi-Ikebukuro, Toshima-ku, Tokyo, 170-8466 Japan
Chisachi Kato
Institute of Industrial Science, The University of Tokyo, 4-6-1 Komaba, Meguro-ku, Tokyo, 153-8505 Japan
Akiyoshi Iida
Faculty of Engineering, Kogakuin University, 2665 Nakano-machi, Hachioji-shi, Tokyo, 192-0015 Japan
Received 16 January 2007; accepted 6 September 2007; published online 23 October 2007
Self-sustained oscillations with acoustic feedback take place in a flow over a two-dimensional
two-step configuration: a small forward-backward facing step, which we hereafter call a bump, and
a relatively large backward-facing step backstep. These oscillations can radiate intense tonal sound
and fatigue nearby components of industrial products. We clarify the mechanism of these
oscillations by directly solving the compressible Navier-Stokes equations. The results show that
vortices are shed from the leading edge of the bump and acoustic waves are radiated when these
vortices pass the trailing edge of the backstep. The radiated compression waves shed new vortices
by stretching the vortex formed by the flow separation at the leading edge of the bump, thereby
forming a feedback loop. We propose a formula based on a detailed investigation of the phase
relationship between the vortices and the acoustic waves for predicting the frequencies of the tonal
sound. The frequencies predicted by this formula are in good agreement with those measured in the
experiments we performed. © 2007 American Institute of Physics. DOI: 10.1063/1.2793170
I. INTRODUCTION
Self-sustained oscillations with tonal sound, such as cav-
ity tones or edge tones, frequently take place when solid
bodies are placed in shear layers. These oscillations are de-
scribed as fluid-acoustic interactions and are interesting phe-
nomena both for aerodynamics and for aeroacoustics. More-
over, the prediction and the suppression of the oscillations
are very important for many practical applications. When the
oscillations are unsuppressed, the noise levels in open cavi-
ties in aircraft can exceed 150 dB.
1
These intense oscillations
can fatigue nearby components such as aircraft wheel-wells
and landing-gear configurations.
2
For decades, many researchers have used various ap-
proaches to investigate the relationship between the flow and
the sound in these phenomena. Brown
3
experimentally found
that the peak frequencies of the edge tones change discon-
tinuously against the position of the edge and the mean flow
velocity. Powell
4
proposed that the interactions of vortices in
a jet with a sharp edge generate acoustic waves that lead to
the formation of new vortices, and the frequencies predicted
by his proposal agree well with Brown’s experimental
results.
3
Meanwhile, Rossiter
5
proposed a similar feedback
loop over open cavities and derived a semiempirical formula,
fL/u
0
=n
/M +1/
, where f is the frequency of the
radiated tonal sound, L is the cavity length, u
0
is the free
stream velocity, n is a positive integer, M is the Mach num-
ber based on the free stream velocity,
is the ratio of the
convection velocity of vortices to the free stream velocity,
and
is a constant for the phase correction. The frequencies
predicted by this formula agree with those measured in
experiments.
6
However, this model does not provide the val-
ues of
and
theoretically, treating them instead as empiri-
cal constants to be determined by the best fit with measured
data.
In theoretical studies, the shear layer was idealized as a
vortex sheet, and the phase distributions of the vortices and
the sound were estimated for edge tones, cavity tones, and
aperture tones. The results enabled researchers to predict the
frequencies of the tonal sound without using empirical or
adjustable parameters Bilanin and Covert,
7
Tam and Block,
8
Crighton,
9
and Howe
10
. However, the phase distributions
estimated by these studies have not been validated by experi-
ments or computations. In recent years, the direct numerical
simulation DNS, where both vortices and acoustic waves
are directly simulated, has provided a means of studying the
self-sustained oscillation mechanism in detail
11,12
see the re-
view of Colonius and Lele
13
. However, few studies have
been done about the detailed mechanism of vortex shedding
and acoustic radiation, including the phase relationship be-
tween the vortices and the sound in the self-sustained oscil-
lations. Therefore, the physical meaning of constant
in
Rossiters formula
5
remains unknown.
The self-sustained oscillations with tonal sound also take
place in a flow over a two-step configuration: a small
forward-backward facing step, which we hereafter call a
a
Electronic mail: yokoyama@iis.u-tokyo.ac.jp
PHYSICS OF FLUIDS 19, 106104 2007
1070-6631/2007/1910/106104/8/$23.00 © 2007 American Institute of Physics19, 106104-1
Downloaded 20 Nov 2007 to 133.11.199.19. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
bump, and a relatively large backward-facing step or back-
step Fig. 1. This configuration appears in various industrial
devices, for example, the small bump on the surface of an
automobile side mirror. The tonal sound caused by the oscil-
lations is uncomfortable for the passengers, so the mecha-
nism of the oscillations needs to be clarified so that the
sounds can be predicted and reduced. However, few studies
have been done on the self-sustained oscillations in this con-
figuration, and there are many open questions about the for-
mation of vortices around the bump and about the acoustic
radiation. The objective of the present work is to answer
these questions, focusing on the phase relationship between
the vortices and the sound.
This work mainly consists of computational studies, but
we also conducted wind-tunnel experiments to validate the
computational results. In Sec. II, we explain the flow con-
figurations, the numerical methods, and the experimental
setup. In Sec. III A, we discuss the pressure fluctuation at a
point far from the backstep and confirm that tonal sound is
radiated. In Sec. III B, we present instantaneous vortices and
acoustic waves in detail. In Sec. III C, acoustic sources are
clarified. In Sec. III D, the phase relationship between the
vortices and the acoustic waves are clarified. Based on this
relationship, we propose a formula for predicting the fre-
quencies of the radiated tonal sound in Sec. III E. The fre-
quencies predicted by this formula are compared with those
measured in the experiments that we performed.
II. METHODOLOGY
A. Flow configurations
The baseline flow configurations and parameters for the
present work are shown in Fig. 1 and Table I, respectively.
The bump is placed in a laminar boundary layer. The
Reynolds number Re
x
based on the value of x measured from
the virtual origin is 4.610
5
at the leading edge of the
bump. Mach number M based on free-stream velocity u
0
is
0.088 in the experiments, but in the computations we inves-
tigate the flow with higher Mach numbers M =0.3, 0.4, and
0.5 to reduce the computational costs.
Boundary layer thickness is 2.4 h at the leading edge of
the bump, and bump length L
b
is 14.3 h, where h is bump
height. Distance L
1
between the trailing edge of the bump
and that of the backstep is 14.3 h. We have clarified in pre-
liminary experiments that intense tonal sound is radiated
from the flow with these configurations. Backstep height H is
21.6 h in the computations and 71.4 h in the experiments. In
the computations, the flow fields are developed more quickly
by shortening backstep height, which also determines the
necessary streamwise length of the computational domain.
To clarify the effect of backstep height on the mechanism of
the self-sustained oscillations, we compare the computational
results of the baseline configurations with those where back-
step height H is 43.1 h in Sec. III E.
B. Numerical methods
1. Governing equations
We simulate both flow and acoustic fields by directly
solving the three-dimensional compressible Navier-Stokes
equations in a conservative form, which are written as
U
t
+ E E
v
x
+ F F
v
y
+ G G
v
z
=0, 1
where U is the vector of the conservative variables, E, F, and
G are the inviscid fluxes, and E
v
, F
v
, and G
v
are the viscous
fluxes. The spatial derivatives are evaluated by the sixth-
order-accurate compact finite difference scheme fourth-
order accurate on the boundaries.
14
Time integration is per-
formed by a third-order accurate Runge-Kutta method. To
suppress the numerical instabilities associated with the cen-
tral differencing in the compact scheme, we use the tenth-
order-accurate spatial filter shown below,
15
f
ˆ
i−1
+
ˆ
i
+
f
ˆ
i+1
=
n=0
5
a
n
2
i+n
+
in
, 2
where
is a conservative quantity and
ˆ
is the filtered quan-
tity. Coefficients a
n
are the same as the values used by
Gaitonde and Visbal,
16
and parameter
f
is 0.45. We have
clarified in preliminary tests that we can capture acoustic
waves by 10 points per wave ppw using the above compu-
tational methods.
2. Computational grid
To evaluate the effect of the three-dimensional distortion
of relatively large vortices in the wake of the backstep on the
radiated sound, we use a three-dimensional computational
grid for the present simulations Fig. 2. The spanwise extent
of the computational domain is L
z
=3H, and the cross section
in the x-y plane is shown in Fig. 3. We use 11 grids in the
spanwise direction and investigate grid dependence by
changing the number of the grids in the spanwise direction in
Sec. III A. The grid resolution we used for acoustic propaga-
tion was more than 40 grids per wavelength of the funda-
mental frequency to suppress the dissipation of the acoustic
waves, and 11 grids per bump height h so that small flow
structures generated by the bump can be captured. The
present computational domain consists of approximately
2.5 10
5
structured grids.
FIG. 1. Flow configurations.
TABLE I. Parameters.
Boundary layer Re
x
L
b
/hL
1
/h
/hH/h
Laminar 4.610
5
14.3 14.3 2.4 21.6
106104-2 Yokoyama et al. Phys. Fluids 19, 106104 2007
Downloaded 20 Nov 2007 to 133.11.199.19. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
3. Initial conditions, boundary conditions,
and buffer zone
All the simulations were initiated by the uniform flow
with free-stream velocity u
0
over the bump and the backstep.
Figure 3 shows the boundary conditions used in the
present studies. The artificial boundaries must allow vortices
and acoustic waves to pass smoothly with minimal distur-
bances. In this study, the nonreflecting boundary conditions
based on the characteristic wave relations
1719
are used at the
inflow, upper, and outflow boundaries. Nonslip and adiabatic
boundary conditions are used at the wall. The periodic
boundary conditions are used in the spanwise direction.
To minimize errors associated with the artificial bound-
aries, buffer zones where we stretch the grid to reduce the
amplitudes of the acoustical and vortical fluctuations before
they reach the boundaries are used near the nonreflecting
boundaries. Moreover, we must keep boundary layer thick-
ness
constant at the leading edge of the bump because the
radiated sound strongly depends on the boundary layer thick-
ness. To achieve desired thickness
at the leading edge of
the bump, Eq. 1 is replaced in the upstream buffer zone by
U
t
+ E E
v
x
+ F F
v
y
+ G G
v
z
=−
0
a/L
b
2
U U
target
, 3
where
0
is 0.05, a is the speed of sound, L
b
is the length of
the buffer zone,
is the nondimensional distance from the
beginning of the buffer zone to the inflow boundary 0
1, and U
target
is the target vector of the conservative
variables.
20,21
U
target
is obtained from the calculation for the
flat-plate laminar boundary layer.
C. Experimental setup
We conducted our experiments using the suction-type
low-noise wind tunnel shown in Fig. 4. The test section is
composed of a backstep accompanied by a small bump
placed upstream of the backstep. In the spanwise direction,
the test section is terminated by two end walls. One of these
walls is made of a porous plate to suppress the reflections of
the sound, and the other is made of an acrylic plate, which
allows optical observations. The velocity profile and the
sound pressure level are measured, respectively, with a hot
wire anemometer and a sound level meter with a nondirec-
tional 1/2 in. microphone Rion NL-31.
At a wind speed of 30 m/s, the turbulence intensity is
less than 0.7% and the nonuniformity of the mean flow ve-
locity is less than 0.1%. The velocity profile at the point of
the leading edge of the bump is measured without the bump.
The profile agrees with the Blasius profile, so the boundary
layer is apparently laminar Fig. 5. At a wind speed of
30 m / s, the 99% boundary layer thickness, the momentum
thickness, the displacement thickness, and the shape factor
are 1.7 mm, 0.23 mm, 0.58 mm, 2.6, respectively. The test
section is placed in an anechoic chamber that is covered with
50 mm thick sound-absorbing material, and the background
noise level is suppressed to less than 64 dB at a wind speed
of 30 m / s, which is much lower than the sound pressure
level of the tonal sound of our interest.
To account for errors and uncertainties, we estimated
that the error of the measurement of the streamwise velocity
is 0.2 m/ s at a wind speed of 30 m / s and that of the bump
height is 0.1 mm. We found that the uncertainty about the
FIG. 2. Computational grid every fifth grid line is shown for clarity.
FIG. 3. Boundary conditions.
FIG. 4. Experimental setup.
FIG. 5. Velocity profile of boundary layer. Measured data is compared
with the Blasius profile .
106104-3 Self-sustained oscillations with acoustic feedback Phys. Fluids 19, 106104 2007
Downloaded 20 Nov 2007 to 133.11.199.19. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
peak frequency of the tonal sound is about 40 Hz when the
peak frequency is 2370 Hz. To evaluate the repeatability of
the experiments, we repeated the experiments and measured
the frequency spectra 10 times at a frequency resolution of
10 Hz. We found that the 95% uncertainty range of the peak
frequency of the radiated sound was 24 Hz.
III. RESULTS AND DISCUSSION
A. Self-sustained oscillations with tonal sound
Figure 6 shows the fluctuation of the pressure coefficient
Cp =p /
0
u
0
2
/2兲兴 for M =0.4, where
0
is the density of the
ambient air at the observation point Fig. 3 after 20 flow-
through times have elapsed since the start of the computa-
tion. This shows that the flow is fully developed from the
initial condition and that the oscillations are occurring. We
also investigated the oscillations when either the bump or the
backstep had been removed Fig. 7. Figure 8 shows a com-
parison of the frequency spectra of the pressure fluctuations
at the observation point. We averaged the results of the Fou-
rier transform 20 times and the duration of the transform is
about 1300 h / u
0
. This duration corresponds to about 30
periods of the fundamental mode. The results clearly indicate
that the tonal sound with the fundamental frequency
St= fh/u
0
=0.024 is radiated only when both the bump and
the backstep exist. We also found that the fundamental fre-
quencies of the tonal sound are St= 0.025 and 0.023, respec-
tively, for M =0.3 and 0.5 for the baseline configuration. We
investigated the effect of the grid resolution and the width of
the computational domain by conducting the computations
with 21 grids over Lz=3H the original width and Lz=6H
for M =0.4. Figure 9 shows that the frequency spectra do not
change from that of the original spectrum. Therefore, we
concluded that the original resolution and width are suffi-
cient to capture the effect of the distortion of the large vor-
tices in the wake of the backstep on the radiated sound.
B. Instantaneous distributions of vortices
and acoustic waves
We investigated the vortical structures and the acoustic
fields in oscillations when the flow is fully developed from
the initial condition. To elucidate the flow structures, the sec-
ond invariant of the velocity gradient tensor Q=
2
S
2
was computed, where and S are, respectively, the antisym-
metric and symmetric parts of the velocity gradient tensor.
Regions with Q0 represent vortex tubes. We present the
acoustic fields using density fluctuations. Using these meth-
ods of visualization, we were able to clearly capture the flow
structures and the acoustic waves in the near field. Figure 10
shows instantaneous flow structures and acoustic waves for
M =0.4 from t =0 to t=0.81T, where T represents the period
of the fundamental frequency. A small vortex “Vortex 3” at
t=0 is shed from the leading edge of the bump and con-
vected downstream. While being convected from the trailing
edge of the bump to the backstep, the vortex merges with its
FIG. 6. Fluctuation of pressure coefficient Cp at the observation point
shown in Fig. 3.
FIG. 7. Configurations without bump top and without backstep bottom.
FIG. 8. Power spectrum density of pressure coefficient Cp for baseline
configuration compared with a simple backstep without a bump ---
and a bump on a flat plate without a backstep ¯·.
FIG. 9. Power spectrum density of pressure coefficient Cp for the original
case of 11 grids over Lz =3H compared with the case of 21 grids over
Lz=3H --- and with that of 21 grids over Lz=6H ¯·.
106104-4 Yokoyama et al. Phys. Fluids 19, 106104 2007
Downloaded 20 Nov 2007 to 133.11.199.19. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
upstream vortex and becomes a larger vortex “Vortex 2” at
t=0.54T. When it passes the trailing edge of the backstep,
the vortex splits into a fast vortex and a slow vortex “Vortex
1-1” and “Vortex 1-2” at t= 0.27T, and the convection ve-
locity differs between the two vortices depending on the dis-
tance from the wall. A compression wave “Compression
wave 2” at 0.27T is radiated between this slow vortex
and its upstream vortex. Meanwhile, an expansion wave
“Expansion wave 2” at 0.27T is radiated from the vortices
distorted in the wake of the backstep. The behaviors of the
acoustic wave fronts show that the acoustic waves are propa-
gated in the upstream direction along the wall at a velocity of
about a-u
0
.
Figure 11 shows the trajectories of the vortices y =1.6 h
and those of compression waves y = 3.6 h represented by
the local maxima of the second invariant and density, respec-
tively. This figure shows that the shed of a vortex coincides
with the crossing of the compression wave. Figure 12 shows
the vortices and the acoustic waves around the leading edge
of the bump from t = 0.54T to t = 0.81T, including the vectors
of velocity fluctuations. The velocity fluctuation virtually
represents the sound particle velocity, although it is affected
by the vortices near the wall. Due to acoustic expansion, the
gradient of the sound particle velocity in the streamwise di-
rection changes suddenly from a negative to a positive value
just after the most compressed wave local maximum line of
density passes. This is because the gradient of pressure fluc-
tuation is very steep around the maximum point, as shown in
Fig. 6. As a result, the vortex formed by the flow separation
at the leading edge of the bump begins to stretch and finally
a new vortex, “Vortex 4,” is shed at t = 0.81 T when the
compression wave “Compression wave 2” passes this vor-
tex in the upstream direction. It also shows that there are
trajectories of modes higher than the fundamental mode of
St= 0.24.
C. Acoustic sources
To identify the source positions of the radiated tonal
sound, we investigated the amplitudes of the pressure fluc-
tuations for the fundamental frequency for M =0.4 Fig. 13.
The local maxima of the fluctuations appear in both the up-
stream and downstream regions of the backstep. The up-
stream local maximum represents the source of the compres-
sion waves radiated between the vortex at the trailing edge of
the backstep and its upstream vortex t = 0.27 T in Fig. 10.
The distance between this point and the trailing edge of the
backstep is L
3.6h. The same value of L
is also obtained
for M = 0.3 and 0.5. This value is important for predicting the
frequencies of the radiated tonal sound because the compres-
sion waves cause new vortices at the leading edge of the
bump. Meanwhile, the local maximum downstream of the
FIG. 10. Contours of density fluctuations
/
0
and isosurfaces of
Q/ u
0
/
2
=0.03. Contour levels range from −5 10
−3
dark to 5 10
−3
light.“T represents the period of fundamental frequency.
FIG. 11. Trajectories of vortices and compression waves over the
bump.
FIG. 12. Contours of density fluctuations
/
0
and isosurfaces of
Q/ u
0
/
2
=0.03. Contour levels range from −5 10
−3
dark to 5 10
−3
light. Vectors represent fluctuations of velocity. T represents the period
of fundamental frequency.
106104-5 Self-sustained oscillations with acoustic feedback Phys. Fluids 19, 106104 2007
Downloaded 20 Nov 2007 to 133.11.199.19. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
backstep represents the source of the expansion waves,
which are radiated from the vortices distorted in the wake of
the backstep t=0.27 T in Fig. 10.
D. Phase distributions of vortices
and acoustic waves
We investigated phase distributions of the second in-
variant of the velocity gradient tensor and fluid density. Fig-
ure 14 shows the phase distributions along the observation
line shown at the top for the fundamental frequency for
M =0.3, 0.4, and 0.5. This observation line approximates the
trajectory of the vortex because the amplitude of the second
invariant fluctuations along this line takes its maximum in
the wall normal direction Fig. 15. The reference point is
above the trailing edge of the backstep, and the phase of the
second invariant at this point is the reference for both the
phase distributions of the second invariant and those of fluid
density.
The phase distributions of the second invariant represent
the convection of the vortex, while those of the fluctuating
density around the leading edge of the bump virtually repre-
sent the upstream propagation of the radiated sound waves
because the density fluctuations caused by the convection of
the vortices are negligibly small around this edge. Figure 14
shows that the phase of the second invariant coincides with
that of the fluctuating fluid density at a downstream point of
the leading edge of the bump. At this point, the compression
waves cause the vortex formed by the flow separation at the
leading edge of the bump to begin to stretch. The distance
between this point and the leading edge of the bump is
L
5h for all the investigated Mach numbers.
The gradient of the phase distributions of the second
invariant also gives the convection velocity of the vortices.
We obtained the averaged convection velocity u
cb
=0.29u
0
on
the bump and u
c1
=0.19u
0
from the trailing edge of the bump
to the backstep. The convection velocity u
cb
on the bump is
approximately confirmed by the trajectories of the vortices in
Fig. 11. The vortices are convected very close to the wall,
and the boundary layer is laminar. As a result, these vortices
are strongly affected by the wall, and the convective veloci-
ties of these vortices are much slower than those of vortices
in free shear layers such as deep cavities. Moreover, the
phase distributions of the second invariant also show that the
distance between the vortex at the trailing edge of the back-
step and its upstream vortex is very narrow. This is because
one of the splitting vortices over the trailing edge of the
backstep “Vortex 1-2” in Fig. 10 is convected with a much
slower velocity as already mentioned in Sec. III B. A com-
pression wave is radiated between this slow vortex and its
upstream vortex.
E. Formula for frequencies of the tonal sound
The discussion so far focuses on the self-sustained oscil-
lation mechanism and the phase relationship between the
vortices and the acoustic waves. The cycle of the self-
sustained oscillations consists of the following three parts:
FIG. 13. Amplitudes of pressure coefficient Cp for the fundamental fre-
quency. Contour levels range from 10
−10
dark to 10
−4
light with the
logarithmic scale.
FIG. 14. Phase changes of second invariant solid line and fluctuating fluid
density dashed line along the observation line for fundamental frequency.
106104-6 Yokoyama et al. Phys. Fluids 19, 106104 2007
Downloaded 20 Nov 2007 to 133.11.199.19. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
1 A small vortex is shed and begins to be convected at
position L
from the leading edge of the bump when a com-
pression wave passes this point.
2 The vortex is convected downstream at velocity u
cb
on the bump and u
c1
from the trailing edge of the bump to
the backstep.
3 The above-mentioned vortex splits into a fast vortex
and a slow vortex when it passes the backstep. Between this
slow vortex and its upstream vortex, a compression wave is
radiated at position L
from the trailing edge of the backstep.
This wave is propagated at a velocity of a-u
0
in the upstream
direction.
The phase variation of the above processes must be 2
n,
where n is a positive integer. As a result, the frequencies f
n
of the tonal sound can be predicted by the following formula:
f
n
=1/T
n
,
T
n
= 共共L
b
L
/u
cb
+ L
1
/u
c1
+ L
b
+ L
1
L
L
/a u
0
兲兲/n, 4
L
= 5.0h, L
= 3.6h, u
cb
= 0.29u
0
, u
c1
= 0.19u
0
.
To be exact, the point of the vortex shed is slightly dif-
ferent from the point of the isophase for
and Q, when Fig.
14 is compared with Fig. 11. Therefore, we need to introduce
the phase delay into this formula to predict the exact peak
frequency. However, it is difficult to estimate the delay theo-
retically and the delay is not large less than 10% of the
fundamental period. The main objective of this paper is to
clarify the self-sustained oscillations around the backstep
with a bump and this simple formula is sufficient for this
objective.
The frequencies predicted by this formula are compared
with those measured in the experiments. As already men-
tioned, the backstep heights for the computational and the
experimental configurations are different. We therefore esti-
mated the effect of the backstep height on the radiated sound.
Figure 16 compares the sound radiated from the original con-
figurations with backstep height H = 21.6 h with that from the
configurations with H = 43.1 h. Figure 16 shows that the dif-
ference in the frequency spectrum is small. We concluded
that the phenomena of the self-sustained oscillations around
the bump do not depend on the backstep height as long as the
backstep height is sufficiently larger than the bump height.
Figure 17 compares the frequency spectra of the pressure
fluctuations at the observation point in the computations with
the sound pressure level measured by the microphone in the
experiment. Mode n = 6 is observed in both the experiment
and the computations, but the lower mode, n = 3, is only ob-
served in the computations. We think this is because of the
difference in Mach numbers. In fact, the amplitude of n =3
decreases and that of n = 6 increases as the Mach number
decreases in our computations. Also, Fig. 18 shows the
changes in measured peak frequency for bump length L
b
where M =0.088 in the experiments. The predicted frequen-
cies are in good agreement with those measured in the ex-
periments. This agreement validates the self-sustained oscil-
lation mechanism clarified by the present computational
FIG. 15. Amplitudes of second invariant fluctuations Q
/u
0
/h
2
near the
bump. Contour levels range from 10
−8
dark to 10
0
light with the loga-
rithmic scale. Dashed line --- represents the observation line for phase
investigation.
FIG. 16. Power spectrum density of pressure coefficient Cp for baseline
configuration compared with that for the bump with a higher backstep
---.
FIG. 17. Power spectrum density of pressure coefficient Cp in computations
compared with sound pressure level measured in experiment circles are
every tenth point.
FIG. 18. Change in fundamental frequency for L
b
. Fundamental frequencies
predicted by Eq. 3 are compared with experimental results .
106104-7 Self-sustained oscillations with acoustic feedback Phys. Fluids 19, 106104 2007
Downloaded 20 Nov 2007 to 133.11.199.19. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
study. In the present paper, we did not discuss the absolute
values of the sound pressure level. For the quantitative pre-
diction of the sound pressure level, it is necessary to exactly
simulate the boundary conditions for both acoustic propaga-
tion and flow convection.
IV. CONCLUSION
We numerically investigated flow structures and radiated
sound in a flow over a backstep with a small upstream bump
by directly solving the three-dimensional compressible
Navier-Stokes equations. The results showed that a small
vortex is shed from the leading edge of the bump and con-
vected downstream to the backstep. When this vortex passes
the backstep, it splits into a fast and a slow vortex. A com-
pression wave is radiated between this slow vortex and its
upstream vortex. Meanwhile, expansion waves are radiated
from the vortices distorted in the wake of the backstep. The
compression waves propagated toward the leading edge of
the bump in turn shed new vortices by stretching the vortex
formed by the flow separation at this edge. We also investi-
gated the acoustic sources and the phase distributions of the
second invariant of the velocity gradient tensor and fluctuat-
ing fluid density. As a result, we were able to identify the
phase relationship between the vortices and the acoustic
waves. Based on this relationship, we proposed a formula for
predicting the frequencies of the radiated tonal sound. The
frequencies predicted by this formula are in good agreement
with those measured in the experiments. This agreement
validated the present computational results.
The methods that we used to identify the phase relation-
ship between the vortices and the acoustic waves in this pa-
per are also useful for the clarification of other phenomena
associated with fluid-acoustic interactions. Moreover, we be-
lieve that this clarification leads to an improved design guide
for various industrial products for the suppression of the
trouble caused by fluid-acoustic interactions discussed here.
1
H. H. Heller and W. M. Dobrzynski, “Sound radiation from aircraft wheel-
well/landing-gear configurations,” J. Aircr. 14, 768 1977.
2
O. W. Mcgregor and R. A. White, “Drag of rectangular cavities in super-
sonic and transonic flow including the effects of cavity resonance,” AIAA
J. 8, 1959 1970.
3
G. B. Brown, “The vortex motion causing edge tones,” Proc. Phys. Soc.
London 49, 493 1937.
4
A. Powell, “On the edge tone,” J. Acoust. Soc. Am. 33,3951961.
5
J. E. Rossiter, “Wind-tunnel experiments on the flow over rectangular
cavities at subsonic and transonic speeds,” Aero. Res. Counc. R&M, No.
3438 1964.
6
M. A. Kegerise, E. F. Spina, S. Garg, and L. N. Cattafesta, “Mode-
switching and nonlinear effects in compressible flow over a cavity,” Phys.
Fluids 16, 678 2004.
7
A. J. Bilanin and E. E. Covert, “Estimation of possible excitation frequen-
cies for shallow rectangular cavities,” AIAA J. 11, 347 1973.
8
C. K. W. Tam and P. J. W. Block, “On the tones and pressure oscillations
induced by flow over rectangular cavities,” J. Fluid Mech. 89, 373 1978.
9
D. G. Crighton, “The jet edge-tone feedback cycle; linear theory for the
operating stages,” J. Fluid Mech. 234, 361 1992.
10
M. S. Howe, “Edge, cavity and aperture tones at very low Mach num-
bers,” J. Fluid Mech. 330,611997.
11
L. Larcheveque, P. Sagaut, I. Mary, O. Labbe, and P. Comte, “Large-eddy
simulation of a compressible flow past a deep cavity,” Phys. Fluids 15,
193 2003.
12
C. W. Rowley, T. Colonius, and A. J. Basu, “On self-sustained oscillations
in two-dimensional compressible flow over rectangular cavities,” J. Fluid
Mech. 455, 315 2002.
13
T. Colonius and S. K. Lele, “Computational aeroacoustics: progress on
nonlinear problems of sound generation,” Prog. Aerosp. Sci. 40, 345
2004.
14
S. K. Lele, “Compact finite difference schemes with spectral-like resolu-
tion,” J. Comput. Phys. 103,161992.
15
K. Matsuura and C. Kato, “Large-eddy simulation of compressible transi-
tional flows in a low-pressure turbine cascade,” AIAA J. 45,4422007.
16
D. V. Gaitonde and M. R. Visbal, “Pad-type higher-order boundary filters
for the Navier-Stokes equations,” AIAA J. 38, 2103 2000.
17
K. W. Thompson, “Time dependent boundary conditions for hyperbolic
systems,” J. Comput. Phys. 68,11987.
18
T. J. Poinsot and S. K. Lele, “Boundary conditions for direct simulations
of compressible viscous flows,” J. Comput. Phys. 101,1041992.
19
J. W. Kim and D. J. Lee, “Generalized characteristic boundary conditions
for computational aeroacoustics,” AIAA J. 38, 2040 2000.
20
J. B. Freund, “Proposed inflow/outflow boundary condition for direct com-
putation of aerodynamic sound,” AIAA J. 35,7401997.
21
J. Larsson, L. Davidson, M. Olsson, and L. Eriksson, “Aeroacoustic in-
vestigation of an open cavity at low mach number,” AIAA J. 42,2462
2004.
106104-8 Yokoyama et al. Phys. Fluids 19, 106104 2007
Downloaded 20 Nov 2007 to 133.11.199.19. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp
... for the upstream damping region shown in Fig. 3, where a is the speed of sound, L d is the streamwise length of the damping region, x d is the non-dimensional distance from the downstream end of the damping region (0 ≤ x d ≤ 1), and Q target is the target vector of the conserved variables. (13) Coefficient σ 0 is set to 0.05. It was demonstrated in a previous study that acoustic reflection does not occur with this setting. ...
... It was demonstrated in a previous study that acoustic reflection does not occur with this setting. (13) At the inflow boundary of the computational domain, the Reynolds number, Re x (i.e., ρ ∞ u 1∞ x/µ, where x is the distance measured from the origin of the laminar boundary layer), and boundary-layer thickness, δ int /L, are assumed to be 9.0×10 3 and 0.016, respectively. ...
Article
We investigate the fluid-acoustic interactions in fluid-dynamic oscillations of a flow over a twodimensional cavity by directly solving the compressible Navier-Stokes equations. The upstream boundary layer is turbulent and the depth-to-length ratio of the cavity is 0.5. To clarify the effects of the freestream Mach number on the fluid-dynamic oscillations, we perform computations for the two freestream Mach numbers of 0.15 and 0.3. The oscillations occur with the Mach number of 0.3, while the oscillations do not occur with the Mach number of 0.15. For the Mach number of 0.3, phase-averaged flow fields show that large-scale vortices form in the shear layer. When the largescale vortex collides with the downstream wall, the low-pressure region spreads along the downstream wall. As a result, the local pressure gradient induces a local downstream velocity, causing the upstream fluid to expand. Finally, an expansion wave propagates to the outside of the cavity. In order to clarify the mechanism for the formation of the large-scale vortices, we also perform the computations of backward-facing step flows with an artificial acoustic source. As a result, it is shown that the large-scale vortices originate from the convective disturbances, which are generated by the acoustic waves. The disturbances grow due to the Kelvin-Helmholtz instability, similar to the growth of the disturbances in a laminar cavity flow. Also, these computations of the backwardfacing step flows clarify the reason why oscillations do not occur in the cavity flow for the Mach number of 0.15.
... For flow oscillations in the region downstream of the step, e.g. vortex shedding and flapping of bubbles, acoustic radiations have been suggested as a possible mechanism in compressible flow (Yokoyama et al. 2007). For incompressible flow, Wee et al. (2004) found that at Re = 5550, the Strouhal number of the self-sustained vortex shedding is St = O(0.1), ...
Article
Full-text available
Amplifications of flow past a backward-facing step with respect to optimal inflow and initial perturbations are investigated at Reynolds number 500. Two mechanisms of receptivity to inflow noise are identified: the bubble-induced inflectional point instability and the misalignment effect downstream of the secondary bubble. Further development of the misalignment results in decay of perturbations from \x=28\ onwards (the step is located at \x=0\), as has been observed in previous non-normality studies (Blackburn et al., J. Fluid Mech., vol. 603, 2008, pp. 271–304), and eventually limits the receptivity. The receptivity is found to be maximized at an inflow perturbation frequency of \{\it\omega}=0.50\ and a spanwise wavenumber of \{\it\beta}=0\, where the inflow noise takes full advantage of both mechanisms and is amplified over two orders of magnitude in terms of the velocity magnitude. In direct numerical simulations (DNS) of the flow perturbed by optimal or random inflow noise, vortex shedding, flapping of bubbles, three-dimensionality and turbulence are observed in succession as the magnitude of the inflow noise increases. Similar features of linear and nonlinear receptivity are observed at higher Reynolds numbers. The Strouhal number of the bubble flapping is 0.08, at which the receptivity to inflow noise reaches a maximum. This Strouhal number is close to reported values extracted from DNS or large eddy simulations (LES) at larger Reynolds numbers (Le et al., J. Fluid Mech., vol. 330, 1997, pp. 349–374; Kaiktsis et al., J. Fluid Mech., vol. 321, 1996, pp. 157–187; Métais, New Trends in Turbulence, 2001, Springer; Wee et al., Phys. Fluids, vol. 16, 2004, pp. 3361–3373). Methods to further clarify the mechanisms of receptivity and to suppress the noise amplifications by modifying the base flow using a linearly optimal body force are proposed. It is observed that the mechanisms of optimal noise amplification are fully revealed by the distribution of the base flow modification, which weakens the bubble instabilities and misalignment effects and subsequently reduces the receptivity significantly. Comparing the base flow modifications with respect to amplifications of inflow and initial perturbations, it is found that the maximum receptivity to initial perturbations is highly correlated with the receptivity to inflow noise at the optimal frequency \{\it\omega}=0.50\, and the correlation reduces as the inflow frequency deviates from this optimal value.
... Substituting Eqs. (10) and (11) into Eq. (9), we obtain the following equation. ...
Article
The main steam stop valve (MSV) in thermal or nuclear power plants is liable to undergo excitation by sound having a specific frequency. The purposes of this article are to clarify the mechanism of the sound generation in the MSV using numerical analysis and to propose a suppression method for the sound generation. The analysis results were validated experimentally using a 1/5-scale model of a typical MSV. The analysis results showed good agreement with the experimental data. From the analysis results, it was clarified that the sound of the MSV was generated by the hole-tone in the opening and the acoustic resonance in the vertical direction of the MSV. A method was proposed to install triangular or slanted tabs at the inlet seat of the MSV in order to prevent formation of vortex rings in the opening which leads to the hole-tone. Effects of the proposed method were tested experimentally. Eight triangular tabs having a height of 15% of the inner diameter in the seat reduced the peak value of the sound pressure by 60%. Eight slanted tabs having a height of 7.5% of the inner diameter in the seat suppressed generation of the hole-tone in the MSV. These results showed that the slanted tabs are effective in suppression of the hole-tone compared with the triangular tabs.
Chapter
This paper is on the effects of a novel span vortex generator on the periodic components of the turbulent shear layers and the Reynolds stresses. The turbulence modelling approach is 3D simulations IDDES, a hybrid RANS/LES technique. The geometry for the study is taken from the experimental configurations for the case. The case comprises a turbulent flow over a backward facing step (BFS), where separation is induced after the step edge. The results from the simulations are compared to the experimental data with and without control. Spanwise vortex generators consist of a strip of magnets placed along the span of the BFS upstream of step and the device oscillates at a given frequency of 280 Hz and amplitude of 0.002m. Turbulent structures, Reynolds stresses, skin friction distributions and velocities are analysed and compared to the experimental measurements. A remarkable effect of the device is observed especially in the reattachment length which is considerably reduced. Experimental measurements for the baseline case were available and a comparison with the available data is performed.
Article
Aerodynamic noise generated around a step such as front pillars of automobiles, which is transmitted into the cabin through the window glass, can negatively affect the comfort of the passengers. For the reduction of this noise, it is necessary to clarify the acoustic source and process of the propagation of the acoustic waves from the step. To do this, direct aeroacoustic simulations were performed along with wind tunnel experiments for flows around a forward step in a turbulent boundary layer, where the freestream Mach number was 0.1 and the Reynolds number based on the incoming boundary layer thickness, δ, and the freestream velocity was 2.73×10⁴. The computations were performed for the height of h/δ = 0.9 and 3.4, where the effects of the step height on the radiated sound were investigated. The radiated sound becomes more intense for the higher step of h/δ = 3.4 particularly in a low frequency range of St ≤ 1.5, where St is the non-dimensional frequency based on the step height and freestream velocity, due to the occurrence of the large-scale vortices in the separated flow. To elucidate the radiation of the acoustic waves, the pressure fluctuations were decomposed into non-radiating convective and radiating acoustic components by usage of two methods. One is utilizing wavenumber-frequency spectra and the other one is spatial Gaussian filtering for the fluctuations at each frequency, which is proposed in this paper. The results by the former method present that the level of convective components is larger than the acoustic components by 20 - 40 dB. Moreover, the radiating components by the proposed spatial filtering method show that the radiation occurs around the step edge and the reattachment point. It is indicated that the proposed method is effective for the investigation of the acoustic radiation for the pressure fluctuations composed of convective and acoustic components.
Article
Full-text available
We investigate the fluid-acoustic interactions in fluid-dynamic oscillations of a flow over a two-dimensional cavity by directly solving the compressible Navier-Stokes equations. The upstream boundary layer is turbulent and the depth-to-length ratio of the cavity is 0.5. To clarify the effects of the freestream Mach number on the fluid-dynamic oscillations, we perform computations for the two freestream Mach numbers of 0.15 and 0.3. The oscillations occur with the Mach number of 0.3, while the oscillations do not occur with the Mach number of 0.15. For the Mach number of 0.3, phase-averaged flow fields show that large-scale vortices form in the shear layer. When the large-scale vortex collides with the downstream wall, the low-pressure region spreads along the downstream wall. As a result, the local pressure gradient induces a local downstream velocity, causing the upstream fluid to expand. Finally, an expansion wave propagates to the outside of the cavity. In order to clarify the mechanism for the formation of the large-scale vortices, we also perform the computations of backward-facing step flows with an artificial acoustic source. As a result, it is shown that the large-scale vortices originate from the convective disturbances, which are generated by the acoustic waves. The disturbances grow due to the Kelvin-Helmholtz instability, similar to the growth of the disturbances in a laminar cavity flow. Also, these computations of the backward-facing step flows clarify the reason why oscillations do not occur in the cavity flow for the Mach number of 0.15.
Article
Full-text available
To clarify the mechanism of acoustic radiation in flows around a cascade of flat plates, fluid structures and acoustic fields were elucidated by direct simulations. The simulations were mainly performed for flows around 5 parallel plates and the separation-to-thickness ratio s/d was 6.0. The freestream velocity was changed from 30 m/sec to 60 m/sec, and the acoustic resonance occurs between plates at the freestream velocity of 44 m/sec. At that velocity, the Reynolds number based on the chord length and the freestream velocity was 8.7×104 and that based on the plate thickness was 5.8×103. Computational results were validated by the experimental results performed in the present research. The computational results showed that large-scale vortices composed of fine-scale vortices are shed in the wakes of the plates independently of the acoustic resonance. When the large-scale vortex is shed from the upper or lower face, an expansion wave is radiated around the downstream edge of the upper or lower face, respectively. The compression wave is radiated around the downstream edge of the opposite face. The simulation for the flow around a single plate was also performed, and the results confirmed the above-mentioned acoustic radiation mechanism. For the flows around a cascade of flat plates, the sheddings of the vortices from neighboring plates are synchronized when the acoustic resonance occurs. It was also clarified that the mode of the synchronization is an anti-phase mode and the standing waves generated between plates are reinforced.
Article
Full-text available
Intense tonal sound radiates from flows around a trailing edge with an upstream kink shape such as found in an automotive body. To clarify the acoustic radiation mechanism and the effects of the kink shape angle on tonal sound, direct numerical simulations (DNS) of flows and sound fields and wind tunnel experiments were performed. As a result, the feedback loop of the sound generation was clarified as below. A vortex is shed between the kink shape and the trailing edge, and convects downstream. The acoustic wave radiates from distortion of the vortex due to the interference with the wall around the trailing edge. The acoustic waves are propagated and lead to the vibrating of shear layers around the kink shape. The disturbances due to this vibrating are amplified in the downstream of the reception point of sound, and a vortex is again shed. The reception point is located slightly downstream of the flow separation point. The positions of the acoustic source and the acoustic reception in this feedback loop were also clarified. Meanwhile, the tonal sound dose not radiate for smaller kink shape angle, where the vortex formation does not occur around the kink shape by the lack of flow separation. Also, for the larger kink shape angle, the large separation prompts turbulent transition and the tonal sound does not radiate.
Article
Several studies relating to strongly resonating cavity shear layers in subsonic or supersonic flows are reviewed and discussed. Special consideration is given to cavity configurations which Rockwell and Naudascher (1978) branded "whistle-type" cavities, deep cavities and cavities mounted in the wall of an otherwise solid-walled duct (such as a wind tunnel). At certain flow conditions, each configuration can promote a strong, acoustic resonance which tends to favor a single dominant frequency in the pressure spectrum. The resultant shear layer motions and unsteady pressure fluctuations may be of large amplitude at the peak frequency. The empirical constants in Rossiter's formula and a similar frequency prediction formula for partially-covered cavities are also discussed.
Article
抄録 In order to calculate aerodynamic noise radiated from a complex-shaped body such as a pantograph of Shinkansen, an algorithm of the boundary element method for solving shape-adapted Green's functions is parallelized. The dense matrix of pluralistic simultaneous linear equations is divided and reallocated artfully on local memories assigned to each CPU of a massively parallel computer, which enable to handle numerical models with a large number of boundary elements. After modification, bench mark tests demonstrate high parallel performance with hundreds of processes. Then incompressible flow around the pantograph is calculated by the large eddy simulation (LES), and its results are coupled with shape-adapted Green's functions to evaluate aerodynamic noise based on formulas of vortex sound theory. By extracting the net contribution from diffracted and scatted sound sources that are distributed downstream of shafts and covers, sound pressure in the far field proves to agree well with wind tunnel experimental data.
Article
Full-text available
Numerical simulations are used to investigate the resonant instabilities in two-dimensional flow past an open cavity. The compressible NavierHelmholtz mode. The wake mode is characterized instead by a large-scale vortex shedding with Strouhal number independent of Mach number. The wake mode oscillation is similar in many ways to that reported by Gharib & Roshko (1987) for incompressible flow with a laminar upstream boundary layer. Transition to wake mode occurs as the length and/or depth of the cavity becomes large compared to the upstream boundary-layer thickness, or as the Mach and/or Reynolds numbers are raised. Under these conditions, it is shown that the Kelvin–Helmholtz instability grows to sufficient strength that a strong recirculating flow is induced in the cavity. The resulting mean flow is similar to wake profiles that are absolutely unstable, and absolute instability may provide an explanation of the hydrodynamic feedback mechanism that leads to wake mode. Predictive criteria for the onset of shear-layer oscillations (from steady flow) and for the transition to wake mode are developed based on linear theory for amplification rates in the shear layer, and a simple model for the acoustic efficiency of edge scattering.
Article
Full-text available
Multiple distinct peaks of comparable strength in unsteady pressure autospectra often characterize compressible flow-induced cavity oscillations. It is unclear whether these different large-amplitude tones (i.e., Rossiter modes) coexist or are the result of a mode-switching phenomenon. The cause of additional peaks in the spectrum, particularly at low frequency, is also unknown. This article describes the analyses of unsteady pressure data in a cavity using time-frequency methods, namely the short-time Fourier transform (STFT) and the continuous Morlet wavelet transform, and higher-order spectral techniques. The STFT and wavelet analyses clearly show that the dominant mode switches between the primary Rossiter modes. This is verified by instantaneous schlieren images acquired simultaneously with the unsteady pressures. Furthermore, the Rossiter modes experience some degree of low-frequency amplitude modulation. An estimate of the modulation frequency, obtained from the wavelet analysis, matches the low-frequency peak seen in the autospectrum. Higher-order spectral methods were employed to investigate potential quadratic nonlinear interactions between the Rossiter modes and to determine if they are responsible for the low-frequency mode present in the autospectrum. In turn, this low-frequency mode could interact with the Rossiter modes to modulate their amplitude. Significant nonlinearities, in the form of sum and difference frequencies of the Rossiter modes, are present in the /d=2 cavity at M∞=0.4, while nonlinear effects are much smaller in the /d=4 at M∞=0.6. The bispectral analysis indicates that quadratic interactions between Rossiter modes in the near-field pressure are not responsible for the observed low-frequency peak in the pressure autospectrum. Furthermore, the low-frequency mode does not exhibit a strong nonlinear coupling with the Rossiter modes.
Article
The use of procedures based on higher-order finite-difference formulas is extended to solve complex fluid-dynamic problems on highly curvilinear discretizations and with multidomain approaches. The accuracy limitations of previous near-boundary compact filter treatments are overcome by derivation of a superior higher-order approach. For solving the Navier-Stokes equations, this boundary component is coupled to interior difference and filter schemes with emphasis on Pade-type operators. The high-order difference and filter formulas are also combined with finite-sized overlaps to yield stable and accurate interface treatments for use with domain-decomposition strategies. Numerous steady and unsteady, viscous and inviscid flow computations on curvilinear meshes with explicit and implicit time-integration methods demonstrate the versatility of the new boundary schemes.
Article
The drag of relatively short rectangular cavities (length-to-depth ratios of 0.50-3.0) with turbulent shear layers has been measured at transonic and supersonic Mach numbers (0.3- 3.0). Of particular importance is the effect of pressure oscillations within the cavity (commonly referred to as cavity resonance) which is shown to increase the drag as much as 250%. Cavity resonance is found to occur over the entire Mach number range investigated and other work has shown it to occur at both lower and higher Mach numbers. The frequency of the pressure oscillations is best predicted by the vortex-wave interaction model presented by Rossiler. The effects of external reinforcement of resonance by reflection of radiated pressure waves are examined both experimentally and analytically. Existing methods for predicting cavity drag are inadequate to cope with this phenomena. A new model for predicting the lower bound of cavity drag (nonresonating cavity) is presented and agrees well with experimental data. Analytical considerations are utilized to indicate the qualitative effect of resonance on cavity drag. © 1970, American Institute of Aeronautics and Astronautics, Inc., All rights reserved.
Article
An experimental program was initiated to determine the noise radiation from landing gear/wheel well configurations of large commercial aircraft. Scale models of typical nose gear and main gear (synthesized from different type aircraft) were exposed to flow of typical landing approach speeds (up to 65 m/sec) on a stationary outdoor wall jet flow facility and attached to the wings of an aerodynamically very clean glider (SB 10). Landing gear noise is composed of sound generated by the interaction of flow with: the wheel well volume; and the external gear equipment. Wheel well related sound is characterized by discrete tones whose intensity is heavily damped by the spoiling effect of the gear. Externally generated landing gear noise is mostly of broad band character. The contributions of some dominant features of a gear (shaft, struts, actuators, doors, wheels) to the total sound signature were determined and normalized nose gear and main gear spectra were developed that predict measured full scale landing gear noise fairly well.
Article
Large-eddy simulation of compressible transitional flows in a low-pressure turbine cascade is performed by using sixth-order compact difference and a 10th-order filtering method. Numerical results without freestream turbulence and those with about 5% of freestream turbulence are compared. In these simulations, separated flows in the turbine cascade accompanied by laminar-turbulent transition are realized, and the present results closely agree with past experimental measurements in terms of the static pressure distribution around the blade. In the case where no freestream turbulence is taken into account, the unsteady pressure field essentially differs from that with strong freestream turbulence. In the no freestream turbulence case, pressure waves that propagate from the blade's wake region have noticeable effects on the separated-boundary layer near the trailing edge and on the neighboring blade. Also, based on the snapshot proper orthogonal decomposition analysis, dominant behaviors of the transitional I p boundary layers are investigated.
Article
The laminar flow and the near-field acoustics of an open cavity at Mach number 0.15 is computed by direct solution (two-dimensional) of the compressible Navier-Stokes equations. The radiated sound is also computed by an acoustic analogy, a modified version of Curle's equation, and compared to the directly computed sound field. The agreement is found to be good. The contributions from the various source terms in Curle's equation are quantified, and the terms involving wall-pressure fluctuations are found to account for most of the radiated intensity. These main source terms are investigated further, and it is found that they are especially large in the downstream part of the cavity, and for about two cavity lengths downstream of the cavity. The upstream dominance of the radiated sound is explained by the fact that the downstream cavity wall contributes primarily to the upstream direction. Correlations between the radiated sound and the source terms, coupled with where the source terms are large, suggest that the near-field terms display a stronger upstream dominance than the far-field terms, and hence the far-field directivity is expected to be flatter.