ArticlePDF Available

Fluid-acoustic interactions in self-sustained oscillations in turbulent cavity flows. I. Fluid-dynamic oscillations

AIP Publishing
Physics of Fluids
Authors:

Abstract and Figures

The fluid-acoustic interactions in a flow over a two-dimensional rectangular cavity are investigated by directly solving the compressible Navier–Stokes equations. The upstream boundary layer is turbulent. The depth-to-length ratio of the cavity is 0.5. Phase-averaged flow fields reveal the mechanism for the acoustic radiation. Large-scale vortices form in the shear layer that separates from the upstream edge of the cavity. When a large-scale vortex collides with the downstream wall, the low-pressure fluid in the vortex spreads along the downstream wall. As a result, a local downward velocity is induced by the local pressure gradient, causing the upstream fluid to expand. Finally, an expansion wave propagates to the outside of the cavity. The large-scale vortices originate from the convective disturbances that develop in the shear layer. The disturbances grow due to the Kelvin–Helmholtz instability, similar to the growth of those in a laminar cavity flow. To clarify the mechanism for the generation of the initial convective disturbances, computations for backward-facing step flows with an artificial acoustic source are also performed. As the artificial acoustic waves become more intense, the initial convective disturbances in the shear layer become more intense while the spatial growth rate of these disturbances does not change. This means that the initial convective disturbances in the shear layer are induced by the acoustic waves. © 2009 American Institute of Physics.
Content may be subject to copyright.
Fluid-acoustic interactions in self-sustained oscillations in turbulent cavity
flows. I. Fluid-dynamic oscillations
Hiroshi Yokoyama1,aand Chisachi Kato2
1Graduate School of Engineering, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku,
Tokyo 113-8656, Japan
2Institute of Industrial Science, The University of Tokyo, 4-6-1 Komaba, Meguro-ku, Tokyo 153-8505, Japan
Received 25 March 2009; accepted 18 September 2009; published online 23 October 2009
The fluid-acoustic interactions in a flow over a two-dimensional rectangular cavity are investigated
by directly solving the compressible Navier–Stokes equations. The upstream boundary layer is
turbulent. The depth-to-length ratio of the cavity is 0.5. Phase-averaged flow fields reveal the
mechanism for the acoustic radiation. Large-scale vortices form in the shear layer that separates
from the upstream edge of the cavity. When a large-scale vortex collides with the downstream wall,
the low-pressure fluid in the vortex spreads along the downstream wall. As a result, a local
downward velocity is induced by the local pressure gradient, causing the upstream fluid to expand.
Finally, an expansion wave propagates to the outside of the cavity. The large-scale vortices originate
from the convective disturbances that develop in the shear layer. The disturbances grow due to the
Kelvin–Helmholtz instability, similar to the growth of those in a laminar cavity flow. To clarify the
mechanism for the generation of the initial convective disturbances, computations for
backward-facing step flows with an artificial acoustic source are also performed. As the artificial
acoustic waves become more intense, the initial convective disturbances in the shear layer become
more intense while the spatial growth rate of these disturbances does not change. This means that
the initial convective disturbances in the shear layer are induced by the acoustic waves. © 2009
American Institute of Physics.doi:10.1063/1.3253326
I. INTRODUCTION
Self-sustained oscillations with fluid-acoustic interac-
tions in a flow over a cavity often radiate intense tonal
sound. If unsuppressed, the sound levels in open cavities can
reach 160 dB at a freestream Mach number of 0.8.1These
strong oscillations can cause fatigue in nearby components
such as aircraft wheel wells and landing gear compartments.
The intense tonal sound radiated by the deep cavity of a
train-car gap is a problem to be overcome in the develop-
ment of faster trains.2Large-amplitude pressure pulses can
occur in gas transport systems with closed side branches
even when the freestream Mach number is lower than 0.2.3,4
In short, prediction of oscillations during the design stage
and invention of ways to suppress them are important in the
development of many flow-related industrial products.
Many researchers over the past 50 years have investi-
gated the mechanism of self-sustained oscillations in a cavity
flow. Rossiter5described an oscillation mechanism similar to
that presented for edge tones by Powell.6In this mechanism,
the interactions of vortices with the downstream edge radiate
acoustic waves, which leads to the formation of new vortices
at the upstream edge. Rossiter derived a semiempirical for-
mula to explain this: fL/u1=n
/M+1/
, where fis
the frequency of the radiated tonal sound, Lis the cavity
length, u1is the freestream velocity, nis a positive integer,
Mis the freestream Mach number,
is the ratio of the con-
vection velocity of the vortices to the freestream velocity,
and
is a constant for the phase correction. The frequencies
predicted by this formula agree with the peak frequencies of
the tonal sound measured in his experiments for a cavity
flow with an upstream turbulent boundary layer turbulent
cavity flowat high Mach numbers M0.4.5East7mea-
sured pressure fluctuations in a deep cavity with a large
depth-to-length ratio D/L1in a turbulent boundary layer
at low Mach numbers M0.2. He found that intense os-
cillations occur only when the shear layer instability caused
by the fluid-acoustic interactions described by Rossiter5is
coupled with the acoustic resonance of the cavity.
Rockwell and Naudascher8classified the self-sustained
oscillations in a cavity flow into two categories. They called
oscillations due to shear layer instability without the reso-
nance described by Rossiter5“fluid-dynamic oscillations”
and those due to the coupling between the shear layer insta-
bility and the acoustic resonance of the cavity “fluid-resonant
oscillations.” For fluid-dynamic oscillations in a laminar cav-
ity flow, Knisely and Rockwell9found by experiments that
disturbances generated at the upstream edge by acoustic
feedback are amplified in the shear layer by the Kelvin–
Helmholtz KHinstability, resulting in the formation of vor-
tices. On the basis of the feedback loop model described by
Rossiter,5Rowley et al.10 suggested a criterion for the onset
of fluid-dynamic oscillations in a laminar cavity flow, which
takes into account the cavity configuration and the
freestream Mach number. This criterion agrees with their
two-dimensional computational results obtained using the
compressible Navier–Stokes equations.10 They did not dis-
cuss oscillations in a turbulent cavity flow. With regard to
aElectronic mail: yokoyama@iis.u-tokyo.ac.jp.
PHYSICS OF FLUIDS 21, 105103 2009
1070-6631/2009/2110/105103/13/$25.00 © 2009 American Institute of Physics21, 105103-1
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
fluid-resonant oscillations in a turbulent cavity flow, by using
particle image velocimetry, Oshkai and Yan3showed that
large-scale vortices form in the shear layer and that they
become acoustic sources. However, the mechanism for the
formation of these vortices and that for the acoustic radiation
have not been described in detail.
Our objective is to clarify in detail the mechanism for
the formation of the large-scale vortices by acoustic feedback
and that for the acoustic radiation by these vortices with
regard to the self-sustained oscillations of a turbulent cavity
flow by directly solving the compressible Navier–Stokes
equations. As mentioned, it is important to suppress the os-
cillations of turbulent cavity flows in various industrial prod-
ucts. In this paper, the fluid-dynamic oscillations are focused
on. In a subsequent paper, the fluid-acoustic interactions of
fluid-resonant oscillations will be focused on.
II. NUMERICAL METHODS
A. Flow configurations
The flow over a two-dimensional cavity, as shown in
Fig. 1, is investigated. The incoming boundary layer is tur-
bulent. The parameters are shown in Table I. The freestream
Mach number Mis 0.3. The ratio of the momentum thickness
to the cavity length
/L, where
is calculated from a sepa-
rate computation for the flow over a flat plate without a cav-
ity, is 0.04 at the position of the upstream edge of the cavity.
The ratio of the momentum thickness of the shear layer to
the cavity length
m/Lis 0.06 at x1/L=0.5, where the origin
of the coordinate is at the position of the upstream edge of
the cavity. The momentum thickness is defined as
m=
ymin
ymax u1av
u1
1−u1av
u1
dx2,1
where u1av is the time- and spanwise-averaged streamwise
velocity. The boundary ymin/L= −0.1 is adopted to remove
the effect of the inverse flow in the cavity. Also, ymax/Lis set
to 1.0 because the u1av/u1is approximately 1.0 at this po-
sition. The parameters of Mand
/Lhave the same values as
those used by Mizushima et al.2in their investigation of a
cavity flow. In their experiment, they measured sound radi-
ating from the flow over the cavity of a train-car gap, for
which D/Lranged from 0.5 to 2.6 around the train car, and
found that self-sustained oscillations occurred.
Preliminary computations of cavity flow were performed
for different values of D/L0.5, 0.9, 1.3, 1.7, 2.1, and 2.5
while keeping the cavity length and incoming boundary layer
thickness constant. The results showed that fluid-dynamic
oscillations occurred when D/Lwas 0.5 and that fluid-
resonant oscillations occurred when it was 0.9.
Here the fluid-dynamic oscillations are clarified for a
cavity flow when D/Lis 0.5 L/D=2.0,
/D=0.08.A
subsequent paper will focus on the fluid-acoustic interactions
of the fluid-resonant oscillations for deep cavities with
D/L0.9. The Reynolds number ReL=
u1L/
, where
and
are, respectively, the density and viscosity of the
free stream,is3.0104in the present computation, whereas
it was 6.6105in Mizushima’s experiment.2The effects of
this difference in the Reynolds number on the fluid-acoustic
interactions in fundamental oscillations are small, as ex-
plained in Sec. III D.
Computations for a backward-facing step backstep
flow, as shown in Fig. 2, are also performed. In the backstep
flow, tonal sound is not radiated by itself. To clarify the
mechanism for the formation of large-scale vortices in the
shear layer of the cavity flow due to acoustic feedback, the
effects of artificial acoustic waves on the shear layer of the
backstep flow are investigated. The boundary layer at the
trailing edge is the same as that at the upstream edge of the
cavity in the cavity flow. The backstep height His the same
as the cavity depth of 0.5L.
B. Governing equations and finite difference
formulation
Both flow and acoustic fields are simultaneously simu-
lated by directly solving the three-dimensional compressible
Navier–Stokes equations in the conserved form
Qt+
xk
FkFvk=0, 2
where Qis the vector of the conserved variables, Fkis the
inviscid flux vector, and Fvkis the viscous flux vector.
The spatial derivatives are evaluated using the sixth-
order-accurate compact finite difference scheme fourth or-
der accurate on the boundaries.11 The time integration is
done using the third-order-accurate Runge–Kutta method.
Tur
b
u
l
ent
b
oun
d
ary
l
ayer
D
L
θ
x1
x2
x
3
FIG. 1. Configuration for flow over two-dimensional cavity.
TABLE I. Parameters.
Label MD/LReL
/L
m/L
Present computation 0.3 0.5 3.01040.04 0.06
Train-car gap experimenta0.3 0.5–2.6 6.61050.04 ¯
aReference 2.
Tur
b
u
l
ent
b
oun
d
ary
l
ayer
H
θ
x1
x2
x3
Acoustic sourc
e
FIG. 2. Configuration for backward-facing step flow.
105103-2 H. Yokoyama and C. Kato Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
To reduce the computational cost, large-eddy simulations
LESare performed in the present work. Nevertheless, the
present computational grid incorporated with the above-
mentioned numerical methods adequately resolves the small-
est active vortices in the boundary layer and in the shear
layer in the cavity. To derive the governing equations for
LES, any variable
is decomposed into its grid-scale GS
¯
and subgrid-scale SGS
components,
=
¯
+
.3
In a compressible flow, the GS quantity is recast in terms of
a Favre-filtered variable
˜
=
¯
.4
The GS components of the conserved variables and the
fluxes are given as
Q=1
J
¯
,
¯
u
˜
1,
¯
u
˜
2,
¯
u
˜
3,
¯
e
˜
t,5
Fk=1
J
¯
u
˜
k
¯
u
˜
1u
˜
k+p
¯
¯
u
˜
2u
˜
k+p
¯
¯
u
˜
3u
˜
k+p
¯
¯
e
˜
+p
¯
u
˜
k
,6
Fvk=1
J
0
˜
k1
˜
k2
˜
k3
u
˜
j
˜
kj +
T
˜
xk
,7
˜
kl =
T
˜
u
˜
k
xl
+
u
˜
l
xk
2
3
kl
u
˜
n
xn
,8
=
T
˜
C/Pr, 9
where
kl is the viscous stress tensor,
is the thermal con-
ductivity, and
kl is the Kronecker’s delta function. The fluid
is assumed to be the standard air. Sutherland’s formula is
used for the viscosity coefficient
. It is assumed that the
specific heat Cis 1004 J kg−1 K−1 and that the Prandtl num-
ber Pr is 0.72. No explicit SGS model is used here: the
turbulent energy in the GS that should be transferred to SGS
eddies is dissipated by the tenth-order spatial filter, as de-
scribed below. It was shown by many researchers1214 that
the above-mentioned method combining the low-dissipation
discretization schemes and the explicit filtering correctly re-
produces turbulent flows such as a cavity flow and a round
jet. This filter removes numerical instabilities at the same
time,15
f
ˆi−1 +
ˆi+
f
ˆi+1 =
n=0
5an
2
i+n+
in,10
where
is a conserved quantity and
ˆis the filtered quantity.
The coefficients anhave the same values as those used by
Gaitonde and Visbal,16 and the parameter
fis 0.45. In Sec.
III, these computational methods are validated.
C. Computational grid
Figure 3shows the computational grids for the cavity
and backstep flows. The origins of the coordinates are at the
position of the upstream edge of the cavity and that of the
trailing edge of the backstep, respectively. The spanwise ex-
tents of these computational domains Lware 0.5Land Hfor
the cavity and backstep flows, respectively. This extent cor-
responds to Lw
+=700, where the wall units are based on the
skin-friction coefficient computed for the boundary layer
over a flat plate at the streamwise position corresponding to
the upstream edge of the cavity and the trailing edge of the
backstep, Rex=4.3105. Therefore, approximately six tur-
bulent streaks lie in the spanwise extent. The cross sections
in the x1-x2plane are shown in Fig. 4, where length is non-
dimensionalized by cavity length Land backstep height H
for the cavity and backstep flows, respectively. The effects of
the extent of the computational domain on the results are
discussed in Sec. III C.
To capture the longitudinal vortices in the turbulent
boundary layer, the grid resolutions are set near the wall as
x1
+=38, x2
+=1, and x3
+=14, where the wall units are
based on the above-mentioned skin-friction coefficient. To
capture the acoustic propagation accurately, more than 20
grid points are used per acoustic wavelength of the funda-
mental frequency of the oscillations even in the field far from
the cavity. It is clarified in Sec. III that the present computa-
tional grid adequately resolves both the longitudinal vortices
in the turbulent boundary layer and the acoustic waves cre-
ated by the fundamental oscillations. The present computa-
tional domain consists of approximately 5.0106structured
grid points.
FIG. 3. Computational grids for cavity flow leftand backward-facing step
flow right. Every tenth grid line is shown for clarity.
105103-3 Fluid-acoustic interactions in self-sustained oscillations Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
D. Boundary conditions
Figure 4also shows the boundary conditions for the cav-
ity and backstep flows. The inflow and outflow boundaries
are artificial, so that they must allow vortices and acoustic
waves to pass smoothly with minimal numerical distur-
bances. Nonreflecting boundary conditions based on the
characteristic wave relations1719 are used at these bound-
aries. Nonslip and adiabatic boundary conditions are used at
the wall. The periodic boundary condition is used in the
spanwise x3direction. The use of the periodic condition is
appropriate in the present study because the disturbances in
the shear layer two dimensionally grow to large-scale vorti-
ces due to the KH instability and radiate two-dimensional
acoustic waves in the fundamental oscillations, as explained
in Sec. IV.
To achieve the desired momentum thickness of the
boundary layer at the upstream edge of the cavity, Eq. 2is
replaced with
Qt+
xk
FkFvk=−
0a/Ldxd
2QQtarget11
for the upstream damping region shown in Fig. 4, where ais
the speed of sound, Ldis the streamwise length of the damp-
ing region, xdis the nondimensional distance from the down-
stream end of the damping region 0xd1, and Qtarget is
the target vector of the conserved variables.20,21 The coeffi-
cient
0is set to 0.05. It was demonstrated in a previous
study that acoustic reflection does not occur with this
setting.22 At the inflow boundary of the computational do-
main, the Reynolds number Rex=
u1x/
, where xis the
distance measured from the origin of the laminar boundary
layerand boundary layer thickness
int /Lare assumed to be
9.0103and 0.016, respectively.
Laminar-turbulent transition is prompted by super-
imposing homogeneous freestream turbulence on the solu-
tions for the compressible laminar boundary layer QBlasius for
1.25x2/
int 21.3 0.02x2/L0.34in the damping re-
gion −16.0x1/L−15.7, where x2= 0 corresponds to the
position of the wall. That is,
Qtarget =QBlasiusx2/
int 1.25,21.3 x2/
int,
12
Qtargett=QBlasius +Qfstt兲共1.25 x2/
int 21.3.
The freestream turbulence Qfsttis obtained by sliding an
instantaneous freezed flow field of homogeneous turbulence
computed by a separate LES. In the simulation of the homo-
geneous turbulence, the initial random field has the Karman
spectrum.23 The turbulence intensity is 6%, and the longitu-
dinal integral scale is about 0.24L. This integral scale corre-
sponds to about 0.06, where is the acoustic wavelength of
the fundamental frequency of the sound radiating from the
cavity flow, and the effect of the freestream turbulence on the
propagation of the radiated acoustic waves is negligibly
small. Also, it was confirmed that the amplitude of the spu-
rious noise generated by the superimposition of this
freestream turbulence is less than 10% of that of the sound
generated by the oscillations on the bottom of the cavity,
where we observe the pressure fluctuations in Sec. IV B.
Therefore, it is concluded that the effects of the freestream
turbulence on the oscillations of the cavity flow are also
small. It is demonstrated in Sec. III B that the boundary layer
becomes a fully developed turbulent state by the upstream
edge of the cavity.
E. Initial flow fields
The computational result for the turbulent flow over a
flat plate is used as the initial flow field for the present com-
putation of the turbulent cavity flow. In the cavity region
0x1Land x20, the velocity is set to 0, and the
pressure and density are set to the same values as those at
x2=0.
III. VALIDATION OF COMPUTATIONAL METHODS
A. Propagation of acoustic waves
The effects of the grid resolution on the propagation of
acoustic waves were preliminarily investigated by using an
artificial monopole. As a result, it was clarified that five grid
points per one wavelength accurately capture the acoustic
waves. Therefore, the present computational grids for the
main computations, for which more than 20 grid points are
used per one acoustic wavelength, sufficiently resolve the
propagation of the acoustic waves for the fundamental
oscillations.
B. Turbulent boundary layer
To confirm that the boundary layer has become a fully
developed turbulent state by the upstream edge of the cavity,
the flow over a flat plate without a cavity is calculated. Fig-
ure 5shows the variation in the computed time- and
spanwise-averaged skin-friction coefficient cfcompared with
x1/L
x2
/
L
0
10.0
0 1.0 7.5
-0.5
Damping region
-16.0
Outflow
Inflow
Adiabatic wall
-15.7
Outflow
x1/H
x2/H
0
2
0.0
0 14.0
-1.0
Damping region
-32
Outflow
Inflow
Adiabatic wall
-31.4
Outflow
Acoustic
source
2.0
FIG. 4. Computational domains and boundary conditions for cavity flow
topand backward-facing step flow bottom.
105103-4 H. Yokoyama and C. Kato Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
that measured in the past experiment24 with the same degree
of freestream turbulence intensity as that of the present com-
putation. The computed skin-friction coefficients are in good
agreement with those measured in the experiment.24 The
boundary layer becomes a fully developed turbulent state at
Rex=4.3105, where the upstream edge of the cavity is to
be located for the main computations.
For the main computation of the cavity flow, the span-
wise cross-correlation coefficient Rof u2along x2
+=15atthe
upstream position of the cavity Rex=3.4105is shown in
Fig. 6. The spanwise extent of the computational domain Lw
+
corresponds to 720. This coefficient is computed from the
flow fields sampled during T+=300 using Eq. 13. The span-
wise periodicity is used when x3+x3is larger than Lw,
Rx3=
t=0
T+=300x3=0
x3=Lwu2
x3u2
x3+x3dx3dt
t=0
T+=300x3=0
x3=Lwu2
x3u2
x3dx3dt
,13
where u2
is the difference between the instantaneous and
time-averaged velocity components in the x2direction. In
Fig. 6, the local minimum of the coefficient, which appears
at around x3
+=60, shows that the mean spanwise streak
spacing is about 120. This value is in good agreement with
that presented in the past research.25 This indicates that the
longitudinal vortices, which generate the streaks in the tur-
bulent boundary layer, are adequately captured in the present
computation.
C. Effects of computational domain
It is of critical importance that the computational domain
does not affect the results. To confirm that it does not, two
additional computations for the cavity flow with computa-
tional domains different from the original domain are also
performed.
First, the effects of the vertical length Lhand the stream-
wise length Ldfrom the downstream edge of the cavity to the
downstream outflow boundary are investigated. The calcula-
tions with a domain Lh/L=15.0, Ld/L= 9.7 , Lw/L= 0.5
larger than the original domain Lh/L=10.0, Ld/L
=6.5, Lw/L= 0.5are performed. Then, the effects of the
spanwise length Lware investigated by performing the cal-
culation for a cavity flow with a domain with a twice-
extended spanwise length Lh/L=10.0, Ld/L= 6.5 , Lw/L
=1.0. The grid resolutions of the two extended computa-
tional domains are the same as that of the original computa-
tional domain for the corresponding regions.
Figure 7compares for these three domains the frequency
spectra of the velocity fluctuations at x1/L=0.5 and x2=0.
Frequency fis normalized as StfL/u1. The duration of
the Fourier transform is about 45L/u1, and the results are
averaged 20 times temporally as well as spanwisely. This
duration corresponds to about 36 periods of the fundamental
mode St=0.8 for the cavity flow. Figure 7shows that both
the frequency and the amplitude of the fundamental oscilla-
tions with the two extended domains are virtually identical to
those with the original domain. Therefore, it is concluded
that the original computational domain is sufficiently large
and wide and does not affect the present results.
D. Comparison to experiments
As shown in Table I, the flow parameters in the present
computation have the same values as those in Mizushima’s
experiment2except for the cavity depth and the Reynolds
number. In the experiment, D/Lvaried from 0.5 to 2.6. The
fundamental frequency St=0.8 in the present computation
for the cavity flow closely agrees with that measured in the
experiment St=0.78, while oscillations at St= 0.4 also oc-
curred in the experiment. Preliminarily, it was demonstrated
that the oscillations at St=0.4 occur for a cavity flow with
FIG. 5. Variations in time- and spanwise-averaged skin-friction coefficients
computed for flat plate boundary layer without cavity.
FIG. 6. Spanwise cross-correlation coefficient of vertical velocity u2along
x2
+=15atRe
x=3.4105for cavity flow.
FIG. 7. Power spectral densities of vertical velocity u2/u1at x1/L=0.5
and x2=0 for cavity flow with original computational domain Lh/L
=10.0, Ld/L= 6.5 , Lw/L= 0.5, with larger domain Lh/L= 15.0 , Ld/L
=9.7, Lw/L= 0.5, and with wider domain Lh/L=10.0, Ld/L
=6.5, Lw/L= 1.0.
105103-5 Fluid-acoustic interactions in self-sustained oscillations Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
D/L=1.3 and that they correspond to the fluid-resonant os-
cillations. The fluid-acoustic interactions of the fluid-
resonant oscillations will be investigated in a subsequent pa-
per. Meanwhile, the mode at St= 0.8 corresponds to the fluid-
dynamic oscillations, which are the primary focus of this
paper.
As mentioned, the effect of the difference in the
Reynolds number ReLbetween the present computation and
Mizushima’s experiment2on the fluid-acoustic interactions
in the fundamental oscillations is presumably small for the
following reason. Namely, in the present computation for the
cavity flow, the two-dimensional disturbances in the shear
layer grow to large-scale vortices due to the KH instability of
the inflection point of the averaged velocity profile, as clari-
fied in Sec. IV. This indicates that the interactions between
the large-scale vortices and the turbulent fine-scale vortices
are small, and therefore the fluid-acoustic interactions in the
fundamental oscillations do not strongly depend on the Rey-
nolds number. Hence it can be mentioned that the present
computation captures the fluid-acoustic interactions in the
fluid-dynamic oscillations observed in the experiment.
IV. RESULTS AND DISCUSSION
A. Analysis methods
1. Spatial variations in power spectral density
and phase
The spatial variations in the power spectral density
PSDand the phases of the variables are computed from the
results of the Fourier transform of spanwise-averaged data.
The duration of the Fourier transform is 10L/u1, which
corresponds to eight periods of the fundamental frequency
for the cavity flow. The variations are averaged 20 times
temporally.
2. Linear stability analysis
Linear stability analysis LSAis performed in a stan-
dard way10,26,27 and thus is described only briefly here. A
compressible, inviscid, locally parallel formulation is used.
This means that the time-averaged velocity profile slowly
varies in the streamwise direction and that the disturbances
are parallel at each streamwise location. To confirm that the
disturbances for the fundamental frequency are parallel in the
shear layer of the cavity flow, we investigated the spatial
distributions of the phase of velocity u2for the fundamental
frequency along x2/L=0.025, 0, and 0.025, as shown in
Fig. 8, where the reference of the phases is the phase at
x1/L=0.5 and x2/L= 0. As a result, it was clarified that the
three phase distributions approximately agree with each
other and therefore, the disturbances are parallel. The slight
difference near the upstream edge is due to the effects of the
particle velocity of the acoustic waves on the velocity u2.
Also, the normal modes of the form
gx1,x2,t=g
ˆx2ei
x1+2
ft+ c.c. 14
are assumed, where fis a frequency,
is a complex eigen-
value, gis any flow variable, g
ˆis a complex eigenfunction,
and c.c. denotes the complex conjugate. The modes are in-
serted into the compressible inviscid equations linearized
about the parallel flow, and the eigenvalues and eigenfunc-
tions are found using a Runge–Kutta shooting method. The
present LES data are used to determine the mean flow for the
LSA. Boundary conditions are set such that the disturbances
are exponentially dampened for large +x2, and zero vertical
velocity is imposed at the bottom wall of the cavity.
3. Spanwise cross correlation
The spanwise cross-correlation coefficient Rfof u2for
the fundamental frequency is computed using Eq. 15. The
spanwise periodicity is used when x3+x3is larger than Lw,
Rfx3=
t=0
t=20Tx3=0
x3=Lwu2fx3u2fx3+x3dx3dt
t=0
t=20Tx3=0
x3=Lwu2fx3u2fx3dx3dt ,15
where u2f is the velocity fluctuation for the fundamental fre-
quency and computed using the Fourier transform. The du-
ration of the transform is 10L/u1. This coefficient Rfrep-
resents the two dimensionality of the flow for the
fundamental frequency.
4. Phase-averaging process
In a phase-averaging process, the variables at the same
phase of the fundamental mode are averaged 45 times
temporally for the cavity flow. It has been preliminarily clari-
fied by using a wavelet analysis that the period of the oscil-
lations varies by about 5% temporally. Considering this
variation in the period, the pressure fluctuation on the bottom
FIG. 8. Spatial distributions of the phase in velocity u2along x2/L= 0.025,
0, and 0.025.
FIG. 9. PSD of vertical velocity u2/u1at x1/L=0.5 and x2/L=0 for cavity
flow and at x1/H=1.0 and x2/H=0 for backward-facing step flow.
105103-6 H. Yokoyama and C. Kato Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
x1/L=0.5, x2/L= −0.5is used as the reference signal. The
phase at each time is calculated assuming that the period of
the oscillations is the time interval from a peak to a next
peak of this pressure fluctuation. The phase-averaged data
are further averaged in the spanwise direction.
B. Self-sustained oscillations
Figure 9shows the frequency spectrum of the velocity
u2in the shear layer at x1/L=0.5 and x2= 0 for the cavity
flow and that at the corresponding position for the backstep
flow without an artificial acoustic source. Frequency fis nor-
malized as St2fH/u1for the backstep flow, where 2H
corresponds to cavity length L. Self-sustained oscillations
occur at St=0.8 in the cavity flow while oscillations do not
occur in the backstep flow. Note also that the amplitudes of
the fluctuations at other frequencies for the cavity flow are
approximately the same as those for the backstep flow. This
means that the self-sustained oscillations do not strongly af-
fect the amplitude of the turbulent kinetic energy at other
frequencies in the shear layer.
Figure 10 shows the frequency spectrum of the pres-
sure fluctuations on the bottom wall at x1/L=0.5 and
x2/L=−0.5 for the cavity flow and that at x1/H= 1.0 and
x2/H=−1.0 for the backstep flow. The pressure is normalized
as pressure coefficient Cpp/
u1
2/2. Tonal sound is ra-
diated in the self-sustained oscillations at St=0.8 only for the
cavity flow.
C. Vortices
1. Fine-scale vortices
Instantaneous vortices for the cavity flow are illustrated
in Fig. 11. To elucidate the vortices, the second invariant of
the velocity gradient tensor q=2S2is computed,
where and Sare, respectively, the asymmetric and sym-
metric parts of the velocity gradient tensor. Regions with
q0 represent vortex tubes. Before the separation at the
upstream edge, fine-scale longitudinal vortices are active in
the turbulent boundary layer. Soon after the separation, these
vortices become free from the blocking effect of the wall and
become more active than those in the upstream turbulent
boundary layer.
2. Large-scale vortices
In Fig. 12, the contours and vectors represent the phase-
averaged pressure coefficients and velocity vectors, respec-
tively, with the time-averaged components at each position
subtracted. Large-scale vortices are apparent in the shear
layer and form low-pressure regions. These large-scale vor-
tices are composed of the above-mentioned fine-scale
vortices.
Figure 13 shows the spanwise cross-correlation coeffi-
cients Rfof u2for the fundamental frequency along
x1/L=0.25, 0.50, and 0.75 x2=0for the cavity flow. The
coefficients are high at all positions, which means that the
large-scale vortices are coherent in the spanwise direction.
These large-scale vortices form by the growth of the
convective disturbances in the shear layer. In Sec. IV E, it
will be clarified that the acoustic waves only generate the
initial convective disturbances and it does not contribute to
the growth of these disturbances. To clarify the mechanism
of the growth of these convective disturbances, the rate of
the streamwise amplification of the fluctuations in u2for the
fundamental frequency is compared with that predicted by
the LSA in Fig. 14. Note that velocity u2includes a convec-
tive component and a propagative component particle veloc-
ity. As will be clarified in Sec. IV E, the propagative distur-
FIG. 10. PSD of pressure coefficient Cpat x1/L=0.5 and x2/L=−0.5 for
cavity flow and at x1/H=1.0 and x2/H=−1.0 for backward-facing step flow.
FIG. 11. Instantaneous isosurfaces of q/u1/L2= 25 for cavity flow.
FIG. 12. Color onlinePhase-averaged pressure coefficients and velocity
vectors for cavity flow, where Tis the time period of fundamental
oscillations.
105103-7 Fluid-acoustic interactions in self-sustained oscillations Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
bances of the acoustic waves are negligible in comparison to
the convective disturbances in the shear layer of x10.1L.
Therefore, this region is focused on. The amplification
rate computed from the present LES data approximately
agrees with that predicted by the LSA although at around
x1/L=0.2 the rate becomes gradually saturated as the aver-
age velocity gradient becomes more moderate. This means
that the convective disturbances of the fundamental fre-
quency in the shear layer grow into large-scale vortices due
to the KH instability, similar to those in a laminar cavity
flow.9,10 Meanwhile, the mechanism of the generation of the
convective disturbances near the upstream edge x10.1L
is not clear because the propagative disturbances of the
acoustic waves are comparable to the convective distur-
bances near the upstream edge. In Sec. IV E, the mechanism
of the generation of the initial convective disturbances by the
acoustic feedback is discussed by decomposing the flow field
into the convective and propagative components for the
backstep flow with an artificial acoustic source.
D. Acoustic radiation mechanism
The contours in Fig. 15 represent the phase-averaged
pressure coefficients with the time-averaged components
subtracted. Figure 15 shows that an acoustic wave is radiated
from the cavity and propagates to the outside of the cavity.
As already mentioned with relation to Fig. 12, large-
scale vortices are apparent in the shear layer. Low-pressure
regions are formed in these large-scale vortices. When the
large-scale vortices are being convected in the shear layer,
the pressure gradient balances with the centrifugal force act-
ing on the rotating fluid in the large-scale vortices. However,
when a large-scale vortex collides with the downstream wall
of the cavity, it is distorted and the low-pressure fluid in the
vortex spreads along the downstream wall, as shown in Fig.
12 t=0. At this time, the low-pressure fluid is not rigidly
rotating, in contrast to the fluid in the large-scale vortices
being convected in the shear layer. This means that the local
pressure gradient no longer balances with the centrifugal
force. As a result, a local downward velocity is induced due
to the local pressure gradient, as shown in Fig. 16, where the
detailed phase-averaged flow field is presented at t=0 in Fig.
12. As shown in Fig. 16, this local downward velocity is
weak in the far upstream region, while it is intense near the
downstream wall. Therefore, the fluid in the upstream region
of the downstream wall expands, an expansion wave is radi-
ated, and finally the radiated expansion wave is propagated
to the outside of the cavity. To the best of the authors’ knowl-
edge, this is the first time that the mechanism of the acoustic
radiation is clarified for the cavity flow in detail.
To roughly estimate the position of the acoustic source
for the cavity flow, we compute the acoustic propagation
from an artificial source placed on the downstream wall
without a free stream, as shown in Fig. 17. Also, as shown
later in Sec. IV E 3, the effects of the mean flow on the
acoustic propagation are negligibly small at the present
Mach number of 0.3. The acoustic source is assumed to be
two dimensional and coherent in the x3spanwisedirection.
FIG. 13. Spanwise cross-correlation coefficient Rfof vertical velocity u2for
fundamental frequency along x1/L=0.25, 0.50, and 0.75 x2=0for cavity
flow.
FIG. 14. Growth of PSD of vertical velocity u2/u1along x2= 0 for cavity
flow compared with that predicted by LSA.
FIG. 15. Color onlinePhase-averaged pressure coefficients for cavity flow,
where Tis the time period of fundamental oscillations.
FIG. 16. Color onlinePhase-averaged pressure coefficients and velocity
vectors near the bottom wall for cavity flow at t=0 in Fig. 12 the region in
this figure is indicated in Fig. 12.
105103-8 H. Yokoyama and C. Kato Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
The acoustic source is located at two different positions of
x1a /L,x2a /L,1.0, 0.5and 1.0, 0, and both the results
are compared with the result of the cavity flow for M= 0.3.
For doing so, Eq. 1is replaced with Eqs. 1618
in the regions 0.8x1/L1.0,−0.5 x2/L−0.3and
0.8x1/L1.0,−0.2 x2/L0.2,28
Qt+
xk
FkFvk=S,16
S=
S
0
0
0
5
2a2
S
,17
S=Aexp
− ln 2.0
x1x1a2+x2x2a2
0.1L2
sin 2
ft,
18
where A/
is 3.010−5, and the frequency fis the same as
the fundamental frequency for the cavity flow. The amplitude
of the pressure fluctuations on the bottom wall at x1/L=0.5
and x2/L=−0.5 is approximately the same as that for the
cavity flow for M= 0.3. The spatial variations in the phase of
the pressure on the bottom wall x2/L=−0.5and on the
upstream wall x1=0are shown in Figs. 18 and 19, respec-
tively. Both the spatial variations in the phase for the artifi-
cial acoustic source at x1a /L,x2a /L=1.0,−0.5without a
free stream are in good agreement with the present compu-
tational results for M= 0.3. However, Fig. 18 shows that the
slope of the pressure phase for the artificial acoustic source at
x1a /L,x2a /L=1.0,0is different near the downstream wall
x1/L=1.0from that of the cavity flow for M= 0.3. This
means the acoustic waves are radiated from x1/L,x2/L
=1.0,−0.5rather than x1/L,x2/L=1.0 , 0in the cavity
flow. However, it should also be mentioned that the acoustic
source in the cavity flow is more complex than a monopole
source. Therefore, we only roughly estimate the position of
the acoustic source here.
The slope of the phase in Fig. 18 shows that the acoustic
waves radiated around the downstream wall x1/L=1.0
are propagated in the upstream direction x1to about
x1/L=0.4 at the sound speed. However, the slope of the
phase becomes approximately flat near the upstream wall
x1=0. This means that the propagation direction of the
acoustic waves changes from the −x1direction to the +x2
direction near the upstream wall. The slope of the phase in
Fig. 19 shows that the acoustic waves are propagated in the
+x2direction along the upstream wall at the sound speed.
E. Backward-facing step flow with artificial acoustic
source
1. Artificial acoustic source
To clarify the mechanism for the generation of the initial
convective disturbances in the shear layer of the cavity flow
by the acoustic feedback, the effects of artificial acoustic
waves on the shear layer of a backstep flow are also inves-
tigated. As shown in Fig. 4, the artificial acoustic source is
placed at x1a /H,x2a /H=2.0,−1.0, where the backstep
height Hcorresponds to 0.5Lfor the cavity flow. This posi-
tion corresponds to the position of the acoustic source of the
cavity the bottom of the downstream wall, which was
roughly estimated in Sec IV D. The acoustic source is added
No flow
x1
2
Acoustic source
O
L
x1
x2
Acoustic sourc
O
L
No flow
FIG. 17. Flow configurations for acoustic propagation from artificial source
without free stream.
FIG. 18. Phase variations in pressure on bottom x2= −0.5Lfor fundamen-
tal frequency for cavity flow of M= 0.3 and for two acoustic fields from
artificial acoustic source around x1/L=1.0 and x2/L=−0.5 and around
x1/L=1.0 and x2/L=0 without a free stream.
FIG. 19. Phase variations in pressure on upstream wall x1=0for funda-
mental frequency for cavity flow of M= 0.3 and for two acoustic fields from
artificial acoustic source around x1/L=1.0 and x2/L=−0.5 and around
x1/L=1.0 and x2/L=0 without a free stream.
105103-9 Fluid-acoustic interactions in self-sustained oscillations Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
into the backstep flow by replacing Eq. 1in the regions
1.6x1/H2.4 and −1.0x2/H−0.6 with Eqs. 16,
17, and 19,28
S=Aexp
− ln 2.0
x1−2H2+x2+H2
0.2H2
sin 2
ft,
19
where Ais the amplitude of the density fluctuations at the
streamwise center of the acoustic source. In order to clarify
the effects of the intensity of the acoustic waves on the
acoustic feedback, three computations with A/
=0,
1.010−5, and 3.010−5 are performed, among which the
amplitude of the pressure fluctuations with A/
=3.0e−5 on
the bottom wall at x1/H=1.0 and x2/H= −1.0 is approxi-
mately the same as that at the corresponding position for the
cavity flow. The excitation frequency fis the same as the
fundamental frequency for the cavity flow.
2. Comparison to cavity flow
Figure 20 shows the frequency spectra of the pressure
fluctuations at x1/H=1.0 and x2/H= −1.0 for the backstep
flows with and without the artificial sound of A/
=3.0
10−5 and the pressure fluctuation at the corresponding po-
sition for the cavity flow. It is confirmed that the amplitude
of the artificial sound of A/
=3.010−5 is approximately
the same as that of the tonal sound in the cavity flow. Also,
the spatial variation in the phase of pressure on the vertical
wall x1=0for the excitation frequency for the backstep
flow with the acoustic source of A/
=3.010−5 is com-
pared with the corresponding variation for the cavity flow in
Fig. 21. Figure 21 shows that the spatial variation in pressure
phase is also in good agreement with that for the cavity flow.
Therefore, it is concluded that the acoustic field in the cavity
flow is reproduced in the backstep flow with the artificial
acoustic source of A/
=3.010−5.
In Fig. 22, the contours and vectors represent the phase-
averaged pressure coefficients and velocities, respectively,
for the backstep flow with the artificial acoustic source of
A/
=3.010−5. By adding the artificial acoustic source,
large-scale vortices become apparent in the shear layer and
form low-pressure regions like in the cavity flow Fig. 12.
In Fig. 23, the spanwise cross-correlation coefficients
Rfof u2for the excitation frequency along x1/H=0.5 and
x2=0 for the backstep flow with and without the artificial
acoustic source of A/
=3.010−5 are compared with the
corresponding coefficient for the cavity flow. The cross-
correlation coefficient becomes as high as 0.9 for the back-
step flow with the artificial acoustic source, although it is
almost zero if the artificial acoustic source is not added. This
is because the above-mentioned large-scale vortices are co-
herent in the spanwise direction like in the cavity flow. Note
also that the cross-correlation coefficient for the backstep
flow with the artificial acoustic source of A/
=3.010−5 is
higher than that for the cavity flow. This is because the arti-
ficial acoustic source is perfectly coherent in the spanwise
FIG. 20. PSD of pressure coefficient Cpat x1/H= 1.0 and x2/H= 1.0 for
backward-facing step flows with and without artificial acoustic source and at
x1/L=0.5 and x2/L=−0.5 for cavity flow.
FIG. 21. Phase variations in pressure along x1= 0 for backward-facing step
flow with artificial acoustic source and cavity flow.
FIG. 22. Color onlinePhase-averaged pressure coefficients and velocity
vectors for backward-facing step flow with artificial acoustic source.
FIG. 23. Spanwise cross-correlation coefficients Rfof vertical velocity u2
along x1/H=0.5 and x2=0 for backward-facing step flows with and without
acoustic source and that along the corresponding line for cavity flow.
105103-10 H. Yokoyama and C. Kato Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
direction while the large-scale vortices, which become the
acoustic sources for the cavity flow, are not perfectly coher-
ent, as shown in Fig. 13.
In Fig. 24, the spatial variations in the amplitude of the
fluctuations in u2for the excitation frequency for the back-
step flows with and without the artificial acoustic source of
A/
=3.010−5 are compared with the corresponding
variation for the cavity flow. It is clarified that the spatial
variation in the region of 0x1/2H0.3 is approximately
the same as that in the corresponding region for the cavity
flow. Therefore, it is concluded that the acoustic feedback in
the cavity flow is also reproduced in the backstep flow with
the artificial acoustic source of A/
=3.010−5 in addition
to the acoustic field, for which a detailed discussion is given
in Sec. IV E 3.
3. Acoustic feedback
Figure 25 shows the variation in the PSD of u2c and that
of u2p for the excitation frequency along x2=0 for the
backstep flow with A/
=3.010−5, where u2c and u2p
represent the convective and propagative components of u2,
respectively,
u2=u2c +u2p.20
The propagative component u2p is assumed to be the same as
particle velocity, which was calculated for the acoustic
propagation around the backstep from the same artificial
acoustic source without flow. The convective component u2c
is obtained by subtracting the propagative component from
the computed velocity for the backstep flow of M= 0.3. In
Fig. 26, the spatial variation in the pressure phase on the
vertical wall x1=0for this backstep flow is compared with
that of the acoustic field around the backstep from the same
acoustic source without flow. The acoustic field of the back-
step flow with M= 0.3 is in good agreement with that without
flow. This confirms that the effects of the mean flow on the
acoustic propagation are negligibly small. Figure 25 shows
that the propagative disturbances caused by the acoustic
waves are negligibly small in comparison to the convective
disturbances caused by the large-scale vortices in the region
x1/H0.2.
Figure 27 shows the variations in the PSD of u2c for the
excitation frequency along x2=0 for backstep flows with and
without an artificial acoustic source compared with that pre-
dicted by the LSA. For the backstep flow without an acoustic
source, u2c is equal to u2. Only the curve predicted by the
LSA for the backstep flow without an acoustic source is
shown because the result of the LSA is virtually independent
of the strength of the acoustic source. As shown in Fig. 27,
the convective disturbances become more intense as the ar-
tificial sound becomes more intense. However, all the three
amplification rates computed from the present LES data ap-
proximately agree with that predicted by the LSA in the re-
gion 0.2x1/H0.4. This means that the growth of the con-
FIG. 24. Variations in PSD of vertical velocity u2/u1along x2= 0 for
backward-facing step flows with and without artificial acoustic source com-
pared with that for cavity flow.
FIG. 25. Variations in PSD of convective component u2c/u1and propaga-
tive component u2p/u1of vertical velocity along x2= 0 for backward-facing
step flows with acoustic source.
FIG. 26. Phase variations in pressure along x1= 0 for backward-facing flows
with artificial acoustic source for M= 0.3 and no flow.
FIG. 27. Variations in PSD of convective component u2c/u1of vertical
velocity along x2=0 for backward-facing step flows with and without acous-
tic source.
105103-11 Fluid-acoustic interactions in self-sustained oscillations Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
vective disturbances is due to the KH instability and not
affected by the acoustic waves. Meanwhile, the acoustic
waves induce the initial convective disturbances as men-
tioned below. As already shown in Fig. 21, the acoustic
waves are propagated around the upstream edge in the +x2
direction. As a result, the shear layer is moved up and down
due to the particle velocity of the acoustic waves and the
initial convective disturbances are generated in the shear
layer. Therefore, the initial convective disturbances become
more intense as the acoustic waves become more intense, as
shown in Fig. 27.
Since the acoustic field and the acoustic feedback are
reproduced in the backstep flow with the artificial acoustic
source, the above discussion also holds for the acoustic feed-
back in the cavity flow. In the cavity, the initial convective
disturbances are generated by the acoustic waves radiated
due to the collision of the vortices on the downstream wall.
These initial convective disturbances are coherent in the
spanwise direction since the acoustic waves are coherent.
These convective disturbances grow into the coherent large-
scale vortices due to the KH instability independent of the
acoustic waves and the fine-scale vortices, similar to those in
a laminar cavity flow.9,10
V. CONCLUSION
The fluid-acoustic interactions in a flow over a two-
dimensional rectangular cavity were investigated by directly
solving the compressible Navier–Stokes equations. The up-
stream boundary layer was turbulent. The depth-to-length ra-
tio of the cavity was 0.5. Moreover, to clarify the acoustic
feedback for the cavity flow in detail, the effects of artificial
acoustic waves on the shear layer of a backstep flow with the
same Mach number were also investigated.
Phase-averaged flow fields revealed the mechanism of
the acoustic radiation in detail. Two-dimensional large-scale
vortices composed of turbulent fine-scale vortices form in
the shear layer that separates from the upstream edge of the
cavity. When a large-scale vortex collides with the down-
stream wall, the low-pressure fluid in the vortex spreads
along the downstream wall. As a result, a local downward
velocity is induced by the local pressure gradient, and the
upstream fluid expands. In this way, an expansion wave is
radiated.
The large-scale vortices originate from the convective
disturbances that are developed in the shear layer. The am-
plification rate of the convective disturbances agrees with
that predicted by LSA. This means that the disturbances
grow due to KH instability, similar to the growth of those in
a laminar cavity flow. To clarify the mechanism for the gen-
eration of the initial convective disturbances, computations
for backward-facing step flows with an artificial acoustic
source were also performed. As the artificial acoustic waves
became more intense, the initial convective disturbances in
the shear layer became more intense while the amplification
rate was not changed. This means that the initial convective
disturbances in the shear layer are induced by the acoustic
waves and developed into the large-scale vortices due to the
KH instability independent of the acoustic waves and the
fine-scale vortices. This mechanism of the formation of the
large-scale vortices also holds for the cavity flow.
The present study has provided much deeper understand-
ing of the fluid-acoustic interactions in turbulent cavity
flows, which should lead to improved design of various in-
dustrial products and the suppression of the effects of such
interactions.
ACKNOWLEDGMENTS
This research has been supported by a global COE pro-
gram “Global Center of Excellence for Mechanical Systems
Innovation” of the University of Tokyo and by a research
program “Revolutionary Simulation Software” supported by
the Ministry of Education, Culture, Sports, Science and
Technology of Japan MEXT.
1K. Karamcheti, “Acoustic radiation from two-dimensional rectangular cut-
outs in aerodynamic surfaces,” NACA Report No. TN 3487, 1955.
2F. Mizushima, H. Takakura, T. Kurita, C. Kato, and A. Iida, “Experimental
investigation of aerodynamic noise generated by a train-car gap,” J. Fluid
Sci. Technol. 2, 464 2007.
3P. Oshkai and T. Yan, “Experimental investigation of coaxial side branch
resonators,” J. Fluids Struct. 24,5892008.
4J. C. Bruggeman, A. Hirschberg, M. E. H. van Dongen, and A. P. J.
Wijnands, “Self-sustained aeroacoustic pulsations in gas transport sys-
tems: Experimental study of the influence of closed side branches,” J.
Sound Vib. 150, 371 1991.
5J. E. Rossiter, “Wind-tunnel experiments on the flow over rectangular
cavities at subsonic and transonic speeds,” Aeronautical Research Council
Report No. 3438, 1964.
6A. Powell, “On the edge tone,” J. Acoust. Soc. Am. 33,3951961.
7L. F. East, “Aerodynamically induced resonance in rectangular cavities,”
J. Sound Vib. 3, 277 1966.
8D. Rockwell and E. Naudascher, “Review—self-sustaining oscillations of
flow past cavities,” ASME Trans. J. Fluids Eng. 100,1521978.
9C. Knisely and D. Rockwell, “Self-sustained low-frequency components
in an impinging shear layer,” J. Fluid Mech. 116 , 157 1982.
10C. 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.
11S. K. Lele, “Compact finite difference schemes with spectral-like reso-
lution,” J. Comput. Phys. 103,161992.
12D. P. Rizzetta and M. R. Visbal, “Large-eddy simulation of supersonic
cavity flow fields including flow control,” AIAA J. 41, 1452 2003.
13C. Bogey and C. Bailly, “Large eddy simulations of round free jets using
explicit filtering with/without dynamic Smagorinsky model,” Int. J. Heat
Fluid Flow 27, 603 2006.
14C. Bogey and C. Bailly, “Turbulence and energy budget in a self-
preserving round jet: Direct evaluation using large eddy simulation,” J.
Fluid Mech. 627, 129 2009.
15K. Matsuura and C. Kato, “Large-eddy simulation of compressible transi-
tional flows in a low-pressure turbine cascade,” AIAA J. 45,4422007.
16D. V. Gaitonde and M. R. Visbal, “Padé-type higher-order boundary filters
for the Navier–Stokes equations,” AIAA J. 38, 2103 2000.
17K. W. Thompson, “Time dependent boundary conditions for hyperbolic
systems,” J. Comput. Phys. 68,11987.
18T. J. Poinsot and S. K. Lele, “Boundary conditions for direct simulations
of compressible viscous flows,” J. Comput. Phys. 101,1041992.
19J. W. Kim and D. J. Lee, “Generalized characteristic boundary conditions
for computational aeroacoustics,” AIAA J. 38, 2040 2000.
20J. B. Freund, “Proposed inflow/outflow boundary condition for direct com-
putation of aerodynamic sound,” AIAA J. 35,7401997.
21J. Larsson, L. Davidson, M. Olsson, and L. Eriksson, “Aeroacoustic in-
vestigation of an open cavity at low Mach number,” AIAA J. 42,2462
2004.
22H. Yokoyama, Y. Tsukamoto, C. Kato, and A. Iida, “Self-sustained oscil-
lations with acoustic feedback in flows over a backward-facing step with a
small upstream step,” Phys. Fluids 19, 106104 2007.
105103-12 H. Yokoyama and C. Kato Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
23J. O. Hinze, Turbulence, 2nd ed. McGraw-Hill, New York, 1975,
pp. 175–320.
24J. Coupland, “Transition modeling for turbomachinery flows, T3 test
cases,” ERCOFTAC Bulletin, No. 5, 1990.
25S. K. Robinson, “Coherent motions in the turbulent boundary layer,”
Annu. Rev. Fluid Mech. 23, 601 1991.
26T. L. Jackson and C. E. Grosch, “Inviscid spatial stability of a compress-
ible mixing layer,” J. Fluid Mech. 208, 609 1989.
27M. Zhuang, T. Kubota, and P. E. Dimotakis, “Instability of inviscid, com-
pressible free shear layers,” AIAA J. 28, 1728 1990.
28C. K. W. Tam and J. C. Hardin, Second Computational Aeroacoustics
CAAWorkshop on Benchmark problems, NASA, CP 3352 1997.
105103-13 Fluid-acoustic interactions in self-sustained oscillations Phys. Fluids 21, 105103 2009
Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp
... Many researchers [3][4][5][6][7][8][9] have investigated the acoustic radiation from a flow over a two-dimensional rectangular cavity. Rossiter proposed the mechanism of self-sustained oscillations in such a cavity flow. ...
... This mechanism was confirmed by direct aeroacoustic simulations based on the compressible Navier-Stokes equations. 4,5 Moreover, the self-sustained oscillations can be intensified by the coupling of the above-mentioned fluid-acoustic interactions in the shear layer with acoustic resonance in the cavity. [6][7][8] Rockwell and Naudascher 9 classified these self-sustained oscillations as fluid-resonant oscillations. ...
... As shown in Fig. 9 (t/T ¼ 0.25), the vortex with the low-pressure region (Vortex 3) starts to be collided with the downstream wall of the expanding pipe and deformed with the spread of the low-pressure regions, which radiates an expansion wave. 5 The expansion wave is propagated inside the cavity, leading to the formation of standing waves and the occurrence of the expansion in the entire region of the third cavity. After the collision of the vortex, the pressure becomes lowest at the bottom of the third cavity at t/T ¼ 0.38. ...
Article
Full-text available
Expanding pipes with orifice plates are often utilized as silencers for fluid machinery. However, intense tonal sounds can be generated from a flow through such expanding pipes. To clarify the mechanism of tonal sound from a flow through a circular expanding pipe with two orifice plates and the conditions for intense acoustic radiation, the flow and acoustic fields are directly solved based on the compressible Navier–Stokes equations. Phase-averaged flow fields indicate the occurrence of periodic vortex shedding in the free shear layers of the expanding pipe, resulting in acoustic radiation. The effects of the orifice radius and freestream Mach number on the acoustic radiation are focused on. The computational results demonstrate that vortex rings or spiral vortices are generated in the cavities formed by the orifice plates, where the primary vortical shape changes, depending on the freestream Mach number and orifice radius. The collision of the vortex ring and spiral vortex with the orifice plate or downstream edge of the expanding pipe leads to the occurrence of circumferentially in-phase and one-wavelength-mode pressure fluctuations, respectively. The orifice radius also affects the convective velocity of vortices and the position of the acoustic source, varying the frequency of the acoustic radiation. The findings of this research provide the first clarifications of fluid–acoustic interactions in an expanding pipe with orifice plates.
... Two actions may occur near the impingement point, that is, some large-scale vortices eject through the slat gap, whereas several vortices are entrapped into the recirculation zone. For the Low-Re the large-scale vortices in the shear layer tend to be more two-dimensional because higherfrequency shear layer instabilities that breakdown the larger-scale vortices are suppressed [39]. ...
... Low pressure values are present to be surrounding the core of the vortex S and S , and high pressure values appear in the connection regions of two neighboring vortices. The pressure gradients in the shear layer are balanced with the centrifugal force of large-scale vortices[38,39]. Figs. ...
Article
Implicit large-eddy simulations are performed to investigate the acoustic resonances from a leading-edge slat at two different Reynolds numbers. Three types of acoustic waves are identified in the vicinity of the slat, but only one is found to be amplified by fluid-acoustic feedback mechanism. With phase-averaged flow-fields, we clearly identify that the vortical structures in the slat cove are formed at the same frequencies of the observed acoustic tones. A fluid-acoustic feedback loop is confirmed between the formation of the discrete vortices and the radiation of low-frequency acoustic waves. A theoretical model is proposed and validated to predict the frequencies of low-frequency resonances at three angles of attack. High-frequency acoustic waves are also found to emanate from the slat-wake region and the suction side of the main wing.
... Rockwell and Naudascher (1978) defined the self-sustained oscillations caused by this coupling of the acoustic and shear layer modes as fluid-resonant oscillations. Recently, the aeroacoustic feedback mechanism of the shear layer mode proposed by Rossiter (1964) has been confirmed using direct aeroacoustic simulations based on the compressible Navier-Stokes equations (Rowley et al., 2002;Yokoyama and Kato, 2009). Terao et al. (2011) experimentally evaluated the times for vortex convection and acoustic propagation based on flow visualizations using acoustic measurements, and the constant of γp in Rossiter's equation (Eq. ...
Article
Full-text available
Expanding pipes with orifice plates are often utilized as silencers to reduce the noise in fluid machineries. However, intense aerodynamic tonal sound can be generated from flows through such expanding pipes. To clarify the generation mechanism and conditions for intense tonal sound, sound measurements and flow visualization were performed for a flow through an expanding pipe with two orifice plates, where three circular cavities were formed in the expanding pipe. The effects of the freestream Mach number and orifice radius on the flow and sound were analyzed. The variation in acoustic radiation with freestream Mach number demonstrated that the Strouhal number of the most intense tonal sound based on the cavity length changes discretely at particular freestream Mach numbers, where a shear layer mode number corresponding to the number of vortices in the cavity is varied. The tonal sound becomes intense owing to the coupling of the shear layer and circumferential acoustic modes. The rotational and stational acoustic modes were identified based on the circumferential phase distributions of pressure fluctuations in the expanding pipe, whereas in-phase pressure fluctuations occurred at a relative higher Mach number. The visualized vortical structures corresponded to these phase distributions of the pressure fluctuations. The sound measurements with different orifice radii showed that the tonal sound occurred at a lower Strouhal number with a larger orifice radius, which indicates weaker effects of the orifice plate on the vortex shedding in the expanding pipe. At a smaller orifice radius, acoustic radiation occurred at a higher Strouhal number owing to the flow acceleration through the orifices, which resulted in intensified sound propagation into the outside of the expanding pipe along the effects of the acoustic mode particularly at a high Mach number.
... Pressure fluctuations within cavities are produced by several mechanisms, which include the Rossiter feedback loop, turbulence within the shear layer, and the transport and production of turbulence due to recirculation within the cavity. The Rossiter feedback loop 1-3 is a selfsustaining noise generation mechanism produced by Kelvin-Helmholtz type vortices 4 shed from the cavity upstream edge. When the vortices impinge on the downstream wall, pressure waves are produced 5,6 that perturb the shear layer, which further produces vortex shedding. ...
Article
Microphone measurements in a closed test section wind tunnel are affected by turbulent boundary layer (TBL) pressure fluctuations. These fluctuations are mitigated by placing the microphones at the bottom of cavities, usually covered with a thin, acoustically transparent material. Prior experiments showed that the cavity geometry affects the propagation of TBL pressure fluctuations toward the bottom. However, the relationship between the cavity geometry and the flowfield within the cavity is not well understood. Therefore, a very large-eddy simulation was performed using the lattice Boltzmann method. A cylindrical, a countersunk and a conical cavity are simulated with and without a fine wire-cloth cover, which is modeled as a porous medium governed by Darcy's law. Adding a countersink to an uncovered cylindrical cavity is found to mitigate the transport of turbulent structures across the bottom by shifting the recirculation pattern away from the cavity bottom. Covering the cavities nearly eliminates this source of hydrodynamic pressure fluctuations. The eddies within the boundary layer, which convect over the cover, generate a primarily acoustic pressure field inside the cavities and thus suggesting that the pressure fluctuations within covered cavities can be modeled acoustically. As the cavity diameter increases compared to the eddies' integral length scale, the amount of energy in the cut-off modes increases with respect to the cut-on modes. Since cut-off modes decay as they propagate into the cavity, more attenuation is seen. The results are in agreement with experimental evidence.
... This finding indicates that the turbulent intensity decreased due to the decrease in the flow rate, and periodic vortex structures were strongly observed at the constriction (Fig. 11). Yokoyama and Kato (2009) reported that a tonal noise was produced by fluid-acoustic interactions in a rectangular cavity along a uniform flow. The periodic sound of the simplified model was also produced by the interaction between the acoustic resonance of the model's geometry (4 kHz) and periodic vortex generation at the constriction inlet (Figs. ...
Article
Full-text available
To elucidate the linguistic similarity between the alveolo-palatal sibilant [ɕ] and palatal non-sibilant [ç] in Japanese, the aeroacoustic differences between the two consonants were explored via experimentation with participants and analysis using simplified vocal tract models. The real-time magnetic resonance imaging (rtMRI) observations of articulatory movements demonstrated that some speakers use a nearly identical place of articulation for /si/ [ɕi] and /hi/ [çi]. Simplified vocal tract models were then constructed based on the data captured by static MRI, and the model-generated synthetic sounds were compared with speaker data producing [ɕ] and [ç]. Speaker data demonstrated that the amplitude of the broadband noise of [ç] was weaker than that of [ɕ]; the characteristic peak amplitude at approximately 4 kHz was greater in [ç] than in [ɕ], although the mid-sagittal vocal tract profiles were nearly identical for three of ten subjects in the rtMRI observation. These acoustic differences were reproduced by the proposed models, with differences in the width of the coronal plane constriction and the flow rate. The results suggest the need to include constriction width and flow rate as parameters for articulatory phonetic descriptions of speech sounds.
... Illingworth, Morgans & Rowley 2012), and with computational aeroacoustics methods, which are based on direct numerical simulations and large eddy simulations (LES) of the compressible Navier-Stokes equations (e.g. Rowley, Colonius & Basu 2002;Gloerfelt, Bailly & Juvé 2003;Yokoyama & Kato 2009) or on the linearization of these equations around a given base flow (e.g. Yamouni, Sipp & Jacquin 2013;Sun et al. 2017). ...
Article
In this work, Mach 1.8 flow over a shallow open cavity is studied through time-resolved schlieren images as well as unsteady pressure measurements respectively. Additionally, the cavity shear layer dynamics are explored using modal decomposition techniques namely; proper orthogonal decomposition (POD) and spectral proper orthogonal decomposition (SPOD). The results display that the cavity undergoes a self-sustaining oscillation with a number of modes/tones that are closely related to the modified Rossiter relation. Depending on the cavity oscillation, the time-resolved images reveal various complex flow features. Through modal analysis, the shear layer structures which are responsible for the cavity oscillation are identified. It is understood that POD requires several hundreds of modes to capture 67.3% of the total intensity for resembling the cavity flow field, while SPOD shows 65.7% of the total intensity is available in the 1st mode to reconstruct the cavity flow field. Further, modal analyses have demonstrated that cavity oscillatory frequencies are very similar to unsteady pressure measurement and the modified Rossiter relation. We confirmed that, both these techniques can be successfully applied to cavity flows and are complementary.
Article
Computations of coupled phenomena between the fluid and acoustic interactions in the cavity flow and heat conduction in the flat plates (stack) installed in the cavity were performed to investigate the flow conditions for an effective thermoacoustic heat pump driven by acoustic radiation in cavity flows. This research presents the first computational approach for an aeroacoustically excited thermoacoustic heat pump. The effects of the freestream Mach number on the thermoacoustic heat pump were investigated by conducting computations with Mach numbers 0.087 and 0.26, where intense self-sustained oscillations occurred. Sound pressure levels of the cavity tone and power of the related velocity fluctuations in the shear layer were reduced by the installation of the stack compared with the cavity flow without a stack, particularly at a lower Mach number. A steeper temperature gradient along the stack was acquired at a higher Mach number, and thermoacoustic heat pump effects were confirmed in the predicted acoustic oscillatory flow between the stack plates. The computational results present that the interaction between the recirculation flows in the cavity and the stack cause the heating of the flow around the cold top end of the stack. The thermoacoustic heat pump effects are reduced due to the interaction. The present computational method is expected to be useful for investigating the utilization of aerodynamical sound energy through conversion to heat energy.
Article
Strong wall pressure fluctuations are generated mostly by a high-speed flow passing over a cavity and should be accurately predicted and controlled for the structural design of aircraft. In this work, a spatiotemporal analysis was performed to estimate wall pressure fluctuations and reveal the mechanism of flow control by a sawtooth-like vortex generator installed on the leading edge or by the use of a rounded trailing edge in suppressing wall pressure fluctuations. Both the cavity tones and overall sound pressure level were analyzed under different flow speeds and flow control approaches. To investigate the power transfer in the temporal and spatial domains, power spectral density contours were drawn and discussed. For the mode separation, the phased array technique with an improved beamforming algorithm was adopted to calculate the wavenumber map and then to separate the acoustic modes from the convective modes by utilizing the spatial correlation of the wall pressure fluctuations. The mode separation results show that the convective modes contribute mainly to the low-frequency pressure fluctuations, while the acoustic modes contribute to both the low-frequency and high-frequency pressure fluctuations. Flow control by the rounded trailing edge approach is more effective than that by the serrated structure approach in reducing the cavity acoustic environment, which enables wall pressure fluctuations to transfer from the high-frequency range to the low-frequency range compared with the “no control” case.
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
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.
Article
An extended conservative formalism of the characteristic boundary conditions is presented on the basis of the generalized coordinates for practical computational aeroacoustics. The formalism is derived for solving the entire conservative form of the compressible Euler or Navier-Stokes equations on the body-fitted grid mesh system by using the high-order and high-resolution numerical schemes. It includes the matrices of transformation between the conservative and the characteristic variables, which were already derived in the literature to analyze the eigenvalue-eigenvector modes in an arbitrary direction. The conservation-form governing equations with their full terms are solved at the boundaries, and no kind of extrapolation or simplification of the equations is included in this formalism. Additional correction terms are devised to preserve the conservative form of flux derivative terms in the generalized coordinates. Especially, the soft inflow conditions are presented to keep the nonreflecting features, as well as to maintain the mean value of inflow velocity at the inlet boundary. These boundary conditions are applied to the actual computation of two-dimensional viscous cylinder flows with Reynolds number of 400 on the grid meshes clustered on the cylinder surface and the downstream region. The Strouhal number due to von Karman vortex streets, root-mean-square lift coefficient, and mean drag coefficient are evaluated correctly in comparison with experimental data. The far-field sound pressure levels are measured directly in this computation, and the accuracy is validated by an analytic formula derived in the literature.
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
Large-eddy simulations of supersonic cavity flowfields are performed using a high-order numerical method. Spatial derivatives are represented by a fourth-order compact approximation that is used in conjunction with a sixth-order nondispersive filter. The scheme employs a time-implicit approximately factored finite difference algorithm, and applies Newton-like subiterations to achieve second-order temporal and fourth-order spatial accuracy. The Smagorinsky dynamic subgrid-scale model is incorporated in the simulations to account for the spatially underresolved stresses. Computations at a freestream Mach number of 1.19 are carried out for a rectangular cavity having a length-to-depth ratio of 5:1. The computational domain is described by 2.06 x 16(7) grid points and has been partitioned into 254 zones, which were distributed on individual processors of a massively parallel computing platform. Active flow control is applied through pulsed mass injection at a very high frequency, thereby suppressing resonant acoustic oscillatory modes. Features of the flowfields are elucidated, and comparisons are made between the unsuppressed and suppressed cases and with available experimental data that were collected at a higher Reynolds number.
Article
To investigate the mechanism of noise generation by a train-car gap, which is one of a major source of noise in Shinkansen trains, experiments were carried out in a wind tunnel using a 1/5-scale model train. We measured velocity profiles of the boundary layer that approaches the gap and confirmed that the boundary layer is turbulent. We also measured the power spectrum of noise and surface pressure fluctuations around the train-car gap. Peak noise and broadband noise were observed. It is found that strong peak noise is generated when the vortex shedding frequency corresponds to the acoustic resonance frequency determined by the geometrical shape of the gap, and that broadband noise is generated at the downstream edge of the gap where vortexes collide. It is estimated that the convection velocity of the vortices in the gap is approximately 45% of the uniform flow velocity.