Sei sulla pagina 1di 10

Aerospace Science and Technology 53 (2016) 85–94

Contents lists available at ScienceDirect

Aerospace Science and Technology


www.elsevier.com/locate/aescte

Aerodynamic sound from a square cylinder with a downstream wedge


Siti Ruhliah Lizarose Samion a,∗ , Mohamed Sukri Mat Ali a , Aminudin Abu a , Con
J. Doolan b , Ric Zong-Yang Porteous c
a
Malaysia–Japan International Institute of Technology, Universiti Teknologi Malaysia, 54100 Kuala Lumpur, Malaysia
b
School of Mechanical and Manufacturing Engineering, The University of New South Wales, Sydney, 2052, Australia
c
School of Mechanical Engineering, The University of Adelaide, South Australia, 5005, Australia

a r t i c l e i n f o a b s t r a c t

Article history: The effect of placing a wedge in the wake of a square cylinder (side length D) at a Reynolds number of
Received 21 September 2015 22,000 is numerically investigated. In particular, the effect of the wedge on aerodynamic noise is observed
Received in revised form 8 March 2016 along with its effect on the flow field. Wedge base height (h) and its gap distance (G) downstream of
Accepted 10 March 2016
the cylinder are systematically varied. Flow simulations are carried out using an unsteady RANS model
Available online 21 March 2016
employing the k–ω SST turbulence model, whereas the calculation of aerodynamic noise radiated from
Keywords: the flow is solved using Curle’s equation. A special correction technique is applied to consider spanwise
Bluff body effects on noise production and validation is provided using new aeroacoustic data for a square cylinder
Passive flow control in cross-flow. It is found that the flow behavior can be divided into two main regimes (regime I and
Passive sound control regime II), with a linking transition regime. For regime I, the generated sound is lower than that of the
High Reynolds number isolated square cylinder case. The thinnest wedge produced the best sound reduction (11.79 dB) when
the wedge is placed at G = 2D. For regime II, the calculated sound level is higher than the case of
an isolated square cylinder. This is because the sound emitted from both bodies have about the same
magnitude and are in phase. For this case, the maximum increase of sound pressure is 6.24 dB, when
the medium wedge is at G = 2.5D.
© 2016 Elsevier Masson SAS. All rights reserved.

1. Introduction [14] are among the experimental studies, while Okajima [15], Shi-
mada and Ishihara [16] are numerical studies; each investigated
Bluff bodies create flow separation over a significant part of the flow over a rectangular cylinder with different aspect ratios.
their surfaces in their wakes. For a fixed flow condition, the aero- However, these investigations are limited to the flow field and did
dynamic forces are mainly influenced by the geometry of the not examine aerodynamic sound.
bluff bodies [3–5]. This is due to the different flow structures, Furthermore, almost all practical applications occur at high
even though the vortex generation process is topologically similar. Reynolds number, but very few investigations of aerodynamic
Therefore, different types of geometry of bluff bodies can create sound are performed at high Reynolds number. In bluff body flow
different kinds of sound [6]. However, published flow noise data problems, a Reynolds number of 22,000 is commonly taken as rep-
are mostly focused on circular cylinders [7–9] and there are very resentative of a high Reynolds number. For example, Sohankar [17]
few reported investigations of aerodynamic sound from square studies the flow over a square cylinder with moderate to high
cylinders [17]. The study of aerodynamic sound from a square Reynolds numbers ranging from 1 × 103 to 5 × 106 and classi-
cylinder is of interest as the flow separation is fixed and gives dif-
fies Re > 20,000 as a high Reynolds number. For a fixed geometry,
ferent flow behavior than that of the circular cylinder.
a change in Reynolds number can affect the vortex shedding fre-
Bearman [10] and Lyn et al. [11] are among the few experi-
quency and wake.
mental studies concerning the wake and surface pressures on a
Concerning aerodynamic sound for a square cylinder in cross
rectangular cylinder. Other investigations of flow over rectangular
flow, there are very limited validation data for the case of high
cylinder have focused on the effect of the aspect ratio on the vor-
Reynolds numbers. Fujita et al. [18] have investigated the sound
tex shedding pattern. Nakaguchi et al. [12], Norberg [13] and Ohya
generated by flow over a square cylinder when the angle of at-
tack is altered at a fixed Re = 1.3 × 104 . Nakato et al. [19] is the
* Corresponding author. sole previous investigation of the aerodynamic sound generated by
E-mail address: lializarose@gmail.com (S.R.L. Samion). flow around a rectangular cylinder at a range of high Reynolds

http://dx.doi.org/10.1016/j.ast.2016.03.007
1270-9638/© 2016 Elsevier Masson SAS. All rights reserved.
86 S.R.L. Samion et al. / Aerospace Science and Technology 53 (2016) 85–94

Nomenclature

D diameter of square cylinder [m] y vertical coordinate


f frequency [Hz] λ sound pressure wavelength [m]
h height of wedge [m] streamwise direction index
i
Lr length of reattachment [m]
j vertical direction index
Re Reynolds number
St Strouhal number 0 initial value according to ambient pressure
u velocity [m/s] rms root-mean-squared value
x streamwise coordinate mean time-averaged value

number, i.e. Re = 6.4 × 103 to 2.3 × 104 . Margnat [20] numerically stream body shape on the wake of a square cylinder. For the case
investigated the flow and noise created by rectangular cylinders of a wedge, the separated shear layers were able to approach each
at incidence but at low Reynolds number. Rokugou et al. [21] in- other and interact more rapidly than without the wedge. However,
vestigated the relationship between force fluctuations of a square no investigation on the sound generation was presented.
cylinder at high Reynolds number with the aerodynamic sound Prasath et al. [33] presented a complete explanation of the
generation. flow physics of a single wedge with different aspect ratios (AR =
There are also previous studies that consider the spanwise Length/Height) and orientations with respect to the flow. Due to
length effect on the production of aerodynamic noise. Seo and the direct impact of eddies on the surface of base-facing wedge,
Moon [22] develop a correction that allows the prediction of noise the lift coefficients of the wedge with its base facing the flow was
from long-span bodies using the numerical simulation of only a much higher compared with a wedge with its apex facing the flow.
small span-wise portions of the body. Casalino and Jacob [23] Also, as AR increases, a delay in the onset of vortex shedding was
use an ad-hoc statistical model to account for random phase dif- observed, especially for the case when the base faced the flow. If
ferences along the span of cylinders with turbulent wakes. Also, seen from the perspective of vortex shedding frequency, Iungo et
Doolan [24] predicts spanwise-corrected noise using a statistical al. [34] found that the vortex shedding frequency increases upon
treatment on the sound source itself. decreasing AR of the wedge.
Passive flow control is an attractive method to control aero- From the review, AR and gap distance may alter the flow struc-
dynamic sound [25]. The use of a downstream object, placed in tures in the wake, which may also affect noise generation. There-
the wake of the upstream body, is one technique that can af- fore, it is desirable to extend previous work on passive sound
fect the aerodynamic sound level. A critical study by Leclercq and control using a downstream wedge as a passive control device, so
Doolan [26] on the vortex–wake interaction of two-tandem rect- that it can contribute to more complete understanding and even-
angular blocks finds that the unsteady force amplitude and phase tual practical solution to industrial noise problems.
differences can either reinforce or cancel noise in the far-field. The current study has two objectives. First, to investigate the
Another type of geometry (for the downstream, secondary body) applicability of 2D Unsteady Reynolds Averaged Navier–Stokes
that has been studied is the thin flat plate, either detached or (URANS) to study noise from square cylinder in high Reynolds
attached (as a splitter plate) to the base of the upstream body. number (i.e. Re = 22,000). Second, to investigate the use of a
You et al. [27] studied the effect of splitter plate length on the downstream wedge as a passive noise control device in more de-
sound radiation for a circular cylinder at low Reynolds number tail than has been previously reported upon. This paper presents a
(Re = 100). Ali et al. [28,29] then adopt the splitter plate for the procedure to predict noise from two-dimensional bluff bodies us-
case of square cylinder at Re = 150. Shear layer reattachment was ing a URANS flow solver and an acoustic analogy based on Curle’s
found to significantly alter sound generation. The sound cancel- theory. The methodology is validated against new experimental
lation mechanism was also investigated by Ali et al. [30] at low data obtained for the case of a single square cylinder in cross flow.
Reynolds number (Re = 150). They found that there were two dis- Then the numerical methodology is used to investigate the effect
tinct regimes of sound generation. In regime I, a sound reduction of a downstream wedge on noise production.
of 2.3 dB is achieved when the gap distance of plate is G = 0.
While in regime II, as the plate length is altered to 0.26D and 2. Numerical procedure
placed at location where plate lift is out-of-phase with the cylinder
lift (5.6D), a significant reduction of 6.3 dB was obtained. How- The aerodynamic noise calculation starts with simulating a low
ever, due to differences in the acoustical waveforms generated by Mach number (Ma = 0.09) turbulent flow using an incompressible
the cylinder and plate, the maximum amount sound reduction is URANS method. Using this flow simulation, the sound source (in
limited. These differences are due to the non-linear fluid dynam- this case the time gradient of the lift force) is obtained. Then, after
ics experienced by the plate inserted in the cylinder wake. Hence, the source is obtained, the sound propagation is then calculated
there is an opportunity to investigate new downstream body ge- using Curle’s equation. Later, a spanwise correction is used to fur-
ometries that may improve the sound cancellation effect observed ther improve the accuracy of the noise prediction by considering
in previous seminal research reviewed above. the spanwise effects of the flow field.
There are some evidences that different downstream geome-
tries may be effective for passive noise control. A study by Uffin- 2.1. Flow simulations
ger et al. [31] correlated the flow and sound fields from square
cylinders with wedge or elliptical bodies placed in their wake. The governing equations for flow are given by
A square cylinder with a wedge in its wake shows a 4.5 dB lower
∂ ui
sound pressure level when compared with a downstream elliptical =0 (1)
body. No significant parameters such as the gap distance between ∂ xi
bodies were investigated in this study, leaving these particular as- ∂ ui 1 ∂p ∂ 2ui ∂  
=− +ν − u i u j − u i u j (2)
pects unexplored. Cheng et al. [32] also studied the effect of down- ∂t ρ ∂ xi ∂x j∂x j ∂xj
S.R.L. Samion et al. / Aerospace Science and Technology 53 (2016) 85–94 87

where subscripts i , j = 1, 2. Here x1 and x2 denote streamwise Only the small deviation between the lines in the upstream di-
and cross-stream directions, respectively, and u 1 and u 2 are the rection is due to the Doppler effect and is not taken into account
corresponding time-averaged velocity components; u i depicts the in the current study as the Mach number for the current study is
fluctuating part of the velocity; ρ is the density of fluid and p small (Ma = 0.09).
represents the fluid pressure; u i u j is the Reynolds stress term. The
k–ω SST model (Wilcox [35]) is used to provide closure to the sys- 3. Experimental method
tem of equations formed by Reynolds averaging. The open source
computational fluid dynamics (CFD) code OpenFOAM is used for For the experimental validation, a square cross-sectioned cylin-
simulating the fluid. The temporal term is discretized using second der with a diameter of 10 mm was tested in an anechoic wind
order backward scheme and the convection term is discretized us- tunnel at the University of Adelaide. The tunnel has internal di-
ing second order QUICK scheme. mensions of 1.4 m × 1.4 m × 1.6 m. The facility contains a rect-
angular contraction outlet with a height of 75 mm and a width of
2.2. Acoustic simulations 275 mm. The wind tunnel walls are lined with foam wedges which
provide an anechoic environment above 300 Hz. The schematic di-
The sound propagation is governed by the equations from the agram of the experimental setup is provided in Fig. 1.
Lighthill acoustic analogy [2] The cylinder had a span of 400 mm (so that it spanned the
  entire width of the contraction) and was mounted 30 mm from
∂2 2 2 ∂2Tij
2
− c 0 ∇ (ρ − ρ0 ) = (x, t ) (3) the exit plane of the nozzle. The mounting plate was the same
∂t ∂ xi ∂ x j that was used by Moreau et al. [37]. The cylinder was mounted
T i j = ρ u i u j − τi j + δi j (( p − p 0 ) − c 02 (ρ − ρ0 )) (4) in such a manner so that the ends were located in a zero velocity
environment. For the experiment, the free-stream velocity of the
Equation (3) is an inhomogenous wave equation derived from re- jet was set at 32 m/s.
arrangement of Navier–Stokes equations [2]. T i j is Lighthill’s stress A single B&K 4190 microphone was used to measure the flow-
tensor, where the first term at the right hand side of Eq. (4) induced noise. The microphone’s frequency range is ±1 dB be-
(ρ u i u j ) is the Reynolds stress tensor; τi j denotes viscous stress, tween 50–20000 Hz as stated in the transducer documentation.
δi j is Kronecker delta, ρ and p are the instantaneous density and The microphone was located 0.5 m from the cylinder axis at a
pressure, respectively. Subscript ‘0’ represents reference value of 90 degree angle to the flow direction directly above the cylinder.
the parameter. The ambient pressure is the reference value in the The acoustic pressure was sampled at 21 6 Hz for 10 seconds us-
current study. ing a DAQ with an automatic anti-aliasing filter. The microphone
In the current study, considering a compact body present in records were not adjusted for interference from the shear layer.
the flow (i.e. the square cylinder), the free-field Green’s function
is used to solve Lighthill’s equations. This method is introduced by 4. Validation test case
Curle and explained in Curle’s theory [1].
As the dimension of the body is very small compared to the
A single square cylinder of diameter D in cross flow is used
wavelength (ratio of wavelength to body dimension is 88.2), the
as a validation test case, of which the numerical prediction of the
sound source are assumed compact. In the case of a compact, fixed,
emitted sound is compared with the experimental measurement.
and rigid body, emission time variation along the body can be ne-
The computational domain size is 31.5D × 21D and the coordi-
glected. Hence, r ≈ |x|. Therefore, taking P i j as pressure, y as the
nate origin is at the center of the square cylinder (see Fig. 2).
point on the rigid surface, and instantaneous force F i of fluid on
The upstream, top, and bottom boundaries are 10D away from
the body (i.e. lift and drag) as
cylinder. The boundary conditions are as shown in Fig. 2. The in-

let boundaries are constant velocity boundary conditions set to
F i (tr ) ≈ [ P i j ]τ =tr n j dS( y ) (5) the free-stream value (Dirichlet BC), with a zero pressure gradi-
S ent (Neumann BC) conditions also applied. The outlet boundary
the Curle’s solution for a fixed rigid compact body is has a zero velocity gradient and a fixed pressure value. A uniform
 structured mesh is made surrounding the cylinder until 1D from
1 xi x j the surface of square cylinder, where the mesh is highly refined.
c 02 [
ρ (x, t ) − ρ0 ] = T i j (y, t − r /c 0 ) dV( y )
4π r 2 rc 02 This type of mesh is adapted from the grid convergence study of
V Ali et al. [38] for the case of square cylinder. Fig. 3 shows the mesh
1 xj ∂ adopted in current study.
− F j (t − r /c 0 ) (6) A grid refinement study assessing three levels of mesh density
4π r 2 c 03 ∂ t
was performed and the results are compared with similar previous
Hence, the radiated sound is obtained by a quadrupolar volume in- studies. Table 1 shows the description of the cells with compari-
tegration and a dipolar surface integration. Inoue and Hatakeyama son to previous information from the literature. Here, y w / D refers
[36] showed that the magnitude of sound pressure can be roughly to the unit mesh size in the buffer zone near the surface (the
estimated as p q /ρ U ∞
2
∝ AMa7/2 /r 1/2 and pd /ρ U ∞
2
∝ Ma5/2 /r 1/2 , smallest cell size). The values of y + are also taken into account to
where subscripts ‘q’ and ‘d’ are for quadrupole and dipole, re- investigate about the best grid for the simulation. y + denotes non-
spectively, and A is a constant. Hence, as the quadrupole sound dimensionalized distance of grid points from the cylinder wall and
is small compared to the dipole sound, it is neglected in this is defined as y + = u τ y /ν , where u τ is the friction velocity, and y
study. This leaves the following as solution to calculate the radi- is the distance from the body wall [35]. y + min
and y +max are values of
ated sound. the smallest and the biggest y + value respectively, while the aver-
    age value of y + is shown in the last column. Murakami et al. [39]
1 ∂ Fi 1 xi ∂ Fi
p  (x, t ) = − = (7) and Arslan et al. [40] performed LES simulations of a rectangular
4π ∂ x i r 4π c 0 r 2 ∂t
cylinder at similar high Reynolds number flow and are compared
Good agreement is found between the calculations using Curle’s in Table 1. The fine mesh has similar levels of refinement to the
solution with the DNS results as has been proved by Ali et al. [28]. previously published LES studies.
88 S.R.L. Samion et al. / Aerospace Science and Technology 53 (2016) 85–94

Fig. 1. Schematic diagram of the experimental setup.

Table 1
Computational meshed used in this study, compared with those in the literature.

Case No. of cells y w /D y+


min
y+
max y+
average

Coarse 158 × 120 0.0500 9.50 67.86 31.56


Medium 211 × 165 0.0285 4.85 38.58 17.36
Fine 353 × 306 0.0200 2.78 24.17 11.26
Murakami et al. [39] 104 × 69 0.022 ( y + ≤ 11.81)
Arslan et al. [40] 1 × 106 0.025 101

on the calculation using Eq. (7), the root mean square value of
sound pressure, P rms for numerical simulation is 8.85 × 10−5 Pa,
i.e. 12.9 dB when measured at R = 50D, θ = 90◦ above the cylin-
der. Sound pressure level (SPL) is calculated using Eq. (8) in which
the P ref is taken as 2 × 10−5 [Pa].

P rms
Fig. 2. Schematic diagram of computational domain and the boundary conditions SPL = 20 log( ) (8)
applied. P ref
The simulation acoustic result is validated by comparing it with
Numerical data used for calculating global, averaged results are that of experiment’s. Following the spanwise correction technique
taken when the solution has reached a statistically stable state. as has been suggested by Seo and Moon [22] and Doolan [24], the
This was done by taking data from the non-dimensional time acoustic power spectral density (PSD) of the current results is used
tU / D = 150 for at least 10 vortex shedding cycles. Table 2 shows for comparison. To calculate the value to be corrected to the sound
comparison of global results obtained using the fine grid with pressure level (SPL) of the simulation, the following Eq. (9) is used.
other previous studies. Almost identical results are obtained. Thus √
SPLc = 10 log( L c / L s ) + 10 log( π N) (9)
the fine mesh density (353 × 306) is used for the results presented
here. Here, L c represents the spanwise coherence length, L s denotes
The sound calculated is compared with the experimental re- the simulated span length as defined as L s = L / N, in which con-
sults obtained in the anechoic wind tunnel. The results show that sidering a long-span body with length L divided into N subsec-
the sound radiated due to the drag is small and the dominant tions. Since the vortex shedding in flow over circular cylinder at
sound source is due to the lift fluctuations, a similar finding by Re = 22,000 is correlated over a distance of about 6.5D, the value
Ali et al. [28]. Therefore, only the sound source due to the un- of L c in current study is taken as 5D, following the procedure used
steady lift force is considered in this study. The location to measure by Doolan [24] and Casalino and Jacob [23]; here, N = 10 giving
the sound is chosen to be at R = 50D, θ = 90◦ from the cen- L s = 2.75D. The numerical and experimental acoustic power spec-
ter of the square cylinder, the same as that of experiment. Based tral densities are compared in Fig. 4. The simulated PSD is able to
S.R.L. Samion et al. / Aerospace Science and Technology 53 (2016) 85–94 89

Fig. 3. Mesh of the computational domain.

Table 2
Comparison of global results of the current study with the literature.

No Author Method Re C D ,mean C L ,rms St


1 Present (Fine) k–ω SST 22,000 2.10 1.43 0.126
2 Tian et al. [41] k–ω SST 21,400 2.06 1.492 0.138
3 Bosch and Rodi [47] k–ε 22,000 2.108 1.012 0.146
4 Lyn [11] LDV Exp. 21,400 2. 1 – 0.13
5 Bearman [10] – 22,000 2.1 1.2 0.13
6 Shimada and Ishihara [16] k–ε 22,000 2.05 1.43 0.141
7 Murakami and Mochida [39] 3D LES 22,000 2.09 1.60 0.132
8 Arslan et al. [40] 3D LES 22,000 2.25 1.43 0.127

Fig. 5. Problem geometry for the case of a downstream wedge. The parameters that
are varied are height, h and gap distance, G.

Fig. 4. Comparison of computed and experimental acoustic power spectral densities.


Experimental results (black solid line) are compared with the corrected numerical
calculation results (blue dotted line). (For interpretation of the references to color
in this figure legend, the reader is referred to the web version of this article.)

reasonably reproduce the Aeolian tone frequency and magnitude.


The broadband levels at higher frequency are not well predicted
and this may be due to the URANS methodology not able to rep-
resent smaller eddies in the near wake as well as a breakdown in
Fig. 6. Root-mean-square sound pressure for various gap distances for all three cases
the assumption of acoustic compactness. As this study is concerned
compared with the isolated square cylinder.
with the flow and noise at the Aeolian tone only, the methodology
is considered sufficiently accurate to proceed with a fundamental
h = 1D (thick wedge), h = 0.5D (medium wedge) and h = 0.25D
investigation of passive flow and noise control using a downstream
(thin wedge) known here as Case 1, Case 2 and Case 3, respec-
wedge.
tively. The gap distance is varied 0 ≤ G ≤ 7D so that the sensitivity
of the gap distance with the flow and noise generation can be in-
5. Investigation of the effect of a downstream wedge
vestigated.

5.1. Test case description 5.2. Summary of flow regimes and sound generation

The effect of gap distances in the downstream (G) and wedge Fig. 6 shows the variations of root-mean-square sound pressure
height (h) are the important factors relating to the sound gener- (p rms /ρ U ∞
2
) at different gap distances. It is found there are two
ation being investigated in this study. Fig. 5 shows the problem types of regimes (regime I and regime II) present in the change of
geometry. Three types of wedge height (h) were used in this study; sound with gap distance. There is also a small transition regime
90 S.R.L. Samion et al. / Aerospace Science and Technology 53 (2016) 85–94

cylinder) returns to the vicinity of that of single square cylin-


der.

5.3. Lift

As flow past over a bluff body, shear layers from two sides of
the body are shed alternately downstream the body. When one
side sheds a shear layer, it will experience a low pressure and the
opposite side will experience a high pressure. Hence, the fluctuat-
ing forces are experienced by the bluff bodies (square cylinder and
wedge) are the result of the alternation of periodic pressure fluc-
tuations from both sides of the body. This unsteady force happens
Fig. 7. Sound source, ∂∂Ft 2 for each test case. at the frequency of the vortex shedding.
In regime I, the lift acting on the square cylinder and wedge
are sensitive to the gap distance as shown in Fig. 9. Especially
between regimes I and II. The classification of regime is based on
in medium and thin wedge case, there is a gradual reduction in
the flow behavior properties and flow patterns.
lift of square cylinder with gap distance, but the magnitude of
Regime I shows a significant sound reduction compared with
lift acting on the wedge increases. The same pattern was found
the isolated square cylinder. The highest reduction is achieved
by Ali et al. [43]. This can be explained by the wake–wedge in-
when the wedge is h = 0.25D at G = 2D i.e. 11.79 dB. On the
teraction. As the wedge is placed nearer to the core of growing
other hand, in regime II, the magnitude of sound at first exceeds,
vortex, the velocity fluctuations around the wedge intensify and
but returning to the vicinity of sound magnitude of that of sin- consequently the surface pressure fluctuations on wedge increase.
gle square cylinder case when approaching G = 6D. There is also Thus, the lift acting on the wedge increases due to this wake–
sound reduction obtained in regime II cases but at the gap distance wedge interaction. In regime II, the lift acting on square cylinder
G ≥ 6D. is weakened before it slowly returns to its original value while
The two regimes are due to the effect of the wedge which has the lift acting on the wedge at first is approximately the same
altered the sound generation mechanism [28]. The sound source magnitude as that of square cylinder but gradually decreases as
(i.e. the time gradient of the lift, ∂∂Ft2 ) for both the square cylin- the gap distance increases. As the wedge is placed further from
der and the wedge are shown in Fig. 7. It is observed that there the core of the growing vortex of the upstream body, the mag-
is an extreme difference of root-mean-square of lift gradient be- nitude of the velocity fluctuations reduce. Thus, at small gaps in
tween that in regime I and regime II, for both the square cylin- regime II, both the lift on the square (cylinder) and wedge de-
der and the wedge. The downstream flow structures are governed crease. However, as the wedge is placed further downstream, the
by the gap distance between the square cylinder and the wedge. presence of the wedge has a diminished effect on the lift. There-
Time-averaged streamlines shown in Fig. 8 illustrate the differ- fore, the value returns to the value of C L ,rms of the isolated square.
ent regimes observed. In regime I, the saddle point is found be- The gradual changes of the lift on the wedge relate to the roll-up
hind the wedge. This happens when the separated shear layer of the shear layer from the square cylinder. This has more chance
from the square cylinder reattaches behind the wedge and a re- to dissipate when the wedge is placed further downstream. Thus
circulation region is formed in the wake of the wedge. In regime the wedge is in the vicinity of a weaker vortex thus experienc-
I, the square cylinder and wedge experience reduced magnitude ing a lower unsteady aerodynamic force. Furthermore, a thinner
of lift acting upon them. Hence, the sound source in regime I wedge experiences lower lift than on the thicker wedge in regime
shows reduced value leading to the reduced magnitude of root- II. This relates with the increasing of vortex diffusion and the loss
mean-square sound pressure. On the contrary in regime II, the of flow symmetry as the wedge becomes thinner. A similar pat-
recirculation region forms within the gap, and shows the same tern found in Iungo et al. [44] where the lift acting on the body
pattern with the case of single square cylinder. Therefore, the is reduced when the aspect ratio (length/height) of the wedge is
sound source (gradient of lift force experienced by the square higher.

Fig. 8. Streamline (time-averaged velocity) for case 3 at representative gap distances of different regimes. (a) Single square cylinder, (b) regime I, G = 0, (c) transition regime,
G = 2.26D, (d) regime II, G = 2.76D.
S.R.L. Samion et al. / Aerospace Science and Technology 53 (2016) 85–94 91

as the wedge is placed far from the square cylinder, the wake–
wedge interaction becomes weak. Thin wedge cases show the low-
est drag value followed by medium wedge case when compared
to regime II. Thus it can be concluded that the thinner the wedge
gets, the lower the drag acting on the wedge.

5.5. Vortex shedding

Figs. 11(a), 12(a) and 13(a) show instantaneous vorticity con-


tours at a time when the lift is maximum for G = 0 for cases 1, 2
and 3 respectively. The vortex shedding pattern from square cylin-
der with wedge in regime I is the same as the vortex shedding of a
low Re case as has been found in Inoue et al. [36,42], Ali et al. [43]
and Doolan [46]. Figs. 11(b), 12(b) and 13(b) represent instanta-
neous vorticity contours in the transition regime for thick, medium
Fig. 9. Root-mean-square lift coefficient for various gap distances for all test cases. and thin wedge cases respectively. Transition regime flow depicts
both the characteristics of regime I and regime II. While, Fig. 11(c),
12(c) and 13(c) show instantaneous vorticity contours for regime II
for cases 1 to 3 respectively. Broadly, regime I is characterised by
vortex formation occurring behind the wedge, regime II is charac-
terised by vortex formation within the gap between the cylinder
and wedge, while the transition region is more complicated in
that there exists an unsteady flow region between the cylinder and
wedge, flow reattachment on the wedge that affects vortex shed-
ding in the wake.
The Strouhal number (St = f D /U , where f is vortex shedding
frequency) is presented in Fig. 14. The vortex shedding frequency is
obtained from power spectrum density of lift fluctuations. Almost
all cases have values of St lower than that of isolated square. This
is expected as the formation of the vortex has been disturbed by
the presence of the wedge.
Fig. 10. Mean drag coefficient for various gap distances for all test cases.

5.6. Sound radiation


5.4. Drag
Figs. 15 to 17 show contours of sound pressure, (that is cal-
The two regimes can also be defined by mean drag coefficient. culated using Eq. (7)) for of regime I, the transition regime, and
Generally, the magnitude of mean drag acting on the square cylin- regime II for thick, medium and thin wedge cases respectively.
der and wedge decrease in regime I. When the wedge is placed in Sound radiates in the direction of 90◦ from the free stream, in a
regime I, the wake width and the vortex formation length increase, dipolar pattern, which is as expected as the source of noise is un-
resulting in lower drag. steady lift. The estimated wavelength of an isolated square cylin-
In regime II, the magnitude of mean drag acting on square der is λ ≈ D /(Ma × St ) ≈ 88.2D (St = 0.0113). In regime I, thick
cylinder increases and returns to the value of an isolated square wedge case generates a sound wave with a shorter wavelength
cylinder. The drag acting on the wedge also increases and stays than the isolated square cylinder, i.e. λ ≈ 75.3D (St = 0.0133) for
constant at a fixed value. When the wedge is placed in regime II, G = 0. In the transition regime, the sound wavelength is λ ≈ 83.3D
the vortex formation length returns to the same length with that (St = 0.012) and regime II the sound wavelength is λ ≈ 88.1D
of the isolated square cylinder, i.e. the development of the vor- (St = 0.0114) approaching that value of isolated square cylinder
tex in the wake of square cylinder is no longer affected by the case. For medium thin wedge case, the placement of the wedge
presence of the wedge when it is placed further downstream in in regime I and the transition regime leads to an increase in the
regime II. Thus the mean drag acting on the square cylinder re- acoustic wavelength. The medium wedge in regime I gives out
turns the same as the drag acting on isolated square as can be sound of wavelength of λ ≈ 107.1D (St = 0.0093), while in tran-
seen in Fig. 10. Also, the reattachment of the shear layers that oc- sition regime (G = 1.5D) the sound wavelength is λ ≈ 102.4D
curs within the gap has constrained the recirculation region. This (St = 0.0098). For thin wedge case, regime I has an acoustic wave-
leads the base pressure magnitude experienced by the wedge to be length of λ ≈ 95.2D (St = 0.0105) and the transition regime, gen-
constant, so the drag acting on the wedge is constant. However, the erates sound with a wavelength of approximately λ ≈ 107.1D (St =
fixed magnitude of drag is lower for the thinner wedge. In Ganga 0.0093). Both medium and thin wedges, for regime II, are found to
Prasath et al. [45] who studied the effect of aspect ratio of trian- radiate the wavelength that approaches the value of the isolated
gular prism, it is found that the drag coefficient decreases with the square cylinder. They radiate sound with a wavelength λ ≈ 83.3D
increase of aspect ratio (the triangular prism becomes longer) due (St = 0.012) for both medium and thin wedge cases.
to the separated flow zone shrinking. This relates to current study; Fig. 18 shows the phase difference between the sound emitted
when the wedge is thinner (aspect ratio increases), the drag acting from the square cylinder and the sound emitted from the wedge
on the wedge itself is lower. for various gap distances for all three cases. The phase lag becomes
Generally in regime II, lift and drag acting on the square cylin- linear once the gap distance is sufficient to establish vortex shed-
der gradually return to their value the same as that of the isolated ding between the cylinder and the wedge (regime II). This linearity
square case. Different behavior is found for the wedge. Lift of the can be explained by the nearly constant convection velocity of the
wedge in regime II gradually decreases, while drag stay constant shed vortices from the upstream cylinder in regime II. In contrast,
at fixed value when the wedge is in regime II. This is because the phase difference is non-linear in regime I and the transition
92 S.R.L. Samion et al. / Aerospace Science and Technology 53 (2016) 85–94

Fig. 11. Instantaneous vorticity contours for thick wedge case (case 1) taken at time when lift is at maximum. Red-lines represent clockwise direction with contour level
ranging −10 ≤ D /U ∞ ≤ −0.3. Green lines represent anticlockwise direction with contour level ranging 0.3 ≤ D /U ∞ ≤ 10. : spanwise vorticity. (For interpretation of
the references to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 12. Instantaneous vorticity contours for medium wedge case (case 2) taken at a time when the lift is at maximum. Red-lines represent clockwise direction with contour
level ranging −10 ≤ D /U ∞ ≤ −0.3. Green lines represent anticlockwise direction with contour level ranging 0.3 ≤ D /U ∞ ≤ 10. : spanwise vorticity. (For interpretation
of the references to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 13. Instantaneous vorticity contours for thin wedge case (case 3) taken at a time when the lift is at maximum. Red-lines represent clockwise direction with contour level
ranging −10 ≤ D /U ∞ ≤ −0.3. Green lines represent anticlockwise direction with contour level ranging 0.3 ≤ D /U ∞ ≤ 10. : spanwise vorticity. (For interpretation of
the references to color in this figure legend, the reader is referred to the web version of this article.)

G = 1D, while medium wedge case starts at G = 3.5D and thin


wedge case starts at G = 2.5D.

6. Conclusion

This paper has presented a method for calculating the flow


and noise from bluff bodies at high Reynolds number using two-
dimensional URANS flow simulations and Curle’s acoustic anal-
ogy. The method was validated using new experimental data for
the test case of a single square cylinder in a cross flow with
Re = 22,000 and Ma = 0.09.
The validated methodology was then used to investigate the ef-
fect of a downstream, detached wedge as a form of passive flow
and noise control. It was found that the flow behavior can be di-
Fig. 14. Strouhal number for various gap distances, for all test cases. vided into two regimes based on the analysis done on the flow
properties and flow patterns. There is also the presence of a tran-
sition regime in which the flow behaves as if it is about to change
regime when constant vortex shedding has not been established from regime I to regime II. Noise generated in regime I showed a
within the gap. The pattern is similar to the study of Ali et al. [30] reduction in level compared with the single square cylinder. How-
who investigated effect of placing a detached plate in the wake of ever, noise generated in regime II case showed an increase in the
a square cylinder. Linearity relation for thick wedge case starts at sound level. The reduction on noise level is due to the reduction
S.R.L. Samion et al. / Aerospace Science and Technology 53 (2016) 85–94 93

Fig. 15. Case 1: Contours of sound pressure (p  /ρ U ∞


2
). Blue-lines represent positive signs. Red-dashed-lines represent negative signs. Contour levels are from 1.0 × 10−9 to
8.0 × 10−9 with constant increment of 1.0 × 10−10 . (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 16. Case 2: Contours of sound pressure (p  /ρ U ∞


2
). Blue-lines represent positive signs. Red-dashed-lines represent negative signs. Contour levels are from 1.0 × 10−9 to
8.0 × 10−9 with constant increment of 1.0 × 10−10 . (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Fig. 17. Case 3: Contours of sound pressure (p  /ρ U ∞


2
). Blue-lines represent positive signs. Red-dashed-lines represent negative signs. Contour levels are from 1.0 × 10−9 to
8.0 × 10−9 with constant increment of 1.0 × 10−10 . (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

and wedge causing an intense vortex–wedge interaction, causing


high levels of noise radiation. For the current test cases, it was
found that the wedge cannot completely cancel the sound created
by a square cylinder in high Reynolds number flow; however, mod-
ification of the shape and position of the downstream wedge may
allow a significant reduction of sound, if the phase and strength of
the noise source of the wedge can be controlled in such a man-
ner so that it can significantly cancel the noise generated by the
upstream cylinder.

Conflict of interest statement


Fig. 18. Sound phase lag for various gap distance in different cases.
The article is the authors’ original work, has not received prior
in aerodynamic forces on the cylinder and wedge in regime I. In publication and is not under consideration for publication else-
regime II, vortex shedding occurs in the gap between the cylinder where.
94 S.R.L. Samion et al. / Aerospace Science and Technology 53 (2016) 85–94

Acknowledgements [23] D. Casalino, M. Jacob, Prediction of aerodynamic sound from circular rods via
spanwise statistical modelling, J. Sound Vib. 262 (2003) 815–844.
[24] C.J. Doolan, Computational bluff body aerodynamic noise prediction using a
The authors gratefully acknowledge the use of the services and
statistical approach, Appl. Acoust. 71 (2010) 1194–1203.
facilities of the HPC-UTM at the Universiti Teknologi Malaysia, [25] M. Gad-El-Hak, Flow Control: Passive, Active and Reactive Flow Management,
FRGS PY/2015/05383 and TWAS 13-272 research grants and MJIIT Cambridge University Press, New York, 2000, p. 14.
fellowship award to the first author. Also, the support from the [26] C.J. Doolan, D.J.J. Leclercq, The interaction of a bluff body with a vortex wake,
University of Adelaide for the experimental results is gratefully ac- J. Fluids Struct. 25 (5) (2009) 867–888.
[27] D.H. You, H. Choi, M.R. Choi, S.H. Kang, Control of flow-induced noise behind a
knowledged.
circular cylinder using splitter plates, AIAA J. 36 (11) (1998) 1961–1967.
[28] M.S.M. Ali, C.J. Doolan, V. Wheatley, The sound generated by a square cylin-
References der with a splitter plate at low Reynolds number, J. Sound Vib. 330 (2011)
3620–3635.
[1] N. Curle, The influence of solid boundaries upon aerodynamic sound, Proc. R. [29] M.S.M. Ali, C.J. Doolan, V. Wheatley, Aeolian tone generated by a square cylin-
Soc. Lond. Ser. A, Math. Phys. Sci. 231 (1187) (1955) 505–514. der with a splitter plate, in: Proceedings of 20th International Congress on
[2] M.J. Lighthill, On sound generated aerodynamically: I. General theory, Proc. R. Acoustics, ICA, Sydney, Australia, 23–27 August, 2010, pp. 1–8.
Soc. Lond. Ser. A, Math. Phys. Sci. 211 (1107) (1952) 564–587. [30] M.S.M. Ali, C.J. Doolan, V. Wheatley, Aeolian tones generated by a square cylin-
[3] A. Goldburg, B.H. Florsheim, Transition and Strouhal number for the incom- der with a detached flat plate, AIAA J. 51 (2) (2013) 291–301.
pressible wake of various bodies, Phys. Fluids 9 (1) (1966) 45–50. [31] T. Uffinger, S. Becker, D. Antonio, Investigations of the flow field around dif-
[4] A. Tamura, T. Miyagi, The effect of turbulence on aerodynamic forces on a ferent wall-mounted square cylinder stump geometries, in: 14th Int Symp on
square cylinder with various corner shapes, J. Wind Eng. Ind. Aerodyn. 83 (1–3) Applications of Laser Techniques to Fluid Mechanics, Lisbon, Portugal, vol. 20,
(1999) 135–145. 2008, pp. 1–13.
[5] M. Ozogen, Flow structure in the downstream of square and circular cylinder, [32] M. Cheng, G. Liu, Effects of afterbody shape on flow around prismatic cylinders,
Flow Meas. Instrum. 17 (4) (2006) 225–235. J. Wind Eng. Ind. Aerodyn. 84 (2000) 181–196.
[6] O. Inoue, W. Iwakama, N. Hatakeyama, Aeolian tones radiated from flow past [33] S. Ganga Prasath, M. Sudharsan, V. Vinodh Kumar, S.V. Diwakar, T. Sundarara-
two square cylinder in a side-by-side arrangement, Phys. Fluids 28 (4) (2006) jan, Shaligram Tiwari, Effects of aspect ratio and orientation on the wake
046101. characteristics of low Reynolds number flow over a triangular prism, J. Flu-
[7] M.H. Wu, C.Y. Wen, R.H. Yen, M.C. Weng, A.B. Wang, Experimental and numer- ids Struct. 46 (2014) 59–76.
ical study of the separation angle for flow around a circular cylinder at low [34] G.V. Iungo, G. Buresti, Experimental investigation on the aerodynamic loads
Reynolds number, J. Fluid Mech. 515 (1) (2004) 233–260. and wake flow features of low aspect-ratio triangular prisms at different wind
[8] A. Sohankar, L. Davidson, C. Norberg, Numerical simulation of unsteady flow directions, J. Fluids Struct. 25 (7) (2009) 1119–1135.
around a square two-dimensional cylinder, in: 12th Australian Fluids Mechan- [35] D.C. Wilcox, Turbulence Modeling for CFD, 2nd ed., DCW Ind. Inc., La Canada,
ics Conference, University of Sydney, Australia, 1995. CA, 1998.
[9] A. Sohankar, L. Davidson, C. Norberg, Numerical simulation of unsteady low- [36] O. Inoue, N. Hatakeyama, Sound generation by a two-dimensional circular
Reynolds number flow around rectangular cylinders at incidence, J. Wind Eng. cylinder in a uniform flow, J. Fluid Mech. 471 (2002) 285–314.
Ind. Aerodyn. 69–71 (1997) 189–201. [37] D.J. Moreau, L.A. Brooks, C.J. Doolan, Broadband trailing edge noise from a
[10] P.W. Bearman, E.D. Obasaju, An experimental study of pressure fluctuations sharp-edged strut, J. Acoust. Soc. Am. 129 (2011) 2820.
on fixed and oscillating square-section cylinders, J. Fluid Mech. 119 (1982) [38] Mohamed Sukri Mat Ali, C.J. Doolan, Vincent Wheatley, Grid convergence study
297–321. for a two-dimensional simulation of flow around a square cylinder at a low
[11] D.A. Lyn, S. Einav, W. Rodi, J.H. Park, A laser-Doppler velocimetry study of Reynolds number, in: Seventh International Conference on CFD in the Minerals
ensemble-averaged characteristic lf turbulent near wake of a square cylinder, and Process Industries, Melbourne, Australia, 2009, pp. 1–6.
J. Fluid Mech. 304 (1995) 285–319. [39] S. Murakami, A. Mochida, On turbulent vortex shedding flow past 2D square
[12] H. Nakaguchi, H. Hashimoto, K. Muto, An experimental study on aerodynamic cylinder predicted by CFD, J. Wind Eng. Ind. Aerodyn. 54–55 (1995) 191–211.
drag of rectangular cylinders, J. Jpn. Soc. Aerosp. Sci. 16 (5) (1968). [40] T. Arslan, G.K. El Khoury, B. Pettersesn, H.I. Andersson, Simulations of flow
[13] C. Norberg, Flow around rectangular cylinder: pressure forces and wake fre- around a three-dimensional square cylinder using LES and DNS, in: The Sev-
quencies, J. Wind Eng. Ind. Aerodyn. 49 (1993) 187–196. enth International Colloquium on Bluff Body Aerodynamics and Applications,
[14] Y. Ohya, Note on discontinuous change in the wake pattern for a rectangular Shanghai, China, 2012, pp. 909–918.
cylinder, J. Fluids Struct. 8 (1994) 325–330. [41] X. Tian, M.C. Ong, J. Yang, D. Myrhaug, Unsteady RANS simulations of flow
[15] A. Okajima, Numerical simulation of flow around rectangular cylinders, J. Wind around rectangular cylinder with different aspect ratios, Ocean Eng. 58 (2013)
Eng. Ind. Aerodyn. 33 (1990) 171–180. 208–216.
[16] K. Shimada, T. Ishihara, Application of a modified k–epsilon model to the pre- [42] O. Inoue, W. Iwakami, N. Hatakeyama, Aeolian tones radiated from flow past
diction of the aerodynamic characteristics of rectangular cross-section cylin- two square cylinders in a side-by-side arrangement, Phys. Fluids 18 (4) (2006)
ders, J. Fluids Struct. 16 (4) (2002) 465–485. 046104.
[17] A. Sohankar, Flow over a bluff body from moderate to high Reynolds numbers [43] Mohamed Sukri Mat Ali, Con J. Doolan, Vincent Wheatley, Low Reynolds num-
using large eddy simulation, Computers and Fluids 35 (206), 1154–1168. ber flow over a square cylinder with a detached flat plate, Int. J. Heat Fluid
[18] H. Fujita, W. Sha, H. Suzuki, Experimental investigations and prediction of aero- Flow 36 (2012) 133–141.
dynamic sound generated from square cylinders, AIAA Paper 98 (2369) (1998) [44] G.V. Iungo, G. Buresti, Experimental investigation on the aerodynamic loads
942. and wake flow features of low-aspect ratio triangular prisms at different wind
[19] S.N. Nakato, Characteristics of aerodynamic sound from rectangular cylinder directions, J. Fluids Struct. 25 (2009) 1119–1135.
with various angle of attack (in Japanese), J. Jpn. Soc. Civ. Eng. 696 (2002) [45] S. Gangga Prasath, M. Sudharsan, V.V. Kumar, S.V. Diwakar, T. Sundararajan, S.
145–155. Tiwari, Effects of aspect ratio and orientation on the wake characteristics of
[20] F. Margnat, Hybrid prediction of the aerodynamic noise radiated by a rectan- low Reynolds number flow over a triangular prism, J. Fluids Struct. 46 (2014)
gular cylinder at incidence, Comput. Fluids 109 (2015) 13–26. 59–76.
[21] Akira Rokugou, Okajima, Numerical analysis of aerodynamic sound radiated [46] C.J. Doolan, V. Choley, J. Crespel, Vortex shedding during the interaction of a
from rectangular cylinder, J. Wind Eng. Ind. Aerodyn. 96 (2008) 2203–2216. turbulent wake with a cylinder, J. Fluids Struct. 31 (2012) 141–146.
[22] J.H. Seo, Y.J. Moon, Aerodynamic noise prediction for long span bodies, J. Sound [47] G. Bosch, W. Rodi, Simulations of vortex shedding past a square cylinder with
Vib. 306 (2007) 564–579. different turbulence models, Int. J. Numer. Methods Fluids 28 (1998) 601–616.

Potrebbero piacerti anche