Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/368542646

Experimental near-field analysis for flow induced noise of a structured


porous-coated cylinder

Article in Journal of Sound and Vibration · May 2023


DOI: 10.1016/j.jsv.2023.117611

CITATIONS READS

7 304

4 authors, including:

Reza Maryami Elias Arcondoulis


Southern University of Science and Technology University of Bristol
26 PUBLICATIONS 258 CITATIONS 66 PUBLICATIONS 597 CITATIONS

SEE PROFILE SEE PROFILE

Yu Liu
Southern University of Science and Technology
121 PUBLICATIONS 1,239 CITATIONS

SEE PROFILE

All content following this page was uploaded by Yu Liu on 09 April 2023.

The user has requested enhancement of the downloaded file.


Please cite this article as: Maryami R, Arcondoulis E, Liu Q, Liu, Y (2023). Experimental
near-field analysis for flow induced noise of a structured porous-coated cylinder, Journal of Sound

©
and Vibration, 551, 117611. https://doi.org/10.1016/j.jsv.2023.117611
2022, Elsevier. This manuscript version is made available under the CC-BY-NC-ND 4.0 license
http://creativecommons.org/licenses/by-nc-nd/4.0/

Experimental near-field analysis for flow induced noise of a structured


porous-coated cylinder

Reza Maryami, Elias J. G. Arcondoulis, Qian Liu, Yu Liu∗


Department of Mechanics and Aerospace Engineering, Southern University of Science and Technology,
Shenzhen,Guangdong 518055, China

Abstract
Porous coating on cylinders reduces the aerodynamic noise due to vortex shedding. A
key understanding of how porous media suppress vortex shedding is related to near-field
hydrodynamics. In this paper, an experimental study was undertaken over the Reynolds number
range of 0.52 × 105 ≤ Re ≤ 1.74 × 105 to measure outer diameter surface pressure fluctuations
of a structured porous-coated cylinder (SPCC). The near- field pressure and far-field noise were
measured simultaneously to obtain a deep understanding of the noise-reduction mechanism of the
SPCC. The results showed that the SPCC causes a delay in the boundary layer separation compared
to the bare cylinder. While a substantial reduction in far-field noise was provided by the SPCC,
a high-frequency noise attributed to the non-propagating hydrodynamic energy field within the
SPCC was observed. The use of an SPCC led to a reduction in the surface pressure fluctuations
over the entire frequency range, acting as a noise reduction mechanism. It was shown that the
surface pressure fluctuations and acoustic signals are strongly correlated at the vortex shedding
frequency for the bare cylinder, while it was insignificant for the SPCC, indicating the strong role
of the SPCC in reducing the surface pressure energy content so that it cannot propagate to the
far-field. The evaluation of vortex shedding in the time-frequency domain was carried out with the
aid of wavelet transform. It was observed that amplitude modulation in time was verified only for
the bare cylinder at the fundamental vortex shedding frequency in the pre- and post-separation
regions.
Keywords: aeroacoustics, passive vortex shedding control, porous media, circular cylinder,
remote-sensing method


Corresponding author
Email address: liuy@sustech.edu.cn (Yu Liu)

Article published by Journal of Sound and Vibration 15 February 2023


1. Introduction

The aerodynamics of circular cylinders placed in a laminar or a turbulent incident flow has been
the subject of interest for many years [1–6] as it is of relevance to a large number of flows found
in engineering applications including high-speed train pantographs, wind turbine pylons, landing
gear struts, bridges, chimneys, and risers in marine engineering. It is well known that above a
critical Reynolds number, vortex shedding occurs in the wake of a cylinder, resulting in significant
structural vibrations, noise, and an increase in force fluctuations acting on the cylinder. The onset
of vortex shedding and its frequency can be shifted and even suppressed by various techniques
classified under two main categories, namely passive [7–13] and active [14–19] methods. Passive
methods are more sophisticated in terms of their reliability, cost, and maintenance capabilities and
have been examined in the general areas of aerodynamics and aeroacoustics.
The application of porous materials to robust control of flow-induced noise, turbulence
stabilization, and significant aerodynamic improvement of the circular cylinders has received
considerable attention in a relatively large body of studies [13, 20–27]. The results in the literature
showed that the effectiveness of porous materials in noise reduction depends heavily on the
geometrical and mechanical properties of the material, namely porous thickness, porosity, spacings
of the pores, and permeability coefficients [13, 20, 22, 26, 28, 29]. Liu et al. [22] showed that the noise
level decreases by increasing the porosity, pore density, and porous layer thickness, but it is more
sensitive to porosity than the other two parameters. Furthermore, variations in circumferential
and spanwise porosity were shown by Arcondoulis et al. [27, 30] to have little effect on tonal noise
reduction yet influence high frequency contributions. They designed an SPCC and demonstrated
that the overall porosity determines the vortex shedding tone magnitudes and also the overall
sound pressure level (OASPL), regardless of its orientation to the flow.
In addition to understanding the optimal design parameters, the studies on the PCCs focus
mainly on vortex shedding, noise control, and aerodynamic loads. One of the pioneers in this
field may be Sueki et al. [13], who reported a delay in the boundary layer separation point of
the PCC and a reduction in the aerodynamic sound. The underlying mechanism for controlling
the wake through the porous layer was numerically investigated by Bhattacharyya and Singh [31]
and it was shown that the vortex shedding is delayed by a porous treatment. Flow around a
PCC was studied by Naito and Fukagata [21] and the results showed that the vortex shedding is
completely suppressed by porous media and the shear layers are nearly symmetric. Liu et al.[22, 32]
showed that the porous cover can reattach the recirculation zone, which consequently results in
the reduction of noise and aerodynamic forces acting on the cylinder. The effect of porous media
on the wake of a cylinder has also been analyzed by Xia et al. [23]. They showed that the porous
media weakens the lateral oscillation of the vortex shedding and hence the vortex region is shifted
further downstream.
It is well known that the near-field pressure is affected by the vortex roll-up [1–3] and vortex
shedding is the main mechanism of the far-field noise [33–36], yet the correlation between near-
and far-field pressure is crucial to understanding the underlying noise reduction mechanism of
a porous coating. As reviewed above, the flow field and far-field noise measurements have been
carried out for any type of PCC over the years [13, 23, 25–27, 30]. While Klausmann and Ruck [29]
measured static pressure on the solid surface beneath a porous layer, to date, no experimental
studies have been conducted to measure the pressure fluctuations at the interface between the fluid
and porous media synchronised with far-field noise, leading to a gap in the literature. This study
aims to fill this gap by measuring the unsteady surface pressure and far-field noise simultaneously.
2
Figure 1: Schematic diagram depicting the experimental setup: (a) the contraction nozzle, the endplates, the
placement of the cylinder in the aeroacoustic wind tunnel, and the location of the far-field microphone (b) the
xy-plane of the cylinder.

To do so, a remote-sensing method was used to measure surface pressure fluctuations around the
circumference of an SPCC, which has not been conducted to date for any type of PCC. The SPCC
was used as the porous media due to its ability to fit remote-sensing pressure taps through the
pores in a special pattern, that would otherwise be more difficult using randomized foams. The
current paper is organized as follows. Section 2 describes the experimental setup including the
experimental facility, the cylinder design, and the measurement approach. Section 3 presents the
results of near- and far-field noise measurements and major findings concerning the underlying
mechanism of the far-field noise reduction by the porous coating are summarized in Sec. 4.

2. Experimental setup

2.1. Experimental facility


The experiments were carried out in a closed-circuit type open-jet anechoic wind tunnel at
Southern University of Science and Technology (SUSTech). The aeroacoustic facility includes an
anechoic chamber with internal dimensions of 3.8 m×5.7 m×3 m. The test section corresponding
to the nozzle exit has a rectangular cross-section with a width of 600 mm and a height of 550
mm as shown in Fig. 1. The wind tunnel is capable of producing flow speeds up to 70 m/s with
an incoming flow turbulence intensity of 0.15%. The cylinders were mounted between the top and
bottom laser-cut plates with a length of 800 mm attached to the nozzle exit. According to the
coordinate system with an origin located at the midspan, and the streamwise (x), the wall-normal
(y), and the spanwise (z) axises, the nozzle exit is located at x = −300 mm.

2.2. Cylinder design


The measurements were carried out using a cylinder with an outer diameter of D = 64.7 mm
and a span of L = 550 mm. The geometrical description of the cylinder coated by a structured
porous layer is presented in Fig. 2(a). The cylinder was made of three different parts, consisting of
one middle section with pressure taps instrumentations and two side extension parts. The blockage
ratio of the cylinder is found to be less than 12%. The cylinder also has an aspect ratio of about
3
Figure 2: (a) Schematic of the cylinder coated with structured porous layer, (b) the sensing area on the cylinder
equipped with pressure taps distributed around the circumference of the cylinder at midspan, and (c) the assembly
view of the remote-sensing microphone holder.

8.5, and thus the endplates’ effect on the flow over the middle section can be negligible [37].
Experiments were carried out at seven different free-stream velocities, U∞ = 15, 25, 30, 35, 40, 45,
and 50 m/s, corresponding to the diameter-based Reynolds numbers of Re = 0.52 × 105 , 0.86 × 105 ,
1.04 × 105 , 1.21 × 105 , 1.39 × 105 , 1.56 × 105 , and 1.74 × 105 , respectively. The middle section was
coated by a structured porous layer and has a span of 7D = 453 mm. The spanwise coherence
length of a bare cylinder extends to 3D [38] or 3.72D [4], and therefore a bare cylinder with 7D
span length would be sufficient to allow the development of a three-dimensional (3-D) flow field
along the span of the cylinder. In the case of the SPCC, it is difficult to specify an adequate span
length to yield 3-D flow features as the spanwise coherence length for permeable media changes
according to different porous structures and properties. In the case of the randomized PCC with
higher porosity, Geyer [26] showed that spanwise coherence is in excess of 7D. In this study, the
SPCC has a design similar to that published by Arcondoulis et al. [25] as shown in Fig. 2(b). The
SPCC possesses four porous layers with a constant porosity of ϕ = 81%. According to earlier
studies [13, 20, 21, 39], effective flow stabilization can be achieved using porous materials with
high porosity (ϕ ≥ 80%). Therefore, an SPCC with a porosity of 81% can be regarded as capable
of suppressing vortex shedding. The SPCC has an inner diameter of d = 40 mm and a porous layer
thickness of t = 12.35 mm and thus the porous coating thickness to the inner cylinder diameter
ratio is t/d ≈ 0.3, consistent with other studies [22, 25]. The PPI of the SPCC is about 9 close to
the effective values in aerodynamic sound reduction [40, 41].

2.3. Unsteady pressure measurements setup


In order to measure steady pressure, the cylinder was instrumented with 20 pressure taps,
which are distributed with non-uniform angular spacing over the circumference of the cylinder at
4
Table 1: Peripheral position of the pressure taps in the middle section at z/D = 0. θ is defined in Figs. 1(b) and
2(a).
Pressure ports θ Pressure ports θ
P1 0◦ P11 180◦
P2 16.36◦ P12 −163.63◦
P3 40.9◦ P13 −130.9◦
P4 57.27◦ P14 −122.72◦
P5 73.63◦ P15 −106.36◦
P6 90◦ P16 −90◦
P7 106.36◦ P17 −73.63◦
P8 130.9◦ P18 −40.9◦
P9 147.27◦ P19 −32.72◦
P10 163.63◦ P20 −16.36◦

midspan. The labeling of the pressure taps and their corresponding angular distances from the
stagnation point (θ = 0◦ ) are presented in Table 1. The pressure taps consist of a brass tube with
the inner and outer diameters of 0.8 mm and 1.6 mm, respectively, which passes through the void
in the structured porous media and is flushed with the outer diameter of the SPCC as shown in Fig.
2(b). The brass tubes may have intrusive effects, such as the changes in porosity yet Arcondoulis et
al. [27, 30] demonstrated that small variations in circumferential and spanwise porosity of an SPCC
have little effect on tonal noise reduction and wake velocity profiles. Furthermore, Klausmann and
Ruck [29] showed that the brass tubes passed through a porous medium are not subject to additional
flow disturbances. The brass tubes are connected to the pressure scanner ports by polyurethane
tubes with an inner and outer diameter of 0.8 mm and 4 mm, respectively. Two 16-channel
Model 9116 Intelligent Pressure Scanners with a full-scale pressure range from 2.5 kPa to 5200
kPa were used to measure the static pressure. The pressure scanners are capable of accuracies up
to ±0.05%. In this study, the cylinder was rotated three times by θ = 8◦ to provide data at 45
peripheral locations.
Near-field pressure was measured using remote-sensing microphone probes via the same pressure
taps as in the steady pressure measurement, as shown in Fig. 2(c). Each remote sensor consists of
a microphone-holder assembly and a long termination tube to eliminate the spurious tones related
to the standing waves in the flexible tubing [19, 42, 43]. In the microphone-holder, a Panasonic
microphone (series WM-61A) is attached to a pinhole with a diameter of 0.4 mm. The transducer
has a diameter of 6 mm, a height of 3.4 mm, and a circular sensing hole diameter of 2 mm. The
same type of microphone was previously used by many [4–6, 19, 42–47] which has been shown
to produce reliable pressure data over the frequency range of interest (100 Hz ≤ f ≤ 10 kHz). To
account for any possible amplitude damping of the pressure fluctuations and phase delay due to the
remote-sensing setup, all remote-sensing microphone probes were calibrated by a setup similar to
that published by Maryami et al. [4]. A reference 1/2 inch Brüel Kjær (B&K) free-field microphone
4966 and each remote-sensor were subjected to a white noise signal to produce the transfer function
based on the detailed calibration approach described in Ref. [48]. For the current remote-sensing
setup, the coherence between the reference and remote-sensing microphone probes was near 1 up
to 6 kHz (St ≈ 7.8 at U∞ = 50 m/s) and the frequency response showed a smooth trend within

5
this range of frequency, which was consistent with other investigations [49, 50]. Note that the
pressure fluctuations in the outer diameter of the porous coating are approximated by the pressure
measured from the brass pressure taps.

2.4. Acoustic measurements and instrumentation


The far-field noise was measured using a B&K microphone, as depicted in Fig. 1. The
microphone was placed at an angle of 90◦ to the flow and a distance of y = −1.5 m away
from the cylinder axis. The far-field microphone was calibrated using B&K calibrator 4231 before
far-field measurements. The near- and far-field noise measurements were acquired simultaneously
at a sampling frequency of 51.2 kHz for 10 seconds using two 24-bit synchronized NI PXI-4496 data
acquisition cards mounted in a National Instruments PXI-10420 chassis. The power spectral density
(PSD) was calculated using the pwelch function and Hamming windowing with 50% overlap is
applied in the data postprocessing. To reduce the statistical convergence error, the pressure spectra
with a frequency resolution of 2 Hz were calculated based on the average spectra of individual data
obtained from dividing the time series pressure data into a sequence of blocks. These data were
then converted into the near-field pressure and acoustic PSD, ϕpp,s and ϕpp (dB/Hz), respectively
using a reference pressure of p0 = 20 µPa.

3. Results

3.1. Aerodynamic characteristics


The distribution of the mean pressure coefficient (Cp ) and the root-mean-square (RMS) of
the instantaneous pressure coefficients (Cp,rms ) are presented in Figs. 3(a) and 3(b), respectively
for Reynolds numbers within Re = 0.52 × 105 –1.74 × 105 . The peripheral location of the pressure
coefficient minimum, the boundary layer separation, and the starting location of the base region are
denoted by θm , θs , and θb for the bare cylinder and θm,s , θs,s , and θb,s for the SPCC, respectively.
For all Reynolds numbers in Fig. 3(a), it is observed that the absolute value of Cp in the case
of the SPCC is smaller than that of the bare cylinder beyond θ ≈ 32◦ . In the favorable pressure
gradient region, i.e., from the stagnation point (θ = 0◦ ) to θm and θm,s , the minimum Cp for the
bare cylinder occurs at θm ≈ 65◦ yet it shifts to θm,s ≈ 81◦ for the SPCC. From θm and θm,s
up to θb ≈ 90◦ and θb,s ≈ 122◦ , i.e., in the adverse pressure gradient region, the distribution of
Cp for both cylinders is followed by a region where the boundary layer separation occurs [51–53].
The SPCC shows a longer angular extent in this region (∆θbm,s = θb,s − θm,s ≈ 41◦ ) compared
to the bare cylinder (∆θbm = θb − θm ≈ 25◦ ), which indicates a delay in the boundary layer
separation from the SPCC surface as will be further discussed in Fig. 3(b). In the base region, i.e.,
within θ ≈ 90◦ –270◦ and ditto θ ≈ 122◦ –245◦ at the leeward side of the bare cylinder and SPCC,
respectively, an almost constant negative Cp is observed for both cylinders yet the SPCC reveals
a shorter base region compared to the bare cylinder. It is also observed that the SPCC decreases
the base pressure from the bare cylinder, which indicates that the drag imposed to the SPCC is
smaller in magnitude than that of the bare cylinder.
By observation in Fig. 3(b), the Cp,rms -distribution is slightly asymmetric which may be
explained by the asymmetric transition to turbulence on either side of the cylinder, as observed
via experiment by others in the Reynolds number range [54–56]. In general, the SPCC produces
smaller Cp,rms compared to the bare cylinder at all Reynolds numbers considered, which indicates
that porous coating reduces the energy content of the surface pressure fluctuations. It is observed

6
Figure 3: (a) Mean and (b) RMS pressure coefficients, Cp and Cp,rms , respectively, for Reynolds number within
Re = 0.52 × 105 − 1.74 × 105 . The Bare cylinder and SPCC are denoted by the circle and square symbols, respectively.

that the Cp,rms is zero in magnitude at θ = 0◦ and then gradually increases to a maximum at
θs ≈ 73◦ and θs,s ≈ 98◦ for the bare cylinder and SPCC, receptively. Since the separation angle
can be determined at the location where Cp,rms has its maximum [51, 57], it can be concluded that
the SPCC causes a delay in the boundary layer separation, consistent with experimental findings
in Ref. [13]. In Fig. 3(b), it can be seen that Cp,rms gradually decreases beyond θs and θs,s and
reaches a minimum near θb and θb,s . In the base region, Cp,rms exhibits overall constant values for
both cylinders.
In terms of the variation of Cp and Cp,rms with respect to the Reynolds number, a significant
reduction in magnitude appears for the bare cylinder from Re = 0.52 ×105 to 1.04 ×105 specifically
in the base region, while a little change can be seen between Re = 1.04 × 105 and 1.74 × 105 , which
is consistent with the results in the literature [58]. As the range of Reynolds numbers considered in
this study can be matched to either the subcritical regime [51, 53, 59, 60] or both subcritical and
critical regimes [59], the variation of Cp and Cp,rms with Reynolds number for the bare cylinder may
be due to the different flow regimes. For the SPCC, Cp decreases gradually from Re = 0.52 × 105
and 1.74 × 105 yet Cp,rms does not show significant variation with Reynolds number.

3.2. Far-field noise spectra


Figure 4 presents the acoustic PSD as a function of Strouhal number (St = f D/U∞ ) for
Reynolds numbers Re = 0.52 × 105 , 1.04 × 105 , 1.39 × 105 , and 1.74 × 105 . At Re = 0.52 × 105 , the
fundamental vortex shedding tone of the bare cylinder (f1 -tone) is observed at St ≈ 0.19, which
lies within the published Strouhal number envelope for that Reynolds number [38]. In addition to
the f1 -tone, the second and third harmonics, i.e., f2 = 2f1 , and f3 = 3f1 , respectively, are also
visible yet the bare cylinder does not show the tonal harmonics at other Reynolds numbers tested.
As expected, the SPCC presents a significant reduction in far-field noise produced by the bare
cylinder yet it decreases gradually by increasing Reynolds number. It is observed that the SPCC
suppresses the tonal noises completely and it is more pronounced with Reynolds number. At
Re = 0.52 × 105 , however, the weak f1 - and f2 -tones can be seen, which is consistent with other
studies [25, 61]. The f1 -tone occurs at a slightly lower Strouhal number compared to that of the
bare cylinder, which is consistent with many investigations [22, 25, 61]. This Strouhal number shift

7
Figure 4: Acoustic PSD obtatined from the far-field microphone. (a) Re = 0.52 × 105 , (b) Re = 1.04 × 105 , (c)
Re = 1.39 × 105 , and (d) Re = 1.74 × 105 .

can be regarded as evidence of a delay in the separation of the boundary layer as earlier seen in
Fig. 3. Furthermore, it suggests that the SPCC produces a thicker boundary layer and wider wake
region than the bare cylinder, such that the typical vortex formation region is likely to emerge
further downstream [23]. By observation in Fig. 4 (a), the f2 -tone has a magnitude close to the
f1 magnitude, which indicates that the drag fluctuations produced at Re = 0.52 × 105 have high
energy content compared to other Reynolds numbers as will be further shown in Fig. 7(b).
The SPCC exhibits a strong noise at higher Strouhal numbers, centered about St ≈ 8, which
is herein referred to as the high frequency (fHF ) noise. It is observed that the amplitude and
the fHF -noise increases with Reynolds number. Because of the regular structure and pore size of
the SPCC, it seems that the fHF -noise is partly due to the small-scale instabilities generated in
separated boundary layers immediately downstream of each structural member of the SPCC [25].
The results in Fig. 8 will clearly show that the small-scale instabilities produce a non-propagating
hydrodynamic field and are different in terms of the amplitude in the windward and leeward sides
of the cylinder.

3.3. Near-field pressure spectra


To gain further insights into the noise generation mechanism of the bare cylinder and SPCC,
the near-field pressure PSD was calculated at several peripheral locations. For brevity, only some
locations (i.e., θ = 0◦ , 40.9◦ , 90◦ , 130.9◦ , and 180◦ ) are presented in Fig. 5 at Re = 0.52 × 105 ,

8
Figure 5: Near-field pressure PSD measured using the remote-sensing microphone probes at several angles.(a1 -a5 )
Re = 0.54 × 105 , (b1 -b5 ) Re = 1.04 × 105 , (c1 -c5 ) Re = 1.39 × 105 , and (d1 -d5 ) Re = 1.74 × 105 .

1.04×105 , 1.39×105 , and 1.74×105 . From Figs. 5(a1 ), 5(b1 ), 5(c1 ), and 5(d1 ), i.e., at θ = 0◦ , it can
be seen that the bare cylinder typically displays the broadband spectral content, which increases
in magnitude with increasing Reynolds number. At θ = 0◦ , the weak yet non-negligible f1 - and
f2 -tones are observed and they do not show significant variation in amplitude from Re = 0.52 × 105
to 1.74×105 . Away from the stagnation point at θ = 40.9◦ , the near-field pressure PSD indicates the

9
emergence of the f1 - and f3 -tones at Re = 0.52×105 as shown in Fig. 5(a2 ). By observation in Figs.
5(a2 ), 5(b2 ), 5(c2 ), and 5(d2 ), the f1 -tone increases in magnitude with Reynolds number, whereas
the f3 -tone is seen to disappear completely and the f2 -tone becomes visible instead. Further away
from the stagnation point at θ = 90◦ and 130.9◦ , only the f1 -tone is observed with a similar trend
to θ = 40.9◦ in terms of the f1 magnitude variations with respect to the Reynolds number. In the
base region at θ = 180◦ as shown in Figs. 5(a5 ), 5(b5 ), 5(c5 ), and 5(d5 ), the f1 -tone disappears
completely and the f2 -tone remains the only noticeable tonal peak, which is consistent with other
research [1, 4, 5]. This finding shows that tonal peaks are generated by separate mechanisms as
supported by many [1, 5, 6, 62], who showed that the f1 - and f2 -tones are associated with lift and
drag fluctuations of the bare cylinder, respectively, as will be further discussed in Sec. 3.4.
In general, the SPCC exhibits a significant reduction in tonal and broadband energy contents
for all Reynolds numbers considered. However, on the windward side of the cylinder specifically at
θ = 40.9◦ , the strong broadband contribution appears for St ⪆ 0.4, which can be ascribed to some
flow transition processes occurring at this location. At Re = 0.52×105 , the SPCC possesses the f1 -
and f2 -tones yet at other Reynolds numbers the SPCC has generally a broadband behaviour and
the tonal peaks are not observed for all peripheral angles. The SPCC exhibits the emergence of the
fHF contributions at St ≈ 8, consistent with observations in Fig. 4. At Re = 0.52 × 105 , this peak
is predominantly discernible at all peripheral locations, except θ = 90◦ . At the crossing angle of
θ = 90◦ , the fHF -peak disappears completely and shows broadband behaviour. At other Reynolds
numbers, the fHF characteristics are very similar to those of Re = 0.52 × 105 . An exception to
this similarity occurs at θ = 90◦ , as shown in Figs. 5(b3 ), 5(c3 ), and 5(d3 ), where a weak fHF -peak
would tend to appear. With increasing Reynolds number, the fHF -peak magnitude increases and
the broadband contribution exceeds the bare cylinder broadband energy content at high Reynolds
numbers.
In order to better understand peripheral variation of the surface pressure PSD, the magnitude
of the f1 - and f2 -tones are presented in Figs. 6(a) and Figs. 6(b), respectively for the presented
range of Reynolds number (Re = 0.52 × 105 − 1.74 × 105 ). At Re = 0.52 × 105 , in Fig. 6(a),
the f1 -tone of the bare cylinder has a magnitude near 75 dB at θ = 0◦ and increases quickly to
about 101 dB at θ = 16◦ . The f1 -tone follows a gradually increasing trend beyond θ = 16◦ to
reach a maximum around 117 dB at θ ≈ 90◦ , and then decreases for θ ⪆ 90◦ until it becomes
broadband completely at θ = 180◦ as previously shown in Fig. 5. At other Reynolds numbers,
the f1 -tone shows similar directivity patterns to Re = 0.52 × 105 . Significant differences are
related to the f1 magnitude, which gradually increases with the Reynolds number. For the SPCC,
a similar trend to the bare cylinder is observed approximately in terms of the variation of the f1
magnitude with respect to the peripheral angle and Reynolds number. However, the SPCC shows
within Re ≈ 0.52 × 105 –1.04 × 105 that the f1 -tone increases gradually in magnitude to achieve
maximum at θ ≈ 98◦ , slightly away from the location of the maximum f1 magnitude observed for
the bare cylinder (θ ≈ 90◦ ), and then decreases between θ ≈ 98◦ –180◦ . For Re ⪆ 1.04 × 105 , the f1
magnitude increases by around 15 dB between θ = 0◦ and θ ≈ 40.9◦ and remains nearly constant
within θ ≈ 40.9◦ –180◦ . By inspecting the overall trend of the f1 -tone around the bare cylinder
and to some extent SPCC, it can also be found that the directivity pattern of the f1 -tone has a
dipole shape, perpendicular to the flow direction, i.e., dominated by the lift fluctuations, but with
considerable pressure fluctuations at θ = 0◦ and 180◦ due to unsteady drag contribution [5, 6].
It is observed in Fig. 6(b) that the spectral content at the f2 -tone frequency increases uniformly
between θ = 0◦ and 180◦ for both bare cylinder and SPCC. This indicates that the f1 -tone reaches

10
Figure 6: Directivity pattern of the near-field pressure PSD, 10 log10 (ϕpp /p20 ) (dB/Hz), at (a) the f1 -tone, (b) the
f2 -tone, and (c) the fHF -peak frequencies. (d) The overall surface pressure level, OSPL (dB), of the bare cylinder
and SPCC for Reynolds numbers ranging from Re = 0.52 × 105 to Re = 1.74 × 105 . The Bare cylinder and SPCC
are denoted by the dashdoted and dashed lines, respectively.

its peak value near the boundary layer separation location, while the f2 -tone continues to grow into
the fully separated flow and peaks at θ = 180◦ . These observations are consistent with the general
understanding of the unsteady aerodynamic loads acting on the bare cylinders [1, 5, 6]. The results
in Fig. 6(b) also show that the f2 magnitude increases with Reynolds number for the bare cylinder
and SPCC. In general, the SPCC exhibits a lower amplitude for the tonal peaks compared to the
bare cylinder over the whole peripheral locations. The application of the SPCC results on average
a 38 − 41 dB decrease in the f1 magnitude at θ = 90◦ . For the f2 -tone, the difference in magnitude
between the SPCC and the bare cylinder becomes more evident at θ = 180◦ about 20 − 47 dB.
For the fHF -peak as shown in Fig. 6(c), the magnitude increases away from θ = 0◦ and peaks
at θ ≈ 40.9◦ . The fHF magnitude then tends to decrease within 14–35 dB, depending on the
Reynolds number, from θ ≈ 40.9◦ to 180◦ on the top half side of the SPCC. Note that the the
fHF -peak becomes fully broadband at θ = 90◦ , where near the boundary layer separation point, for
low Reynolds numbers (Re = 0.52 × 105 –0.86 × 105 ) as observed earlier in Fig. 5(c). It is observed
that the SPCC produces the fHF -peak with a magnitude greater than that of the bare cylinder

11
PSD on the windward side of the cylinder; the opposite tends to appear on the leeward side of the
cylinder specifically at high Reynolds numbers. Furthermore, the fHF -peak increases in magnitude
with Reynolds number.
The overall surface pressure levels (OSPLs) are presented in Fig. 6(c) by integrating the surface
pressure PSD over the frequency range of interest. It is evident for all Reynolds numbers tested
that the SPCC does not influence significantly the OSPL directivity shape of the bare cylinder.
The OSPL of the bare cylinder indicates the emergence of a maximum value around θ = 90◦ ,
which is consistent with the f1 -tone distribution observed in Fig. 6(a). However, the high OSPL
of the SPCC is localised around θ = 40.9◦ where a high broadband energy content was previously
observed in Figs. 5(a2 ), 5(b2 ), 5(c2 ), and 5(d2 ). The OSPL of the SPCC case is lower than that of
the bare cylinder over all peripheral angles. It can be seen that the minimum difference between
the bare cylinder and SPCC in terms of the OSPL occurs at θ = 0◦ by 4 − 7 dB and increases up to
a maximum of 20 − 34 dB at θ = 180◦ . The results in Fig. 6(c), moreover, show that the OSPL is
sensitive to the Reynolds number and tends to increase gradually for the bare cylinder and SPCC
on average within 2 − 10 dB and 15 − 19 dB, respectively from Re = 0.52 × 105 to 1.74 × 105 .

3.4. Lift and drag spectra


Figures 7(a) and 7(b) show the spectra of the lift and drag coefficients, i.e., ϕC C and ϕC C ,
L L D D
respectively, for all Reynolds numbers tested within Re = 0.52 × 105 − 1.74 × 105 . The unsteady lift
and drag were estimated by integrating the unsteady surface pressure around the cylinders. The
contributions of skin friction and the forces due to the loss of total pressures that may be present
around the cylinders’ surface were neglected in the calculation of the lift and drag PSD profile
[51, 52, 63]. For the bare cylinder and all flow condition cases, the ϕC C and ϕC C present
L L D D
the dominant peaks at the f1 - and f2 - tone frequencies, respectively. The analysis of the ϕC C
D D
against ϕC C shows that the dominant tone frequency of the unsteady drag is twice that of the
L L
unsteady lift, which is consistent with many studies [1, 5, 6, 62, 64]. However, the amplitude of
the lift fluctuations is often larger than that of the drag fluctuations. Interestingly, a comparison
between the surface pressure PSD and the lift and drag coefficient spectra exhibits that the tonal

Figure 7: (a) Lift coefficient PSD, ϕC C , and (b) drag coefficient PSD, ϕC C , as a function of St for Re =
L L D D
0.52 × 105 –1.74 × 105 . The Bare cylinder and SPCC are denoted by the dashdoted and dashed lines, respectively.

12
behaviours observed for the bare cylinder at θ = 90◦ (see Figs. 5(a3 ), 5(b3 ), 5(c3 ), and 5(d3 )) and
180◦ (see Figs. 5(a5 ), 5(b5 ), 5(c5 ), and 5(d5 )) are similar to those of ϕC C and ϕC C , respectively.
L L D D
These observations are consistent with previous findings in the literature [1, 5, 6], i.e., the f1 -
and f2 -tones appeared in the near-field pressure spectra correspond to the vertical oscillations of
the cylinder due to the unsteady lift and axial oscillations due to the drag fluctuations, respectively.
The dominant peaks of ϕC C and ϕC C , i.e., the f1 - and f2 - tones, respectively, are also observed
L L D D
for the SPCC similar to the bare cylinder yet f2 -tone emerges only up to Re = 1.04 × 105 as shown
in Fig. 7(b). The ϕC C is purely broadband for Re > 1.04 × 105 and this adds further evidence
D D
that the tonal noise observed at the f2 -tone frequency in Fig. 4(a) is produced by drag fluctuations.
By observation in Fig. 7, each Reynolds number reveals that the SPCC causes a reduction in the
energy content of lift and drag fluctuations produced by the bare cylinder yet there is a noticeable
exception at the high frequencies around the fHF -peak frequency. It is observed for the bare
cylinder and SPCC that the ϕC C and ϕC C decrease in magnitude with Reynolds number and
L L D D
is more pronounced within Re = 0.52 × 105 –1.04 × 105 for the bare cylinder.

3.5. Coherence analysis


3.5.1. Near-to far-field coherence
To further understand the far-field propagation effects of the identified surface pressure signals,
the magnitude-square coherence, γp2 ,p , between the surface pressure and far-field signals is
i j
calculated using the following equation,

|ϕpi pj (f )|2
γp2 ,p = (1)
i j ϕpi pi (f )ϕpj pj (f )
′ ′
where pi is the reference surface pressure sensor and pj is the far-field acoustic signal. The ϕpi pj
′ ′
denotes the cross-spectrum between the pressure signals pi and pj and ϕpi pi is the auto-spectrum
of each individual signal.
The γp2 ,p of the bare cylinder and SPCC at selected angular positions (θ = 0◦ , 40.9◦ , 90◦ ,
i j
130.9◦ , and 180◦ ) is presented in Figs. 8(a), 8(c), 8(e), and 8(g), respectively. To show the
effects of the SPCC on γp2 ,p , the changes to γp2 ,p , i.e., ∆γp2 ,p = γp2 ,p |SPCC − γp2 ,p |Bare , for all
i j i j i j i j i j
remote-sensing sensors distributed between θ = 0◦ and 180◦ are presented in Figs. 8(b), 8(d), 8(f),
and 8(h), respectively.
It can be seen at Re = 0.52 × 105 that the high coherence level for the bare cylinder occurs
at the f1 -tone frequency and all peripheral angles, except θ = 180◦ . The f1 magnitude is almost
zero at θ = 180◦ , which is consistent with observations in Fig. 5 and the results in Refs. [4, 5]. It
can therefore be explained that the surface pressure fluctuations imposed by vortex shedding at
the f1 -tone frequency have a further contribution to the tonal noise. From Fig. 8(a), it is observed
that the f1 -tone magnitude increases quickly to around 0.7 at θ = 40.9◦ . While the f1 -tone
magnitude decreases to around 0.6 up to θ = 90◦ , it remains constant within θ ≈ 90◦ − 130.9◦
and then goes down to zero at θ = 180◦ . This trend is consistent with previous studies [19],
which demonstrated that the surface pressure fluctuations imposed by vortex shedding mainly in
the pre-and post-separation regions have a further contribution to the tonal noise. For the SPCC,
a weak coherence around 0.2 is observed at the f1 -tone frequency. In addition to the distinct
f1 -tones, the SPCC possesses the f2 -tone regardless of the pressure port locations, whereas there

13
Figure 8: (a), (c), (e), (g) Coherence between the surface pressure fluctuations and far-field noise, γp2i ,pj , where the
bare cylinder and SPCC results are depicted by lines with and without symbols, respectively and (b), (d), (f ), (h)
changes in the near-to far-field coherence, ∆γp2i ,pj , for selected Reynolds numbers. (a, b) Re = 0.54 × 105 , (c, d)
Re = 1.04 × 105 , (e, f ) Re = 1.39 × 105 , and (g, h) Re = 1.74 × 105 .

14
is no significant f2 -tone for the bare cylinder. These results suggest that the hydrodynamic field
responsible for the generation of the f2 -tone in the SPCC case, propagates further from the cylinder
and produces far-field noise as already observed in Fig. 4(a). For the SPCC around θ = 40.9◦ ,
there is no sign of distinct f2 -tone. At θ = 0◦ , 130.9◦ , and 180◦ , the SPCC also shows that the
f1 -tone is smaller in magnitude compared to that of the f2 -tone, indicating that the propagation
of the f2 -hydrodynamic pressure field is likely due to the drag fluctuations.
The results in Fig. 8(b) indicate that the use of SPCC can significantly reduce the coherence
magnitude at the f1 -tone frequency over the whole angles, which is consistent with the low energy
content of the pressure fluctuations observed in Fig. 5. Since the vortex shedding is significantly
attenuated by the PCC [21, 23, 31], it can be explained that the hydrodynamic pressure field
generated by the SPCC dissipates rapidly so that can not propagate to far-field region. At the
f2 -tone frequency, however, the presence of the SPCC increases γp2 ,p over the angle range, except
i j
θ ≈ 32◦ –73◦ where ∆γp2 ,p ≈ 0. For other Reynolds numbers, as shown in Figs. 8(c), 8(e), and 8(g),
i j
similar behaviours as Re = 0.52 × 105 are observed with some notable exceptions. The coherence
at the f1 -tone frequency increases in magnitude up to γp2 ,p ≈ 0.8 with Reynolds number in the
i j
pre-and post-separation regions (θ = 40.9◦ , 90◦ , and 130.9◦ ); the opposite is observed at θ = 0◦
and 180◦ . Moreover, negligible γp2 ,p -values exists for the SPCC over the frequency range. It can
i j
be seen in Figs. 8(d), 8(f), and 8(h) that the coherence level of the SPCC at the f2 -tone frequency
is approximately the same as the bare cylinder, i.e., ∆γp2 ,p ≈ 0.
i j
While both far-field and surface pressure PSD measurements, respectively, in Figs. 4 and 5,
indicate the emergence of the fHF -peak, it can be observed that the coherence level at the fHF -peak
frequency is nearly eliminated for all Reynolds numbers with γp2 ,p < 0.1. Therefore, it can be
i j
explained that fHF -peak is generated by complex interaction between the porous structure and the
free stream flow rather than a propagating hydrodynamic field.

3.5.2. Peripheral coherence


In order to understand the evolution of near-field pressure over the cylinder from θ = 0◦ to

180 , the peripheral coherence (γθ2 ,θ ) between the surface pressure signals was calculated. For
1 2
brevity, only some important peripheral coherence at Re = 0.52 × 105 and 1.74 × 105 between
pressure taps on the top half side of the cylinders are presented in Fig. 9 and 10, respectively. The
γθ2 ,θ was calculated as a function of angular distance (∆θ) between the reference pressure taps at
1 2
θ1 = 0◦ , 40.9◦ , 90◦ , 130.9◦ , and 180◦ and the other peripheral taps (θ2 ).
For the bare cylinder as shown in Figs. 9(a)–9(e), it can be seen that the coherence with
respect to the signals measured at θ = 0◦ (γ02,θ ) is purely tonal at the f1 - and f2 -tone frequencies
2
and is nearly zero at all other frequencies, regardless of the coherence results for ∆θ = 0◦ where
the fully coherence is observed (γθ2 ,θ = 1). The γ02,θ presents the f2 -tone only on the windward
1 2 2
side of the cylinder for ∆θ < 45. The coherence with respect to θ = 180◦ (γ180 2
,θ 2 ) is totally
tonal only at the f2 -tone frequency on the leeward side of the cylinder within ∆θ < −45. Away
from θ = 0◦ and 180◦ , the coherence with respect to θ = 40.9◦ , 90◦ , and 130.9◦ , i.e., pre- and
post-separation locations, shows a broadband spectral content, as well as the f1 -tones which are
broader in frequency. This, of course, makes sense that the surface pressure broadband components
have a low energy hydrodynamic field with local effect, while the vortex shedding tonal components
can propagate upstream as a strong hydrodynamic field and reach even the stagnation point [4, 65].
For the SPCC, γ02,θ has a level approximately zero over the frequency range. In the pre-
2

15
Figure 9: Peripheral coherence, γθ21 ,θ 2 , measured around the circumference of the bare cylinder and SPCC at
Re = 0.52 × 105 . The white horizontal dotted line denotes the angular position of the reference pressure tap. (a-e)
Bare cylinder and (f -j) SPCC.

and post-separation regions, γ40.9 2 2 2


,θ 2 , γ90 ,θ 2 , and γ130.9 ,θ 2 are also decreased in level by the SPCC
from that of the bare cylinder and the f1 -tones are seen to be narrow in frequency. It is more
pronounced for γ40.92
,θ 2 where the broadband spectral content has a level lower than 0.2. As the
surface pressure fluctuations are a result of vortex streets rolling up in the wake [1, 2], it may be
interpreted that the attenuation of the near-field pressure at the f1 -tone frequency is related to
the role of porous coating in suppressing the vortex shedding and moving the vortex formations
region further downstream [23, 31]. By observation in Figs. 9(g)–9(i), the broadband spectral
content becomes stronger for the coherence with respect to the reference pressure taps away from
θ = 40.9◦ , i.e., γ90
2 2
,θ 2 and γ130.9 ,θ 2 . As a shift is observed in the appearance of the strong surface
pressure broadband components in the case of the SPCC compared to the bare cylinder (see
Figs. 9(b)–9(c) vs 9(g)–9(i)), it can be explained that the SPCC boundary layer has a delay
in development and separation, which is consistent with observations in Fig. 3. Interestingly,
the SPCC reveals the f2 -tone for γ90 2 2 2
,θ 2 , γ130.9 ,θ 2 , and γ180 ,θ 2 on the leeward side of the cylinder.
Furthermore, the f2 magnitude increases when the location of the reference pressure tap shifts
from θ1 = 90◦ to 180◦ , i.e., the maximum magnitude appears for γ180 2
,θ 2 . Based on the relationship
between the f2 -tone and drag fluctuations discussed earlier in Fig. 7(b), it can be explained that
the SPCC increases base pressure and drag fluctuations at Re = 0.52 × 105 .
At Re = 1.74 × 105 presented in Fig. 10, the peripheral coherence for the bare cylinder is
similar to that at Re = 0.52 × 105 with some notable exceptions. For example, it is observed that
the f1 -tone is very narrow for γ40.9 2 2 2
,θ 2 , γ90 ,θ 2 , and γ130.9 ,θ 2 . In the case of the SPCC, the tonal peaks
disappear totally and except ∆θ = 0◦ the broadband spectral content is observed only for γ90 2

2

16
Figure 10: Peripheral coherence, γθ21 ,θ 2 , measured around the circumference of the bare cylinder and SPCC at
Re = 1.74 × 105 . The white horizontal dotted line denotes the angular position of the reference pressure tap. (a-e)
Bare cylinder and (f -j) SPCC.

2
and γ130.9 ,θ at St ⪅ 0.2.
2

3.6. Auto-correlation
The temporal characteristics of the flow structures responsible for the most energetic pressure
fluctuations can be acquired from the auto-correlation of the surface pressure fluctuations, defined
as
′ ′
p (t + τ )pi (t)
Rpi ,pi (τ ) = i (2)
p2irms
′ ′
where pi is the surface pressure signal collected at θ, prms is the RMS of the pressure signal, τ
denotes the time delay between the pressure signals and the overbar represents the time averaging.
Figure 11 shows the auto-correlations of the surface pressure fluctuations in terms of the
normalized time delay (τ ∗ = τ U∞ /D). At Re = 0.52 × 105 , the bare cylinder at θ = 0◦ shows
a rapid decay with signs of weak oscillation within −10 < τ ∗ < 10. At θ = 40.9◦ , however, a
slowly decaying periodic behaviour is observed compared to θ = 0◦ . At angles θ = 90◦ and 130.9◦ ,
the Rpi ,pi (τ ∗ ) decay rate increases slightly from that at θ = 40.9◦ and then experiences a rapid
decay at θ = 180◦ , which is consistent with Ref. [4]. The peak-to-peak distance (τ1∗ ) observed
in the Rpi ,pi (τ ∗ ) results for the bare cylinder corresponds to the time delay and vortex shedding
frequency. At θ = 40.9◦ , 90◦ , and 130.9◦ , the time delay is τ1∗ ≈ 5.4, corresponding to the f1 -tone
at St ≈ 0.19 as earlier seen in Fig. 5. At θ = 180◦ , however, the distance between the two peaks is
τ2∗ ≈ 2.8, switching in the periodicity corresponding to the f2 -tone (see Figs. 5(a5 ), 5(b5 ), 5(c5 ), and
5(d5 )). For the SPCC as shown in Fig. 11(a), a very weak periodic shape can be seen compared
to the bare cylinder at θ = 0◦ and 180◦ . The Rpi ,pi (τ ∗ ) curve at θ = 40.9◦ shows a very fast decay
17
Figure 11: Auto-correlation of the surface pressure fluctuations, Rpi ,pi , measured at different angular positions at
(a) Re = 0.52 × 105 , (b) 1.04 × 105 , (c) 1.39 × 105 , and (d) 1.74 × 105 . The bare cylinder and SPCC results are
depicted by lines with and without symbols, respectively.

with a weak periodic behaviour, indicating the emergence of two simultaneous phenomena, namely
an envelope of a very weak periodic hydrodynamic field and a rapid decay in the correlation. At
θ = 40.9◦ , the sharp decay occurs at about τ ∗ = 0, which is associated with the pressure exerted
by the boundary layer structures at the pressure tap location [65]. At θ = 90◦ and 130.9◦ , however,
the Rpi ,pi (τ ∗ ) curves decay over the whole range of τ ∗ and present a lower decay rate compared to
that at θ = 40.9◦ . The periodic behaviours observed at θ = 40.9◦ , 90◦ , and 130.9◦ indicates that
the vortex shedding there still exists, although its hydrodynamic field is significantly weaker than
the bare cylinder and decays faster over the range of τ ∗ .
Finally, the auto-correlation results of the bare cylinder seem to be similar for all Reynolds
numbers. However, the periodic shape of the Rpi ,pi (τ ∗ ) would be weaker with much faster decay
at Re = 1.74 × 105 compared to that of Re = 0.52 × 105 . By increasing the Reynolds number, one
can observe in the case of the SPCC that Rpi ,pi (τ ∗ ) and the hydrodynamic field decay slower over
a larger range of τ ∗ .

3.7. Wavelet transform analysis


Since vortex shedding involves multiscale flow structures that are unevenly spaced and show
variations in their duration, the wavelet transform is a promising tool to identify such intermittent
events. The wavelet transform can be used to analyse the amplitude modulation mechanics
18
Figure 12: Wavelet scalogram, |W |, of a segment of pressure signals measured at Re = 0.52 × 105 . (a-e) Bare cylinder
and (f -j) SPCC.

associated with dominant tones produced by vortex shedding, i.e., the f1 - and f2 -tones, in both time
and frequency domains, which is not possible by the Fast Fourier Transform (FFT). To generate
time-frequency representations of the pressure signals, the continuous wavelet transform (CWT)
method [6, 42, 67–70] has been employed in the present study, defined as
Z +∞  
1 ∗ t−τ
W (a, τ ) = p p(t)ψa,τ dt (3)
Cψ −∞ a

where W (a, τ ) is the CWT of function p(t), a is the scale dilation parameter, τ is the time delay,
ψ(t) is the mother wavelet, a continuous function in both time domain and frequency domain,
ψ ∗ ((t − τ )/a) is the complex conjugate of the dilated and translated ψ(t), and Cψ−0.5 is a constant
that takes the mean value of the ψ(t). Since Morlet is a common mother wavelet [6, 42, 67, 70, 66],
it was chosen to extract the signal information in the present study.
Figure 12 shows the wavelet coefficient (|W |) as function of St and time (t). To keep brevity,
the results are only presented at Re = 0.52 × 105 . For the bare cylinder, at θ = 40.9◦ , 90◦ , and
130.9◦ , the highest level of energy is concentrated around St ≈ 0.19, corresponding to the f1 -tone
frequency. Furthermore, the amplitude modulation in time occurs at the f1 -tone frequency. For the
time range given in Figs. 12(b)–12(d), a high level of energy at the f1 -tone frequency is observed only
within t ≈ 0 − 0.4 and t ≈ 0.8 − 1, which indicates the appearance of large fluctuations and strong
vortex shedding [67]. In the middle range of time (t ≈ 0.4 − 0.8), however, the wavelet coefficient
is low, which is related to the smooth fluctuations with low energy content and weak vortex
shedding [67]. Note that the entire signal strength can not be determined by the wavelet coefficient,
but it is capable to show the most energetic fluctuations in a certain scale and point [67]. While
the wavelet coefficient is well distributed over the frequency range for the pressure signals at
19
θ = 0◦ , the amplitude modulation can not be seen over the whole time range, particularly at the
f1 -tone frequency. At θ = 180◦ , however, the wavelet coefficient presents a significant amplitude
modulation at the f2 -tone frequency with high level of energy occurring in small discrete time
zones only within t ≈ 0 − 0.1, t ≈ 0.2 − 0.4, and t ≈ 0.6 − 0.7. It indicates that vortex shedding
at the f2 -tone frequency behaves differently from that at the f1 -tone frequency over time. As the
dominant peaks of lift and drag fluctuations occurs at the f1 - and f2 -tone frequencies, respectively
(see Fig. 7), it can also be interpreted that the separated shear layers roll up to form vortices that
impact on the rear side of the cylinder at θ = 180◦ to produce drag fluctuations with periodicity
shorter than that of lift fluctuations.
As anticipated, the level of the wavelet coefficient for the SPCC is significantly lower that that
of the bare cylinder. Furthermore, the amplitude modulation and spectral peaks are absent at
the f1 - and f2 -tone frequencies for all peripheral angles over the time range, which indicates the
role of porous coating in suppressing vortex shedding. Therefore, the wavelet transform explicitly
demonstrates that the application of porous coating significantly reduces the tonal peaks, and thus
the noise generated by the bare cylinder by altering the vortex shedding and flow structures in the
wake region.

4. Conclusion

In this study, the correlation between near- and far-field pressure was carried out to
understanding the potential mechanism responsible for noise control of a PCC. A highly
instrumented cylinder with several pressure taps at different peripheral locations was used to
measure pressure fluctuations at the interface between the fluid and an SPCC in synchronisation
with far-field noise.
The SPCC showed a delay in boundary layer separation compared to the bare cylinder. The
results demonstrated that the SPCC totally suppresses the vortex shedding noise and produces
a reduction in the broadband noise of the bare cylinder for all Reynolds numbers yet a strong
fHF -noise appears at higher frequencies. A significant part of this noise reduction was due to the
suppression of the near-field pressure, resulting from the delay in vortex shedding caused by the
SPCC.
Near- to far-field coherence analysis showed the surface pressure fluctuations imposed by
vortex shedding mainly in the pre-and post-separation regions have a further contribution to
the tonal noise. In addition, the SPCC provides a significant reduction of the vortex shedding
peak, which demonstrates that hydrodynamic field pressure generated at the SPCC surface has
low energy content and dissipates rapidly so that can not propagate to far-field region. Unsteady
aerodynamic loads deduced from the distribution of the surface pressure fluctuations exhibited
that the lift and drag fluctuations occur at the fundamental vortex shedding frequency and second
harmonic, respectively and their magnitudes are reduced significantly by applying the SPCC. The
auto-correlation also showed a very weak periodic shape with fast decaying behaviours for the SPCC
compared to the bare cylinder, indicating that the SPCC significantly suppresses the development
of vortex shedding. Based on the wavelet transform, it was shown that vortex shedding at the
fundamental tone frequency has high energy level but it does not appear over the whole time
range, i.e., the amplitude of the spectral peak modulates over time, while this peak was absent for
the SPCC.
As the near-hydrodynamic pressure field varies significantly by the porous material, as
demonstrated in this study, further flow field measurements in synchronisation with surface
20
pressure and acoustic noise can provide more insight into the underlying physics of the flow-porous
interaction phenomena.

Acknowledgments

This research was supported by the National Natural Science Foundation of China (Grant Nos.
12111530102 and 12272163) and the Natural Science Foundation of Shenzhen Municipality (Stable
Support Plan Program, Grant No. 20220814230752003).

References
[1] D. Casalino, M. Jacob, Prediction of aerodynamic sound from circular rods via spanwise statistical modelling,
J. Sound Vib. 262 (4) (2003) 815–844.
[2] C. Norberg, Interaction between freestream turbulence and vortex shedding for a single tube in cross-flow, J.
Wind. Eng. Ind. 23 (1986) 501–514.
[3] Y. Oguma, T. Yamagata, N. Fujisawa, Measurement of sound source distribution around a circular cylinder in a
uniform flow by combined particle image velocimetry and microphone technique, J. Wind. Eng. Ind. 118 (2013)
1–11.
[4] R. Maryami, M. Azarpeyvand, A. Dehghan, A. Afshari, An experimental investigation of the surface pressure
fluctuations for round cylinders, J. Fluids Eng. 141 (6) (2019).
[5] R. Maryami, S. A. Showkat Ali, M. Azarpeyvand, A. Afshari, Turbulent flow interaction with a circular cylinder,
Phys. Fluids 32 (1) (2020) 015105.
[6] R. Maryami, S. A. S. Ali, M. Azarpeyvand, A. A. Dehghan, A. Afshari, The influence of cylinders in tandem
arrangement on unsteady aerodynamic loads, Exp. Therm. Fluid Sci. (2022) 110709.
[7] S.-J. Lee, S.-I. Lee, C.-W. Park, Reducing the drag on a circular cylinder by upstream installation of a small
control rod, Fluid Dyn. Res. 34 (4) (2004) 233.
[8] H. Akilli, B. Sahin, N. F. Tumen, Suppression of vortex shedding of circular cylinder in shallow water by a
splitter plate, Flow Meas. Instrum. 16 (4) (2005) 211–219.
[9] H.-C. Lim, S.-J. Lee, Flow control of a circular cylinder with o-rings, Fluid Dyn. Res. 35 (2) (2004) 107.
[10] S. Kunze, C. Brücker, Control of vortex shedding on a circular cylinder using self-adaptive hairy-flaps, Comptes
Rendus Mécanique 340 (1-2) (2012) 41–56.
[11] H.-C. Lim, S.-J. Lee, Flow control of circular cylinders with longitudinal grooved surfaces, AIAA J. 40 (10)
(2002) 2027–2036.
[12] A. Afshari, M. Azarpeyvand, A. A. Dehghan, M. Szőke, R. Maryami, Trailing-edge flow manipulation using
streamwise finlets, J. Fluid Mech. 870 (2019) 617–650.
[13] T. Sueki, T. Takaishi, M. Ikeda, N. Arai, Application of porous material to reduce aerodynamic sound from
bluff bodies, Fluid Dyn. Res. 42 (1) (2010) 015004.
[14] W.-L. Chen, H. Li, H. Hu, An experimental study on a suction flow control method to reduce the unsteadiness
of the wind loads acting on a circular cylinder, Exp. Fluids 55 (4) (2014) 1–20.
[15] E. Konstantinidis, S. Balabani, M. Yianneskis, The timing of vortex shedding in a cylinder wake imposed by
periodic inflow perturbations, J. Fluid Mech. 543 (2005) 45–55.
[16] G. Artana, R. Sosa, E. Moreau, G. Touchard, Control of the near-wake flow around a circular cylinder with
electrohydrodynamic actuators, Exp. Fluids 35 (6) (2003) 580–588.
[17] A. Glezer, M. Amitay, Synthetic jets, Annu. Rev. Fluid Mech. 34 (1) (2002) 503–529.
[18] K. Roussopoulos, Feedback control of vortex shedding at low Reynolds numbers, J. Fluid Mech. 248 (1993)
267–296.
[19] R. Maryami, E. Arcondoulis, C. Yang, M. Szoke, Z. Xiang, J. Guo, R. Wei, Y. Liu, Application of local blowing
to a structured porous-coated cylinder for flow and noise control, in: 28th AIAA/CEAS Aeroacoustics 2022
Conference, 2022, p. 2921.
[20] C.-H. Bruneau, I. Mortazavi, Numerical modelling and passive flow control using porous media, Comput Fluids
37 (5) (2008) 488–498.
[21] H. Naito, K. Fukagata, Numerical simulation of flow around a circular cylinder having porous surface, Phys.
Fluids 24 (11) (2012) 117102.

21
[22] H. Liu, J. Wei, Z. Qu, Prediction of aerodynamic noise reduction by using open-cell metal foam, J. Sound Vib.
331 (7) (2012) 1483–1497.
[23] C. Xia, Z. Wei, H. Yuan, Q. Li, Z. Yang, Pod analysis of the wake behind a circular cylinder coated with porous
media, J. Vis. 21 (6) (2018) 965–985.
[24] I. Ashtiani Abdi, K. Hooman, M. Khashehchi, A comparison between the separated flow structures near the
wake of a bare and a foam-covered circular cylinder, J. Fluids Eng. 136 (12).
[25] E. J. Arcondoulis, Y. Liu, Z. Li, Y. Yang, Y. Wang, Structured porous material design for passive flow and
noise control of cylinders in uniform flow, Materials 12 (18) (2019) 2905.
[26] T. F. Geyer, Experimental evaluation of cylinder vortex shedding noise reduction using porous material, Exp.
Fluids 61 (7) (2020) 1–21.
[27] E. J. Arcondoulis, T. F. Geyer, Y. Liu, An acoustic investigation of non-uniformly structured porous coated
cylinders in uniform flow, J. Acoust. Soc. Am. 150 (2) (2021) 1231–1242.
[28] G. Ozkan, H. Akilli, B. Sahin, Effect of high porosity screen on the near wake of a circular cylinder, in: EPJ
Web of Conferences, Vol. 45, EDP Sciences, 2013, p. 01071.
[29] K. Klausmann, B. Ruck, Drag reduction of circular cylinders by porous coating on the leeward side, J. Fluid
Mech. 813 (2017) 382–411.
[30] E. J. Arcondoulis, T. F. Geyer, Y. Liu, An investigation of wake flows produced by asymmetrically structured
porous coated cylinders, Phys. Fluids 33 (3) (2021) 037124.
[31] S. Bhattacharyya, A. Singh, Reduction in drag and vortex shedding frequency through porous sheath around a
circular cylinder, Int. J. Numer. Methods Fluids 65 (6) (2011) 683–698.
[32] H. Liu, J. Wei, Z. Qu, The interaction of porous material coating with the near wake of bluff body, J. Fluids
Eng. 136 (2).
[33] T. Von Karman, Uber den mechanismus des flussigkeits-und luftwiderstandes, Phys. Z. (1912) 49–59.
[34] L. Rayleigh, Æolian tones., Mon. Weather Rev. 43 (10) (1915) 511–511.
[35] E. Z. Stowell, A. F. Deming, Vortex noise from rotating cylindrical rods, J. Acoust. Soc. Am. 7 (3) (1936)
190–198.
[36] N. Curle, The influence of solid boundaries upon aerodynamic sound, Proceedings of the Royal Society of
London. Series A. Proc. Math. Phys. 231 (1187) (1955) 505–514.
[37] M. Zdravkovich, Review and classification of various aerodynamic and hydrodynamic means for suppressing
vortex shedding, J. Wind. Eng. Ind. 7 (2) (1981) 145–189.
[38] C. Norberg, Fluctuating lift on a circular cylinder: review and new measurements, J. Fluids Struct. 17 (1)
(2003) 57–96.
[39] M. Zhao, L. Cheng, Finite element analysis of flow control using porous media, Ocean Eng. 37 (14-15) (2010)
1357–1366.
[40] T. Sueki, M. Ikeda, T. Takaishi, Aerodynamic noise reduction using porous materials and their application to
high-speed pantographs, Quarterly Report of RTRI 50 (1) (2009) 26–31.
[41] H. Yuan, C. Xia, Y. Chen, Z. Yang, Flow around a finite circular cylinder coated with porous media, in:
Proceedings of the 8th international colloquium on bluff body aerodynamics and applications, 2016.
[42] H. Kamliya Jawahar, S. A. Showkat Ali, M. Azarpeyvand, Serrated slat cusp for high-lift device noise reduction,
Phys. Fluids 33 (1) (2021) 015107.
[43] S. S. Vemuri, X. Liu, B. Zang, M. Azarpeyvand, On the use of leading-edge serrations for noise control in a
tandem airfoil configuration, Phys. Fluids 32 (7) (2020) 077102.
[44] R. Maryami, S. A. Showkat Ali, M. Azarpeyvand, A. Dehghan, A. Afshari, Experimental study of the unsteady
aerodynamic loading for a tandem cylinder configuration, in: 25th AIAA/CEAS Aeroacoustics Conference,
2019, p. 2742.
[45] R. Maryami, S. Showkat Ali, M. Azarpeyvand, A. Afshari, A. Dehghan, Turbulent flow interaction with a
circular cylinder, in: Proceedings of the 25th AIAA/CEAS Aeroacoustics Conference, American Institute of
Aeronautics and Astronautics Inc. (AIAA), United States, 2019. doi:10.2514/6.2019-2503.
[46] R. Maryami, A. Dehghan, A. Afshari, Experimental investigation of flow induced noise around circular cylinder
by measuring unsteady surface pressures, AJME 51 (2) (2019) 329–346.
[47] R. Maryami, A. A. Dehghan, A. Afshari, Experimental investigation of the turbulence effect of incoming flow
on the unsteady pressure field and the flow noise around circular cylinder, AJME 52 (4) (2018) 923–942.
[48] P. F. Mish, Mean loading and turbulence scale effects on the surface pressure fluctuations occurring on a naca
0015 airfoil immersed in grid generated turbulence, Ph.D. thesis, Virginia Tech. (2001).
[49] M. Gruber, Airfoil noise reduction by edge treatments, Ph.D. thesis, University of Southampton (2012).
[50] A. Garcia-Sagrado, T. Hynes, Stochastic estimation of flow near the trailing edge of a naca0012 airfoil, Exp.

22
Fluids 51 (4) (2011) 1057–1071.
[51] E. Achenbach, Distribution of local pressure and skin friction around a circular cylinder in cross-flow up to
re = 5 × 106 , J. Fluid Mech. 34 (4) (1968) 625–639.
[52] W. Z. Sadeh, D. B. Saharon, Turbulence effect on crossflow around a circular cylinder at subcritical Reynolds
numbers, Tech. rep., NASA (1982).
[53] P. W. Bearman, On vortex shedding from a circular cylinder in the critical reynolds number regime, J. Fluid
Mech. 37 (3) (1969) 577–585.
[54] C. Zhang, S. Moreau, M. Sanjosé, Turbulent flow and noise sources on a circular cylinder in the critical regime,
AIP Advances 9 (8) (2019) 085009.
[55] O. Lehmkuhl, I. Rodrı́guez, R. Borrell, J. Chiva, A. Oliva, Unsteady forces on a circular cylinder at critical
reynolds numbers, Phys. Fluids 26 (12) (2014) 125110.
[56] J. J. Cincotta, G. Jones Jr, R. W. Walker, Aerodynamic forces on a stationary and oscillating circular cylinder
at high Reynolds numbers, Tech. rep. (1969).
[57] G. West, C. Apelt, Measurements of fluctuating pressures and forces on a circular cylinder in the reynolds
number range 104 to 2· 5× 105, J. Fluids Struct. 7 (3) (1993) 227–244.
[58] M. M. Zdravkovich, Flow around circular cylinders: Volume 2: Applications, Vol. 2, Oxford university press,
1997.
[59] E. Achenbach, Total and local heat transfer from a smooth circular cylinder in cross-flow at high reynolds
number, International Journal of Heat and Mass Transfer 18 (12) (1975) 1387–1396.
[60] A. Roshko, On the development of turbulent wakes from vortex streets, Tech. rep. (1954).
[61] T. F. Geyer, E. Sarradj, Circular cylinders with soft porous cover for flow noise reduction, Exp. Fluids 57 (3)
(2016) 1–16.
[62] B. Etkin, G. Korbacher, R. T. Keefe, Acoustic radiation from a stationary cylinder in a fluid stream (aeolian
tones), J. Acoust. Soc. Am. 29 (1) (1957) 30–36.
[63] A. Fage, V. Falkner, Further experiments on the flow around a circular cylinder, Tech. rep., HM Stationery
Office (1931).
[64] L. Qin, Development of reduced-order models for lift and drag on oscillating cylinders with higher-order spectral
moments, Ph.D. thesis, Virginia Tech. (2004).
[65] S. A. Showkat Ali, A. Mahdi, C. R. I. Da Silva, Trailing-edge flow and noise control using porous treatments,
J. Fluid Mech. 850 (2018) 83–119.
[66] L. Li, P. Liu, Y. Xing, H. Guo, Time-frequency analysis of acoustic signals from a high-lift configuration with
two wavelet functions, Appl. Acoust. 129 (2018) 155–160.
[67] M. Farge, Wavelet transforms and their applications to turbulence, Annual Review of Fluid Mechanics 24 (1)
(1992) 395–458.
[68] S. Mallat, A wavelet tour of signal processing, Elsevier, 1999.
[69] H. Li, T. Nozaki, Wavelet analysis for the plane turbulent jet: analysis of large eddy structure, JSME Int. J.
Ser. B 38 (4) (1995) 525–531.
[70] D. A. Lysenko, I. S. Ertesvåg, K. E. Rian, Large-eddy simulation of the flow over a circular cylinder at reynolds
number 2× 104, Flow Turbul. Combust. 92 (3) (2014) 673–698.

23

View publication stats

You might also like