ArticlePDF Available

The influence of jet on aerodynamic noise in the pantograph area at different sinking heights

IOP Publishing
Physica Scripta
Authors:

Abstract and Figures

The influence mechanism of jet on aerodynamic noise control in the pantograph region at different sinking heights was numerically studied using an Improved Delayed Detached Eddy Simulation (IDDES) model and the Ffowcs Williams-Hawkings (FW-H) equation. Active flow control was achieved by setting jet slots at the leading edge of the cavity to predict the noise generated by the jet itself. The results showed that in the pantograph region with a sinking height of 500 mm, the shear layer was lifted by the jet, which prevented high-speed turbulence caused by flow separation from entering the cavity. Therefore, the model with jet control device reduced the overall far-field noise on the side of the pantograph by 1.6 ∼ 3.9 dB. Through flow field analysis, for the pantograph region where flow separation occurs in the front of the cavity, the jet needs to lift the shear layer enough to cross the front of the pantograph and prevent flow separation, thereby reducing the aerodynamic noise. For the pantograph region without flow separation in the front of the cavity, the jet may introduce additional noise sources, deteriorating the aerodynamic noise in the pantograph region.
This content is subject to copyright. Terms and conditions apply.
Physica Scripta
PAPER
The influence of jet on aerodynamic noise in the
pantograph area at different sinking heights
To cite this article: Yangyang Cao
et al
2024
Phys. Scr.
99 105228
View the article online for updates and enhancements.
You may also like
Anomaly Detection of Pantograph Based
on Salient Segmentation and Generative
Adversarial Networks
Ye Wang, Wei Quan, Xuemin Lu et al.
-
Drift characteristics and the multi-field
coupling stress mechanism of the
pantograph-catenary arc under low air
pressure
Zhilei Xu, , Guoqiang Gao et al.
-
A design of novel Gudermannian neural
networks for the nonlinear multi-
pantograph delay differential singular
model
Zulqurnain Sabir and Sharifah E Alhazmi
-
This content was downloaded from IP address 182.139.40.100 on 05/09/2024 at 14:54
Phys. Scr. 99 (2024)105228 https://doi.org/10.1088/1402-4896/ad7413
PAPER
The inuence of jet on aerodynamic noise in the pantograph area at
different sinking heights
Yangyang Cao, Jiye Zhang , Jiawei Shi and Yuzhe Ma
State Key Laboratory of Rail Transit Vehicle System, Southwest Jiaotong University, Chengdu 610031, Peoples Republic of China
E-mail: jyzhang@swjtu.edu.cn
Keywords: pantograph, jet ow, cavity, aerodynamic noise control, noise prediction
Abstract
The inuence mechanism of jet on aerodynamic noise control in the pantograph region at different
sinking heights was numerically studied using an Improved Delayed Detached Eddy Simulation
(IDDES)model and the Ffowcs Williams-Hawkings (FW-H)equation. Active ow control was
achieved by setting jet slots at the leading edge of the cavity to predict the noise generated by the jet
itself. The results showed that in the pantograph region with a sinking height of 500 mm, the shear
layer was lifted by the jet, which prevented high-speed turbulence caused by ow separation from
entering the cavity. Therefore, the model with jet control device reduced the overall far-eld noise on
the side of the pantograph by 1.6 3.9 dB. Through ow eld analysis, for the pantograph region
where ow separation occurs in the front of the cavity, the jet needs to lift the shear layer enough to
cross the front of the pantograph and prevent ow separation, thereby reducing the aerodynamic
noise. For the pantograph region without ow separation in the front of the cavity, the jet may
introduce additional noise sources, deteriorating the aerodynamic noise in the pantograph region.
1. Introduction
High-speed railway, being regarded as an important symbol of railway modernization due to its many
advantages such as safety, speed, punctuality, comfort, and environmental protection, has developed rapidly
worldwide. With the continuous increase in train speed and the densication of railway networks, noise control
has become an important issue for the sustained prosperity and development of high-speed trains.
High-speed trains mainly suffer the wheel-rail noise and aerodynamic noise. The former is proportional to
the third power of the trains speed, while the latter is to the sixth. It is generally believed that when a trains speed
exceeds 300 km h
1
, aerodynamic noise will surpass wheel-rail noise and become the dominant external
radiation noise [13].The pantograph, one of the main aerodynamic noise sources, is generally considered as the
strongest local noise of high-speed trains [4,5]. For the aerodynamic noise generated by the pantograph, Tan
et al [6]conducted a predictive analysis of the pantograph ow eld and far-eld noise using Large Eddy
Simulation (LES)and FW-H equation. The results showed that the sound energy in the bottom region of the
pantograph accounted for over 50% of the total energy. Zhao et al [7]found that dipole sources dominate in the
far-eld radiation noise of pantograph, while quadrupole sources can be ignored, and the contribution of
aerodynamic noise at the pantograph base frame, balance arm, and upper and lower arm rods is higher than
other components. Zhang et al [8]used computational uid dynamics (CFD)/FW-H acoustic analogy method
to predict the far-eld aerodynamic noise of pantographs, and the results indicated that the main aerodynamic
noise of pantographs was generated from the panhead and the base frame. Lei et al [9]numerically predicted the
noise radiation of single-arm pantograph at 300 km h
1
, and found that the aerodynamic noise of high-speed
pantograph was caused by vortex shedding. Sun et al [10]used detached eddy simulation (DES)to numerically
simulate high-speed pantographs. It can be seen from the results that there were larger vortices around the head
and base of the pantograph, which affected the uctuating pressure on the pantograph and resulted in
aerodynamic noise. Shi et al [11]studied the suppression technology of aerodynamic noise for Faiveley CX-PG
RECEIVED
27 April 2024
REVISED
12 August 2024
ACCEPTED FOR PUBLICATION
27 August 2024
PUBLISHED
5 September 2024
© 2024 IOP Publishing Ltd. All rights, including for text and data mining, AI training, and similar technologies, are reserved.
pantograph using Delayed Detached Eddy Simulation (DDES)and FW-H equation. It was found that the
strongest radiated noise in the range above 500 Hz came from the head region of the pantograph. Meskine et al
[12]used the FW-H equation to predict and analyze the aerodynamic noise of pantographs based on both
permeable surface method and impermeable surface method. As can be seen from the results, both methods
matched well with the experimental results.
In addition to the noise generated by the pantograph itself, in the pantograph region with a cavity, the
interaction between the cavity and the pantograph also affects the aerodynamic noise in this region. Kim et al
[13]compared the ow characteristics and radiated noise in the pantograph regions with and without cavities.
The research results indicated that the cavity structure reduces the velocity of the uid to the bottom of the
pantograph, which facilitates noise reduction. Yao et al [14]used Large Eddy Simulation (LES)and Acoustic
Finite Element Method (FEM)to analyze the aerodynamic noise characteristics of high-speed trains with
pantographs installed on different roof bases (at surface and concave surface). The analysis showed that the
pantograph with a concave base outperformed the at structure in terms of aerodynamic noise. In the reference
[15], Kim et al mentioned that due to the ow separation in the cavity, signicant pressure uctuations may
occur on the tail wall of the cavity, which may introduce additional noise sources.
Therefore, for the leading-edge ow of the cavity, Saddington et al [16]experimentally installed a spoiler at
the leading edge of the cavity at Mach 0.71 for the reduction of momentum exchange between the free ow and
the cavity, and noise reduction. Kim [17,18]studied two noise-reduction effects of pantograph cabins on high-
speed trains represented by rectangular cavities (round edges and chamfered edges). It can be seen from the
results that this method can be used to signicantly reduce the unsteady ow above the cavity, thereby reducing
radiated noise. The above measures are not well adapted to changes in ow conditions due to their nature of
passive ow control, since geometric shapes or mechanical devices cannot be changed once determined or
installed. It is therefore meaningful, in this regard, to study the noise reduction control technology in the
pantograph area based on active ow control methods.
Yu et al [19]established a SAS model for active control based on the two-equation SST turbulence model,
and analyzed the suppression effect of jet ow on cavity noise. The research showed that the jet raised the shear
layer, and reduced the high pressure on the back wall of the cavity. Ukai et al [20]investigated the effects of
injecting jets at different positions upstream of the cavity, and found that the separation shock oscillation formed
by jet ow at the leading edge of a cavity was signicant. Zhao et al [21]demonstrated that a planar jet can
effectively eliminate the aerodynamic noise caused by the radiation of a series of rods in crossow, however,
current research on jets mainly focuses on controlling high Mach number cavity noise only, and the cavity does
not contain any other objects. There is still a lack of research on lower Mach and other objects inside the cavity.
Therefore, in this paper, this active ow control technology was applied to the pantograph region of high-speed
trains, revealing its impact mechanism on the ow eld and aerodynamic noise in the pantograph region. The
related research is expected to provide new understanding and insights for the innovation of aerodynamic noise
control technology in the pantograph region of high-speed trains.
2. Numerical methods
2.1. IDDES model
To achieve high computational accuracy at reasonable cost, a suitable turbulence model is necessary to simulate
the turbulent ow eld around the pantograph region, as well as the details of ow separation and shedding
vortices in the turbulent ow eld. Large Eddy Simulation (LES)is not widely used due to its computational
efciency though it can simulate pressure uctuations on the surface of objects well. Detached Eddy Simulation
(DES)is a hybrid of Unsteady Reynolds Average NavierStokes (URANS)and Large Eddy Simulation (LES)
methods. In the near-wall region, URANS with the Spalart-Allmaras (S-A)model is used to represent the ow,
which is useful to relax the strong LES mesh constraints close to solid surfaces. Meanwhile LES with a single-
equation model for the Subgrid-scale (SGS)viscosity is used in separated ow regions dominated by large
turbulence scales. To solve the problem of grid-induced separation caused by this method, Spalart et al [22]
improved the DES method into Delayed Detached Eddy Simulation (DDES)which was later found by Shur et al
[23]to possibly lead to logarithmic-layer mismatch. Therefore they introduced a new subgrid length scale that is
related to grid spacing and the wall distance, namely the Improved Delayed Detached Eddy Simulation (IDDES)
which was used in this paper to simulate the ow eld in the pantograph region based on SST k-w. For more
details and research on IDDES model, please refer to references [24,25].
2.2. FW-H equation
As an exact re-arrangement of the N-S equation, the FW-H equation is directly derived from the N-S equation. It
is very difcult to solve this equation directly, therefore its more practical to think about this equation based on
2
Phys. Scr. 99 (2024)105228 Y Cao et al
Light hill acoustic analogy, that is, to think of the right-hand side as the source term. At this point, the equation is
a typical wave equation. A wave operator is introduced on the left side of the equation, and the three terms at the
right side correspond to the monopole source, dipole source distributed on the surface of the object and
quadrupole source in the uid space, respectively. Thus, the far-eld sound pressure is obtained by
superimposing the contributions of monopole sources, dipole sources, and quadrupole sources, as shown in
equation (1).
¢=¢ +¢ +¢ptp tp tp txx x x,,, , 1
TLQ
() () () () ()
where ¢ptx,,
T()
¢ptx,
L
()
and ¢ptx,
Q()
respectively represent the sound pressure generated by the monopole
source, dipole source, and quadrupole source terms.
References [2628]all indicate that the pantograph quadrupole source can be ignored under ow conditions
with low Mach number. In the reference, Zhao et al analyzed the quadrupole and dipole sources of the
pantograph at a speed of 400 km h
1
. It was found that quadrupole sources were mainly distributed in the place
where the inow is split, with the highest energy reaching 1 ×10
7
. Dipole sources are mainly distributed on the
surfaces of various rods and cavities, with the highest energy of 1 ×10
8
.
The research results [7]showed that, at a speed of 400 km h
1
, the dipole source intensity in the
pantograph region is greater compared to that of quadrupole source. From the perspective of calculation
methods, in this paper, the FW-H equation was used to predict the noise in the pantograph region. Limited by
computing resources, we are unable to consider a quadrupole source for volume integration. Therefore, the
inuence of quadrupole sources was ignored considering the relevant research results and current computing
resources.
The numerical simulation in this paper is made based on the wind tunnel model, and the pantograph region
is a stationary rigid object surface, so the monopole source term is 0, and only the noise contribution from the
dipole source needs to be considered. Farassat proposed a space-time integrated solution of the FW-H equation
for subsonic ow conditions suitable for numerical calculations [29]. The formula for the sound pressure
contributed by a dipole source is expressed as follows:
òò
ò
¢=
p- +p
-
-
+p
+-
-
pt c
L
rM SLL
rM S
c
LrMcMcM
rM S
x,1
41 d1
41 d
1
41 d2
Lr
rret
rM
rret
rr r
rret
0222
0
00
2
23
() () ()
()
() ()
r=+ -LPn uu v 3
iijj inn
() ()
ds=- -Ppp 4
ij ij ij
0
() ()
where
M
i
refers to the Mach number in the x
i
direction and
M
r
stands for the Mach number in the observer
direction, the subscript
r
et
of the integral term represents the evaluation of the relevant variable over time,
c
0
denotes the speed of sound,
r
refers to the distance from a source point to the observer, u
n
represents the uid
velocity component normal to the surface,
v
stands for the surface velocity component normal to the surface,
P
ij
is the compressive stress tensor,
s
ij denotes the viscous stress tensor, and p
0
refers to the ambient pressure.
3. CFD model
3.1. Geometry model
The research object of this paper is the CX-PG type active control pantograph installed on the CR400BF high-
speed train in China, which is a typical single-arm single-sliding plate pantograph, and the computational model
is in the scale of 1:1. Figure 1(c)shows the geometric model of the pantograph used for CFD simulation. It can be
seen from the gure that most of the geometric features of the pantograph are preserved and xed in the
uncovered cavity on the roof through insulators. And a jet slot with a width of 15 mm is set at the leading edge of
the pantograph cavity, with the right end of the jet slot 50 mm away from the front wall of the cavity, as shown in
gure 1(d). To explore the inuence of jet on the aerodynamic noise of pantographs at different sinking heights,
six calculation conditions were set up. The non-jet models were Base model 1, Base model 2, and Base model 3,
respectively, with heights h of 350 mm, 500 mm, and 650 mm. The corresponding jet models were Jet model 1,
Jet model 2, and Jet model 3. And the specic dimensions and related details of the pantograph region are shown
in gures 1(a),(b).
3
Phys. Scr. 99 (2024)105228 Y Cao et al
3.2. Computational domain and boundary conditions
The calculation domain is established for the ow eld simulation of the pantograph model to conform to the
actual operation of high-speed trains as much as possible, as shown in gure 2. The dimensions of the calculation
domain along the x, y, and z directions are 35 m, 20 m, and 50 m, respectively. The size of the region is sufcient
to eliminate blocking effects in computational simulations and ensure the full development of the wake. The lack
of computing resources currently in the laboratory makes it difcult to meet the minimum time step and spatial
grid scale requirement for compressible model. As can be seen from the research results in reference [7], for the
pantograph at a speed of 400 Km h
1
, the presence of local high-Mach numbers in individual components does
not have a signicant impact on the results. Due to the above reasons, an incompressible model is still taken as
Figure 1. Pantograph region model.
4
Phys. Scr. 99 (2024)105228 Y Cao et al
the research content in this paper. The entrance of the computational domain is set as the velocity-inlet, with an
inow velocity of 111.11 m s
1
(400 km h
1
). While the outlet of the computational domain is set as the
pressure-outlet, with a gauge pressure of 0. Set symmetrical boundaries on both sides and the top of the
computational domain. The bottom surface is set as a slip wall, and the slip velocity is consistent with the
velocity-inlet. And the other surfaces are set as non-slip walls. For the modeling of planar jets, the surface of the
jet outlet is set as the velocity-inlet, without modeling the interior of the jet slot. The research results in reference
[30]show that this modeling method has no impact on the jet trajectory. In terms of jet velocity, Zhao et al [21]
found that the shielding region formed by the jet can reduce the aerodynamic noise of objects in it. Therefore, to
ensure that the shielding region meets the requirements to the maximum extent, this situation where the jet
velocity is 111.11 m s
1
is mainly studied in this paper.
3.3. Mesh strategy
The computational domain is discretized based on a hybrid mesh strategy of prism layer mesh and trimmed
volume mesh. The surface mesh size of the pantograph was controlled between 15 mm. To accurately simulate
the ow of the boundary layer on the pantograph surface, 15 layers of ne prism layer mesh with an initial height
of 0.01 mm and a growth rate of 1.2 were generated on the pantograph surface. To verify the effectiveness of the
boundary layer, a convergent steady-state solution was obtained through 4000 steps of RANS simulation, with
its Y+value shown in gure 3. The Y+values in the entire pantograph area are all less than 5, as clearly shown in
the gure. As concluded in reference [25], for Y+value less than 5, full Y+wall treatment (A wall treatment
method of STAR-CCM+)should be adopted in numerical simulation calculations for better numerical
calculation results. Meanwhile, multiple renement zones have been established to locally rene the
Figure 2. Computational domain and boundary conditions of pantograph region.
Figure 3. Y+distribution around the pantograph region.
5
Phys. Scr. 99 (2024)105228 Y Cao et al
computational domain volume mesh, resulting in a nal volume mesh number of approximately 100 million.
Refer to gure 4for the specic distribution of the mesh around the pantograph region.
3.4. Solver setup
The discrete ow control equation is solved by using segregate ow solver based on SIMPLE algorithm. The
convection term is discretized by using a mixed second-order upwind and bounded center difference scheme,
while the diffusion term is discretized by using a second-order scheme. The gradient calculation was conducted
based on the second-order hybrid Gaussian LSQ method [25], while the second-order implicit method was used
for time advancement with an acoustic simulation time step of 5 ×10
5
s. The uctuations of the ow eld
within the target frequency range can be fully and accurately analyzed to ensure the stability of the solution.
Meanwhile, as can be seen from the calculation, the convective CFL number is about 0.926, which can fully
ensure the stability of the pantograph region in the simulation experiment [25]. Before ow eld data collection,
the converged steady-state solution obtained through RANS simulation was used as the initial eld to simulate
the unsteady ow eld, which lasted for 4000 steps. After the full development of the transient ow eld for
0.15 s, the FW-H solver started to calculate the far-eld aerodynamic noise of the research object while solving
the ow eld. The transient simulation in this phase lasted for another 0.21 s. Based on the sampling theorem
(=D
f
t
max
1
2), it is expected that up to 10000 Hz of noise components can be analyzed. In this paper, to maintain
high accuracy of noise capture, as each cycle contains 4 sampling points, it is expected that up to 5000 Hz of noise
components can be analyzed. Meanwhile, if the second-order difference scheme is used for ow eld
calculations, make sure that the minimum wavelength contains at least 8 mesh to capture sufciently small
vortices [31]. In this paper, the analysis showed that the highest frequency of aerodynamic noise in the
pantograph area was 5000 Hz, which corresponds to a wavelength of 68 mm. Therefore, the size of the surface
grid should be less than 8.5 mm to fulll the corresponding grid size requirements.
4. Validation of the numerical methods and mesh strategy
4.1. Cavity case
Limited by research funding and experimental resources, ow eld and aerodynamic noise data of a full-size
pantograph are unavailable through wind tunnel tests. In this paper, the research of Kim et al [13]and the
experiment of Plentovich et al [32]were referenced to validate the mesh strategy and numerical method in the
current work through clean cavity(the cavity does not contain other objects). The cavity has a length of
L
1
=366 mm, a width of W
1
=244 mm, and a depth of D
1
=61 mm. The corresponding length-depth ratio is
6:1, and the width-depth ratio is 4:1. The static pressure distribution at the bottom of the cavity was calculated at
the inow velocity of 0.2 Mach. The simulated computational domain and boundary conditions are shown in
gure 5.
To validate the effect of mesh size on numerical simulation, three sets of meshes of different sizes were
obtained by changing the surface mesh size of the cavity and the volume mesh size in the renement zone, as
Figure 4. Mesh distribution around the pantograph region.
6
Phys. Scr. 99 (2024)105228 Y Cao et al
shown in table 1. The number of volume grids is 3.5 million, 5.2million, and 8.3 million, respectively. The mesh
distribution near the cavity is shown in gure 6.
In terms of the pressure coefcient CP(
=
r
-
¥
¥
C
p
PP
12u
2
/
)at the bottom of the cavity, the comparison and
validation based on three sets of mesh calculations and the experimental results of Plentovich et al are shown in
gure 7. The pressure data is measured on the centerline at the bottom of the cavity. As can be seen from the
gure, in the front and middle of the cavity, the calculated values are in good agreement with the experimental
results. And the pressure coefcient shows a trend of uniform distribution approaching 0. While at the rear of
the cavity bottom, the pressure coefcient increases rapidly with the increase of d/L
1
(where d refers to the
distance from the measurement point to the front wall of the cavity). The simulation results are higher than the
experimental values, but showing the same trend as the latter. This situation has not been signicantly improved
even if the mesh is further rened. This difference is more likely caused by the difference between the simulated
region and the actual testing environment. Overall, the distribution trend of pressure coefcient at the bottom of
the cavity is consistent between the simulation results and the experimental results, and the difference between
Figure 5. Computational domain and boundary conditions of cavity.
Figure 6. Layout of cavity grid renement region.
Table 1. Cavity mesh parameters.
mesh1 mesh2 mesh3
Cavity surface 1.8 mm 1.5 mm 1.2 mm
Rregion1(R1)2.1 mm 1.8 mm 1.5 mm
Rregion1(R2)4.2 mm 3.6 mm 3 mm
Rregion1(R3)8.4 mm 7.2 mm 6 mm
Rregion1(R4)16.8 mm 14.4 mm 12 mm
7
Phys. Scr. 99 (2024)105228 Y Cao et al
the two is also within an acceptable range. Therefore, the mesh parameters used in current research for the ow
eld simulation in the pantograph region are reasonable considering both computational accuracy and resource
consumption.
4.2. Pantograph case
To further verify the rationality of the grid size used in transient numerical simulation, the transient ow eld of
the pantograph at an inow velocity of 200 km h
1
was calculated based on the wind tunnel test results of the
CS-PG pantograph in [33]. The calculation domain is shown in gure 8where the center of the pantograph is
15 m away from the velocity inlet and 35 m away from the pressure outlet. The boundary conditions and mesh
strategy are completely consistent with chapter 3.
To collect aerodynamic noise in numerical simulations, a series of measurement points were established.
The distance from the measuring point to the centerline of the pantograph is 7 m, that to the no-slip wall surface
is 3.5 m, and the distance between adjacent measuring points is 0.8 m. Therefore, 11 symmetrical measuring
points were established around the geometric center of the pantograph. Among them, the strength of sound is
dened by sound pressure level, which is dened as follows:
=SPL P P dB20 log 5
0
() ()/
where SPL represents the sound pressure level,
P
is the sound pressure, and
P
0stands for the reference sound
pressure.
The numerical simulation results were obtained based on the same solution setup as in chapter 3, as shown
in table 2. As can be seen, the simulated values are basically consistent with the experimental values. The error
Figure 7. Simulation and test results of the pressure coefcient at the bottom of cavity.
Figure 8. Computational domain and boundary conditions of pantograph.
8
Phys. Scr. 99 (2024)105228 Y Cao et al
between the two may be due to the environment and experimental conditions during the experiment. Overall,
the numerical simulation results meet the requirements, and the errors are within the allowable range.
Therefore, in this paper, the mesh strategy and calculation method showed high accuracy, and the mesh used in
the pantograph region also has high robustness.
5. Results and discussion
5.1. Noise sources
The uctuation of surface pressure (Standard deviation of pressure)in the pantograph region is shown in
gure 9, which determines the effect of jet introduction on the strength and distribution of surface dipole
sources in this region. As can be seen from the gure, high amplitude pressure uctuations were observed in all
base models at the panhead, base frame, insulators, upper and lower arm bars, and the rear surface of the cavity.
After the introduction of the jet, compared with the corresponding base model, in Jet model 1, the pressure
uctuations on the surface of the pantograph base frame, insulators, lower arm bar, and the back of the cavity are
signicantly increased. While in Jet model 2, the pressure uctuations on the surface of the lower arm bar,
insulators, and the cavity are signicantly suppressed, and those in the region of the base frame are signicantly
enhanced. In Jet model 3, the pressure uctuations in the lower arm and insulator surfaces are enhanced, and the
pressure uctuations in the base frame region and the rear wall of the cavity are somewhat reduced. For all Jet
models, the introduction of jet has no signicant effect on the pressure uctuations at the upper arm and
panhead of the pantograph. As can be seen in gure 9, there is no meaningful change in the pressure amplitude at
Figure 9. Surface pressure uctuation in the pantograph region.
Table 2. Comparison between numerical simulation and experimental results. Units:dB(A).
Measure point 1234567891011
Experiment 79.7 80.4 80.1 80.4 80.4 80.7 80.6 80.3 79.9 80 80.2
Simulation 80.8 81.6 81.2 81.7 81.9 81.8 81.9 81.8 81.6 81.2 81.1
Error 1.1 1.2 1.1 1.3 1.5 1.1 1.3 1.5 1.7 1.2 0.9
9
Phys. Scr. 99 (2024)105228 Y Cao et al
these two locations, which indicates that the jet raising the shear layer has no impact on the ow of air in the
upper part of the pantograph.
5.2. Flow eld analysis
To further explore the impact of introduced jets on the pressure uctuations in the pantograph region, in this
section, the impact of introducing a jet into the pantograph region on the wake ow eld of the pantograph is
further analyzed.
Figure 10 shows the instantaneous vorticity distribution around the pantograph region in the base model
and jet model at three different sinking heights. In the gure, three important vortex triggering sources are
conrmed in the pantograph region, namely, the panhead, knuckle, and base frame. In panhead section, there is
almost no effect on its vorticity after the introduce of the jet. Although the jet raises the shear layer, it affects the
middle and lower regions of the pantograph, but has no signicant impact on the position of the panhead.
Therefore, the analysis of the ow eld in the pantograph region is focused on the middle and lower part of the
pantograph region.
In Base model 1, the ow separation occurred at the leading edge of the cavity, and a portion of gas entered
the cavity from the front, and formed shedding vortex which interacted with the front of the cavity and the
insulator at the bottom of the pantograph before reattaching and moving downstream. Another part of the gas
formed shedding vortex after interacting with the pantograph base, which continued to move downstream and
interact with the lower arm bar and knuckle of the pantograph. Different from Base model 1, in Jet model 1, the
introduced jet raises the shear layer and intensies the shedding vortex formed after the impact of gas on the base
frame, resulting in the increase of pressure uctuations at the pantograph base frame and lower arm. At the same
time, the obstruction at the front of the pantograph prevented the high-speed shedding vortex formed by the jet
from crossing the base frame, instead, it moved towards the bottom of the base frame, causing the high-speed
shedding vortex to impact the front of the cavity. The shedding vortex reattached and moved downstream after
the impact. In the downstream, these shedding vortex structures interacted with the bottom insulator and cavity,
signicantly enhancing their pressure uctuations. The increase in pressure uctuations in these sections can be
clearly seen in gure 9.
In Base model 2, the ow separation occurred at the leading edge of the cavity, and a portion of the gas
entered the cavity from the front and formed shedding vortex. These shedding vortex structures rolled up and
moved downstream after hitting the front of the cavity, and then interacted with the bottom insulator of the
pantograph and the cavity. The other portion of the gas formed shedding vortex by impacting the base frame,
which continued to move downstream and interact with the lower arm and knuckle. In Jet model 2, the jet did
Figure 10. Distribution of instantaneous vorticity around the pantograph region.
10
Phys. Scr. 99 (2024)105228 Y Cao et al
not make a movement towards the bottom of the base frame, instead, it impacted the front of the
pantograph base frame. After the impact, a small portion of the shedding vortex formed by the jet crossed the
base frame and continued to move downstream. Compared to Base Model 2, the impact of the associated vortex
structures on the pantograph region was signicantly weakened, and the formed wake was also signicantly
narrowed. Considering the results of the pressure uctuations shown in gure 9, the pressure uctuations in the
pantograph region in Jet model 2 are signicantly reduced.
In Base model 3, due to the fact that the height of the base frame is already lower than the cavity, the gas did
not undergo ow separation at the leading edge of the cavity. Instead, it impacted the front of the base frame, and
a small portion of the shedding vortex structure formed after the impact, and continued to move downstream
across the base frame and interacted with the pantograph lower arm bar. In Jet model 3, the jet did not interact
with the front of the pantograph region, but crossed the entire base frame region and continued to move
downstream. In downstream, the high-speed shedding vortex formed by the jet interacted with the lower arm of
the pantograph. A portion of the shedding vortex formed after the interaction continued to move downstream,
resulting in increased pressure uctuations at the knuckle. The other part of the shedding vortex moved towards
the bottom of the pantograph, and interacted with the bottom of the cavity and then re-attached and moved
towards the front of the cavity, causing a slight increase in pressure uctuations at the front of the cavity and the
insulator. Based on the analysis shown in gure 9, in Base model 3, the pantograph base frame blocked a portion
of the incoming ow from direct impact with the lower arm of the pantograph and the knuckle. However, the
high-speed shedding vortex formed by Jet model 3 after the introduction of the jet avoided the obstruction of the
pantograph base frame, resulting in increased pressure uctuations in the lower arm bar of the pantograph,
knuckle, insulators, and the front of the cavity. The increase in pressure uctuations in these sections can be
clearly seen in gure 9.
Figure 11 shows the time averaged streamline of the wake in the pantograph region, which better describes
how the jet affects the main vortex structure in the wake of the pantograph region, and claries its mechanism of
action in affecting the aerodynamic noise in the pantograph region. In Base model 1, the incoming air
underwent aerodynamic separation at the leading edge of the cavity, and introduced a recirculation region at the
front of the cavity. The similar separation and recirculation have also been pointed out by Kim [13]and Noger
[34]. At present, a portion of the gas, after the separation of the front edge of the cavity, moved toward the
bottom of the pantograph, and impacted the insulators at the bottom of the pantograph and the bottom of the
cavity. After the impact, the uid velocity decreased signicantly, and the uid continued to move downstream,
and then rolled up at the back of the cavity. In Jet model 1, the jet raised the shear layer, but the same ow
separation occurred caused by the front of the pantograph base frame. A portion of the high-speed incoming air
Figure 11. Time averaged streamline of wake in the pantograph region.
11
Phys. Scr. 99 (2024)105228 Y Cao et al
moved towards the bottom of the cavity, and a recirculation region was introduced in the front of the cavity and
near the pantograph insulator. It can be seen from the trend of the streamline that the shear layer impacts the
pantograph insulators and the bottom of the cavity before rolling up. The tendency to roll upwards is more
pronounced compared to the case of Base model 1, indicating that the shear layer impacts the cavity and the
pantograph more violently, generating larger pressure uctuations.
In Base model 2, the incoming air also underwent ow separation at the leading edge of the cavity. A portion
of the separated gas moved downward, and impacted the front of the cavity. At the same time, a recirculation
region was introduced in the front of the cavity and the front of the pantograph insulator. The impacted gas
continued to move downstream, and reattached and rolled up downstream of the cavity after impacting the
pantograph insulator. It can be seen from the trend of the streamline, the incoming air also impacted the rear
part of the cavity, generating signicant pressure uctuations. In Jet model 2, the introduced jet raised the height
of the shear layer, and therefore, the incoming air did not undergo ow separation due to the front of the
pantograph, instead, it moved downstream after crossing the entire pantograph and introducing a larger
recirculation region at the rear of the cavity. Compared to Base model 2, due to the lack of high-speed uid
entering the bottom of the pantograph from the front of the cavity, the impact of air on the bottom of the
pantograph and the cavity was greatly reduced, including the pressure uctuations caused by it.
In Base model 3, due to the fact that the height of the pantograph base frame is lower than the depth of the
cavity, the incoming air did not undergo ow separation at the leading edge of the cavity, indicating that there
was no high-speed air separated from the leading edge of the cavity and entering the cavity. Instead, it directly
crossed the pantograph base frame and formed a relatively at recirculation region at the rear of the cavity after
interacting with the front of the pantograph. In Jet model 3, the jet raised the shear layer, so that its height
exceeded that of the pantograph base frame. It is worth noting that the raised shear layer caused the velocity of
the incoming air to increase. Therefore, its impact on the lower arm and knuckle will be greater, causing
signicant pressure uctuations here, and the speed of the recirculation region formed by the shear layer
crossing the pantograph is also higher compared to the case of Base model 3. Due to the high speed of ow in the
recirculation region, the impact on the pantograph insulator is greater. The uid in the recirculation region
continued to move towards the front of the cavity after impacting the insulator. It can be seen from gure 11 that
the velocity at the front of the cavity of Jet model 3 is greater than that of Base model 3, resulting in greater impact
of the uid on the front of the cavity, which creates a larger pressure uctuation at the front of the cavity. Figure 9
shows that the pressure uctuations at the lower arm, knuckle, insulator, and the front of the cavity in Jet model
3 have all increased.
As can be seen from gure 12, the distribution of turbulent kinetic energy in the pantograph region is under
various conditions. Compared to Base model 1, Jet model 1 signicantly enhances the amplitude of turbulent
kinetic energy in the lower arm region of the pantograph and the rear of the cavity. The reason is that the high-
speed turbulence is formed after the jet raises the shear layer, and the ow separation occurs in the front of the
pantograph. One part of the high-speed turbulence is pushed to the lower arm bar of the pantograph, while the
other part enters the cavity from the front of the cavity, and interacts with the bottom surface of the cavity. Seen
from gure 9, the amplitude of pressure uctuation in these sections is signicantly increased compared to the
case of Base model 1.
Based on the comparison with Base model 2, it can be clearly seen that due to the introduction of the jet, Jet
model 2 did not undergo ow separation. This also indicates that the high-speed turbulence formed by the jet
which raised the shear layer did not enter the cavity from the leading edge of the cavity, and it is also mostly
canceled by the front of the pantograph. Therefore, the high-speed turbulence in the cavity region and the rear of
the pantograph has been signicantly reduced compared to the case of Base model 2. This can also be
understood as the reason for the reduced amplitude of pressure uctuations in the pantograph region, as shown
in gure 9.
Based on the analysis of Base model 3 and Jet model 3, it can be concluded that no ow separation occurs at
this sink height with or without the presence of the jet. In another word, there is no high-speed turbulence
entering the cavity from the front of the cavity at this sink height. Figure 12 shows that, in Jet model 3 the high-
speed turbulence crosses the front of the pantograph, and directly interacts with the lower arm bar of the
pantograph, which continues to move towards the rear of the cavity, and interact with the back of the cavity. In
Base model 3, the front structure of the pantograph has a certain blocking effect on high-speed turbulence, with
only a small portion of high-speed turbulence crossing the front of the pantograph, resulting in less strong effect
on the lower arm and knuckle of the pantograph. However, due to the lack of jet to raise the shear layer, this part
of the turbulence collided with the upper part of the rear wall of the cavity, as can be seen from gure 9, the
pressure uctuation in this part is large in Base model 3. However, the pressure uctuations in Jet model 3 model
are signicant at the front of the cavity, lower arm, and knuckle.
Through the analysis of the ow eld in the pantograph region, it can be concluded that, at the same speed,
due to the obstruction of the pantograph base frame, in Jet model 1 with shallow cavities, it is difcult for the jet
12
Phys. Scr. 99 (2024)105228 Y Cao et al
to raise the shear layer to the required height. It can cause the high-speed uid formed by the shear layer to enter
the bottom of the cavity from the front edge of the cavity, and result in a signicant increase in pressure
uctuations in Jet model 1 compared to Base model 1. For Jet model 2 and Jet model 3 with deeper cavities, this
situation will not occur. The results indicate that jet can raise the shear layer to a sufcient height. In Jet model 2,
the introduction of jet effectively solves the problem of ow separation that occurs at the leading edge of the
cavity in Base model 2, resulting in a signicant reduction in pressure uctuations in the pantograph region.
However, in Jet model 3, Base model 3 no longer experiences ow separation at the leading edge of the cavity,
and the introduction of a jet at this point may cause additional noise.
Meanwhile, ow eld analysis focused on the potential impact of jet introduction on aerodynamic noise in
the pantograph area, without further analysis of the interaction between the jet and the ow inside the cavity. In
gure 11, it can be clearly seen from Jet model 2 and Jet model 3 that the jet crosses the pantograph base frame,
without entering the cavity from the leading edge of the cavity. Therefore, the ow inside the cavity is similar to
that in an open cavity (L/D<10)[35]. The ow inside the open cavity is less disturbed by the shear layer above.
Therefore, the introduction of jets into jet models 2 and 3 mainly reduced the velocity of the ow inside the
cavity, while the impact of the jet itself on the ow inside the cavity is relatively weak. For more information on
the mechanism of interaction between jets and open cavities, please refer to references [19,36]. In Jet Model 1,
the jet entered the cavity from the leading edge, and interacted with the cavity and pantograph frame, resulting in
increased pressure pulsation of related components and reduction of noise.
5.3. Far-eld noise
To analyze the far-eld noise in the pantograph region, 33 measurement points at different heights were set at
the most important location for evaluating train noise (track side), as shown in gure 13. The pantograph region
was divided into three parts for noise radiation calculation, namely, the overall part (cavity and pantograph),
pantograph part, and cavity part. Figure 14 summarizes the noise reduction effects of these three parts at the
measurement points.
For the three parts, Jet model 1 and Jet model 3 showed no good noise reduction effect, and compared to the
Base model without jet installed, their aerodynamic noise was reduced signicantly. This is consistent with the
results of the analysis in the previous section. In Jet model 2, the aerodynamic noise was reduced by about
1.53.9 dB at all measurement points in the overall section. In the pantograph section, the aerodynamic noise
was reduced by about 0.10.5 dB at most measurement points. In the cavity section, the aerodynamic noise at
the measurement point was reduced by about 4.57.5 dB, showing a signicant noise reduction effect. To
further analyze the noise reduction effect of Jet model 2, gure 15 shows the spectral results of Base model 2 and
Jet model 2 models at B6 and A6, and the corresponding overall sound pressure level (OASPL)is shown in
Figure 12. Turbulent kinetic energy distribution in the wake of the pantograph region.
13
Phys. Scr. 99 (2024)105228 Y Cao et al
Figure 13. Location of side measurement points.
Figure 14. Noise reduction effect at lateral measurement points.
14
Phys. Scr. 99 (2024)105228 Y Cao et al
gure 16(a). Figure 16(b)illustrates the noise reduction effect in the form of one-third octave band. Seen from
gure 15, for both Base model 2 and Jet model 2, their spectral results at B6 and A6 are generally broadband. And
for the entire pantograph region, the noise below 500 Hz is mainly contributed by the cavity, while the noise
above 500 Hz mainly comes from the pantograph. As can be seen from the spectrum of Base model 2, the vortex
shedding frequency and the peak corresponding to its harmonic frequency are clearly visible, and dominate the
total noise. From the spectrum of Jet model 2, the introduction of the jet mainly mitigates the noise in the
Figure 15. Spectral results of B6 and A6.
Figure 16. OASPL results and noise reduction effect of Jet model 2 at B6 and A6.
15
Phys. Scr. 99 (2024)105228 Y Cao et al
frequency range associated with vortex shedding, and the suppression of the cavity is extremely pronounced,
with a noise reduction of about 7 dB realized in the cavity section. The pantograph section achieved a noise
reduction of approximately0.4 dB. Therefore, for the entire pantograph region, OASPL has decreased by
approximately 3.7 dB.
5.4. Estimation of the planar jet noise
Considering that with the addition of the planar jet, it generates certain level of noise itself, which may affect the
noise reduction effect in the pantograph region. It is therefore necessary to evaluate the level of noise generated
by the jet itself to conrm that the noise itself generated by the addition of the jet does not affect the noise
reduction effect on the pantograph region.
Munro and Ahuja conducted aerodynamic-acoustic experiments on planar jets with high spreading ratios
(no cross ow)[37,38]. Rectangular jets with widths h between 0.66 and 1.45 mm and lengths w between 17 and
76 cm were tested, and the aspect ratios of the jets ranged from 100 to 3000 (12 different cases). The jet velocity
was set within the range of 150335 m s
1
, the noise measurements of the jet were made at different observation
angles
q
,
and they found that the noise intensity of the jet with high aspect ratio satises the following
relationship:
r
rqµ-
-
IVL
aR M1cos 6
mjeq
m
c
282
00
52
5
() ()
where
p
refers to the acoustic pressure,
r0
stands for the average density of the environment,
rm
represents the
density of the mixing zone,
a0
is the environmental sound velocity,
L
eq
denotes the equivalent length scale, and
M
c
stands for the convective Mach number. Based on the experimental results, the convective Mach number is
approximated as
=
M
M0.36
cj
where
M
j
refers to the Mach number of the jet. The denition of equivalent
length
L
eq
is expressed as follows:
=Lhw 7
eq 34 14
()
//
It should be noted that the convective Mach number does not work at the viewing angle of 90°. Therefore,
the spectral levels were normalized according to equation (6), and compared with the normalized frequency to
obtain:
q=-
+
ffLMVlog 1 cos 8
eq c j
[( )] ()/
Oerlemans et al [39]combined equations (6)to (8)to give the normalized spectrum used to estimate the jet
noise. With
SPL
nor
m
and
f
nor
m
(the normalized level and frequency as read from the normalized spectrum), and
all other variables in metric units, the dimensional sound level SPL and frequency
f
are determined by the
following equations:
xq=+ - - + - +
gg-
SPL SPL R V h w M20log 10 log 20log 50log 1 cos 98.72 9
mj cnorm 1
() () ( ) ( ) ()
q=-
gg-
ffhwMVlog 1 cos 10
norm cj
1
[( )] ()/
where
x
=8,
g
=0.75.
In the study by Oerlemans et al [39], it was found that at jet velocities below 140 m s
1
(M
j
=0.4), according
to the scale given in the reference [37,38], and the law of 8 powers of the jet velocity, the test results are not well
collapsing. Better results can be obtained if the law of 5 powers of jet velocity is used for normalization instead of
the law of 8 powers.
Therefore, noise estimation for the jet at a velocity of 140 m s
1
is performed based on the results obtained
by Munro et al [37,38]and the normalized spectrum provided by Oerlemans et al [39]in combination with
equations (9)and (10). Then, the results were scaled by using the 5th power law of jet velocity to obtain a jet noise
spectrum at 111.11 m s
1
. To simplify the calculations, only the observation angle of 90°is considered. Figure 17
shows the comparison between the noise spectra of Base model 2 and Jet model 2 at measurement points B6 and
C6 and the estimated jet noise spectra (displayed in the form of 1/3 octave band). As can be seen from the gure,
whether it is the Base model or the Jet model, the noise generated by the jet itself is much lower than that in the
pantograph region at the frequency below 2000 Hz. For frequency noise above 2000 Hz, the noise generated by
the jet itself is slightly greater. However, considering that high-frequency noise decays rapidly when measuring
far-eld noise, its impact on far-eld noise measurement is not signicant. On the other hand, since only a 90°
observation angle was considered during estimation, it can be inferred from equation (9)that the actual
aerodynamic noise generated by the jet is lower than the estimation. Therefore, based on the overall results, the
noise generated by the jet itself will not affect the noise reduction effect.
16
Phys. Scr. 99 (2024)105228 Y Cao et al
6. Conclusions
This paper mainly explored the potential application of introducing a planar jet at the leading edge of the
pantograph cavity to control aerodynamic noise in the pantograph region of high-speed trains through
numerical research. Three different pantograph region models were established with sinking heights of 350 mm,
500 mm, and 650 mm, respectively, namely Base model 1, Base model 2, and Base model 3. On this basis, planar
jets are arranged at 50 mm away from the leading edge of the cavity, and the corresponding jet-
pantograph region models are called Jet model 1, Jet model 2, and Jet model 3, respectively. The jet velocity is set
to be the same as the incoming velocity during the simulation to provide enough momentum to lift the shear
layer. Based on the results, considering the plane jet, the dipole source intensity (pulsating pressure)on the
surface of the pantograph region in Jet model 2 is signicantly reduced compared to the case of Base model 2,
especially at insulators, cavities and lower boom surfaces. And for measurement points on the side of the
pantograph, the overall sound pressure level is reduced by up to 4 dB, showing a signicant noise reduction
effect. Meanwhile, the results of the far-eld noise estimation of the jet itself show that the noise generated by the
jet itself does not affect the noise reduction. In Jet model 1 and Jet model 3, the dipole source intensity on the
surface of the pantograph region increases in different degree compared to the corresponding Base model, which
leads to a deterioration of the overall sound pressure level.
The analysis of the ow elds shows that the wake elds in both Base model and Jet model
pantograph regions are dominated by alternating vortex shedding. In the pantograph region with shallow
cavities (Base model 1 and Base model 2), the ow separation at the leading edge of the cavity mainly accounts for
the increase of the dipole source strength at the bottom of the pantograph and in the cavity. The comparative
analysis shows that the introduction of jet in Jet model 1 cannot solve the problem of ow separation. The reason
is that though the jet raises the shear layer, it is not enough to cross the base frame of the pantograph, instead,
ow separation occurs in the front of the pantograph, consequently, the high-speed turbulence enters the cavity
from the front of the pantograph, and interacts with the cavity and the pantograph to cause large pressure
pulsations. The introduction of the jet in Jet model 2 is a good solution to the problem of ow separation at the
leading edge of the cavity, which provides good noise reduction. For the pantograph region with a deeper cavity
(Base model 3), there is no ow separation at the leading edge of the cavity. If a jet is introduced (such as Jet
model 3), the high-speed turbulence formed by the elevated shear layer will impact the lower arm of the
pantograph, the knuckle, the insulators, and the front part of the cavity, introducing an additional noise source,
thus worsening the aerodynamic noise situation.
In summary, in the pantograph region where ow separation occurs at the leading edge of the cavity, the
introduced jet must raise the shear layer to cross the front of the pantograph, thereby preventing high-speed
turbulence from entering the cavity from the front of the pantograph to achieve noise reduction effect. On the
contrary, in the pantograph region where ow separation does not occur at the leading edge of the cavity, the
introduction of jet will introduce additional noise sources, leading to a reduction of aerodynamic noise in the
pantograph region. However, it should be noted that the current numerical results alone are not sufcient for
the investigators to fully understand the impact of the jet on the pantograph region. Therefore, further research
in this eld is necessary in the future, especially to further describe the optimal jet velocity, position, and shape of
Figure 17. Comparison of spectral results of pantograph noise and jet noise.
17
Phys. Scr. 99 (2024)105228 Y Cao et al
the jet nozzle, and the effect of the jet on the pantograph region when the high-speed train is moving in the other
direction.
Acknowledgments
This work was supported by National Natural Science Foundation of China (12172308).
Data availability statement
The data cannot be made publicly available upon publication because no suitable repository exists for hosting
data in this eld of study. The data that support the ndings of this study are available upon reasonable request
from the authors.
Author contributions
Yangyang Cao:Conceptualization, Methodology, Data curation, Formal analysis, Writing - original draft. Jiye
Zhang: Funding acquisition, Supervision. Jiawei Shi:Writing - review & editing, Supervision. Yuzhe Ma:
Editing, Supervision.
Declaration of competing interest
The authors declare that they have no known competing nancial interests or personal relationships that could
have appeared to inuence the work reported in this paper.
ORCID iDs
Jiye Zhang https://orcid.org/0000-0003-2502-2444
References
[1]Sun Z, Song J and An Y 2012 Numerical simulation of aerodynamic noise generated by high speed trains Engineering Applications of
Computational Fluid Mechanics 617385
[2]Zhang Y et al 2016 Research on aerodynamic noise reduction for high-speed trains Shock and Vibration 2016 121
[3]Qin D et al 2023 Numerical study on aerodynamic drag and noise of high-speed pantograph by introducing spanwise waviness
Engineering Applications of Computational Fluid Mechanics 17 2260463
[4]Meskine M, Perot F and Kim M S 2015 Community noise prediction of digital high speed train using LBM 19th AIAACEAS
Aeroacoustics Conf. and Exhibit, AIAA 117
[5]Thompson D J et al 2015 Recent developments in the prediction and control of aerodynamic noise from high-speed trains International
Journal of Rail Transportation 311950
[6]Tan X M et al 2018 Vortex structures and aeroacoustic performance of the ow eld of the pantograph J. Sound Vib. 432 1732
[7]Zhao Y et al 2020 Analysis of the near-eld and far-eld sound pressure generated by high-speed trains pantograph system Appl. Acoust.
169 107506
[8]Zhang Y D et al 2017 Investigation of the aeroacoustic behavior and aerodynamic noise of a high-speed train pantograph Sci. China
Technol. Sci. 60 56175
[9]Lei S et al 2013 Numerical analysis of aerodynamic noise of a high-speed pantograph 2013 Fourth Int. Conf. on Digital Manufacturing &
Automation. IEEE 83741
[10]Sun X and Xiao H 2018 Numerical modeling and investigation on aerodynamic noise characteristics of pantographs in high-speed
trains Complexity 50 93541
[11]Shi F et al 2022 Numerical study on aerodynamic noise reduction of pantograph Applied Sciences 12 10720
[12]Meskine M, Pérot F and Kim M S 2013 Community noise prediction of digital high speed train using LBM 19th AIAA/CEAS
Aeroacoustics Conf. 2015
[13]Kim H, Hu Z and Thompson D 2020 Numerical investigation of the effect of cavity ow on high speed train pantograph aerodynamic
noise J. Wind Eng. Ind. Aerodyn. 201 104159
[14]Yao Y et al 2019 Analysis of aerodynamic noise characteristics of high-speed train pantograph with different installation bases Applied
Sciences 92332
[15]Kim H, Hu Z and Thompson D 2021 Effect of different typical high speed train pantograph recess congurations on aerodynamic noise
Proc. Inst. Mech. Eng. Part F J. Rail Rapid Transit 235 57385
[16]Saddington A J, Thangamani V and Knowles K 2016 Comparison of passive ow control methods for a cavity in transonic ow J. Aircr.
53 143947
[17]Kim H 2016 Unsteady aerodynamics of high speed train pantograph cavity ow control for noise reduction. 22nd AIAA/CEAS
Aeroacoustics Conf. 2848
[18]Kim H, Hu Z and Thompson D 2020 Effect of cavity ow control on high-speed train pantograph and roof aerodynamic noise Railway
Engineering Science 28 5474
18
Phys. Scr. 99 (2024)105228 Y Cao et al
[19]Yu P X et al 2014 Suppression effect of jet ow on aerodynamic noise of 3D cavity Applied Mechanics and Materials 444 58895
[20]Ukai T et al 2014 Effectiveness of jet location on mixing characteristics inside a cavity in supersonic ow Exp. Therm Fluid Sci. 52 5967
[21]Zhao K et al 2018 Aerodynamic noise reduction using dual-jet planar air curtains J. Sound Vib. 432 192212
[22]Spalart P R et al 2006 A new version of detached-eddy simulation, resistant to ambiguous grid densities Theor. Comput. Fluid Dyn. 20
18195
[23]Shur M L et al 2008 A hybrid RANS-LES approach with delayed-DES and wall-modelled LES capabilities Int. J. Heat Fluid Flow 29
163849
[24]Squires K et al 2002 Progress on detached-eddy simulation of massively separated ows 40th AIAA Aerospace Sciences Meeting &
Exhibit 1021
[25]STAR-CCM+User Guide (Version 12.04), Siemens PLM Sofware, 2017
[26]Zhang Y D, Zhang J Y and Li T 2016 Research on aerodynamic noise source characterization and noise reduction of high-speed trains
vehicle Journal of the China Railway Society 38 409
[27]Zhang Y D, Zhang J Y and Sheng X Z 2020 Study on the ow behaviour and aerodynamic noise characteristics of a high-speed
pantograph under crosswinds Sci. China Technol. Sci. 63 97791
[28]Zhu J Y, Hu Z W and Thompson D J 2014 Flow simulation and aerodynamic noise prediction for a high-speed train wheelset
International Journal of Aeroacoustics 13 53352
[29]Farassat F 2007 Derivation of Formulations 1 and 1A of Farassat [R]
[30]He Y 2023 Aerodynamic Noise Simulation of High-Speed Train Bogie (University of Southampton)
[31]Jian D U, Jianying L and Aiqin T 2015 Analysis of aeroacoustics characteristics for pantograph of high-speed trains Journal of Southwest
Jiaotong University 28 93541
[32]Plentovich E B, Stallings Jr R L and Tracy M B 1993 Experimental cavity pressure measurements at subsonic and transonic speeds
Static-Pressure Results
[33]Xiao-Ming T et al 2018 Vortex structures and aeroacoustic performance of the ow eld of the pantograph J. Sound Vib. 432 1732
[34]Noger C et al 2000 Aeroacoustical study of the TGV pantograph recess J. Sound Vib. 231 56375
[35]Zhang X and Edwards J A 1990 An investigation of supersonic oscillatory cavity ows driven by thick shear layers The Aeronautical
Journal 94 35564
[36]Zhang X 1995 Compressible cavity ow oscillation due to shear layer instabilities and pressure feedback AIAA J. 33 140411
[37]Munro S and Ahuja K 2003 Aeroacoustics of a high aspect-ratio jet 9th AIAA/CEAS Aeroacoustics Conf. and Exhibit 3323
[38]Munro S and Ahuja K 2003 Development of a prediction scheme for noise of high -aspect ratio jets 9th AIAA/CEAS Aeroacoustics Conf.
and Exhibit 3255
[39]Oerlemans S and de Bruin A 2009 Reduction of landing gear noise using an air curtain 15th AIAA/CEAS Aeroacoustics Conf. (30th AIAA
Aeroacoustics Conf.) 3156
19
Phys. Scr. 99 (2024)105228 Y Cao et al
... The research conclusions of references [25,26] indicate that under low Mach number flow conditions, the noise generated by the quadrupole contributes very little to the overall noise, and therefore can be ignored in acoustic simulations of the pantograph region. Meanwhile, Zhao et al. 's [27] research results indicate that for the pantograph area of high-speed trains, the aerodynamic noise contributed by dipoles is much greater than that contributed by quadrupoles. ...
... It is important to note that, while increasing jet velocity results in higher noise levels from the jet slot, the impact on overall cavity noise must be considered. According to findings from references [25,32], at a velocity of 400 km/h, the noise generated by the jet itself has no significant adverse effect on the noise reduction outcomes. Consequently, the observed noise reduction in cavity aerodynamic noise is deemed reliable. ...
Article
Full-text available
A hybrid method incorporating the simulations of noise sources with delayed detached eddy simulation (DDES) and calculations of far-field noise with the Ffowcs Williams–Hawkings (FW-H) equation is used to study the suppression technique for the aerodynamic noise of a Faiveley CX-PG pantograph. Considering that China’s Fuxing bullet trains operate at 350 km/h, the inflow velocity of 350 km/h is applied in this paper. The noise radiated from the panhead area, middle area, and bottom area at an inflow velocity of 350 km/h is distinguished. The noise intensities at the standard observer show that the noise radiated from the panhead area is the strongest, and the sound pressure level spectrum value is larger than the other two in the range above 500 Hz. The influence of applying the wavy rods and modifying the contact strip shape on the aerodynamic noise is discussed in detail. By comparing the acoustic source distribution and the far-field noise intensity, it is found that applying the wavy rods can effectively reduce the panhead noise, especially around the peak frequency. Modifying the shape of the contact strip to a hexagon can suppress the vortex shedding, leading to a lower surface pressure level. Combining the strip modification and wavy rods, the total noise intensity can be diminished by about 3.0 dB.
Article
Full-text available
The pantograph and its recess on the train roof are major aerodynamic noise sources on high-speed trains. Reducing this noise is particularly important because conventional noise barriers usually do not shield the pantograph. However, less attention has been paid to the pantograph recess compared with the pantograph. In this paper, the flow features and noise contribution of two types of noise reduction treatments rounded and chamfered edges are studied for a simplified high-speed train pantograph recess, which is represented as a rectangular cavity and numerically investigated at 1/10 scale. Improved delayed detached-eddy simulations are performed for the near-field turbulent flow simulation, and the Ffowcs Williams and Hawkings aeroacoustic analogy is used for far-field noise prediction. The highly unsteady flow over the cavity is significantly reduced by the cavity edge modifications, and consequently, the noise radiated from the cavity is reduced. Furthermore, effects of the rounded cavity edges on the flow and noise of the pantographs (one raised and one folded) are investigated by comparing the flow features and noise contributions from the cases with and without rounding of the cavity edges. Different train running directions are also considered. Flow analysis shows that the highly unsteady flow within the cavity is reduced by rounding the cavity edges and a slightly lower flow speed occurs around the upper parts of the raised pantograph, whereas the flow velocity in the cavity is slightly increased by the rounding. Higher pressure fluctuations occur on the folded pantograph and the lower parts of the raised pantograph, whereas weaker fluctuations are found on the panhead of the raised pantograph. This study shows that by rounding the cavity edges, a reduction in radiated noise at the side and the top receiver positions can be achieved. Noise reductions in the other directions can also be found.
Article
Full-text available
The high-speed-train pantograph is a complex structure that consists of different rod-shaped and rectangular surfaces. Flow phenomena around the pantograph are complicated and can cause a large proportion of aerodynamic noise, which is one of the main aerodynamic noise sources of a high-speed train. Therefore, better understanding of aerodynamic noise characteristics is needed. In this study, the large eddy simulation (LES) coupled with the acoustic finite element method (FEM) is applied to analyze aerodynamic noise characteristics of a high-speed train with a pantograph installed on different configurations of the roof base, i.e. flush and sunken surfaces. Numerical results are presented in terms of acoustic pressure spectra and distributions of aerodynamic noise in near-field and far-field regions under up- and down-pantograph as well as flushed and sunken pantograph base conditions. The results show that the pantograph with the sunken base configuration provides better aerodynamic noise performances when compared to that with the flush base configuration. The noise induced by the down-pantograph is higher than that by the up-pantograph under the same condition under the pantograph shape and opening direction selected in this paper. The results also indicate that, in general, the directivity of the noise induced by the down-pantograph with sunken base configuration is slighter than that with the flush configuration. However, for the up-pantograph, the directivity is close to each other in Y-Z or X-Z plane whether it is under flush or sunken roof base condition. However, the sunken installation is still conducive to the noise environment on both sides of the track.
Thesis
Aerodynamic noise from high-speed trains becomes more and more important as the train speed increases. Of the various sources, the train bogies contribute significantly to the overall aerodynamic noise, especially the leading bogie. This research aims to reveal the aerodynamic noise generation mechanisms from bogies of high-speed trains and propose suitable noise reduction measures. Numerical simulation has been a great challenge for the simulation of the aerodynamic noise of bogies. This is due to the complex geometry, which makes the discretization very difficult and, meanwhile, the grid will be very large. To overcome these challenges, a hybrid grid system is explored, which can guarantee a high-quality grid in the boundary layer while maintaining the overall number of cells in the grid at an acceptable level; the model size and flow speed are both scaled down to further reduce the number of the grid cells. The Delayed Detached Eddy Simulation is used to investigate the flow and obtain the noise source information to feed into the Ffowcs Williams-Hawkings acoustic analogy for far-field noise prediction. The hybrid grid system and the numerical methods are validated by simulations for a circular cylinder, square cylinder and an isolated wheelset. After that, the hybrid grid system is applied and further developed for the simulations of a bogie in a simplified cavity. The results show that the rear part of the bogie and cavity have strong pressure fluctuations and the noise generated by the cavity is much greater than that by the bogie. A more complex model of a bogie under a leading car is then investigated. It is found that the bottom of the cowcatcher and the bogie, the cavity rear surface, and the side dampers, which are directly flapped by the highly turbulent wake and detached shear layer, form strong pressure fluctuations. The far-field noise levels and the sound power levels emitted by the cavity are greater than that of the bogie. The effect on the aerodynamics and aeroacoustics of the lateral position of the bogie’s side components relative to the car body is investigated. The flow field results show that the protruded side components shield the detached shear layer from upstream, preventing it from impinging on the rear part of the cavity. The pressure fluctuation on the side components increases, while it reduces at the rear surface of the cavity, as a result of which only a small difference in the sound power is found between the various cases. Based on the analysis of the flow field and pressure fluctuation distribution of the simulated cases, a noise control technique based on a dual staggered jet is developed to reduce the noise level of the leading car. The wake at the bottom and the detached shear layer at the two sides of the cavity is pushed away by the jets, which reduces the pressure fluctuations on the bogie and the cavity. A reduction of 2 dB for the sound pressure levels and 3.5 dB for the sound power levels is obtained. Finally, to reduce further the computational cost of the models, a novel decomposition method for CFD simulation is developed. A model of tandem square cylinders is adopted to validate the method showing good agreement between the decomposed model and the complete model. The computational time of the decomposed model is 29.6% less than that of the complete model. The collected inflow data is compressed by a convolutional Variational AutoEncoder neural network and the compression ratio achieves 63.5. The decomposition method then is applied to the simulation of a half-width leading car model.
Article
For high-speed trains, the aerodynamic noise becomes an essential consideration in the train design. The pantograph and pantograph recess are recognised as important sources of aerodynamic noise. This paper studies the flow characteristics and noise contributions of three typical high-speed train roof configurations, namely a cavity, a ramped cavity and a flat roof with side insulation plates. The Improved Delayed Detached-Eddy Simulation approach is used for the flow calculations and the Ffowcs Williams & Hawkings aeroacoustic analogy is used for far-field acoustic predictions. Simulations are presented for a simplified train body at 1/10 scale and 300 km/h with these three roof configurations. In each case, two simplified pantographs (one retracted and one raised) are located on the roof. Analysis of the flow fields obtained from numerical simulations clearly shows the influence of the train roof configuration on the flow behaviour, including flow separations, reattachment and vortex shedding, which are potential noise sources. A highly unsteady flow occurs downstream when the train roof has a cavity or ramped cavity due to flow separation at the cavity trailing edge, while vortical flow is generated by the side insulation plates. For the ramped cavity configuration, moderately large pressure fluctuations appear on the cavity outside walls in the upstream region due to unsteady flow from the upstream edge of the plate. The raised pantograph, roof cavity, and ramped cavity are identified as the dominant noise sources. When the retracted pantograph is located in the ramped roof cavity, its noise contribution is less important. Furthermore, the insulation plates also generate tonal components in the noise spectra. Of the three configurations considered, the roof cavity configuration radiates the least noise at the side receiver in terms of A-weighted level.
Article
Due to the limited understanding of spatial sound field and contribution of pantograph, a deep analysis is presented to extract understanding the aerodynamic noise characteristics of high-speed trains pantograph system. The near-field and far-field noise are predicted by the acoustic perturbation equations and Ffowcs Williams-Hawkings equation, respectively. The spatial sound propagation is analyzed by proper orthogonal decomposition and cluster-based reduced-order modelling. The flow field results predicted by large eddy simulation show that the flow behind pantograph is governed by hierarchical structures, which occurs to be featured with three layers as well as periodic evolution. The dipole source is dominated in far-field radiated noise while the quadrupole source is negligible, since the intensity of the quadrupole source is less than that of the dipole source. The contribution rates of base-frame, pan-head, groove, upper-arm, horn, lower-arm and rod-insulator are higher than the other components, and the radiated sound energy of them accounts for approximately 90% of the total energy. The noise contribution of pan-head exceeds 10% at the frequency of 1000 Hz. Base-frame is the largest contributor, and more than 6% of noise contribution occurs at the frequency of 630 Hz. The spatial sound propagation includes two major aspects: one is mainly reflected upwind from groove, and the other is propagated with the center of groove. The results highlight the possibility to develop a new design of high-speed trains pantographs, in order to obtain an aerodynamic noise reduction.
Article
The aerodynamic noise of high-speed trains increases significantly under crosswinds. Researches have typically focused on the characteristics of aerodynamic loads and the corresponding safety issues, with less attention to flow-induced noise characteristics. In the present paper, the near-field unsteady flow behaviour around a pantograph was analysed using a large eddy simulation. The far-field aerodynamic noise from a pantograph was predicted using the Ffowcs Williams-Hawkings acoustic analogy. The results showed that asymmetric characteristics of the flow field could be observed using the turbulent kinetic energy and the instantaneous vortexes in crosswind conditions. Vortex shedding, flow separation and recombination around the pantograph were the key factors for aerodynamic noise generation. The directivity of the noise radiation was inclined towards the leeward side of the pantograph. The aerodynamic noise propagation pattern can be considered as a typical point source on spherical waves when the transverse distance from the pantograph geometrical centre is farther than 8 m. The sound pressure level grew approximately as the 6th power of the pantograph speed. The peak frequency exhibited a linear relationship with the crosswind velocity. The numerical simulation results and wind tunnel experiments had high consistency in the full frequency domain, namely, the peak frequency distribution range, the main frequency amplitude and the spectral distribution shape.
Article
Reducing train pantograph noise is particularly important. In this paper, the flow behaviour and noise contribution of simplified geometries representing high-speed train pantographs and the roof cavity at 1/10th scale are investigated. The Improved Delayed detached-Eddy Simulation (IDDES) turbulent model is used for the flow field simulation and the Ffowcs Williams & Hawkings aeroacoustic analogy is used for far-field noise prediction. The pantograph recess geometry is simplified to a rectangular cavity and two simplified DSA350 pantographs are included. The effect of the pantograph cavity is studied by comparing the flow behaviour and radiated noise from cases with and without the cavity, and also for different train running directions. When the pantographs are installed in a cavity, the shear layer, separated from the cavity leading edge, interacts with the pantographs, and generates large pressure fluctuations on the pantograph surfaces. In comparison with pantographs installed on a flat train roof, the flow around the pantographs with the cavity has different characteristics in terms of the velocity profile upstream of the pantographs. The study shows that the main noise source is from the panhead of the raised pantograph which produces strong tonal noise and this noise source is affected by the cavity flow.
Article
The object of study in this paper is the Faiveley CX-PG pantograph. We first used the large-eddy simulation (LES) model to simulate the surrounding fluctuating flow field. We then identified the vortex structures in the flow field of the pantograph via the Q criterion, and performed a Fourier transform on the fluctuating pressure. We finally used the Ffowcs Williams-Hawkings (FW-H) equation to predict the far-field radiation noise of the pantograph. Through these steps, we explored the vortex structures in the flow field of the pantograph, aeroacoustic performance of the pantograph's main components, and the relationship between them, and proposed corresponding acoustic optimization countermeasures. The results showed that the vortex structures in the flow field of the pantograph varied with time and had a certain periodicity, and that the sound source intensity of the pantograph was mainly distributed in the bottom frame, three insulators, balance beam, upper arm frame, and lower arm. The sound source energy of these components accounted for approximately 92% of the total energy; the influencing factors for the aerodynamic sound source intensity of the pantograph included the shedding positions and vorticities of the vortex structures as well as whether it was located in the wake of the vortex structures. The aerodynamic noise of the pantograph could be effectively controlled by adjusting the vortex shedding position, reducing the vorticities of the vortex structures, increasing the distance between the mutually interfering components, setting the diversion structure to control the discharge area of the vortex structures. The sound source energy of the bottom frame area accounted for more than 50% of the total energy; a settlement platform or shroud could be installed to effectively control the noise in the area, thereby effectively reducing the noise radiated by the pantograph. The simulation results in this paper were in good agreement with the wind tunnel test results and theoretical results, and can provide a reference for the optimal design of future acoustics for the pantograph.