1-s2.0-S0307904X24002762-main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Applied Mathematical Modelling 134 (2024) 29–49

Contents lists available at ScienceDirect

Applied Mathematical Modelling


journal homepage: www.elsevier.com/locate/apm

Free propagation of elastic waves in small-curvature, damped,


infinite cables
Lijun Li a, b, Xiaohui Zeng a, b, *, Han Wu a, b, Zhehua Cui a, b
a
Institute of Mechanics, Chinese Academy of Sciences, Beijing 100190, China
b
University of Chinese Academy of Sciences, Beijing 100049, China

A R T I C L E I N F O A B S T R A C T

Keywords: Understanding the wave propagation is crucial for dynamic analysis of long elastic cables prone to
Wave propagation in elastic cables high-order vibrations. The coupling of elastic waves in cables is still not sufficiently understood.
Damping However, wave coupling produces attenuation frequency bands, which is useful for vibration
Wave dispersion
control. In this study, the coupled waves equation for a damped cable with a small curvature was
Stopband
Passband
proposed. Using the equation, wave frequencies and velocities were studied. Then the cable re­
Wave dissipation sponses in the stopbands and passbands were examined. Finally, the free responses were analysed
in the spatial domain. The study focused on the influence of curvature and damping on wave
propagation. The results showed that the wave coupling caused by curvature was critical for
producing the attenuation bands and dispersion of waves. It generated the stopbands and pass­
bands for the wave propagation, in which the waves respectively attenuated or propagated.
Furthermore, the coupling effect induced pronounced dispersion in the in-plane longitudinal
waves. As for damping, it not only affected the frequencies of stopbands and passbands, but also
caused wave attenuation. Damping was found to shift the cut-off frequency from the real axis to
the imaginary axis and transform the stopband from a line segment into a quadrilateral area.
Additionally, the relationship between frequency and wave-number evolved from a two-
dimensional curve into a three-dimensional surface. What’s more, damping caused wave dissi­
pation over time.

1. Introduction

Cables are extensively employed in bridges, railway catenary systems, marine mooring, aerospace structures and in optical and
electrical signal transmission systems. The diversity of applications has led to different theories for elastic cables, broadly categorized
into those for parabolic cables with small sags and those for catenary cables with large sags. For cables with small sags, analytical
solutions for the natural frequencies and mode shapes can be obtained by introducing a parabolic approximation to solve the con­
tinuum formulations. In contrast, for cables with large sags, discrete formulations and numerical methods are commonly employed,
including finite difference [1–3], finite element [4,5], lumped parameter and multi-body dynamics methods.
The research on cable continuously evolves over time, advancing with the needs of practical engineering applications. It began with
the model of an inextensible taut string and later developed into elastic cable models that considered small sags [6], which are now
adopted in most engineering applications, even if sometimes the geometric nonlinearity is considered [7]. Although most of these cable

* Corresponding author at: Institute of Mechanics, Chinese Academy of Sciences, Beijing 100190, China.
E-mail address: zxh@imech.ac.cn (X. Zeng).

https://doi.org/10.1016/j.apm.2024.05.043
Received 8 December 2023; Received in revised form 24 May 2024; Accepted 30 May 2024
Available online 3 June 2024
0307-904X/© 2024 Elsevier Inc. All rights are reserved, including those for text and data mining, AI training, and similar technologies.
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

models ignore bending stiffness, some do take it into account [8]. Cable dynamics literature is abundant and includes monographs [9]
and extensive reviews [10–13]. There are numerous phenomena worthy of study in the dynamics of elastic cable. Rega [11] sum­
marizes the dynamic response of elastic cable. Although the analysis almost exclusively focuses on normal cable models, this approach
is only effective for low-order vibration in short cable. However, as the cable adopted for engineering applications become longer,
higher-order vibration become easier to be excited. The focus should thus shift to the elastic wave propagation instead of the modal
description in such cases. In addition, many engineering problems are also closely related to wave propagation. In bridge engineering,
wind-induced vibrations can trigger cable vortex-induced resonance, potentially leading to instability in the tower structure [13,14].
For high-speed railways, the increase in train speed can excite higher-order catenary responses [15]. Moreover, when the moving train
speed is close to the wave propagation speed, the pantograph-catenary interaction performance may deteriorate and produce
abnormal contact forces [16,17].
The wave propagation in elastic cable structures can be researched through theoretical, numerical and experimental methods. In
most studies, cables are simplified into string, rod or beam structures. Tao and Zhang [18] investigate the elastic wave propagation in
an anchor cable of a nonlinear tethered unmanned aerial vehicle (UAV) and finds that the longitudinal and transverse waves are linked
via the equilibrium curvature, leading to their dispersion. Zhang et al. [19] employ a combination of finite element and spectral
element methods to study the propagation of elastic waves in cable structures. They found that local damage has a significant impact
on the propagation of elastic waves in cables. Perkins et al. [20–23] derive the equations of motion for infinite elastic suspended cables
with small sags using the Hamiltonian principle and study the free propagation of waves. According to the results, the curvature
couples transverse and longitudinal waves, producing in dispersive elastic waves in cables. They further use analytical and numerical
methods to investigate the propagation of harmonic waves produced by external forcing in the frequency and time domains. The
results show significant differences in the responses of the cable in the stopbands and passbands. Graff [24] studies theoretically wave
propagation in the mooring line and suspended cable, considering the mass and curvature of the cable. Li et al. [25] consider the
equation of motion of a damped three-dimensional elastic cable. They find that damping plays an important role in wave dispersion
and causes wave dissipation of the crest and changes in edge response.
Sorrentino et al. [26] establish a distributed parameter model for the two-level railway catenary system, where the contact and the
messenger wires are modeled as two straight parallel beams. Based on this model, the wave propagation of the contact wire is studied
and a highly efficient and accurate numerical integration scheme is proposed. Zou et al. [27] investigate the wave propagation in a
three-span contact wire using a combination of experiments and numerical simulations. According to their conclusions, treating the
contact wire as a string or Euler-Bernoulli beam results in an overestimation of the wave speed. Song et al. [28] study the frequency,
reflection and transmission of the transverse waves in a railway catenary system subjected to a moving load by combining analytical
with finite element models. In analytical models, the contact and messenger wires are modeled as straight tensioned cables without
bending stiffness. The results show a resonance may occur when the dominant wave frequency becomes closer to the natural fre­
quency, which may lead to the sudden rise in the contact force fluctuations. Van et al. [29] uses a combination of mass-drop tests,
analytic models and finite element simulations on a catenary allowing a detailed analysis of the influence of train speed. Based on the
string model, the propagation, reflection and transmission of waves on contact wire are theoretically studied. In the numerical and
experimental studies, wave dispersion is found after considering bending stiffness.
Damping and wave coupling induced by curvature have strong effects on the wave propagation in elastic cable. Research indicates
that damping is a critical parameter for energy dissipation [30,31], and the dynamic response and wave propagation in flexible
structure can be significantly impacted by even the weak damping [32,33]. The results reported in this paper also emphasize the
importance of damping for wave frequency and response. Studies have shown that wave coupling has a significant impact on wave
propagation in beams [29,34–37], frames [38], plates [33] or composite structures [39]. For cable structures with curvature, the wave
coupling not only leads to the wave interaction and dispersion, but also affects their frequency [20–22]. The above references indicate
the importance to considering wave coupling in the wave equations. However, it is challenging to solve analytically the coupled wave
equations since they may be highly nonlinear. Thus, the complex coupled waves are usually analyzed by numerical methods, such as
the finite element method [29,38,40,41] and generalized multi-symplectic method [42–44]. In this study, the wave coupling are
considered by introducing cable curvature and damping. The nonlinear coupled wave equations for a cable can be linearized in the
case of small sag, facilitating analytical solution.
Research shows that using the wave theory is promising for vibration control. Jiang et al. [45] demonstrate the possibility of
constructing a wide bandgap by consecutively superposing multiple stopbands generated by cylinders of various lengths. The system is
highly efficient in blocking propagating waves. Lin et al. [46] obtain and verify the dispersion relationships for the periodic flexo­
electric curved beams with intriguing coupling characteristics. The results indicate that the bandgaps are mainly influenced by
flexoelectric effect, while significant geometric and material parameters have the synergistic effects. Darche et al. [47] present a novel
method based on the waveguide modal analysis to evaluate the effectiveness of bounded periodic metamaterials in terms of frequency
bandgaps. The results indicate that the local resonance causes the first frequency bandgap to appear in a very low-frequency range. In
this study, an analytical expression for the cutoff frequency is provided, and the wave propagation in attenuation and propagation
bands is analyzed.
This paper is organized as follows. The coupled wave equations considering curvature and damping were proposed and linearized
in Section 2. Based on the characteristic equation, the parameters of freely propagating waves in cables were discussed in Section 3.
The results show that the relationship between frequency and wave-number changes significantly after considering curvature,
especially in the generation of stopbands and passbands. Furthermore, the stopband changes from a line segment to a plane because of
damping. The free responses in the wave-number domain and in spatial domain were examined in Section 4. The cable responses at
stopbands and passbands are studied in Section 4.1. Then, the responses in the spatial domain are investigated by the inverse Fourier

30
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

transform in Section 0. The different behaviors of responses at stopbands and passbands are found here. In addition, the combined
effects of curvature and damping on the responses were discussed, indicating that the wave dispersion is related to the curvature and
wave dissipation is related to damping. Concluding remarks are provided in Section 5.

2. Infinite cable equation of motion

For slender cable with high initial tension, bending and torsion have little impact on their dynamics [14,48]. Therefore, the cable
was modeled herein as a one-dimensional elastic continuum that had only axial stiffness [22]. The three different configurations of an
elastic cable are depicted in Fig. 1: the initial (unstressed) configuration χi, the static equilibrium configuration under dead load χe, and
the dynamic configuration under the action of external load or other disturbance χd [10]. The study focuses on the displacement from
static equilibrium configuration χe to dynamic configuration χd, denoted by u(s, t) in Fig. 1. This displacement is a vector, which can be
expressed as u(s, t) = Rd(s,t) − Re(s). Here Re(s) and Rd(s,t) are the equilibrium position vector and dynamic position vector corre­
sponding to the same material point. Shown in the blue dashed box in Fig. 1, the Frenet triad (τ, n, b) is defined at the static equilibrium
configuration. Here, τ, nandb are the tangential, normal and bi-normal directions, respectively. Let eτ, en and eb represent the unit
vectors in the τ,n andb-directions, respectively. The displacement can then be expressed as u(s, t) = u1eτ + u2en + u3eb. Here, u1,
u2andu3 are the displacement components in the tangential, normal, and bi-normal directions, respectively. Fig. 1 also defines the
Cartesian system xyz, and e1, e2 and e3 are the unit vectors corresponding to the x, y, and z-directions, respectively.
The dynamic strain in the equilibrium Frenet triad coordinate system is
[( )2 ( )2 ( )2 ]
∂u1 1 ∂u1 ∂u2 ∂u3
εd = − κu2 + − κu2 + + κu1 + (1)
∂s 2 ∂s ∂s ∂s

ρAgP0
where κ = P 2
+(ρAgds)2
is the equilibrium curvature, P0 is the horizontal component of equilibrium tension, and ds is the arc length along
0

the static equilibrium configuration of the cable.


The kinetic energy T of the cable in χd is
∫s
1
T = ρA Vd · Vd ds (2)
2 0

where ρA is the linear density and Vd = dtd u(s, t) represents the velocity of a cable particle associated with χd.
The potential energy V of the cable in χd comprises the gravitational potential energy Πdg and strain energy Πds of the cable in χd,
which are
∫ s ∫ s
Πdg = Πeg + ρgA u1 l1 ds + ρgA u2 l2 ds (3)
0 0

∫ s( )
1
Πds = Πes + Pεd + EAεd 2 ds (4)
0 2
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
where Πeg ,Πes are the gravitational potential energy and strain energy of the cable in χe. P = P0 2 + (ρAgds)2 is the equilibrium tension,
EA is the axial stiffness, and l1 and l2 represent the components of e2 in τ and n, respectively, which are related as follows: l1eτ + l2en =

Fig. 1. Cable configurations: initial configurationχi, static equilibrium configurationχe, and dynamic configuration χd. Three-dimensional
displacement, u = Rd − Re, refers to equilibrium Frenet triad {et,en,eb}.

31
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

e2.
∑k
The external load and damping force are nonconservative forces, and their work j=1 Qj
ʹq is
j

k 3 ∫ s( )
∑ ∑ ∂uj
Qjʹqj = Fj − μj u ds (5)
j=1 j=1 0 ∂t j

∂u
where Qjʹ = Fj − μj ∂tj , qj = uj, Fj is the external load, and μj is the coefficient of damping.
For non-conservative systems, the Hamiltonian principle is expressed as follows:
∫ t2 ( ∑ k
)
δT − δV + Qjʹδqj dt = 0 (6)
t1 j=1

Substituting Eqs. (1)–(5) into Eq. (6), the nonlinear differential equations of motion are obtained as follows: in the tangential
direction τ:
[ ( )] ( )
∂2 u ∂ ∂u1 ∂u2 ∂u1
ρA 21 = − ρgAl1 + (P + EAεd ) 1 + − κu2 − κ(P + EAεd ) + κu1 − μ + F1 (7)
∂t ∂s ∂s ∂s ∂t

in the normal direction n:


[ ( )] [ ( )]
∂2 u2 ∂ ∂u2 ∂u1 ∂u2
ρA = − ρgAl2 + (P + EAεd ) + κu1 + κ (P + EAεd ) 1 + − κu2 − μ + F2 (8)
∂t 2 ∂s ∂s ∂s ∂t

in the bi-normal direction b:


[ ( )]
∂2 u ∂ ∂u3 ∂u3
ρA 23 = (P + EAεd ) − μ + F3 (9)
∂t ∂s ∂s ∂t
The static equilibrium equations are obtained by setting all time derivatives, uj and Fj to zero:
∂P
− ρgAl1 = 0, κP − ρgAl2 = 0 (10)
∂s
Substituting Eq. (10) into Eqs. (7)–(9) yields
[ ( ) ] ( )
∂2 u ∂ ∂u1 ∂u2 ∂u1
ρA 21 = (P + EAεd ) − κu2 + EAεd − κ(P + EAεd ) + κu1 − μ + F1 (11)
∂t ∂s ∂s ∂s ∂t
[ ( )] [ ( ) ]
∂2 u2 ∂ ∂u2 ∂u1 ∂u2
ρA = (P + EAεd ) + κu1 + κ (P + EAεd ) − κu2 + EAεd − μ + F2 (12)
∂t 2 ∂s ∂s ∂s ∂t
[ ( )]
∂2 u3 ∂ ∂u3 ∂u3
ρA = (P + EAεd ) − μ + F3 (13)
∂t 2 ∂s ∂s ∂t
Consider cable structures with a small sag in static equilibrium, such as bridge suspension and stay cable and contact wire in high-
speed railway, where P0=P and κ = ρAg/P0. The curvature of cable in the static equilibrium can now be considered as a small constant.
By ignoring high-order terms at the small curvature, dynamic strain is ε = ∂u1/∂s − κu2. Let ψ = ∂u2/∂s + κu1 and θ = ∂u3/∂s. The
nonlinear equations of motion are linearized about the static equilibrium configuration of the cable:

∂2 u1 ∂ε μ ∂u1
= a1 − κa4 ψ − 1 + f1 (14)
∂t2 ∂s ρA ∂ t

∂2 u2 ∂ψ μ ∂u2
= a4 + κa1 ε − 2 + f2 (15)
∂t2 ∂s ρA ∂ t

∂2 u3 ∂θ μ ∂u3
= a4 − 3 + f3 (16)
∂t2 ∂s ρA ∂t
√̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅
where a1 = c2t + c2l , a4 = c2t , fj = Fj/(ρA), cl = E/ρ andct = P0 /ρA = g/κ are the classical velocities of longitudinal waves in a rod
and transverse waves in a string, respectively. The first two Eqs. (14) and (15) describe the coupled in-plane response including
longitudinal and transverse response, whereas the third Eq. (16) describes the out-of-plane response. The undamped out-of-plane cable
response can be reduced to a classical taut string wave equation whose solutions are readily available in classical mathematical physics
textbooks. For the damped out-of-plane response, the solutions are also available [25]. In this paper, the solutions for the damped
in-plane response were obtained for the parameters listed in Table 1.

32
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

3. Cables’s characteristics of freely propagating waves

3.1. Frequency of elastic waves in cable

For the in-plane motion (Eqs. (14) and (15)), the free vibration problem (fj = 0) was examined to obtain the relationship between
frequency and wave-number and phase and group velocities of the in-plane waves. By introducing ΦT = {u1,u2,ε, ψ}, Φj (s, t) =
Φj (s)e− iωt , j = 1 ∼ 4 and applying the (spatial) Fourier transform to Eqs. (14) and (15), the eigenvalue problem for the in-plane
response can be obtained:
[ ]
iγI + K1 − I ̃ =0
Φ (17)
iωμ + ω I iγA + K2
2

where Φ
̃ is the Fourier transform of Φ, γ is the propagation constant, and the remaining matrices are as follows:

Table 1
Description of model parameters.
Parameters Description

χi Initial configuration
χe Static equilibrium configuration
χd Dynamic configuration
u(s, t) Displacement vector
Re(s) Equilibrium position vector
Rd(s,t) Dynamic position vector
τ, n, b Tangential, normal and bi-normal directions
eτ, en, eb Unit vectors in τ,n and b
u1, u2, u3 Displacement components in τ,n and b
e1, e2, e3 Unit vectors corresponding to x, y, and z-directions of Cartesian coordinate system
εd Dynamic strain
κ Equilibrium curvature
P, P0 Equilibrium tension and its horizontal component, P0=26.460 kN
T Kinetic energy of cable in χd
ρA Linear density of cable, ρA=1.07 kg/m
Vd Velocity of a cable particle associated with χd
V Potential energy V of cable in χd
Πdg , Πds Gravitational potential energy and strain energy of cable in χd
Πeg , Πes Gravitational potential energy and strain energy of cable in χe
EA Axial stiffness of cable (elastic modulus E = 130 GPa, cross-sectional area A = 120 mm2)
l1 , l 2 Components of e2 in τ and n, respectively, l1eτ + l2en = e2
∑k
Work of external and damping forces
j=1 Qj
ʹq
j
Fj, fj External load, fj = Fj/(ρA)
μj Coefficient of damping
ε, ψ, θ Components of vector ∂u/∂s in τ, n and -directions, ε = ∂u1/∂s − κu2, ψ = ∂u2/∂s + κu1, θ = ∂u3/∂s
√̅̅̅̅̅̅̅̅
cl Longitudinal wave velocities in rod cl = E/ρ
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅
ct Transverse wave velocity in string ct = P0 /ρA = g/κ
( ) ( )
a1, a2, a3, a4, a5 Coefficients in equations of motion, a1 = ct + cl , a4 = c2t , a2 = 2c2t + c2l κ, a3 = c2t κ2 and a5 = c2t + c2l (κ)2
2 2

Φ(s,t) Functions of displacement at (s, t)


̃
Φ(s), Φ(γ) Functions of displacement only related to position s and its Fourier transform
γ Propagation constant
ω Frequency
β1′, β2′, β3′, β4′ Coefficients in frequency equation
β1, β2, β3 Coefficients in equation of propagation constant/dispersion equation
ωcf1, ωcf2, ωcf3, ωcf4 four foots of the cut-off frequency
d Band gap
c Phase velocity
ε, ν, λ Perturbation parameter
A
̂1 , Â2 , A
̂3 , A
̂4 Coefficients in perturbation equations
ξ Damping ratio, ξ = 0.01
ωn nth-order natural angular frequency
fn nth-order natural frequency
A Amplitude matrix formed by four frequency components of u1
Ω Conversion matrix between initial conditions u
̃ 0 and amplitude A
rj Coefficient related to u2
u
̃0 Fourier transform of initial conditions
λ0 Wavelength of cosine pulse, λ0 = 20 m
Ac Amplitude of cosine pulse, Ac = 10 mm
γa, ωja, Aja(γ), Bja(γ) Imaginary part of γ, ωj, Aj(γ), Bj(γ)
γb, ωjb, Ajb(γ), Bjb(γ) Real part of γ, ωj, Aj(γ), Bj(γ)

33
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

[ ] [ ] [ ] [ ] [ ]
1 0 a 0 0 − κ 0 − κa4 1 μ1 0
I= ,A = 1 , K1 = , K2 = ,μ =
0 1 0 a4 κ 0 κa1 0 ρA 0 μ2
For non-trivial solutions, one must set determinant of the coefficient matrix equal to zero, resulting in the following characteristic
equation

Q(ω; γ, cl , ct ) = ω4 + β3ʹω3 − β1ʹω2 − β4ʹω + β2ʹ = 0 (18)

where
μ1 μ2 ʹ ( 2 )( )
β1ʹ = (a1 + a4 )γ2 + a3 + a5 +· , β = a1 γ + a3 a4 γ 2 + a5 − a22 γ 2 ,
ρA ρA 2
(μ μ ) [μ μ (μ μ ) ]
β3ʹ = i 1 + 2 , β4ʹ = i 1 a5 + 2 a3 + 2 a1 + 1 a4 γ 2
ρA ρA ρA ρA ρA ρA
( ) ( )
with a2 = 2c2t + c2l κ, a3 = c2t κ2 and a5 = c2t + c2l (κ)2 . By solving Eq. (18) in the main amplitude angle, four roots ω1, ω2, ω3 and ω4
can be obtained, representing four groups of propagating wave frequencies along different directions, including two groups of lon­
gitudinal wave frequencies and two groups of transverse wave frequencies propagating in the positive and negative directions,
respectively. The relationships between the propagation constant and frequencies can be obtained from Eq. (18). It can be seen that the

Fig. 2. In-plane wave frequencies (a: wave propagation constant curves without damping for various curvatures; b: wave propagation surfaces for
damping ratio ξ = 0.01 and curvature κ = 0.005).

34
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

relationships are affected by the cable curvature, material properties and damping.
As shown in Fig. 2(a), the relationship between frequency and wave-number without damping is clear in a two-dimensional plane,
where the stopband appears as a straight line segment. However, there are always imaginary terms in the frequency equation when
considering damping. Therefore, one must investigate frequency in the complex plane. For this reason, the examination of the rela­
tionship between frequency and wave-number must shift from a two-dimensional space to a three-dimensional space. As shown in
Fig. 2(b), the stopband will transform into a quadrilateral and the relationship between frequency and wave-number will transform
into a three-dimensional surface after taking damping into account. It is worth noting that the stopband still appears as a line segment
located on the real axis of frequency when examining the relationship between frequency and wave-number of undamped cable in the
complex plane.
The elastic wave propagation forms are influenced by the stopband and passband separated by the cut-off frequencies. At the
stopband, the response only oscillates in time but no wave propagates in space, i.e., it is a standing wave. At the passband, a wave can
propagate in space, i.e., it is a traveling wave. Mathematically, the cut-off frequencies satisfy Re[γ] = 0. To obtain the cut-off frequency
of the in-plane waves in cable, the frequency Eq. (18) was converted into the equation in the propagation constant g:

β1 γ 4 − β2 γ 2 + β3 = 0 (19)

where
β1 = a1 a4

μ1 ) 2
β2 = (a1 + a4 )ω2 + i a4 ω + 2a1 a3
a1 +
ρA ρA
(μ μ ) ( μ μ ) (μ μ )
β3 = ω4 + i 1 + 2 ω3 − a3 + a5 + 1 · 2 ω2 − i 1 a5 + 2 a3 ω + a3 a5 ,
ρA ρA ρA ρA ρA ρA
For any non-zero complex, its argument is infinite. Therefore, one must limit the arguments of complex solutions to the principal
argument. Other solutions can be obtained by adding an integer multiple of 2π to the basic solution arguments. Four solutions for γ can
be obtained by solving Eq. (19) in the range of principal argument angle. Hereγ1,2 = ±γl are the propagation constants related to
longitudinal waves and γ3,4 = ±γt are the propagation constants related to transverse waves, where

⎛ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅⎞1
2
β2 − β2 2 − 4β1 β3
γl = ⎝ ⎠ (20)
2β1

⎛ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅⎞1
2
β2 + β2 2 − 4β1 β3
γt = ⎝ ⎠ (21)
2β1

√̅̅̅̅ (θ/2)
We will now discuss γ2 at the cut-off frequencies. Represent γ2 in the complex form γ2 = Reiθ, then, γ = Rei , where R is
√̅̅̅̅
magnitude and θ is the argument. At the cut-off frequencies, Re[γ] = Rcos[θ /2] = 0. Two solutions are possible: R = 0 or cos[θ/2] =
0. The former is a special case of cut-off frequencies, which is discussed in detail in Section 4.1.1. The latter leads to sinθ = 0, thus
Im[γ2 ] = Rsinθ = 0. Accordingly, there are only real parts in γ2 at the cut-off frequencies. Additionally, coefficient β1 is a positive real
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
number. Hence, only real parts exist onβ2 ± β2 2 − 4β1 β3 at cut-off frequencies.
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
When β2 + β2 2 − 4β1 β3 is a real number, β2 + β2 2 − 4β1 β3 > 0 is always satisfied. Hence, transverse waves do not have a cut-
off frequency. On the other hand, the longitudinal waves at the cut-off frequency satisfy the following:
[ ( √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅)12 ]
1
Re[γl ] = Re √̅̅̅̅̅̅̅̅ β2 − β2 2 − 4β1 β3 =0 (22)
2β1

which can be simplified to


[( √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅)12 ]
Re β2 − β2 2 − 4β1 β3 =0 (23)

√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
At the cut-off frequencies, only the real parts exists in β2 − β2 2 − 4β1 β3 and Eq. (23) can now be simply stated as

β3 = 0 (24)
The four roots obtained by solving Eq. (24) are as follows:

35
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

( √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅)
1 μ ( μ )2
ωcf1 = − i 1−4a3 − 1
,
2 ρA ρA
( √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅)
1 μ ( μ )2
ωcf2 = − i 1 + 4a3 − 1
,
2 ρA ρA
( √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅) (25)
1 μ2 ( μ )2
2
ωcf3 = − i − 4a5 − ,
2 ρA ρA
( √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅)
1 μ ( μ )2
ωcf4 = − i 2 + 4a5 − 2
,
2 ρA ρA
√̅̅̅̅̅ √̅̅̅̅̅ √̅̅̅̅̅ √̅̅̅̅̅
For an undamped cable, Eq. (25) simplifies to ωcf1 = a3 , ωcf2 = − a3 , ωcf3 = a5 and ωcf4 = − a5 , where all four cut-off
frequencies are located on the real axis. ωcf1 and ωcf2 are the first cut-off frequencies, whereas ωcf3 and ωcf4 are the second cut-off
frequencies. The stopbands extend between the first cut-off frequencies and the second cut-off frequencies, while the remaining re­
√̅̅̅̅̅ √̅̅̅̅̅
gions are the passbands. The length of the stopband is the bandgap, which can be calculated as d = |ωcf3 − ωcf1 | = a5 − a3 for the
undamped cable. For damped cable, the changes in cut-off frequencies and stopband are shown in Fig. 3, where the scattered points are
cut-off frequencies for different damping values and the direction of arrows represents the direction of increasing damping. It can be
seen in Fig. 3 that the cut-off frequencies ωcf1, ωcf2, ωcf3 and ωcf4 move away from the real axis towards the imaginary axis with the
increase in damping. The four cut-off frequencies are used to form a plane in the complex plane, with the stopband inside and the
passband outside.

3.2. Phase and group velocities

Let c = ω/γ denote the phase velocity, the frequency Eq. (18) yields the dispersion equation:

D(c; ω, cl , ct ) = β3 c4 − β2 ω2 c2 + β1 ω4 (26)
Eq. (26) provides four phase velocities c1,2 = ±cʹl and c3,4 = ±cʹt , where
{( √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅)/ }1/2
clʹ = ω β2 − β22 − 4β1 β3 2β3 (27)

{( √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅)/ }1/2
ctʹ = ω β2 + β22 − 4β1 β3 2β3 (28)

are complex in general. In particular, the dispersion Eq. (26) is singular when the propagation constant tends to zero (γ → 0). Hence, a
special solution is needed for the phase velocity in the vicinity of cut-off frequencies. Here, the perturbation method is used to obtain it.
Let ωl = ωcf ± ε by introducing a small perturbation parameter ε and then substitute this into Eq. (26) and retain only the first-order
term:

εA ̂2 + ε A
̂1 c4 + ( A ̂4 + ε A
̂3 )c2 + ( A ̂5 ) = 0 (29)

Fig. 3. Variation in cut-off frequency and stopband with damping.

36
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

where
( )
̂1 = 4ωcf1 3 + 3i μ1 + μ2 ωcf1 2
A
ρA ρA
( μ μ ) ( μ μ )
− 2 a3 + a5 + 1 · 2 ωcf1 − i a3 2 + a5 1
ρA ρA ρA ρA
( )
̂2 = − (a1 + a4 )ωcf1 4 − i a4 μ1 + a1 μ2 ωcf1 3 − 2a1 a3 ωcf1 2
A
ρA ρA
( )
̂3 = − 4(a1 + a4 )ωcf1 3 − 3i a4 μ1 + a1 μ2 ωcf1 2 − 4a1 a3 ωcf1
A
ρA ρA

̂4 = a1 a4 ωcf1 4 , A
A ̂5 = 4a1 a4 ωcf1 3

The highest order of the characteristic Eq. (29) in c corresponds to the highest order of the differential equation and it only contains
small parameters thus, the corresponding differential equation describes a boundary type singular perturbation problem. It can be seen
that when the small parameter ε → 0, Eq. (29) simplifies from a quartic equation to a quadratic equation. Then, the corresponding
differential equation is changed from the fourth order to the second order. At this time, the solution of the phase velocity becomes
discontinuous, and two solutions can be obtained by regular perturbation, while the other two solutions by the singular perturbation
method.
In order to obtain the other two solutions, the transform c2 = εvλ is introduced and substituted into Eq. (29) to obtain
( )
A
̂1 ε1− v + A
2λ 2 ̂3 ε1− λ v + ( A
̂2 ε− λ + A ̂4 + ε A
̂5 ) = 0 (30)

Extract the dominant terms and apply the dominant balance

A
̂ 1 ε1− v +A
2λ 2 ̂ 2 ε− λ v = 0 (31)

where ε1 − 2λ
~ ε− λ. Then, 1 − 2λ = − λ, i.e., λ = 1, can be obtained. By substituting λ = 1 into Eq. (30), one obtains
( )
A
̂1 v2 + ( A ̂3 ε)v + ε A
̂2 + A ̂5 = 0
̂4 + ε2 A (32)
2
Now, expand v as regular perturbation v = v0 + εv1 + ε v2 + …. Substituting this expansion into Eq. (19) and collecting terms of
same order yields

Fig. 4. Phase velocity of transverse waves compared to field test results.

37
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

/
ε0 : A
̂ 1 v2 + A
0
̂ 2 v0 = 0→v(1) = 0,
0 0 = − A2 A1
v(2) ̂ ̂
( )/( )
ε1 : (2 A
̂ 1 v0 + A
̂ 2 )v1 + ( A
̂ 3 v0 + A
̂ 4 ) = 0→v(i) = − A
1
̂ 3 v(i) + A
0
̂4 2A
̂ 1 v(i) + A
0
̂2
(33)
( )2
( ) A
̂ 1 v(i) + A
1
̂ 3 v(i) + A
1
̂5
ε2 : ̂ 1 v2 + 2v0 v2 + A
A 1
̂ 2 v2 + A
̂ 3 v1 + A
̂ 5 = 0→v(i) = −
2
2A
̂ 1 v(i) + A
0
̂2

Thus, the third and fourth roots are


/
c1 2 = v(2) ε = v(2)
0 ε
− 1
+ v(2)
1 + εv2 + …
(2)

/ (34)
c2 2 = v(1) ε = v(1)
0 ε
− 1
+ v(1)
1 + εv2 + …
(1)

Substituting Eq. (33) into Eq. (34) yields

c1 2 = ( − A
̂ 2/A
̂ 1 )ε− 1 + ( − A
̂ 3/A
̂1 + A
̂ 4 /A
̂ 2 ) + v(2) ε + …
2
/ ( [ ]/ 3 ) (35)
c2 2 = − Â4 Â2 − ̂2 − A
̂1A
A ̂2Â3A
̂4 + A ̂2 A
̂5A ̂ ε+…
4 2 2

It can be seen that when parameter ε → 0, the second solution in Eq. (35) approaches the constant − A ̂ 2 , but the first solution c21
̂ 4/A
becomes singular. To verify the solution for phase velocities, we compared the phase velocity solutions of the transverse waves in
elastic cable with the results of the field experiment for a railway contact wire [27,49]. Introduce the relation μ/ρA = 2ξωn to include
damping, where it is assumed that the damping coefficients ξ were equal to 0.01 for all directions [30]. ωn = 2πfn represents the
nth-order natural angular frequency, where fn is the nth-order natural frequency. Generally, the frequency range of concern for
pantograph-catenary dynamics does not exceed 20 Hz [50,51][51], whereas the dominant frequency of contact wire is from 1.02 Hz to
1.2 Hz [15]. In this paper, fn is assumed to be 10 Hz to clarify the influence of damping. After transforming to ωn = 2πfn, the results
below 20 Hz are shown in the Fig. 4. It can be seen that the phase velocity predicted by Eq. (28) is higher than the results of field tests.
The relative errors between the theoretical predictions and the tests are listed in Table 2, where it can be seen the model considering
damping produced more accurate wave velocities than those without damping.
The group velocity can be obtained by differentiating implicitly the frequency equation:

dω (2β1 γ2 − β2 )γ
cg = = (36)
dγ [4ω + 6β3ʹω2 − 2β5ʹω − β4ʹ]
3

( )
where β5ʹ = (a1 + a4 )γ2 − 2 a3 + a5 + μ1 μ2
ρA · ρA.
Eq. (36) provides four group velocities cg|1,2 = ±cgl and cg|3,4 = ±cgt since there are four solutions to γ. Here

(2β1 γ l 2 − β2 )γl
cgl = (37)
[4ω + 6β3lʹω2 − 2β5lʹω − β4lʹ]
3

(2β1 γt 2 − β2 )γ t
cgt = (38)
[4ω3 + 6β3tʹω2 − 2β5tʹω − β4tʹ]

where β3l′, β4l′, β5l′, β4t′ and β5t′ are coefficients related to β3′, β4′ and β5′, which use γland γt rather than γ.

4. Cable response to initial cosine displacement

Consider now the in-plane free response solution for an infinite cable subject to an initial disturbance. The initial conditions are
ui(s,0) and ui,t(s, 0), i = 1, 2. The four solutions ω1, ω2, ω3 and ω4 will generate four propagating waves, two of which are transverse
waves and the remaining two are longitudinal waves. These can be superimposed to form the following responses:

Table 2
Relative errors in wave propagation velocities between proposed theoretical method and experiments.
Experiment (m/s) Eq. (15) Relative error (%) Eq. (15) Relative Reference
(ξ = 0) (m/s) (ξ = 0.01) (m/s) error (%)

1.200 138.9033 154.7794 11.43 153.7851 10.71 [27]


1.207 137.9920 154.8043 12.18 153.8197 11.47 Case 1–1 in [27]
1.199 139.5259 154.7727 10.93 153.7761 10.21 Case 3–1 in [27]
1.235 144 154.9104 7.57 153.9661 6.92 Field test for Jing Jin interurban railway [27]
1.189 138.87 154.7344 11.42 153.7232 10.69 [49]

38
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

4

u1 (γ, t) =
̃ Aj (γ)e− iωj t
(39)
j=1

4

u2 (γ, t) =
̃ rj (γ)Aj (γ)e− iωj t
(40)
j=1

Applying the inverse Fourier transform to Eqs. (39) and (40) leads to the general solutions [25]
∫ ∞ 4 ∫
∑ ∞
u1 (s, t) = u1 (γ, t)eiγs dγ =
̃ Aj (γ)ei(γs− ωj t )
dγ (41)
− ∞ j=1 − ∞

∫ ∞ 4 ∫
∑ ∞
u2 (s, t) = u2 (γ, t)eiγs dγ =
̃ rj (γ)Aj (γ)ei(γs− ωj t )
dγ (42)
− ∞ j=1 − ∞

where rj and Aj(γ), which depend on frequencies and initial conditions, can be represented as follows:
( μ )/
rj (γ) = − a1 γ2 − a3 + ω2j + iωj 2 ia2 γ
ρA
⎤ ⎡
⎡ ⎤ ⎧ ⎫
A1 1 1 1 1 ⎪
⎪ u10 ⎪
̃ ⎪
⎨ ⎬
− 1
⎢ A ⎥
2⎥
⎢ r1 r2 r3 r4 ⎥ u20
̃
A = Ω u0 = ⎣ ⎦, hereΩ = ⎣
̃ ⎢ ⎢ ⎥, u
̃0 =
A3 − iω1 − iω2 − iω3 − iω4 ⎦ ⎪̃u ⎪
⎩ 10,t ⎪
⎪ ⎭
A4 − iω1 r1 − iω2 r2 − iω3 r3 − iω4 r4 u20,t
̃

Consider the case of an initial cosine displacement applied in the transverse direction only in the region − λ0/2 < s < λ0/2. Then the
associated Fourier transforms are:

u10 = ̃
̃ u10,t = ̃
u20,t = 0,
1 1
u20 =
̃ Γ(γ) + [Γ(γ + 2π /λ0 ) + Γ(γ − 2π/λ0 )],
2 4 (43)
( ) ( )/( )
Ac λ0 λ0 γ λ0 γ
Γ(γ) = sin
π 2 2 2

where λ0 = 20m and Ac = 10mm are the wavelength and amplitude of the cosine pulse.
By substituting Eq. (43) into Eqs. (41) and (42), the in-plane response solution can be obtained.

4.1. Cable response at stopband and passband

The frequency analysis indicated that the in-plane longitudinal waves had a clear stopband and passband. The undamped stopband
was a line segment whose width increased with curvature. On the other hand, the damped stopband was a quadrilateral region, the
area of which was related to damping and curvature. On the contrary, the frequency analysis for the transverse waves suggested no
stopband or passband. Based on the frequency analysis for in-plane waves, the propagation of cable waves in stopband and passband
are investigated in this section.
Assuming the wave-number is γ, the responses predicted by Eqs. (41) and (42) are
4

u1 (s, t, γ) = Aj (γ)ei(γs− ωj t )
(44)
j=1

4

u2 (s, t, γ) = rj (γ)Aj (γ)ei(γs− ωj t )
(45)
j=1

Let Bj(γ) = rj(γ)Aj(γ) and represent complex γ, ωj, Aj(γ) and Bj(γ) as γ = γa + iγb, ωj = ωja + iωjb, Aj(γ) = Aja(γ) + iAjb(γ) and Bj(γ) =
Bja(γ) + iBjb(γ), where γa, ωja, Aja(γ) and Bja(γ) are the real parts, whereas γb, ωjb, Ajb(γ) and Bjb(γ) are the imaginary parts. The real parts
of responses are
4
∑ [ ( ) ( )]
Re[u1 (s, t, γ)] = e− e
|γb s| ωj b t
Aja cos γ a s − ωja t − Ajb sin γa s − ωja t (46)
j=1

4
∑ [ ( ) ( )]
Re[u2 (s, t, γ)] = e− e
|γb s| ωj b t
Bja cos γa s − ωja t − Bjb sin γ a s − ωja t (47)
j=1

39
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

Eqs. (46) and (47) are the components of responses at the passband. It is worth noting that only when the imaginary part of a
frequency satisfies ωjb ≤ 0, can a meaningful solution that converges in time be obtained.
At the stopband, γa = 0 and Eqs. (46) and (47) transform into
4
∑ [ ( ) ( )]
Re[u1 (s, t, γ)] = e− e
|γb s| ωj b t
Aja cos ωja t + Ajb sin ωja t (48)
j=1

4
∑ [ ( ) ( )]
Re[u2 (s, t, γ)] = e− e
|γb s| ωj b t
Bja cos ωja t + Bjb sin ωja t (49)
j=1

It can be seen that the responses at the stopband oscillate in time and decay exponentially with the position in space, i.e. they
describe standing waves.
Now, we apply Eqs. (46)–(49) to the overhead contact wire of a high-speed railway. Here, the elastic modulus E is 130 GPa, line
density ρA 1.07 kg/m, cross-sectional area A 120 mm2, initial horizontal tension P0 26.460 kN [27] and curvature κ 0.000,4. The
damping coefficient ξ is assumed as 0.01, consistent with Section 3.

4.1.1. Response at stopband


The responses at stopband predicted by Eqs. (48) and (49) are spatially varying but non-propagating. Two possibilities exist at
stopbands, namely, γb ∕ = 0 and the other is γb = 0. For the former, the in-plane responses are standing waves. As shown in Figs. 5 and 6
(a), the time histories are cosines curves, the frequencies of which are related to the cut-off frequency. As shown in Figs. 5 and 6(b), the
waves attenuate as e− |γb s| with the distance from the origin. For the latter possibility, the longitudinal response is zero, while the
transverse response oscillates in time and has an equal amplitude everywhere in space as shown in Fig. 7. This latter case can be seen as
a special case of the former one when the attenuation factor in space is infinitely close to 1. Similar conclusions from a study on taut
strings are reported in reference [24], where a taut string at the stopband vibrated essentially as a simple spring-mass system with no
spatial variation.
The dotted curves plotted in Figs. 5–7 are the periodic responses of the undamped cable with a frequency related to the cut-off
frequency. Within the observed time range, it can be observed that the time history of the undamped transverse response has a
period of T = 4,110 ms. For the undamped cable, the cut-off frequency is ωcf3 = 0.0626 rad/s and the corresponding periods are Tcf1 =
2π/ωcf1= 4,108 ms and Tcf3 = 2π/ωcf3=100,319 ms. Comparing the above results, it can be found that the observed period T is related
to the cut-off frequency ωcf1. The solid lines in Figs. 5–7(a) are the responses of the damped cable. According to Eq. (25), the damped
cut-off frequency contains an imaginary term, thus the periodicity of responses is disrupted and their amplitude decays continuously.
When the amplitude decays to zero, the cable no longer vibrates. Figs. 5 and 6(b) investigate the cases where γ = 0.0001i. At this wave-
number, the waves attenuate in space slowly and it only decays to zero after reaching several tens of kilometers. The waves will decay
faster in space when the imaginary part of the wave-number increases. However, solutions that do not converge in time will be ob­
tained when the imaginary part of frequency ωjb is large. Accordingly, the imaginary part of the wave-number is limited by the
dispersion relationship and cannot be taken too large.
The time history curves for γb = 0 are plotted in the Fig. 7(a), in which there exist four legends but only two curves. This is because
the pink and green plots coincide. For clarity, the green lines are placed as small images in the top right-hand corner of Fig. 7(a). The
overlapping time histories at different positions indicate that the cable vibrates with the same amplitude throughout the space for γb =
0, as shown in Fig. 7(b).

Fig. 5. Time history (a) and waveform (b) for longitudinal displacement in stopband (γ = 0.0001i).

40
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

Fig. 6. Time history (a) and waveform (b) for transverse displacement in stopband (γ = 0.0001i).

Fig. 7. Time history (a) and waveform (b) for transverse displacement in stopband (γ = 0).

4.1.2. Response at passband


The in-plane waves in the passband were studied based on Eqs. (46) and (47) and the results are shown in Figs. 8 and 9, where can
be seen that the response with damping was lower than that without damping. The damping causes response attenuation, and the
attenuation ratio increases with time. The in-plane waves (both longitudinal and the transverse) were cosine with wavelengths equal to
2π/γ. The time histories of the longitudinal and the transverse responses differed markedly.
It can be seen in Fig. 8(a) that the time history of the longitudinal response comprises the two clear periods. Taking γ = 0.2and ξ =
0 as an example, according to Eq. (18), two positive frequencies ω1 = 764.3036 rad/s and ω3 = 31.3048 rad/s exist. The time history
shown in Fig. 8(a) contains two main periods, denotes as TP1 and TP2, respectively, where, TP1 = 200 ms is related to the frequency of
the transverse component ω3 and TP2= 8 ms to the longitudinal component ω1. For the same propagation constant, the small waves
related to longitudinal wave components change rapidly with a high frequency, while the transverse wave components change slowly
with a low frequency and form the envelope of the small waves. The transverse response is different from the longitudinal response.
The time histories in Fig. 9(a) show that the transverse responses are dominated by the transverse wave components and only their
amplitudes are attenuated by the longitudinal wave components.
For the longitudinal response, A1a and A3a are of the same magnitude, hence, both frequency ω1 related to longitudinal waves and
frequency ω3 related to transverse waves are evident in Fig. 8(a). For the transverse response, one always has r1 ≪ r3, which makes the
transverse wave components dominate the final results in Fig. 9(a).
The waveforms for γ = 0.2 at different times are shown in Figs. 8(b) and 9(b). It can be seen that the waves oscillate in space with
the wavelength λ = 2π/γ ≈ 31.4m. The four small figures on the top illustrate the wave components, where it can be seen that the four
components are equivalent for the longitudinal waves, while only u23 and u24 are dominant in transverse waves. The four component
waveforms vary with time, which indicates that the components at the passbands will propagate in space. As shown on the left of
Figs. 8(b) and 9(b), the superposition of components results in the formation of standing waves. In this case, only the amplitude
fluctuates but the waveforms do not propagate in space.

41
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

Fig. 8. Time histories curve (a) and waveforms (b) of longitudinal displacement in passband.

4.2. Cable response in spatial domain

The free vibration responses of cable in the spatial domain discussed in the authors’ previous work [25] are adopted in this section,
see Eqs. (41) and (42). Similar to reference [25], the considered railway contact line is also linear elastic. Similar conclusions can also
be drawn comparing to reference [25]: Firstly, curvature causes high dispersion of in-plane longitudinal waves, resulting in faster
propagation of their leading edges and slower propagation of the trailing edges, as shown in Fig. 10. Secondly, the waveform of the
in-plane longitudinal waves is anti-symmetric about the origin, as shown in Fig. 10, while the waveform of the in-plane transverse
waves is symmetrical about the origin, as shown in Fig. 12(a) and (b). Thirdly, damping causes the non-zero response at the trailing

42
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

Fig. 9. Time histories: (a) and waveforms (b) of transverse displacement in passband.

edge endpoint of the longitudinal waves, as shown in Fig. 11(c). In addition, the amplitude of damped in-plane transverse waves
decreases with time, as shown in Fig. 12(c) and (d) and Fig. 13(b).
The propagation of waves will vary with different material properties even for the linear elastic constitutive model. This study
compared the free vibration response of railway contact wire and steel suspension cable. The results show that the longitudinal
response of steel suspension cable was larger than that of the railway contact wire. The steel suspension cable had higher longitudinal
wave velocities than the railway contact wire, but their transverse wave velocities were lower. The waves in steel suspension cables
were attenuated faster than in the railway contact wire.
The responses at the stopband and passband of the cables indicated that, for the same propagation constant, the longitudinal

43
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

Fig. 10. Contour plots of longitudinal displacement (a: steel suspension cable, b: overhead contact line).

Fig. 11. Time history and waveform for longitudinal displacement (a: time history at s = 20 m; b: waveforms at 2 ms, 6 ms, 3 ms and 8 ms; c:
waveforms at 64 ms, 192 ms, 74 ms and 204 ms).

response contained both transverse and longitudinal wave components, while the transverse responses were dominated by longitu­
dinal wave components. Similar observations can directly be made in the response contour maps. Fig. 10 shows the contour plots of the
in-plane longitudinal waves within 0–20 ms, where Fig. 10(a) and (b) are applicable to the steel suspension cable and railways contact

44
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

Fig. 12. Contour plots of transverse displacement (ace: steel suspension cable, bdf: railway contact wire; ab: contour plots; cd: contour plots of
amplitude, ef: contour plots of wave edge).

45
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

Fig. 13. Time history and waveform of transverse displacement (a: time history at s = 20 m; b: waveform at 64 ms, 192 ms, 74 ms and 214 ms).

wire, respectively. The horizontal axis is the position measured from the origin and the vertical axis is time. The colored region is the
propagation range of the waves and the color of the contour plot represents the response magnitude. The inverse of the contour plot
slope represents the speed of wave propagation (c = ds/dt), Thus, the smaller the slope, the faster the wave velocity. The colors in the
contour map represent the response amplitude. It can be seen in Fig. 10 that there are fast propagating leading edges and slow
propagating trailing edges, where the wave velocity of the leading edge is related to the longitudinal wave components and the wave
velocity of the trailing edge is related to the transverse wave components. By comparing Fig. 10(a) and (b), the longitudinal wave
velocity at the leading edge of the steel suspension cable (close to 5,000 m/s) can be confirmed as significantly higher than that of the
railway contact wire (close to 4,000 m/s). In addition, the deep red and dark blue parts of the contour plot for the longitudinal response
of the steel suspension cable are more pronounced than those of the contact wire, which means the longitudinal response of the steel
suspension cable was stronger than that of the railway contact wire.
The propagation of longitudinal waves in the cable plane consists of two stages: the first stage is the formation and propagation
stage, and the second stage is the slow propagation stage. To present the propagation characteristics of the longitudinal waves more
clearly, the position at s = 20 m was selected and the corresponding time histories were drawn. Several representative time instances
were also selected from the first and second stages of longitudinal wave propagation to draw the corresponding waveforms as shown in
Fig. 11. It can be observed that the material properties affected the longitudinal responses at the selected time intervals at both stages.
The longitudinal response of the steel suspension cable is significantly larger than that of the railway contact wire. For the position s =
20 m, the longitudinal wave component of the steel suspension cable arrived at T1S=2 ms and left at T2S=6 ms, while the corresponding
times for the railway contact wire were T1O=3 ms and T2O=8 ms, respectively. The transverse wave component of the steel suspension
cable arrived at T3S=74 ms and left at T4S=214 ms, while the corresponding times for the railway contact wire were T3O=64 ms and
T4O=192 ms, respectively.
Taking the railway contact wire as an example, we further investigated the propagation of longitudinal waves using Fig. 11(a)–(c).
The wave propagation characteristics in steel suspension cables are similar to those in the railway contact wire but the time of each
period is different. It can be seen in Fig. 11(a) that at the position s = 20 m, longitudinal waves formed and the leading edge propagated
rapidly before T2O=8 ms. The trailing edges of longitudinal waves propagated from T3O=64 ms to T4O=192 ms. The time from T2O = 8
ms to T3O=64 ms was the transition stage. During this period, the leading edges have left the positions s = 20 m, but the trailing edges
have not yet propagated to s = 20 m. After T4O=192 ms, the endpoints of the trailing edges left the position s = 20 m. At this time, the
response of the undamped cable here returned to zero, while the response of the damped cable did not. After adding damping, T1O, T2O,
T3O and T4O hardly changed but the response magnitudes did.
Fig. 11(b) and (c) shows the waveforms during the first and second stages of longitudinal wave propagation, respectively. It can be
seen that during the period of the formation of longitudinal waves the amplitudes of the waves increased continuously. Then the
leading edges of the waves began to propagate rapidly when the amplitudes of the waves stabilized. As time passed, the leading and
trailing edges of the waves began to separate, resulting in changes in the waveforms and clear dispersion. Observing the wave
propagating forward along the cable by comparing the waveforms shown in Fig. 11(b) and (c) to those in Fig. 11(a), it can be seen that
the endpoint E3 of leading edge R2 reached s = 20 m at T1O=3 ms. The time history increased from zero in Fig. 11a(a). The crest E2 of
the leading edge R2 reached s = 20 m at T2O=8 ms, and the longitudinal response at s = 20 m reached its maximum value. After this
time, the leading edge R2 left s = 20 m, but the longitudinal response at s = 20 m remained unchanged for some time since the trailing
edge R1 did not arrived at s = 20 m until T3O=64 ms. At time T3O=64 ms, the end point E2 on the trailing edge R1 reached s = 20 m,
and thus the longitudinal response began to decrease. At time T4O=192 ms, the end point E1 of the trailing edge R1 left s = 20 m, which
ended the response temporal fluctuations at s = 20 m. As can be seen in Fig. 11(b), the leading edge endpoint E3 located at s = 20 m at
time T1O=3 ms propagated to s = 40 m at T2O=8 ms, i.e., the leading edge of the wave rapidly propagated along the cable at a speed of
4,000 m/s. From Fig. 11(c), it can be seen that the trailing edge endpoint E1 located at s = 20 m at time T3O= 64 ms reached s = 40 m at
time T4O=192 ms, i.e., the trailing edges of the longitudinal waves propagated slowly at a speed of 156.3 m/s. It can also be seen from

46
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

Fig. 11(c) that after adding damping, the responses of the trailing edge endpoints were not equal to zero, and the farther away from the
origin, the greater the response value of the trailing edge endpoint. The response the of trailing edge endpoint E1 increased from
0.0014 mm at s = 20 m at time 192 ms to 0.002 mm at s = 29 m at time 250 ms.
The contour plots of the in-plane transverse waves in the steel suspension cable and railway contact wire are shown in Fig. 12. It can
be seen from the slopes of the contour plots that the leading and trailing edge of the transverse waves always propagated along the
cable with the same wave velocity and that the transverse waves hardly dispersed during propagation. Below the dotted line in Fig. 12
(a) and (b), a of yellow to deep red semicircle can be seen, which corresponds to the first stage of transverse response propagation,
when the transverse waves has not completely separated into two single waves. However, above the dotted line in Fig. 12(a) and (b),
the separated single wave propagated away with the same velocity for the leading and the trailing edge wave. The wave crests are
plotted in Fig. 12(c) and (d). The colors become lighter with time, which means the crest of the transverse waves decayed with time. By
comparing Fig. 12(c)–(d), it appears that the attenuation of the wave crest in the steel suspension cable was faster than that in the
railway contact wire: at 262 ms, the wave crest in the steel suspension cable decreased from 5 mm to 4.15 mm (17.0 %), while the wave
crest in the contact wire decreased from 5 mm to 4.22 mm (15.6 %). In Fig. 12(e) and (f), the color of the trailing edge and the leading
edge of the waves are different, which indicates that the response to the single wave is asymmetric about the wave crest. By comparing
Fig. 12(e) and (f), it can be concluded that the dark blue region for the steel suspension cable was smaller than that for the railway
contact wire, indicating that the response at the edge of the steel suspension cable was larger than that of the railways’ contact wire. In
Fig. 12(a)–(d), the slope of the transverse wave contour plot for the steel suspension cable is larger than that of the railway contact
wire. By calculating the inverse of the slope of the contour plot of the same color, the wave velocities in the steel suspension cable of
142.9 m/s and 156.3 m/s in the railway contact wire were obtained.
Fig. 13 shows the transverse wave time history at s = 20 m and the waveform propagating forward. It can be seen that the material
properties hardly affected the magnitude of the transverse response, but it had a significant impact on the times of arrival and leaving
of transverse wave. The transverse wave in the steel suspension cable arrived at s = 20 m at T3S=74 ms and left at T4S=214 ms, while
the corresponding times for the railway contact wire were T3O=64 ms and T4O=192 ms, respectively.
Taking the railway contact line as an example, we examined the propagation of transverse waves using Fig. 13(a)–(c). Fig. 13(a)
shows the time history of the in-plane transverse waves at s = 20 m, two representative time points, T3O=64 ms and T4O=192 ms, can
be seen. The in-plane transverse waves gradually separated from the initial cosine displacement before time T3O. The second stage
began at time T3O when the initial cosine displacement had separated into two independent waves. Then the leading and trailing edge
of the independent waves propagated along the cable in positive and negative directions with the same wave velocity. Only the wave
propagate in positive direction was shown in Fig. 13. (b). The times T3O and T4O were nearly the same after considering damping, but
the response changed.
Fig. 13(b) and (c) shows the waveforms of the transverse response in the first and second stages, respectively. As shown in Fig. 13
(b), at time T3O=64 ms, the endpoint E3 in the leading edge R2 of the transverse waves reached s = 20 m and thus the response here
increased from zero. At time T5O= 128 ms, the crest E4 reached s = 20 m, and so the response at s = 20 m reached the maximum value
and then began to decrease. At time T4O=192 ms, the trailing edge endpoint E1 reached s = 20 m, and then the normal response at s =
20 m stopped fluctuating with time. It can also be seen in Fig. 13(c) that the endpoint E3 of the leading edge at position s = 20 m and
time T3O=64 ms moved to position s = 40 m at time T4O=192 ms. It can hence be found that the single wave propagated along the cable
at a wave velocity of 156.3 m/s in the second stage.
Then, the effect of the damping on the transverse waves was analyzed. As can be seen in Fig. 13(b) and (c), the waveform after
considering damping almost coincided with the waveform without damping in the first stage, which indicates that damping had little
effect on the transverse wave propagation at this stage. In the second stage of transverse wave propagation, damping caused the
amplitude of the wave to decay continuously. The amplitude of a single wave decreased from 5 mm to 4.491 mm at 192 ms (10.2 %)
after considering damping and this ratio increased with time. In addition to the continuous attenuation of the amplitude of the wave,
the response with damping at the endpoint of the wave trailing edge was also greater than zero.

5. Conclusions

The paper developed a three-dimensional elastic cable equation of motion considering damping and small curvature. The prop­
agation characteristics of free coupled in-plane waves were evaluated. The wave propagation characteristics were examined based on
the frequency analysis. The paper focused on the influence of damping and curvature on the waves. The main conclusions can be
summarized as follows.
The frequency, phase velocity, group velocity and cut-off frequency were investigated based on the linearized equation of the
coupled in-plane waves. The results showed that four cut-off frequencies were generated under the coupling effect of curvature. The
cut-off frequencies divided the frequency domain into stopbands and passbands. Damping changed the cut-off frequencies and
stopbands and passbands: the cut-off frequencies moved from the real axis to the imaginary axis, the stopband transformed from a line
segment into a quadrilateral, and the two-dimensional frequency curve transformed into a three-dimensional surface. Based on the
relationship between frequency and wave-number, the responses of in-plane waves in stopbands and passbands were investigated. The
responses behaved differently at stopbands and passbands. Two possibilities existed at stopbands. The evanescent waves occurred
when the wavenumber was purely imaginary. For the undamped cable, the evanescent waves exponentially decayed with spatial
distance and oscillated periodically in time. The evanescent waves of the damped cable behaved similarly in space, but decayed to zero
with time. When the wavenumber was zero, cable only vibrated transversely with the amplitude constant in space and fluctuating with
time. At passband, the in-plane waves were cosine curves in space. In time, the longitudinal waves were a superposition of cosine

47
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

curves containing the components of transverse and longitudinal waves, whereas the transverse waves were cosine curves dominated
by the transverse waves. Damping attenuated the response at ratio increasing with time. The responses of a railway contact wire and a
steel suspension cable in the spatial domain were calculated and compared to explore the influence of material properties on wave
propagation using a linear elastic constitutive model. The results showed that the steel suspension cable had a larger longitudinal
response magnitude, the wave velocity of longitudinal waves at the leading edge and the velocity of the crest attenuation of the
transverse waves. The transverse wave velocity in the steel suspension cable was lower than in the railway contact wire.

Funding

This study is supported by the National Natural Science Foundation of China (No. 12372051, 11672306), the National Key
Research and Development Program of China (No. 2022YFB4201500).

CRediT authorship contribution statement

Lijun Li: Writing – original draft, Visualization, Validation, Software, Investigation, Formal analysis, Conceptualization. Xiaohui
Zeng: Writing – review & editing, Validation, Supervision. Han Wu: Writing – review & editing, Validation, Supervision. Zhehua Cui:
Writing – review & editing, Supervision.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

References

[1] Q. Wu, K. Takahashi, S. Nakamura, Non-linear vibrations of cables considering loosening, J. Sound Vib. 261 (2003) 385–402, https://doi.org/10.1016/S0022-
460X(02)01090-8.
[2] C. Koh, Y. Rong, Dynamic analysis of large displacement cable motion with experimental verification, J. Sound Vib. 272 (2004) 187–206, https://doi.org/
10.1016/S0022-460X(03)00326-2.
[3] N. Srinil, G. Rega, S. Chucheepsakul, Three-dimensional non-linear coupling and dynamic tension in the large-amplitude free vibrations of arbitrarily sagged
cables, J. Sound Vib. 269 (2004) 823–852, https://doi.org/10.1016/S0022-460X(03)00137-8.
[4] R. Karoumi, Some modeling aspects in the nonlinear finite element analysis of cable supported bridges, Comput. Struct. 71 (1999) 397–412, https://doi.org/
10.1016/S0045-7949(98)00244-2.
[5] Y. Ni, J. Ko, G. Zheng, Dynamic analysis of large-diameter sagged cables taking into account flexural rigidity, J. Sound Vib. 257 (2002) 301–319, https://doi.
org/10.1006/jsvi.2002.5060.
[6] H. Irvine, T. Caughey, Linear theory of free vibrations of a suspended cable, Proc. R. Soc. Lond. Ser. A-Math. Phys. Eng. Sci. 341 (1974) 299–315.
[7] G. Rega, F. Vestroni, F. Benedettini, Parametric analysis of large amplitude free vibrations of a suspended cable, Int. J. Solids Struct. 20 (1984) 95–105, https://
doi.org/10.1016/0020-7683(84)90001-5.
[8] L. Chen, B. Basu, S. Nielsen, A coupled finite difference mooring dynamics model for floating offshore wind turbine analysis, Ocean Eng. 162 (2018) 304–315,
https://doi.org/10.1016/j.oceaneng.2018.05.001.
[9] H. Irvine, Cable Structures, The MIT Press, Cambridge, 1981.
[10] G. Rega, Nonlinear vibrations of suspended cables—Part I: modeling and analysis, Appl. Mech. Rev. 57 (2004) 443–478, https://doi.org/10.1115/1.1777224.
[11] G. Rega, Nonlinear vibrations of suspended cables—Part II: deterministic phenomena, Appl. Mech. Rev. 57 (2004) 479–514, https://doi.org/10.1115/
1.1777225.
[12] R. Ibrahim, Nonlinear vibrations of suspended cables—Part III: random excitation and interaction with fluid flow, Appl. Mech. Rev. 57 (2004) 515–549, https://
doi.org/10.1115/1.1804541.
[13] M. Jafari, F. Hou, A. Abdelkefi, Wind-induced vibration of structural cables, Nonlinear Dyn. 100 (2020) 351–421, https://doi.org/10.1007/s11071-020-05541-
6.
[14] F. Di, L. Sun, L. Chen, Suppression of vortex-induced high-mode vibrations of a cable-damper system by an additional damper, Eng. Struct. 242 (2021) 112495,
https://doi.org/10.1016/j.engstruct.2021.112495.
[15] Y. Song, F. Duan, Z. Liu, Analysis of critical speed for high-speed railway pantograph-catenary system, IEEE Trans. Veh. Technol. 71 (2021) 3547–3555, https://
doi.org/10.1109/TVT.2021.3136920.
[16] W. Zhan, D. Zou, M. Tan, N. Zhou, R. Li, G. Mei, Review of pantograph and catenary interaction, Front. Mech. Eng. 13 (2018) 311–332, https://doi.org/
10.1007/s11465-018-0494-x.
[17] J. Lee, T. Park, H. Oh, Y. Kim, Analysis of dynamic interaction between catenary and pantograph with experimental verification and performance evaluation in
new high-speed line, Veh. Syst. Dyn. 53 (2015) 1117–1134, https://doi.org/10.1080/00423114.2015.1025797.
[18] Y. Tao, S. Zhang, Research on the vibration and wave propagation in ship-borne tethered UAV using stress wave method, Drones-Basel 6 (2022) 349, https://
doi.org/10.3390/drones6110349.
[19] S. Zhang, R. Shen, T. Wang, G. Roeck, G. Lombaert, A two-step FEM-SEM approach for wave propagation analysis in cable structures, J. Sound Vib. 415 (2018)
41–58, https://doi.org/10.1016/j.jsv.2017.11.002.
[20] N. Perkins, J. Mote, Three-dimensional vibration of travelling elastic cables, J. Sound Vib. 114 (1987) 325–340, https://doi.org/10.1016/S0022-460X(87)
80157-8.
[21] S. Cheng, N. Perkins, Free vibration of a sagged cable supporting a discrete mass, J. Acoust. Soc. Am. 91 (1992) 2654–2662, https://doi.org/10.1121/1.402973.
[22] M. Behbahani-Nejad, N. Perkins, Freely propagating waves in elastic cables, J. Sound Vib. 196 (1996) 189–202, https://doi.org/10.1006/jsvi.1996.0476.

48
L. Li et al. Applied Mathematical Modelling 134 (2024) 29–49

[23] M. Behbahani-Nejad, N. Perkins, Harmonically forced wave propagation in elastic cables with small curvature, J. Vib Acoust. 119 (1997) 390–397, https://doi.
org/10.1115/1.2889735.
[24] K. Graff, Wave Motion in Elastic Solids, Clarendon/Oxford University Press, London, 1975.
[25] L. Li, X. Zeng, Z. Cui, H. Wu, Propagation of elastic wave in infinite cable with small sag considering damping, Chin. J. Mech. 55 (2023) 1138–1150, https://doi.
org/10.6052/0459-1879-22-606.
[26] S. Sorrentino, D. Anastasio, A. Fasana, S. Marchesiello, Distributed parameter and finite element models for wave propagation in railway contact lines, J. Sound
Vib. 410 (2017) 1–18, https://doi.org/10.1016/j.jsv.2017.08.008.
[27] D. Zou, N. Zhou, L. Ping, G. Mei, W. Zhang, Experimental and simulation study of wave motion upon railway overhead wire systems, Proc. Inst. Mech. Eng. Part
F-J. Rail Rapid Transit. 231 (2017) 934–944, https://doi.org/10.1177/0954409716648718.
[28] Y. Song, Z. Liu, F. Duan, Z. Xu, X. Lu, Wave propagation analysis in high-speed railway catenary system subjected to a moving pantograph, Appl. Math. Model.
59 (2018) 20–38, https://doi.org/10.1016/j.apm.2018.01.001.
[29] O. Van, J. Massa, E. Balmes, Waves, modes and properties with a major impact on dynamic pantograph-catenary interaction, J. Sound Vib. 402 (2017) 51–69,
https://doi.org/10.1016/j.jsv.2017.05.008.
[30] T. Jiang, R. Anders, Y. Song, G. Frøseth, P. Nåvi, A detailed investigation of uplift and damping of a railway catenary span in traffic using a vision-based line-
tracking system, J. Sound Vib. 527 (2022) 116875, https://doi.org/10.1016/j.jsv.2022.116875.
[31] W. Hu, J. Ye, Z. Deng, Internal resonance of a flexible beam in a spatial tethered system, J. Sound Vib. 475 (2020) 115286, https://doi.org/10.1016/j.
jsv.2020.115286.
[32] S. Bruni, J. Ambrosio, A. Carnicero, Y. Cho, L. Finner, M. Ikeda, S. Kwon, J. Massat, S. Stichel, M. Tur, W. Zhang, The results of the pantograph-catenary
interaction benchmark, Veh. Syst. Dyn. 54 (2015) 412–435, https://doi.org/10.1080/00423114.2014.953183.
[33] W. Hu, C. Zhang, Z. Deng, Vibration and elastic wave propagation in spatial flexible damping panel attached to four special springs, Commun. Nonlinear Sci.
Numer. Simul. 84 (2020) 105199, https://doi.org/10.1016/j.cnsns.2020.105199.
[34] A. Baxy, R. Prasad, A. Banerjee, Elastic waves in layered periodic curved beams, J. Sound Vib. 512 (2021) 116387, https://doi.org/10.1016/j.jsv.2021.116387.
[35] W. Hu, X. Xi, Z. Song, C. Zhang, Z. Deng, Coupling dynamic behaviors of axially moving cracked cantilevered beam subjected to transverse harmonic load,
Mech. Syst. Signal Proc. 204 (2023) 110757, https://doi.org/10.1016/j.ymssp.2023.110757.
[36] W. Hu, M. Xu, J. Song, Q. Gao, Z. Deng, Coupling dynamic behaviors of flexible stretching hub-beam system, Mech. Syst. Signal Proc. 151 (2021) 107389,
https://doi.org/10.1016/j.ymssp.2020.107389.
[37] R. Prasad, A. Banerjee, Flexural waves in elastically coupled telescopic metabeams, J. Vib. Acoust. 143 (6) (2021) 061009, https://doi.org/10.1115/1.4050809.
[38] B. Syam, A. Sarkar, Wave analysis for in-plane vibration of angular and curved frames, J. Vib. Acoust. 143 (6) (2021) 061001, https://doi.org/10.1115/
1.4049627.
[39] W. Hu, M. Xu, F. Zhang, C. Xiao, Z. Deng, Dynamic analysis on flexible hub-beam with step-variable cross-section, Mech. Syst. Signal Proc. 180 (2022) 109423,
https://doi.org/10.1016/j.ymssp.2022.109423.
[40] Y. Yang, M. Kingan, A hybrid wave and finite element/boundary element method for predicting the vibroacoustic characteristics of finite-width complex
structures, J. Sound Vib. 582 (2024) 118402, https://doi.org/10.1016/j.jsv.2024.118402.
[41] Y. Sun, Q. Han, T. Jiang, C. Li, Coupled bandgap properties and wave attenuation in the piezoelectric metamaterial beam on periodic elastic foundation, Appl.
Math. Model. 125 (2024) 293–310, https://doi.org/10.1016/j.apm.2023.09.030.
[42] W. Hu, Z. Deng, S. Han, W. Zhang, Generalized multi-symplectic integrators for a class of Hamiltonian nonlinear wave PDEs, J. Comput. Phys. 235 (2013),
https://doi.org/10.1016/j.jcp.2012.10.032.
[43] W. Hu, Y. Huai, M. Xu, X. Feng, R. Jiang, Y. Zheng, Z. Deng, Mechanoelectrical flexible hub-beam model of ionic-type solvent-free nanofluids, Mech. Syst. Signal
Proc. 159 (2021) 107833, https://doi.org/10.1016/j.ymssp.2021.107833.
[44] W. Hu, Z. Han, T. Bridges, et al., Multi-symplectic simulations of W/M-shape-peaks solitons and cuspons for FORQ equation, Appl. Math. Lett. 145 (2023)
108772, https://doi.org/10.1016/j.aml.2023.108772.
[45] Y. Jiang, E. Kim, J. Yang, J. Rho, Physics-informed discrete element modeling for the bandgap engineering of cylinder chains, Appl. Math. Model. 125 (2024)
571–590, https://doi.org/10.1016/j.apm.2023.09.011.
[46] S. Lin, Q. Han, C. Li, Coupled bandgaps and wave attenuation in flexoelectric curve nanobeam with periodic distribution, Appl. Math. Model. 130 (2024)
419–437, https://doi.org/10.1016/j.apm.2024.03.008.
[47] M. Darche, F. Lopez-Caballero, B. Tie, Modal analysis of waveguide for the study of frequency bandgaps of a bounded periodic medium, J. Sound Vib. 572
(2024) 118158, https://doi.org/10.1016/j.jsv.2023.118158.
[48] X. Zhang, H. Xu, M. Cao, D. Sumara, Y. Lu, J. Peng, In-plane free vibrations of small-sag inclined cables considering bending stiffness with applications to cable
tension identification, J. Sound Vib. 544 (2023) 117394, https://doi.org/10.1016/j.jsv.2022.117394.
[49] X. Chen, Z. Xi, Y. Wang, X. Wang, Improved study on the fluctuation velocity of high-speed railway catenary considering the influence of accessory parts, IEEE
Access 8 (2020) 138710–138718, https://doi.org/10.1109/ACCESS.2020.3011415.
[50] E N British Standard, 50318: 2002, Railway applications-current collection systems-validation of simulation of the dynamic interaction between pantograph and
overhead contact line.
[51] Y. Song, Z. Liu, F. Duan, Z. Xu, X. Lu, Wave propagation analysis in high-speed railway catenary system subjected to a moving pantograph, Appl. Math. Model.
59 (2018) 20–38, https://doi.org/10.1016/j.apm.2018.01.001.

49

You might also like