Keywords

1 Introduction

High-speed railways are currently popular globally. However, there are some problems including passenger riding comfort, noise pollution, and even operational safety (Jin et al. 2013). Rail corrugation , rail welding irregularity , wheel burning, and wheel out-of-roundness (OOR) generate high-frequency components of the dynamic wheel–rail contact forces that contribute significantly to the total wheel–rail contact forces (Nielsen et al. 2003), and reduce the life of the components of track and vehicle, such as wheels, rails, and fasteners. Rail grinding and wheel reprofiling are the most common measures that have been proved to be effective in controlling rail irregularities and wheel OOR. However, these measures lead to notably high maintenance costs. A lot of measurements at the sites and coupling vehicle–track dynamics modeling have been carried out to investigate the mechanism and development of these phenomena. In the vehicle–track dynamics modeling, a rigid multi-body system is often adopted to simulate railway vehicles, based on several commercial codes available for the low-frequency domain, such as GENSYS, NUCARS, SIMPACK, and VAMPIRE. These computer programs are generally used to analyze railway vehicle dynamics responses at frequencies below 20 Hz, where the influence of rigid motions of the vehicle on wheel–rail contact forces is dominant (Nielsen et al. 2005). To analyze the vehicle dynamic responses at mid- and high frequencies, the vehicle structural flexibility should be taken into account in the modeling. It is obvious that wheelset structural flexibility has an influence on wheel–rail contact behaviors at mid- and high frequencies. Different flexible wheelset models have been set up due to various motivations in the past (Chaar 2007).

The methods applied to modeling flexible wheelset can be summarized as three major categories (Chaar 2007). The first is a lumped model developed in a simple and convenient way, in which a wheelset is divided into several parts interconnected with springs and dampers. This model can describe the bending and torsional motions of the wheelset with only a few degrees of freedom, which could not be applied to studying wear phenomena on wheel treads or rails (Popp et al. 1999). The second is a continuous model developed by Szolc (1998a, b), in which the wheelset axle was modeled as a beam, and two wheels and brake disks were modeled as rigid rings attached to the axle through a massless, elastically isotropic membrane. The model can characterize the wheelset dynamic behavior in the frequency range of 30–300 Hz. In the model proposed by Popp et al. (2003), the wheelset axle was considered as a 1D continuum, having the properties of a bar, a torsional rod, and a Rayleigh beam. The wheel was considered as a 2D continuum, having the properties of a disk and a Kirchhoff plate. The third was developed based on finite element method (FEM) , which simulates wheelset flexibility more realistically than the first two categories of model. The wheelset modes and corresponding natural frequencies were obtained through the modal analysis of the finite element (FE) model by using the commercial software, and they were input into the simulation by means of the commercial codes (SIMPACK, NEWEUL) (Meinders and Meinker 2003) or some non-commercial multi-body dynamic system codes. The non-commercial code developed by Fayos et al. (2007) and Baeza et al. (2008, 2011) introduced the Eulerian coordinate system to replace the Lagrangian coordinate system in the flexible wheelset modeling. In this way, it is convenient to obtain the motion of fixed physical nodes and consider the inertial effect due to wheelset rotation. Relying on current computing power, it is feasible to use FEM to consider the effect of flexible wheelset in modeling a railway vehicle coupling with a track.

Regarding the wheel–rail contact treatment in considering flexible wheelset influence, wheel–rail rolling contact condition is simplified based on different prior assumptions, especially in the detection of wheel–rail contact points . This is the prerequisite for the calculation of wheel–rail creepages and contact forces . Baeza et al. (2011) neglected the effect of the high-frequency deformation and the deviation of a rotating flexible wheelset rolling over a flexible track model on the wheel–rail contact point in the investigation into the effect of the rotating flexible wheelset on rail corrugation . Through the detailed calculation, Kaiser and Popp (2006) found that the contact point was in the location where the wheel and the rail had positive penetration maxima, and the penetration direction was orthogonal to the common tangent plane of the wheel and the rail before their deformations . A linear wheel–rail contact model was proposed and used to carry out the detection of wheel–rail contact point and the contact zone’s normal direction (Andersson and Abrahamsson 2002). In the detection, the functions were created using a first-order Taylor expansion around a reference state described by a group of parameters which represent a configuration, in which the train was in static equilibrium and the wheel and the track were free from geometric imperfections. The advantage of this approach is that the contact position and orientation in each time step can be calculated by interpolation replacing iterations, which results in a low computational cost. But the approach is only suitable for the case that the effect of all the parameters is very small on the contact point position and the contact patch orientation around the references is in static equilibrium. The wheel–rail contact point position and the contact patch orientation greatly depend on parameters, such as the curvatures of wheel and rail . In (Torstensson et al. 2012; Torstensson and Nielsen 2011), the contact point detection was done before the simulation and used in the subsequent time integration analysis in the form of lookup table. The commercial software GENSYS allows for such calculations using the preprocessor KPF (from Swedish contact point function). In the KPF, the location and orientation of the contact patch were assumed to be dependent only on the relative displacement in the lateral direction between the wheelset and the rails, and hence, the influence of the wheelset yaw angle was not taken into account. In some other papers, detailed discussions on the wheel–rail contact model were omitted. In this study, the wheel–rail contact model considering the effect of wheelset flexibility (Zhong et al. 2013, 2014) is further improved and the new contact model is suitable for the analysis on the effect of the local higher-frequency deformation of the wheels on the wheel–rail contact behavior .

2 Vehicle–Track Coupling Dynamic System

A flexible wheelset model (to be illustrated in Sect. 2.1) and a suitable wheel–rail contact model (to be discussed in Sect. 2.2) are integrated into the vehicle–track coupling dynamic system model. All parts of the vehicle system, except for its four wheelsets, are considered as rigid bodies. The primary and secondary suspension systems of the vehicle are modeled with spring–damper elements. A triple-layer model of discrete elastic support is adopted to simulate the ballasted track . The rails are modeled as Timoshenko beams. The sleepers are modeled as rigid bodies, and the ballast model consists of discrete equivalent masses. The equivalent spring–damper elements are used as the connections between the rails and the sleepers, the sleepers and the equivalent ballast bodies, and the ballast bodies and the roadbed. Figure 1 shows the vehicle–track coupling dynamic system model. The equations of motion of each component of the vehicle excluding wheelsets and the track are illustrated in detail in (Xiao et al. 2007, 2008, 2010). The parameters and their values describing the dynamic models are given in Appendix A.

Fig. 1
figure 1

Vehicle–track coupling model (elevation)

2.1 Flexible Wheelset Model

The wheelset structural flexibility is considered by modeling the wheelset axle as an Euler–Bernoulli beam in two planes, one perpendicular to the track centerline and the other parallel to the track level. The crossing effect of the bending deformations in the two planes is ignored. In the first two bending modes obtained using the modal analysis of the FE model of a wheelset, two wheels have little deformation (Fig. 2), and their frequencies are in the available frequency range (0–500 Hz) of an Euler–Bernoulli beam model. Therefore, two wheels can be treated as rigid bodies in this study.

Fig. 2
figure 2

First two bending modes obtained using FE model

There are two force systems acting on the wheelset, one is the wheel–rail contact forces and the other is the forces of the primary suspension system (Fig. 3).

Fig. 3
figure 3

Force analysis diagram of the flexible wheelset

In Fig. 3, O fL and O fR are the left and right points on the axle, respectively, where the primary suspension force systems are applied. O CL and O CR are the left and right contact points of wheel–rail, respectively. O indicates the origin of the coordinate system O-XYZ that is a coordinate system with a translational motion along the tangent track centerline at operational speed. If the speed is constant, this coordinate system is an inertial coordinate system and therefore regarded as an absolute coordinate system (geodetic coordinate system).

To analyze the axle’s deformation, the force systems from wheel–rail interaction acting on the left and right wheel treads are translated to the nominal circle centers O L and O R, respectively, and extra moments are produced in the procedure of translating contact forces . Thus, the force systems acting on the axle in the two planes are obtained in Fig. 4.

Fig. 4
figure 4

Force analysis diagram in the plane O-YZ (a) and in the O-XY plane (b)

The notations of the variables and symbols are defined in Table 1. The subscript p denotes the primary suspension, the subscripts x, y, and z denote X-, Y-, and Z-direction, respectively, and A denotes the axle.

Table 1 Notations of the variables

The differential equation for the flexural vibration of an Euler–Bernoulli beam (the axle) in the plane O-YZ is written as

$$ {\text{EI}}_{x} \frac{{\partial^{4} u_{z} (y,t)}}{{\partial y^{4} }} + \rho A\frac{{\partial^{2} u_{z} (y,t)}}{{\partial t^{2} }} = Q_{z} (y,t) - \frac{{\partial M_{x} (y,t)}}{\partial y}, $$
(1)

where

$$ Q_{z} (y,t) = F_{{\text{AL}z}} \delta (y - y_{\text{wL}} ) + F_{{\text{pL}z}} \delta (y - y_{\text{pL}} ) + F_{{\text{AR}z}} \delta (y - y_{\text{wR}} ) + F_{{\text{pR}z}} \delta (y - y_{\text{pR}} ), $$
(2)
$$ \begin{aligned} M_{x} (y,t) & = M_{{\text{AL}}x} \delta (y - y_{\text{wL}} ) \\ & \quad + ( - F_{{\text{pL}}y} u_{{\text{pL}}z} + F_{{\text{pL}}z} u_{{\text{pL}y}} )\delta (y - y_{\text{pL}} ) \\ & \quad + M_{{\text{AR}}x} \delta (y - y_{\text{wR}} ) \\ & \quad + ( - F_{{\text{pR}}y} u_{{\text{pR}}z} + F_{{\text{pR}}z} u_{{\text{pR}}y} )\delta (y - y_{\text{pR}} ). \\ \end{aligned} $$
(3)

The force analysis diagram of the two wheels including the D’Alembert forces is shown in Fig. 5, based on which differential equations of motion of the two wheels are written as

Fig. 5
figure 5

Force analysis diagram of the two wheels

$$ m_{\text{w}} \frac{{{\partial }^{2} }}{{{\partial }t^{2} }}u_{z} (y_{\text{w(L,R)}} ,t) = m_{\text{w}} g - F_{{{\text{A(L,R)}}z}} - F_{{{\text{wr(L,R)}}z}} , $$
(4)
$$ \begin{aligned} J_{\text{w}} \frac{{{\partial }^{2} }}{{{\partial }t^{2} }}u^{\prime}_{z} (y_{\text{w(L,R)}} ,t) & = - F_{{{\text{wr(L,R)}}y}} u_{{{\text{c(L,R)}}z}} \\ & \quad - F_{{{\text{wr(L,R)}}z}} u_{{{\text{c(L,R)}}y}} - M_{{{\text{A(L,R)}}x}} . \\ \end{aligned} $$
(5)

Note that the lateral accelerations of the wheels are assumed to be the same as the wheelset axle so there is no relative motion between wheels and axle.

Substituting the expressions of F A(L,R)z and M A(L,R)x obtained through Eqs. (4) and (5) into Eqs. (2) and (3), respectively, we can obtain:

$$ \begin{aligned} Q_{z} (y,t) & = F_{{\text{pL}z}} \delta (y - y_{\text{pL}} ) \\ & \quad + \left( {m_{\text{w}} g - F_{{{\text{wrL}}z}} - m_{\text{w}} \frac{{{\partial }^{2} }}{{{\partial }t^{2} }}u_{z} (y_{\text{wL}} ,t)} \right)\delta (y - y_{\text{wL}} ) \\ & \quad + \left( {m_{\text{w}} g - F_{{{\text{wrR}}z}} - m_{\text{w}} \frac{{{\partial }^{2} }}{{{\partial }t^{2} }}u_{z} (y_{\text{wR}} ,t)} \right)\delta (y - y_{\text{wR}} ) \\ & \quad + F_{{\text{pR}z}} \delta (y - y_{\text{pR}} ), \\ \end{aligned} $$
(6)
$$ \begin{aligned} M_{x} (y,t) & = ( - F_{{{\text{pL}}y}} u_{{\text{pL}z}} + F_{{{\text{pL}}z}} u_{{\text{pL}y}} )\delta (y - y_{\text{pL}} ) \\ & \quad + \left( { - F_{{{\text{wrL}}y}} u_{{{\text{cL}}z}} - F_{{{\text{wrL}}z}} u_{{{\text{cL}}y}} - J_{\text{w}} \frac{{\partial^{2} }}{{\partial t^{2} }}u_{z}^{{\mathbf{\prime }}} (y_{\text{wL}} ,t)} \right)\delta (y - y_{\text{wL}} ) \\ & \quad + ( - F_{{{\text{pR}}y}} u_{{\text{pR}z}} + F_{{{\text{pR}}z}} u_{{\text{pR}y}} )\delta (y - y_{\text{pR}} ) \\ & \quad + \left( { - F_{{{\text{wrR}}y}} u_{{{\text{cR}}z}} - F_{{{\text{wrR}}z}} u_{{{\text{cR}}y}} - J_{\text{w}} \frac{{\partial^{2} }}{{\partial t^{2} }}u_{z}^{{\mathbf{\prime }}} (y_{\text{wR}} ,t)} \right)\delta (y - y_{\text{wR}} ), \\ \end{aligned} $$
(7)
$$ \begin{aligned} & EI_{x} \frac{{\partial^{4} u_{z} (y,t)}}{{\partial y^{4} }} + \rho A\frac{{\partial^{2} u_{z} (y,t)}}{{\partial t^{2} }} + m_{\text{w}} \frac{{\partial^{2} }}{{\partial t^{2} }}u_{z} (y,t)\delta (y - y_{\text{wL}} ) \\ & \quad + m_{\text{w}} \frac{{\partial^{2} }}{{\partial t^{2} }}u_{z} (y,t)\delta (y - y_{\text{wR}} ) - J_{\text{w}} \frac{{\partial^{2} }}{{\partial t^{2} }}u^{\prime}_{z} (y,t)\delta (y - y_{\text{wL}} ) \\ & \quad - J_{\text{w}} \frac{{\partial^{2} }}{{\partial t^{2} }}u^{\prime}_{z} (y,t)\delta (y - y_{\text{wR}} ) = W_{0} , \\ \end{aligned} $$
(8)

where

$$ \begin{aligned} W_{0} & = (m_{\text{w}} g - F_{{{\text{wrL}}z}} )\delta (y - y_{\text{wL}} ) + F_{{\text{pL}z}} \delta (y - y_{\text{pL}} ) \\ & \quad + (m_{\text{w}} g - F_{{{\text{wrR}}z}} )\delta (y - y_{\text{wR}} ) + F_{{\text{pR}z}} \delta (y - y_{\text{pR}} ) \\ & \quad - \frac{\partial }{\partial y}[( - F_{{{\text{wrL}}y}} u_{{{\text{cL}}z}} - F_{{{\text{wrL}}z}} u_{{{\text{cL}}y}} )\delta (y - y_{\text{wL}} ) \\ & \quad + ( - F_{{{\text{pL}}y}} u_{{\text{pL}z}} + F_{{{\text{pL}}z}} u_{{\text{pL}y}} )\delta (y - y_{\text{pL}} ) \\ & \quad + ( - F_{{{\text{wrR}}y}} u_{{{\text{cR}}z}} - F_{{{\text{wrR}}z}} u_{{{\text{cR}}y}} )\delta (y - y_{\text{wR}} ) \\ & \quad + ( - F_{{{\text{pR}}y}} u_{{\text{pR}z}} + F_{{{\text{pR}}z}} u_{{\text{pR}y}} )\delta (y - y_{\text{pR}} )]. \\ \end{aligned} $$
(9)

Consider a solution of Eq. (8) in the form:

$$ u_{z} (y,t) = U_{z} (y)\sin (\omega t + \sigma ). $$
(10)

Using the calculus of variation (Qiu et al. 2009), the modal function satisfies:

$$ \begin{aligned} m_{ij} & = \int\limits_{0}^{L} {\rho A(U_{zi} U_{zj} ){\text{d}}y} \\ & \quad + m_{\text{w}} \left( {U_{zi} (y_{\text{wL}} )U_{zj} (y_{\text{wL}} ) + U_{zi} (y_{\text{wR}} )U_{zj} (y_{\text{wR}} )} \right) \\ & \quad + J_{\text{w}} \left( {U_{zi}^{\prime } (y_{\text{wL}} )U_{zj}^{\prime } (y_{\text{wL}} ) + U_{zi}^{\prime } (y_{\text{wR}} )U_{zj}^{\prime } (y_{\text{wR}} )} \right) = \delta_{ij} , \\ \end{aligned} $$
(11)
$$ k_{ij} = \int\limits_{0}^{L} {EI_{x} \left( {U_{zi}^{\prime \prime } U_{zj}^{\prime \prime } } \right)} {\text{dy}} = \omega_{j}^{2} \delta_{ij} $$
(12)
$$ \begin{aligned} & EI_{x} \int\limits_{0}^{L} {U_{zj} U_{zi}^{{{\prime \prime \prime \prime }}} {\text{d}}y} \\ & \quad + J_{\text{w}} \left( {U_{zj}^{{\prime }} (y_{\text{wL}} )U_{zi}^{{\prime }} (y_{\text{wL}} ) + U_{zj}^{{\prime }} (y_{\text{wR}} )U_{zi}^{{\prime }} (y_{\text{wR}} )} \right) \\ & \quad + J_{\text{w}} \left( {U_{zj} (y_{\text{wL}} )U_{zi}^{{\prime }} (y_{\text{wL}} ) + U_{zj} (y_{\text{wR}} )U_{zi}^{{\prime }} (y_{\text{wR}} )} \right) = \omega_{i}^{2} \delta_{ij} , \\ \end{aligned} $$
(13)

where δ ij is the Kronecker delta. For i = j, Eq. (11) can be written as

$$ \begin{aligned} m_{jj} & = \int\limits_{0}^{L} {\rho A(U_{zj}^{2} )\text{d}y} + m_\text{w} \left( {U_{zj}^{2} (y_{\text{wL}} ) + U_{zj}^{2} (y_{\text{wR}} )} \right) \\ & \quad + J_\text{w} \left( {U_{zj}^{\prime 2} (y_{\text{wL}} ) + U_{zj}^{\prime2} (y_{\text{wR}} )} \right) = 1. \\ \end{aligned} $$
(14)

To obtain the mode shape functions with the wheelset axle modeled as a uniform Euler–Bernoulli beam carrying two particles (wheels), the segment of the beam from the left end to the first particle is referred to as the first portion, in between the two particles as the second portion, and from the second particle to the right end as the third portion. The beam mode shape will be the superposition of the mode shapes of the three portions. The derivation of the mode shape functions is presented in Appendix B. The first three modes have the frequencies of f 1 = 111 Hz, f 2 = 245 Hz, and f 3 = 547 Hz, respectively. These mode shape functions are normalized so as to satisfy Eq. (14), as shown in Fig. 6. The third mode is not in the frequency range of 0–500 Hz where the Euler–Bernoulli beam is available to analyze the system. Hence, the effect of the first two modes on dynamic responses is conducted in this study.

Fig. 6
figure 6

First three bending mode shapes of the wheelset

According to the modal analysis, we let the solution of Eq. (8) have the form:

$$ u_{z} = \sum\limits_{i = 1}^{N} {U_{zi} q_{zi} } . $$
(15)

Substituting Eq. (15) into Eq. (8), the differential equation can be written as

$$ \begin{aligned} & {EI}_{x} \sum\limits_{i = 1}^{N} {U_{zi}^{{{\prime \prime \prime \prime }}} q_{zi} } + \rho A\sum\limits_{i = 1}^{N} {U_{zi} \ddot{q}_{zi} } \\ & \quad + m_{\text{w}} \sum\limits_{i = 1}^{N} {U_{zi} (y_{\text{wL}} )\ddot{q}_{zi} } + m_{\text{w}} \sum\limits_{i = 1}^{N} {U_{zi} (y_{{\text{wR}}} )\ddot{q}_{zi} } \\ & \quad - J_{\text{w}} \sum\limits_{i = 1}^{N} {U_{zi}^{{\prime }} (y_{\text{wL}} )\ddot{q}_{zi} } - J_{\text{w}} \sum\limits_{i = 1}^{N} {U_{zi}^{{\mathbf{\prime \prime }}} (y_{\text{wR}} )\ddot{q}_{zi} } = W_{0} . \\ \end{aligned} $$
(16)

Multiplying both sides of Eq. (16) by \( U_{zj} \) and integrating over the domain 0 < y < L, we can obtain:

$$ \begin{aligned} & {{EI}}_{x} \sum\limits_{i = 1}^{N} {\left[ {q_{zi} \int\limits_{0}^{L} {U_{zj} U_{zi}^{{{\prime \prime \prime \prime }}} {\text{d}}y} } \right]} + \rho A\sum\limits_{i = 1}^{N} {\left[ {\ddot{q}_{zi} \int\limits_{0}^{L} {U_{zj} U_{zi} {\text{d}}y} } \right]} \\ & \quad \cdot m_{\text{w}} \sum\limits_{i = 1}^{N} {\left[ {\ddot{q}_{zi} \int\limits_{0}^{L} {U_{zj} U_{zi} \delta (y - y_{\text{wL}} ){\text{d}}y} } \right]} \\ & \quad + m_{\text{w}} \sum\limits_{i = 1}^{N} {\left[ {\ddot{q}_{zi} \int\limits_{0}^{L} {U_{zj} U_{zi} \delta (y - y_{\text{wR}} ){\text{d}}y} } \right]} \\ & \quad - J_{\text{w}} \sum\limits_{i = 1}^{N} {\left[ {\ddot{q}_{zi} \int\limits_{0}^{L} {U_{zj} U_{zi}^{{\prime }} \delta (y - y_{\text{wL}} ){\text{d}}y} } \right]} \\ & \quad - J_{\text{w}} \sum\limits_{i = 1}^{N} {\left[ {\ddot{q}_{zi} \int\limits_{0}^{L} {U_{zj} U_{zi}^{{\prime }} \delta (y - y_{\text{wR}} ){\text{d}}y} } \right]} = W_{1j} . \\ \end{aligned} $$
(17)

Using the orthogonality of the modal shape function as expressed in Eqs. (11) and (13), Eq. (17) can be written as

$$ \begin{aligned} & \sum\limits_{i = 1}^{N} {\left\{ {q_{zi} \left[ {\omega_{i}^{2} \delta_{ij} - J_{\text{w}} \left( {U^{\prime}_{zj} (y_{\text{wL}} )U^{\prime}_{zi} (y_{\text{wL}} )} \right.} \right.} \right.} \\ & \quad + \left. {U^{\prime}_{zj} (y_{\text{wR}} )U^{\prime}_{zi} (y_{\text{wR}} )} \right) - J_{\text{w}} \left( {U_{zj} (y_{\text{wL}} )U^{\prime}_{zi} (y_{\text{wL}} )} \right. \\ & \quad + \left. {\left. {U_{zj} (y_{\text{wR}} )U^{\prime}_{zi} (y_{\text{wR}} )} \right)} \right] \\ & \quad + \ddot{q}_{zi} \left[ {\delta_{ij} - J_{\text{w}} \left( {U^{\prime}_{zj} (y_{\text{wL}} )U^{\prime}_{zi} (y_{\text{wL}} )} \right.} \right. \\ & \quad + \left. {U^{\prime}_{zj} (y_{\text{wR}} )U^{\prime}_{zi} (y_{\text{wR}} )} \right) - J_{\text{w}} \left( {U_{zj} (y_{\text{wL}} )U^{\prime}_{zi} (y_{\text{wL}} )} \right. \\ & \quad + \left. {\left. {\left. {U_{zj} (y_{\text{wR}} )U^{\prime}_{zi} (y_{\text{wR}} )} \right)} \right]} \right\} = W_{1j} , \\ \end{aligned} $$
(18)

where

$$ \begin{aligned} & W_{1j} = \int\limits_{0}^{L} {U_{zj} W_{0} {\text{d}}y} \\ & = (m_{\text{w}} g - F_{{{\text{wrL}}z}} )U_{zj} (y_{\text{wL}} ) + F_{{\text{pL}z}} U_{zj} (y_{\text{pL}} ) \\ & \quad + (m_{\text{w}} g - F_{{{\text{wrR}}z}} )U_{zj} (y_{\text{wR}} ) + F_{{\text{pR}z}} U_{zj} (y_{\text{pR}} ) \\ & \quad + ( - F_{{{\text{wrL}}y}} u_{{{\text{cL}}z}} - F_{{{\text{wrL}}z}} u_{{{\text{cL}}y}} )U^{\prime}_{zj} (y_{\text{wL}} ) \\ & \quad + ( - F_{{{\text{pL}}y}} u_{{\text{pL}z}} + F_{{{\text{pL}}z}} u_{{\text{pL}y}} )U^{\prime}_{zj} (y_{\text{pL}} ) \\ & \quad + ( - F_{{{\text{wrR}}y}} u_{{{\text{cR}}z}} - F_{{{\text{wrR}}z}} u_{{{\text{cR}}y}} )U^{\prime}_{zj} (y_{\text{wR}} ) \\ & \quad + ( - F_{{{\text{pR}}y}} u_{{\text{pR}z}} + F_{{{\text{pR}}z}} u_{{\text{pR}y}} )U^{\prime}_{zj} (y_{\text{pR}} ). \\ \end{aligned} $$
(19)

Equation (18) can be expressed as:

$$ \begin{aligned} & \ddot{q}_{zj} + \omega_{j}^{2} q_{zj} - J_{\text{w}} \sum\limits_{i = 1}^{N} {\left\{ {q_{zi} \left[ {U^{\prime}_{zj} (y_{\text{wL}} )U^{\prime}_{zi} (y_{\text{wL}} )} \right.} \right.} \\ & \quad + U^{\prime}_{zj} (y_{\text{wR}} )U^{\prime}_{zi} (y_{\text{wR}} ) \\ & \quad + \left. {U_{zj} (y_{\text{wL}} )U^{\prime}_{zi} (y_{\text{wL}} ) + U_{zj} (y_{\text{wR}} )U^{\prime}_{zi} (y_{\text{wR}} )} \right] \\ & \quad + \ddot{q}_{zi} \left[ {U^{\prime}_{zj} (y_{\text{wL}} )U^{\prime}_{zi} (y_{\text{wL}} ) + U^{\prime}_{zj} (y_{\text{wR}} )U^{\prime}_{zi} (y_{\text{wR}} )} \right. \\ & \quad + \left. {\left. {U_{zj} (y_{\text{wL}} )U^{\prime}_{zi} (y_{\text{wL}} ) + U_{zj} (y_{\text{wR}} )U^{\prime}_{zi} (y_{\text{wR}} )} \right]} \right\} = W_{1j} . \\ \end{aligned} $$
(20)

Equation (20) can be written in the matrix form:

$$ {\varvec{M}}_{1} \left[ {\ddot{q}_{zj} } \right] + {\varvec{M}}_{2} \left[ {q_{zj} } \right] = \left[ {W_{1j} } \right], $$
(21)

where

$$ \begin{aligned} ({\varvec{M}}_{11} )_{(i,j)} & =({\varvec{M}}_{21} )_{(i,j)} \\ & = - J_{\text{w}} \left[ {U^{\prime}_{zj} (y_{\text{wL}} )U^{\prime}_{zi} (y_{\text{wL}} ) + U^{\prime}_{zj} (y_{\text{wR}} )U^{\prime}_{zi} (y_{\text{wR}} )} \right. \\ & \quad + U_{zj} (y_{\text{wL}} )U^{\prime}_{zi} (y_{\text{wL}} ) + \left. {U_{zj} (y_{\text{wR}} )U^{\prime}_{zi} (y_{\text{wR}} )} \right], \\ {\varvec{M}}_{12} & = {\varvec{I}},\quad {\varvec{M}}_{22} = \left[ {\omega_{j}^{2} } \right]{\varvec{I}}, \\ {\varvec{M}}_{1} & = {\varvec{M}}_{11} + {\varvec{M}}_{12} ,\quad {\varvec{M}}_{2} = {\varvec{M}}_{21} + {\varvec{M}}_{22} . \\ \end{aligned} $$
(22)

The explicit integral method illustrated in Zhai (2007) is used to obtain the vector \( \left[ {\ddot{q}_{zj} } \right] \) of each acceleration coordinate.

For the vibration in the plane YOX, the differential equation expressed with respect to \( \left[ {\ddot{q}_{xy} } \right] \) can be written as:

$$ \begin{aligned} & \ddot{q}_{xj} + \omega_{xj}^{2} q_{xj} - J_{\text{w}} \sum\limits_{i = 1}^{N} {\left\{ {q_{i} \left[ {U^{\prime}_{xj} (y_{\text{wL}} )U^{\prime}_{xi} (y_{\text{wL}} )} \right.} \right.} \\ & \quad + U^{\prime}_{xj} (y_{\text{wR}} )U^{\prime}_{xi} (y_{\text{wR}} ) \\ & \quad + \left. {U_{xj} (y_{\text{wL}} )U^{\prime}_{xi} (y_{\text{wL}} ) + U_{xj} (y_{\text{wR}} )U^{\prime}_{xi} (y_{\text{wR}} )} \right] \\ & \quad + \ddot{q}_{i} \left[ {U^{\prime}_{xj} (y_{\text{wL}} )U^{\prime}_{xi} (y_{\text{wL}} ) + U^{\prime}_{xj} (y_{\text{wR}} )U^{\prime}_{xi} (y_{\text{wR}} )} \right. \\ & \quad + \left. {\left. {U_{xj} (y_{\text{wL}} )U^{\prime}_{xi} (y_{\text{wL}} ) + U_{xj} (y_{\text{wR}} )U^{\prime}_{xi} (y_{\text{wR}} )} \right]} \right\} = W_{1j}^{xoy} . \\ \end{aligned} $$
(23)

The derivation of Eq. (23) is similar to that of Eq. (20) and omitted here. Equation (23) can be expressed in matrix form:

$$ {\varvec{M}}_{1}^{xoy} \left[ {\ddot{q}_{xj} } \right] + {\varvec{M}}_{2}^{xoy} \left[ {q_{xj} } \right] = \left[ {W_{1j}^{xoy} } \right], $$
(24)

where

$$ \begin{aligned} & ({\varvec{M}}_{11} )_{(i,j)}^{xoy} = ({\varvec{M}}_{21} )_{(i,j)}^{xoy} \\ & \quad = - J_{\text{w}} \left( {U^{\prime}_{xj} (y_{\text{wL}} )U^{\prime}_{xi} (y_{\text{wL}} )} \right. + U^{\prime}_{xj} (y_{\text{wR}} )U^{\prime}_{xi} (y_{\text{wR}} ) \\ & \quad \quad + U_{xj} (y_{\text{wL}} )U^{\prime}_{xi} (y_{\text{wL}} ) + \left. {U_{xj} (y_{\text{wR}} )U^{\prime}_{xi} (y_{\text{wR}} )} \right), \\ & {\varvec{M}}_{12}^{xoy} = {\varvec{I}},\quad {\varvec{M}}_{22}^{xoy} = \left[ {\omega_{j}^{2} } \right]{\varvec{I}}, \\ & {\varvec{M}}_{1}^{xoy} = {\varvec{M}}_{11}^{xoy} + {\varvec{M}}_{12}^{xoy} ,\quad {\varvec{M}}_{2}^{xoy} = {\varvec{M}}_{21}^{xoy} + {\varvec{M}}_{22}^{xoy} . \\ \end{aligned} $$
(25)

2.2 Wheel–Rail Contact Model

As mentioned in Sect. 2.1, the main concern in this work is the wheelset axle bending. The wheels are assumed to be rigid, and their nominal rolling circles are always perpendicular to the deformed wheelset axle at their interference fit surfaces. Figure 7 shows that the flexible wheelset moves from its initial reference state (O 1(t 1)) to its t 2 status (O 2(t 2)), which is described in the plane of O-YZ. O 1 is the center of the undeformed wheelset at t 1, and O 2 is the center of the deformed wheelset at any time t. O 1 O 2 is the displacement vector of the wheelset center due to its rigid motion, and ϕ R1 is the roll angle due to the wheelset rigid motion. The auxiliary line, \( A_{\text{L}}^{0} A_{\text{R}}^{0} , \) is the central line of the undeformed wheelset axle , \( A_{\text{L}}^{1} A_{\text{R}}^{ 1} \) is obtained by moving \( A_{\text{L}}^{0} A_{\text{R}}^{0} \) from O 1(t 1) to O 2(t 2), and \( A_{\text{L}}^{2} A_{\text{R}}^{2} \) is obtained through rotating \( A_{\text{L}}^{1} A_{\text{R}}^{1} \) by ϕ R1. \( A_{\text{L}}^{2} A_{\text{R}}^{2} \) is actually the central line of the rigid wheelset axle at t 2. Figure 7 shows that the wheels are assumed to be rigid and always perpendicular to the deformed axle line at their connections at any time t 2.

Fig. 7
figure 7

A flexible wheelset moving from its initial reference state (O 1(t 1)) to its any status (O 2(t)) in the plane of O-YZ

To clearly describe the new wheel–rail contact model, the dummies of the two rigid half wheelsets, as shown in Fig. 8, are employed to describe wheel–rail rolling contact behavior affected by the wheelset bending. The two dummies are indicated by DWL and DWR, respectively, and the wheels of the DWL and the DWR are assumed to overlap the left and right wheels of the flexible wheelset, respectively; all the time, namely, the motion of the assumed rigid wheels of the flexible wheelset can be described by the DWL and the DWR (Fig. 8). ϕ R2 is the roll angle of the right wheel due to the bending deformation of the flexible wheelset. It is exactly the included angle between the line \( A_{\text{L}}^{2} A_{\text{R}}^{2} \) and the axle line of the right wheel or the wheel of the DWR.

Fig. 8
figure 8

Relationship between the two rigid half wheelset dummies and the flexible wheelset

It is not difficult to calculate the wheel–rail contact geometry considering the effect of the flexible deformation of the wheelset or the local high-frequency deformations of the wheels if the spatial positions of the DWL and the DWR are determined. Determining the spatial positions of the DWL and the DWR involves calculating their motion parameters, such as the lateral displacements of the centers of the wheels of the DWL and the DWR, indicated by y DWL and y DWR, respectively, the vertical displacements, z DWL and z DWR, the roll angles, ϕ DWL and ϕ DWR, and the yaw angles, ψ DWL and ψ DWR. These parameters are key to calculating the contact geometry of the flexible wheelset in rolling contact with a pair of rails by using this new wheel–rail contact model. This will now be demonstrated in detail.

Figure 8 describes the motion of the DWL and the DWR influenced by the wheelset bending and its rigid motion in the O-YZ plane only. After the rigid wheelset moves with the center displacement of O 1 O 2 and the rolling angle of ϕ R1 in the O-YZ plane of the global reference, O-XYZ, its center position O 1(t 1) reaches the position O 2(t 2) and \( A_{\text{L}}^{0} A_{\text{R}}^{0} \) reaches (or becomes) \( A_{\text{L}}^{2} A_{\text{R}}^{2} \). Note that the vector O 1 O 2 and the roll angle ϕ R1 around axis X are described in the O-YZ plane. The dash-dot line \( A_{\text{L}}^{1} A_{\text{R}}^{1} \) is through point O 2(t 2) and parallel to \( A_{\text{L}}^{0} A_{\text{R}}^{0} \). From Fig. 6, it is obvious that the rolling angle of the DWR caused by the wheelset rigid motion is just ϕ R1 and that caused by the wheelset bending deformation is ϕ R2, so the total rolling angle of the DWR is ϕ DWR = ϕ R1 + ϕ R2, as shown in Fig. 6.

In addition, the displacement of the DWR is the vector O 1 O 3R, which could be written as:

$$ {\varvec{O}}_{ 1} {\varvec{O}}_{{ 3 {\text{R}}}} = {\varvec{O}}_{ 1} {\varvec{O}}_{ 2} + {\varvec{O}}_{ 2} {\varvec{O}}_{{ 3 {\text{R}}}} . $$
(26)

In Fig. 8, the vector O 2 A 1 is parallel to O 3R A 2 with the same length l 0. l 0 is actually the distance between the center of the wheel nominal circle and the center of the undeformed wheelset. The vector O 2 O 3R is parallel to A 1 A 2, with the same length. Thus, O 1 O 3R can be written as:

$$ {{\varvec{O}}_{ 1}} {{\varvec{O}}_{ 3 {\text{R}}}} = {{\varvec{O}}_{ 1}} {{\varvec{O}}_{ 2}} + {{\varvec{A}}_{ 1}} {{\varvec{A}}_{ 2}} = {{\varvec{O}}_{ 1}} {{\varvec{O}}_{ 2}} + \left( {{\varvec{O}}_{ 2}} {{\varvec{A}}_{ 2}} - {{\varvec{O}}_{ 2}} {{\varvec{A}}_{ 1}} \right). $$
(27)

Moreover, the vector O 2 A 1 is described by {x 1 y 1 z 1}[i j k]T in O-XYZ and can be obtained by rotating the vector {0 l 0 0}[i j k]T (coinciding with the line \( A_{\text{L}}^{1} A_{\text{R}}^{ 1} \)) about the X-axis by ϕ DWR. O 2 A 1 is written as:

$$ \begin{aligned} {\varvec{O}}_{2} {\varvec{A}}_{ 1} & = \left\{ {\begin{array}{*{20}c} {x_{1} } & {y_{1} } & {z_{1} } \\ \end{array} } \right\}\left[ {\begin{array}{*{20}c} {\varvec{i}} & {\varvec{j}} & {\varvec{k}} \\ \end{array} } \right]^{\text{T}} \\ & = \left\{ {\begin{array}{*{20}c} 0 \\ {l_{0} } \\ 0 \\ \end{array} } \right\}^{\text{T}} \left[ {\begin{array}{*{20}c} 1 & 0 & 0 \\ 0 & {\cos (\phi_{\text{R1}} + \phi_{\text{R2}} )} & {\sin (\phi_{\text{R1}} + \phi_{\text{R2}} )} \\ 0 & { - \sin (\phi_{\text{R1}} + \phi_{\text{R2}} )} & {\cos (\phi_{\text{R1}} + \phi_{\text{R2}} )} \\ \end{array} } \right]\left[ {\begin{array}{*{20}c} {\varvec{i}} \\ {\varvec{j}} \\ {\varvec{k}} \\ \end{array} } \right]. \\ \end{aligned} $$
(28)

The curve (Fig. 6) is the deformed axle centerline of the wheelset, which does not consider the influence of the rotation caused by the wheelset rigid motion. The point B R is the center of the right nominal circle. The axle centerline of the deformed wheelset can be obtained by rotating about the X-axis by ϕ R1. According to the definition of the curve , the vector O 2 B R is defined as:

$$ {\varvec{O}}_{2} {\varvec{B}}_{\text{R}} = \left\{ {\begin{array}{*{20}c} {x_{2} } \\ {y_{2} } \\ {z_{2} } \\ \end{array} } \right\}^{\text{T}} \left[ {\begin{array}{*{20}c} {\varvec{i}} \\ {\varvec{j}} \\ {\varvec{k}} \\ \end{array} } \right] = \left\{ {\begin{array}{*{20}c} {\Delta x_{2} } \\ {\Delta y_{2} + l_{0} } \\ {\Delta z_{2} } \\ \end{array} } \right\}^{\text{T}} \left[ {\begin{array}{*{20}c} {\varvec{i}} \\ {\varvec{j}} \\ {\varvec{k}} \\ \end{array} } \right], $$
(29)

where {Δx 2 Δy 2 Δz 2}[i j k]T is the displacement vector of the center of the right nominal circle due to the axle bending. Then, the vector O 2 A 2 is defined as {x 3 y 3 z 3}[i j k]T and can be written as:

$$ {\varvec{O}}_{2} {\varvec{A}}_{ 2} = \left\{ {\begin{array}{*{20}c} {x_{3} } \\ {y_{3} } \\ {z_{3} } \\ \end{array} } \right\}^{\text{T}} \left[ {\begin{array}{*{20}c} {\varvec{i}} \\ {\varvec{j}} \\ {\varvec{k}} \\ \end{array} } \right] = \left\{ {\begin{array}{*{20}c} {x_{2} } \\ {y_{2} } \\ {z_{2} } \\ \end{array} } \right\}^{\text{T}} \left[ {\begin{array}{*{20}c} 1 & 0 & 0 \\ 0 & {\cos \phi_{\text{R1}} } & {\sin \phi_{\text{R1}} } \\ 0 & { - \sin \phi_{\text{R1}} } & {\cos \phi_{\text{R1}} } \\ \end{array} } \right]\left[ {\begin{array}{*{20}c} {\varvec{i}} \\ {\varvec{j}} \\ {\varvec{k}} \\ \end{array} } \right], $$
(30)

which is obtained according to the relationship between and or obtained by rotating by ϕ R1. The wheelset center displacement vector O 1 O 2 is defined as {x 0 y 0 z 0}[i j k]T.

Substituting Eqs. (28) and (30) and the expression of O 1 O 2 into Eq. (27), the vector O 1 O 3R can be written as:

$$ \begin{aligned} {\varvec{O}}_{1} {\varvec{O}}_{{ 3 {\text{R}}}} & = \left( {\left\{ {\begin{array}{*{20}c} {x_{0} } \\ {y_{0} } \\ {z_{0} } \\ \end{array} } \right\}^{\text{T}} + \left\{ {\begin{array}{*{20}c} {\Delta x_{2} } \\ {\Delta y_{2} + l_{0} } \\ {\Delta z_{2} } \\ \end{array} } \right\}^{\text{T}} {\varvec{M}}_{1} - \left\{ {\begin{array}{*{20}c} 0 \\ {l_{0} } \\ 0 \\ \end{array} } \right\}^{\text{T}} {\varvec{M}}_{2} } \right)\left[ {\begin{array}{*{20}c} {\varvec{i}} \\ {\varvec{j}} \\ {\varvec{k}} \\ \end{array} } \right], \\ {\varvec{M}}_{1} & = \left[ {\begin{array}{*{20}c} 1 & 0 & 0 \\ 0 & {\cos \phi_{\text{R1}} } & {\sin \phi_{\text{R1}} } \\ 0 & { - \sin \phi_{\text{R1}} } & {\cos \phi_{\text{R1}} } \\ \end{array} } \right], \\ {\varvec{M}}_{2} & = \left[ {\begin{array}{*{20}c} 1 & 0 & 0 \\ 0 & {\cos (\phi_{\text{R1}} + \phi_{\text{R2}} )} & {\sin (\phi_{\text{R1}} + \phi_{\text{R2}} )} \\ 0 & { - \sin (\phi_{\text{R1}} + \phi_{\text{R2}} )} & {\cos (\phi_{\text{R1}} + \phi_{\text{R2}} )} \\ \end{array} } \right]. \\ \end{aligned} $$
(31)

Similarly, when considering the wheelset bending deformation in the plane O-XY, the vector O 1 O 3R should be given as

$$ \begin{aligned} & {\varvec{O}}_{1} {\varvec{O}}_{{ 3 {\text{R}}}} = \left( {\left\{ {\begin{array}{*{20}c} {x_{0} } \\ {y_{0} } \\ {z_{0} } \\ \end{array} } \right\}^{\text{T}} + \left\{ {\begin{array}{*{20}c} {\Delta x_{2} } \\ {\Delta y_{2} + l_{0} } \\ {\Delta z_{2} } \\ \end{array} } \right\}^{\text{T}} {\varvec{M}}_{1} {\varvec{M}}_{3} - \left\{ {\begin{array}{*{20}c} 0 \\ {l_{0} } \\ 0 \\ \end{array} } \right\}^{\text{T}} {\varvec{M}}_{2} {\varvec{M}}_{4} } \right)\left[ {\begin{array}{*{20}c} {\varvec{i}} \\ {\varvec{j}} \\ {\varvec{k}} \\ \end{array} } \right], \\ & {\varvec{M}}_{3} = \left[ {\begin{array}{*{20}c} {\cos \psi_{\text{R1}} } & {\sin \psi_{\text{R1}} } & 0 \\ { - \sin \psi_{\text{R1}} } & {\cos \psi_{\text{R1}} } & 0 \\ 0 & 0 & 1 \\ \end{array} } \right], \\ & {\varvec{M}}_{4} = \left[ {\begin{array}{*{20}c} {\cos (\psi_{\text{R1}} + \psi_{\text{R2}} )} & {\sin (\psi_{\text{R1}} + \psi_{\text{R2}} )} & 0 \\ { - \sin (\psi_{\text{R1}} + \psi_{\text{R2}} )} & {\cos (\psi_{\text{R1}} + \psi_{\text{R2}} )} & 0 \\ 0 & 0 & 1 \\ \end{array} } \right], \\ \end{aligned} $$
(32)

where ψ R1 and ψ R2 are the yaw angles caused by the rigid motion and the bending deformation in the plane O-XY, respectively.

ψ DWR = ψ R1 + ψ R2 is the total yaw angle of the DWR. Similarly, the position of the DWL can be obtained. When the positions of the two dummies are known at t 2, the wheel–rail contact geometry can be calculated. Then, the positions of the wheel–rail contact points are easily found, and the wheel–rail contact forces can be calculated. The normal wheel–rail contact forces are calculated by the Hertzian nonlinear contact spring model, and the tangent contact forces and spin moments are calculated by means of the model by Shen et al. (1983). Compared with the conventional wheel–rail contact model (Wang 1984; Zhai 2007), this new wheel–rail contact model can characterize the independent high-frequency deformations of the two wheels of the flexible wheelset more conveniently.

3 Results and Discussion

When a vehicle is running on an ideal track , it is only excited by sleepers. Note that the “flexible” wheelset model used in this section denotes the model considering the first two bending modes. The dynamic system with flexible wheelset models is used in the simulation on an ideal track at the speed of 300 km/h. Figure 9a, b shows the vertical forces in the frequency domain in steady and unsteady stages, respectively. In the unsteady stage, the peaks appear not only at a set of harmonic frequencies nf s (n = 1, 2, 3, …) produced by passing sleeper but also at f b1, while the influence of the second bending mode is small since there is no peak at f b2. In the steady stage, the contribution of the component at f b1 is weakened and only the peaks at nf s (n = 1, 2, 3, …) remain. These results are reasonable because when a system comes to a steady stage, its responses only contain the component at the excitation frequency.

Fig. 9
figure 9

Vertical contact force in steady (a) and unsteady (b) stages

Based on a large range of site measurements, the components of roughness on rails mostly appear in the range of 1–20 m. The natural frequencies of the first two bending modes are below 250 Hz, meaning the available frequency of this model is limited. Therefore, the components of the random irregularity on the rails are mainly in the frequency range of 0–150 Hz at the speed of 300 km/h. Figure 10a presents the local section of 900–950 m in the time domain, and Fig. 10b shows the irregularity in the frequency domain. Note that the results below are from the steady stage.

Fig. 10
figure 10

Random irregularity in the time domain (a) and frequency domain (b)

Figures 11 and 12 show the wheel–rail contact forces acting on the rigid and flexible wheelsets in the time and frequency domains , respectively. As shown in Fig. 11a, the average of the oscillation of the lateral contact force acting on the flexible model is little smaller than that on the rigid wheelset model, and the shapes of the oscillation are different. As shown in Fig. 11b, the vertical contact forces acting on the two models oscillate around a similar average, while their shapes are different. These differences are caused by the wheelset flexibility.

Fig. 11
figure 11

Lateral contact forces (a) and vertical contact forces (b) in the time domain

Fig. 12
figure 12

Lateral contact forces (a) and vertical contact forces (b) in the frequency domain

In the frequency domain, the distributions of the components contained in both the lateral and vertical contact forces are in the excitation frequency range of the random irregularity. A peak at frequency 2 f s appears in Fig. 11a, b. The contribution of the component at frequency f s is overwhelmed by the effect of the irregularity . In addition, the uniform distribution in 0–150 Hz of the irregularity results in the non-uniform distribution of contact forces . As shown in Fig. 12a, b, the components in 80–150 Hz are higher than those in 0–80 Hz. This shows that under this present irregularity, this dynamic system is more sensitive to the excitation in 80–150 Hz than to that in 0–80 Hz.

In the frequency domain , the component at f b1 of the lateral contact force acting on the flexible model is a little larger than that on the rigid model , as marked using the arrow in Fig. 12a. This shows that the first bending mode is excited, and the availability of the model to characterize the wheelset bending is proved. However, there is no evident difference at f b1 for vertical contact forces acting on the two models. This shows that the wheelset bending deformation has a stronger effect on the lateral contact force than on the vertical contact force.

The wheel–rail contact force is affected by the position of the lateral contact points. Figure 13a, b shows the oscillations of the contact points in lateral direction described in the body coordinate system attached to the rail cross section in the time and frequency domains, respectively. The average of the magnitudes of oscillation of the contact points on the flexible model in the time domain is larger than that on the rigid model . This is caused by the wheelset bending. Moreover, it can weaken the relative movement between rail and wheel caused by the irregularity . Therefore, it is one cause of the smaller average of the lateral contact force acting on the flexible model (Fig. 11a). As shown in Fig. 13b, the difference of the components between the two models at f b1 is evident. This explains the difference in the time domain (Fig. 13a) and again shows the effectiveness of the proposed model.

Fig. 13
figure 13

Oscillation of contact points in lateral direction in the time domain (a) and frequency domain (b)

4 Conclusions

In this study, a new wheel–rail contact model is integrated into the high-speed vehicle–track coupling dynamics system model, which takes into account the effect of wheelset structural flexibility. Based on the new vehicle–track model, the effect of the first two bending modes of the wheelset on wheel–rail contact behavior is analyzed under the random irregularity in a frequency range of 0–150 Hz. The numerical results of the rigid wheelset model and the flexible wheelset model are compared in detail. The following conclusions can be drawn from the results:

  1. 1.

    The present vehicle–track model considering flexible wheelsets can very well characterize the effect of the flexible wheelset on wheel–rail dynamic behavior.

  2. 2.

    Under the excitation, the shapes of the oscillations of the wheel–rail contact forces and contact points for the new and conventional vehicle–track models are different. The difference is caused by the excited first bending mode of the wheelset.

For future work, the first improvement to be considered is to model a wheelset using the FEM or the Timoshenko beam theory to broaden the model’s available frequency range. This could allow it to help investigate the mechanisms behind the generation and development of wheel–rail wear and noise.