Nothing Special   »   [go: up one dir, main page]

Next Article in Journal
Mechanical Property Degradation of Transmission Wire Rope Caused by Different Wear Evolution
Next Article in Special Issue
Improved Calculation of Dynamic Load Capacity for Cylindrical Roller Thrust Bearings: Numerical Update of the Lifetime Reduction Factor η for Bearings with Small and Medium Spin-to-Roll Ratios
Previous Article in Journal
Investigation of the Contact Characteristics of a Single-Nut Ball Screw Considering Geometric Errors
Previous Article in Special Issue
Analysis of Chaotic Features in Dry Gas Seal Friction State Using Acoustic Emission
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dynamic Response Analysis of Ballastless Tracks Considering the Temperature-Dependent Viscoelasticity of Cement-Emulsified Asphalt Mortar Based on a Vehicle–Track–Subgrade Coupled Model

School of Rail Transportation, Soochow University, Suzhou 215131, China
*
Authors to whom correspondence should be addressed.
Lubricants 2025, 13(2), 58; https://doi.org/10.3390/lubricants13020058
Submission received: 5 December 2024 / Revised: 27 January 2025 / Accepted: 28 January 2025 / Published: 30 January 2025
(This article belongs to the Special Issue Recent Advances in Lubricated Tribological Contacts)
Figure 1
<p>CRTS I ballastless track.</p> ">
Figure 2
<p>The FM model.</p> ">
Figure 3
<p>Zener model.</p> ">
Figure 4
<p>Fitting curves of the creep compliance.</p> ">
Figure 5
<p>Finite element model for the compressive creep experiment with CA mortar.</p> ">
Figure 6
<p>Simulation results of the creep experiment.</p> ">
Figure 7
<p>Vertical vehicle–track–subgrade system model [<a href="#B30-lubricants-13-00058" class="html-bibr">30</a>].</p> ">
Figure 8
<p>Vertical track irregularity.</p> ">
Figure 9
<p>The VTS model.</p> ">
Figure 10
<p>Track subsystem simulation results. (<b>a</b>) Rail displacement; (<b>b</b>) slab displacement; (<b>c</b>) fastener force; (<b>d</b>) CA mortar stress.</p> ">
Figure 11
<p>Comparison of the simulation results with the relevant literature. (<b>a</b>) Slab displacement and rail displacement; (<b>b</b>) fastener force and CA mortar stress.</p> ">
Figure 12
<p>Comparison of time history curves of dynamic responses of the track subsystem. (<b>a</b>) Rail displacement; (<b>b</b>) slab displacement; (<b>c</b>) CA mortar displacement; (<b>d</b>) compressive displacement of CA mortar; (<b>e</b>) CA mortar stress; (<b>f</b>) base displacement.</p> ">
Figure 12 Cont.
<p>Comparison of time history curves of dynamic responses of the track subsystem. (<b>a</b>) Rail displacement; (<b>b</b>) slab displacement; (<b>c</b>) CA mortar displacement; (<b>d</b>) compressive displacement of CA mortar; (<b>e</b>) CA mortar stress; (<b>f</b>) base displacement.</p> ">
Figure 13
<p>Comparison of the maximum simulation outcomes. (<b>a</b>) Rail and slab displacement; (<b>b</b>) CA mortar and concrete base displacement; (<b>c</b>) compressive displacement of CA mortar; (<b>d</b>) CA mortar stress.</p> ">
Figure 14
<p>Comparison of the maximum dynamic response results of the vehicle subsystem. (<b>a</b>) Vertical acceleration of the vehicle body; (<b>b</b>) wheel–rail force.</p> ">
Figure 15
<p>Growth rate of dynamic responses at different temperatures.</p> ">
Figure 16
<p>Comparison of the maximum dynamic response results of the track subsystem. (<b>a</b>) Slab displacement; (<b>b</b>) CA mortar displacement; (<b>c</b>) compression displacement of CA mortar; (<b>d</b>) CA mortar stress.</p> ">
Figure 17
<p>Comparison of the maximum dynamic response results of the track subsystem. (<b>a</b>) Rail displacement; (<b>b</b>) CA mortar displacement; (<b>c</b>) CA mortar acceleration; (<b>d</b>) CA mortar stress.</p> ">
Figure 18
<p>Position diagram for the CA mortar layer.</p> ">
Figure 19
<p>Comparison of the maximum dynamic response results of CA mortar. (<b>a</b>) CA mortar displacement; (<b>b</b>) CA mortar stress.</p> ">
Figure 20
<p>CA mortar layer compressive stress cloud diagram.</p> ">
Versions Notes

Abstract

:
This study aims to explore the dynamic response of ballastless tracks under various temperatures of the cement-emulsified asphalt (CA) mortar layer and other environmental factors. CA mortar is the key material in the ballastless track structure, exhibiting notably temperature-dependent viscoelastic properties. It can be damaged or even fail due to the continuous loads from trains. However, the dynamic behaviors of ballastless tracks considering the temperature-dependent viscoelasticity of CA mortar have been insufficiently studied. This paper captures the temperature-dependent viscoelastic characteristics of CA mortar by employing the fractional Maxwell model and applying it to finite element simulations through a Prony series. A vehicle–track–subgrade (VTS) coupled CRTS I ballastless track model, encompassing Hertz nonlinear contact and track irregularity, is established. The model is constrained symmetrically on both of the longitudinal sides, and the bottom is fixed on the infinite element boundary, which can reduce the effects of reflected waves. After the simulation outcomes in this study are validated, variations in the dynamic responses under different environmental factors are analyzed, offering a theoretical foundation for maintaining the ballastless tracks. The results show that the responses in the track subsystem will undergo significant changes as the temperature rises; a notable effect is caused by the increase in speed and fastener stiffness on the entire system; the CA mortar layer experiences the maximum stress at its edge, which makes it highly susceptible to damage in this area. The original contribution of this work is the establishment of a temperature-dependent vehicle–track–subgrade coupled model that incorporates the viscoelasticity of the CA mortar, enabling the investigation of dynamic responses in ballastless tracks.

1. Introduction

High-speed railways have swiftly advanced worldwide for their exceptional efficiency and excellent safety, as evidenced by notable developments in countries such as China, Germany, and Japan [1]. Ballastless tracks, serving as the carrier for high-speed trains, have emerged as the premier choice in contemporary railway construction [2] for high smoothness and low maintenance demands. In China, CRTS I ballastless tracks, which are shown in Figure 1, have surpassed 2700 km in total operating mileage, and the cement-emulsified asphalt (CA) mortar layer is a crucial structure throughout the construction and operation period, performing functions such as leveling and vibration reduction. Its mechanical performance significantly affects the operation safety of trains [3]. A CRTS I ballastless track will undergo damage or even failure under dynamic loads from the train, primarily manifested in the large-scale cracking and peeling of CA mortar layers, which constitutes a serious risk to train operation [4]. Therefore, evaluating how CA mortar’s mechanical characteristics effect the dynamic behaviors of ballastless tracks is of great importance.
Given that CA mortar is mainly constituted by emulsified asphalt, the mechanical behaviors of the CA mortar layer demonstrate typical viscoelastic characteristics. This time-dependent viscoelasticity has made it difficult to study the dynamic responses of ballastless tracks [5,6]. Studies have indicated a significant relationship between temperature and CA mortar viscoelastic properties. Fu et al. [7] conducted relaxation tests on CA mortar and established a stress relaxation function with temperature-sensitive factors. Wang et al. [8] carried out compressive tests on CA mortar and derived the stress–strain relationship at different temperatures. Zhu et al. [9] found that CA mortar’s creep behavior was aggravated when the temperature rose. Yuan et al. [10] examined the dynamic physical reactions of CA mortar and obtained viscoelastic curves in relation to temperature. You et al. [11] discovered that the elastic modulus of asphalt materials decreases as the temperature rises. Additionally, Xu et al. [12] developed a molecular chain network model to analyze how temperature variation influences the mechanical behavior of viscoelastic materials from a micro-level perspective. The above research mainly focused on physical experiments with CA mortar and connected different mechanical behaviors with the temperature, while the alterations in dynamic responses of the track system caused by this require further investigation. As the structure most prone to damage, the CA mortar layer gains considerable attention in dynamics studies of the track system. For instance, Dai [13] constructed a model of ballastless tracks in ABAQUS and carried out comparative analyses of the displacement and stress in slab when CA mortar was regarded as a viscoelastic or elastic material. Ren et al. [14] investigated the impact of the cracking area in the CA mortar layer on track structures with a vertical finite element model. However, this study regarded CA mortar as an elastic material. In Ref. [15], Ren simulated the deformation process of CA mortar and established a track structure model with the help of the time-hardening rate viscoelastic model in ABAQUS. Subsequently, they set different Young’s moduli for CA mortar to analyze the pattern of displacement variations. Deng et al. [16] wrote a subroutine to simulate the failure evolution process of CA mortar and then studied how the cracking area influences the fatigue accumulation of the CA mortar layer via a full-scale model in ABAQUS. Chen et al. [17] investigated the distortion between the layers of ballastless tracks and predicted the fatigue life of the ballastless tracks. Nevertheless, these studies generally use point loads to simulate the wheel–rail forces, neglecting track irregularity and wheel–rail contact, which leads to errors in load transfer. In analyzing the wheel–rail forces, Hertz nonlinear theory is typically employed for calculations [18]. Based on this, researchers have established corresponding models for different adhesion conditions between the wheel and rail to obtain the response of creeping forces [19,20]. Additionally, the aforementioned studies did not consider CA mortar’s temperature-dependent viscoelastic properties. This research aims to investigate how temperature variation in CA mortar affects the dynamic responses of ballastless tracks. In Section 2, the fractional Maxwell model is introduced to capture the creep curves, and we obtain the viscoelastic parameters required in ABAQUS to realize the creep progress of CA mortar using the Prony series. Then, a coupled vehicle–track–subgrade model is built. We employ Hertz nonlinear contact theory to calculate the vertical wheel–rail force. Additionally, the tangential friction coefficient is introduced to determine the lateral force. After comparing and verifying the simulation data with actual values, discrepancies in the dynamic responses under different environmental factors are further explored in Section 3. Section 4 presents the conclusions of this study.

2. Materials and Methods

Previous studies have often described the viscoelastic behavior of materials using the Zener model, Burgers model, or more complex classical models [21,22]. It was reported that fractional derivative models can describe the mechanical response of viscoelastic materials more accurately with fewer parameters [23]. The viscoelastic curves of CA mortar are captured by the fractional derivative model in this section.

2.1. The Fractional Maxwell (FM) Model

We can see from Figure 2 that by substituting the Newtonian dashpot in the Maxwell model with an Abel dashpot, a fractional Maxwell (FM) model can be obtained, and the constitutive relation is described by [24]
ε = ε e + ε v e ε e = σ / E σ = η D α ε v e
where εe and εve are the strains of the elastic and viscoelastic bodies, E represents the elastic modulus, η represents the viscosity coefficient, and Dα denotes the fractional derivative, which is defined in the Riemann–Liouville form [25]:
D α f t = Γ 1 α 1 d / d t 0 t f τ t τ α d τ
Researchers have established a bridge in the Laplace domain between different viscoelastic behaviors with the help of the transfer function T(s); this can be achieved using the ratio of stress to strain [26], which has been employed in viscoelastic models [27]. Table 1 presents the transfer functions of basic mechanical components.
According to the definition, the transfer function T(s) of the FM model is achieved:
T s = E η s α / ( η s α + E )
In addition, the creep compliance J(t) is achieved from Equation (6) [26]:
J t = L 1 1 / s / T s
For the FM model, the creep compliance is obtained by
J t = 1 / E + 1 / η Γ 1 + α

2.2. Viscoelastic Description of CA Mortar Based on Creep Experimental Data

We adopt the results of the creep test on CA mortar with the temperature ranging from −20 to 40 °C in Ref. [28] and then employ the FM and Zener model to capture the creep curves. The Zener model consists of 2 springs and a Newtonian dashpot, as shown in Figure 3, and the creep compliance includes both elastic and viscoelastic components:
J t = 1 / E 1 + 1 / E 2 1 e E 2 t / λ
where E1 and E2 represent the elastic modulus of the springs.
The fitting results are presented in Figure 4, and Table 2 lists the fitting parameters of the two models. It can be found that under the same condition of using three parameters, the FM model describes the creep behavior much better than the Zener model, with a goodness of fit no less than 0.95. Moreover, compared to other models with more parameters, such as the Burgers model, the advantage of having fewer parameters is that it avoids overfitting and is less likely to produce fitting results that lack physical meaning, making it more beneficial for researchers to analyze the relationship between parameters and environmental factors [24].
In ABAQUS, the viscoelastic properties of materials are expressed by the generalized Maxwell model, and we can simulate the creep progress of viscoelastic materials by its relaxation modulus, the Prony series [28]:
G t = G 0 G 0 i = 1 k g i 1 exp t / τ i
where G0 represents the instantaneous modulus, k is the number of the Maxwell model, gi is the relative modulus, and τi is the relaxation time for the i-th Maxwell model.
For the purpose of describing the viscoelasticity of CA mortar in ABAQUS, the FM model is converted to the Prony series in this section. By performing the Laplace transform in Equation (9), the transfer function with parameters of the Pony series is obtained:
T s = G 0 1 i = 1 k g i 1 + τ i s 1
According to Equation (2), we can obtain the parameters in Equation (7) by numerical inverse Laplace transform [29], and the parameters are optimized by a fitting program. The parameters are presented in Table 3, and k is set to 7, which is accurate enough to capture the creep curves.
To confirm the accuracy of the fitting parameters, referring to the specimen size in Ref. [28], a φ110 × 110 mm finite element model is established to simulate the uniaxial compression creep experiment in ABAQUS, as presented in Figure 5. The model adopts C3D8 elements with a total number of 13,328. The stress applied at the top of the model is 0.1 MPa, and three displacement degrees of freedom are fixed at the other end. The results of the simulation are presented below.
Figure 6 shows that despite the presence of diverse errors in the simulation of compressive creep under 4 temperatures, the accuracy is high enough to provide a favorable foundation for constructing a model of ballastless tracks, and the maximum error is 1.97%.

2.3. Establishing a Vehicle–Track–Subgrade (VTS) Finite Element Model

An accurate model that is congruent with reality is the basis for conducting dynamics studies. We constructed a VTS coupled model in ABAQUS. In this study, the vehicle subsystem is regarded as comprising rigid bodies to enhance the computational efficiency, and they are linked by the connection sections in ABAQUS. Stiffness and damping values are assigned to the connection sections to imitate the suspension system.
Figure 7 presents the vertical schematic diagram of this model, and the vertical motion equation in the vehicle subsystem is obtained by D’Alembert’s principle [18].
Heaving motion of vehicle body:
M v Z ¨ v + C v b 2 Z ˙ v Z ˙ b 1 Z ˙ b 2 + K v b 2 Z v Z b 1 Z b 2 = M v g
Nodding motion of vehicle body:
J v β ¨ v + C v b l b 2 β ˙ v + Z ˙ b 1 Z ˙ b 2 + K v b l b 2 β v + Z b 1 Z b 2 = 0
Heaving motion of front bogie:
M b Z ¨ b 1 + C b w 2 Z ˙ b 1 Z ˙ w 1 Z ˙ w 2 + C v b 2 Z ˙ b 1 Z ˙ v l v β ˙ v + K b w 2 Z b 1 Z w 1 Z w 2 + K v b 2 Z b 1 Z v Z v β v = M b g
Nodding motion of front bogie:
J b β ¨ b 1 + l b C b w 2 l b β ˙ b 1 Z ˙ w 1 + Z ˙ w 2 + l b K b w 2 l b β b 1 Z b 1 + Z b 2 = 0
Heaving motion of rear bogie:
M b Z ¨ b 2 + C b w 2 Z ˙ b 2 Z ˙ w 3 Z ˙ w 4 + C v b 2 Z ˙ b 2 Z ˙ c l c β ˙ c + K b w 2 Z b 2 Z w 3 Z w 4 + K v b 2 Z b 2 Z c Z c β c = M b g
Nodding motion of rear bogie:
J b β ¨ b 2 + l b C b w 2 l b β ˙ b 2 Z ˙ w 3 + Z ˙ w 4 + l b K b w 2 l b β b 2 Z b 3 + Z b 4 = 0
Heaving motion s of four wheelsets:
M w 1 Z ¨ w 1 + C b w Z ˙ w 1 Z ˙ b 1 + l b β b 1 + K b w Z w 1 Z b 1 + l b β b 1 = M w 1 g + F r w , 1
M w 2 Z ¨ w 2 + C b w Z ˙ w 2 Z ˙ b 1 l b β b 1 + K b w Z w 2 Z b 1 l b β b 1 = M w 2 g + F r w , 2
M w 3 Z ¨ w 3 + C b w Z ˙ w 3 Z ˙ b 2 + l b β b 2 + K b w Z w 3 Z b 2 + l b β b 2 = M w 3 g + F r w , 3
M w 4 Z ¨ w 4 + C b w Z ˙ w 4 Z ˙ b 2 l b β b 2 + K b w Z w 4 Z b 2 l b β b 2 = M w 4 g + F r w , 4
The meanings of the symbols in the formula above are shown in Table 4.
To reduce the boundary effect, this model has 21 track slabs, and the total length is 106 m. C3D8R hexahedral solid elements with eight nodes are employed to model the track and subgrade subsystem, with a total number of 100,440. The fastener system is replaced with connection sections, simulating the vibration reduction effect by defining stiffness and damping coefficients on connection sections. In actual engineering projects, the relative displacements between the various structures of the ballastless tracks are very small, so the layers are bound by tie connections to lower computational expenses. Since both the base and the subgrade are longitudinally continuous structures, symmetric constraints are applied at both ends of these structures to ensure that the model does not experience rigid body displacement [13,30]. The bottom is fixed at the infinite element boundary, which helps to reduce the effects of reflected waves, especially when the train is moving at high speeds [31].
The vehicle connects to the rail via the interaction between them, and the Hertz nonlinear theory provides a formula for calculating the vertical force [18]:
q t = Δ Z t / G 3 / 2
where G represents the wheel–rail contact constant, and ∆Z(t) is the total compression amount in both of them.
For high-speed trains, worn-profile tread wheels are used, and G is calculated as follows:
G = 3.86 R 0.155 × 10 8
where R represents the wheel radius.
The lateral friction coefficient µ determines the tangential wheel–rail force F(t), and it is set to 0.3.
F t = μ × q t
Track irregularities will cause the train to vibrate irregularly, presenting a risk to the operational safety of the train. This paper employs the spectral density function for the American sixth-level track to simulate the vertical track irregularity; the formula is given by
S v ϕ = j A v ϕ c 2 / ϕ 2 ϕ 2 + ϕ c 2
where j is a constant coefficient that is taken as 0.25, Av represents the coefficient of roughness (cm2·rad·m−1), and ϕ and ϕc are the spatial frequency and the cut-off frequency (cm2·rad·m−1), respectively, as shown in Table 5.
The data for vertical track irregularity along the advancing direction are obtained after numerical transformation [32] and are shown in Figure 8.
We can modify the coordinates of the rail elements to generate the track irregularity in ABAQUS. The constructed finite element model is shown in Figure 9.

2.4. Model Verification

In this model, the vehicle adopted is a power-distributed train developed in China. Table 6 lists the parameters of the vehicle and fastener [33].
The track–subgrade subsystem is simulated by C3D8R solid elements, and its physical property parameters are derived from actual values [34,35], as presented in Table 7. Additionally, the viscoelastic parameters listed in Table 3 are assigned to the CA mortar.
We utilize a computer with an Intel(R) Core(TM) i7-12700 CPU to perform calculations for this model following explicit dynamic steps. The simulation results in the vertical direction for the 4th fastener at the 11th slab at a speed of 300 km/h and temperature of 20 °C are presented in Figure 10.
Table 8 lists the calculated results (taking the maximum value) from this paper and compares them with relevant literature [36] and actual measurement data [37].
From the data presented, we can see that the results in this paper are similar to measured values from the actual lines, while the rail displacement is slightly larger than the measured value. This is because the simulation environment cannot fully reproduce the measured conditions, and both have certain discrepancies, leading to unavoidable errors in various parameters, but the errors are within an acceptable range. Furthermore, from the comparison in Figure 11, it can be found that the results of this paper are significantly better than those in the literature [36]; thus, it can be concluded that it is feasible to use this model for dynamics analysis.

3. Results and Discussion

3.1. Effect of Temperature on the Dynamic Responses

Applying the parameters in Table 3 to the CA mortar, Figure 12 and Figure 13 show the comparison of the track subsystem dynamic responses in the vertical direction at 300 km/h when the temperature ranges from −20 °C to 40 °C.
Figure 12a,f indicates that changes in temperature have a relatively small influence on the rail and base displacement. This may be because the rail is distant from the CA mortar layer and bears the moving load directly, while the base is continuous and has a mass that is much larger than that of the above two structures. Therefore, the vertical displacements of these two structures are less affected.
It can be observed from Figure 12b–e and Figure 13 that there are obvious differences in the dynamic responses of the CA mortar itself and in the slab displacement due to temperature variations. As the temperature rises, the elastic modulus decreases, leading to an increase in strain when the train passes, which is consistent with Ref. [11]. From −20 °C to 40 °C, the maximum displacement in the CA mortar layer rises from 0.316 mm to 0.352 mm, with a growth rate of approximately 11%; the maximum compression displacement rises from 0.020 mm to 0.055 mm, and growth rate reaches 175%. Since the two structures are adjacent to each other, the displacement in the slab rises from 0.320 mm to 0.357 mm, with a growth rate of 12%, which mirrors that of the CA mortar layer. We can see that the variation laws presented by them are identical: after undergoing an approximately linear increment ranging from −20 °C to 20 °C, the stress rises rapidly; the CA mortar stress diminishes when the temperature ascends, and the stress decreases from 54.878 kN to 41.199 kN, with a reduction rate of 25%. As observed in Ref. [38], a significant reduction in the stress growth rate of CA mortar occurs as temperature rises when subjected to pressure. Moreover, it can be seen that when the train is not passing, the stress in CA mortar is independent of temperature and approaches zero. Therefore, it can be deduced that higher temperatures result in less stress on the CA mortar when the train passes (in agreement with the calculated results in Figure 12e).
The results show that as the train moves over the track, the dynamic responses of track structures are notably influenced by the mechanical characteristics of CA mortar at various temperatures. Therefore, it is necessary to consider temperature as a factor in the mechanical analysis.

3.2. Effect of Speed on Dynamic Responses

Increasing train speed will intensify the wheel–rail interaction, leading to heightened vibrations throughout the entire system. Figure 14 presents the variations in vehicle body acceleration and wheel–rail force under four speeds. It can be observed from the results that the temperature exerts an extremely limited influence on these two variables. When the speed rises from 200 km/h to 350 km/h, the maximum vertical acceleration of the vehicle body at four temperatures rises linearly, with an increment of approximately 60%; the wheel–rail force growth pattern is analogous, and the growth rate is approximately 5%. The increase in wheel–rail force is not significant. This is because the wheelsets are regarded as rigid bodies in this simulation, thereby neglecting their compressive deformation, leading to slightly smaller calculated values using Hertz nonlinear theory.
Figure 15 and Figure 16 show a comparison of some dynamic responses of the track subsystem under different vehicle speeds. The results indicate that, when compared to the vehicle subsystem, temperature variation has a more remarkable impact on the track subsystem; additionally, the effect of vehicle speed is quite significant. When the speed rises from 200 to 350 km/h, the vertical displacements of the track slab and the CA mortar at four temperatures increase significantly by about 24%; the compression displacement and vertical stress of the CA mortar enhance at an accelerating rate with the increase in speed, attaining peak increments of approximately 16% and 18%, respectively.

3.3. Effect of Fastener Stiffness on Dynamic Responses

The fastener system undergoes aging after long-term exposure to train loads, resulting in an increase in its stiffness. Therefore, examining the impact of the vertical stiffness of fasteners on the track structure is very crucial, especially under high-speed conditions. We set the fastener stiffness values to 40 kN/mm, 50 kN/mm, and 60 kN/mm, and the speed was set to 350 km/h. According to the simulation data presented in Figure 17, we can observe that as the stiffness increases, both the rail and CA mortar layer displacements show a decreasing trend. The displacement reduction for the rail reaches 8%, while the change in displacement for the CA mortar is relatively small. Notably, the vertical acceleration and stress of the CA mortar exhibit opposite trends compared to the displacement: the maximum acceleration increases from 5.5 g to 5.8 g, an increase of 6%; the maximum stress rises from 48.4 kPa to 57.1 kPa, an increase of 18%. We can infer that the aging of the fastener system exacerbates the vibrations in track structures, which may reduce the fatigue life of ballastless tracks.

3.4. Effect of Position on Dynamic Responses

This paper analyzes the simulation results of the CA mortar layer at various positions, which are presented in Figure 18 and Figure 19, when the temperature is set to 20 °C. It is not difficult to see that the vertical displacement exhibits a tendency of first decreasing and then increasing from the center to the end, ultimately reaching a maximum value. The vertical stress shows a steady rise. Starting from the third fastener, its growth rate increases sharply. The vertical stress rises by about 72.4 kN in total. Therefore, we can infer that the CA mortar layer is vulnerable to be damaged at its end, thereby triggering catastrophes such as cracks and block shedding.
Throughout the entire dynamic simulation process, stress concentration frequently occurs in the part of the CA mortar layer near the lug boss, as shown in Figure 20. Therefore, we speculate that the displacement restriction of the lug boss is one of the reasons for the increased stress at the edges of the CA mortar layer. Creating a larger contact area between the lug boss and the CA mortar layer, such as by increasing the size of the lug boss or, alternatively, changing its shape to an ellipse, is likely to reduce the magnitude of the CA mortar edge stress.

4. Conclusions

This study establishes a vehicle–track–subgrade coupled model of a CRTS I ballastless track, taking the CA mortar’s temperature-dependent viscoelastic characteristics into account, and investigates the dynamic responses under different environmental factors. First, the creep curves of CA mortar under different temperatures are fitted by the fractional Maxwell (FM) model; subsequently, the creep parameters are transformed into an applicable Prony series in ABAQUS through the transfer function. Based on the Hertz nonlinear theory, a CRTS I finite element model is built in ABAQUS, and then we further analyze the influence of different environmental factors on the dynamic responses of the whole system. The main conclusions are outlined below:
  • The fractional Maxwell (FM) model accurately captures the creep curves of CA mortar, with a fitting degree of excellence above 0.95; moreover, the generalized Maxwell model parameters obtained from the transfer function can be conveniently and accurately implemented for the creep simulation of viscoelastic materials in finite element software.
  • The dynamic responses of the track are significantly impacted by the temperature-dependent viscoelastic characteristics of CA mortar, while the vehicle subsystem is slightly affected. All structural layers deform more as the temperature increases. The changes in the slab and the CA mortar layer are particularly significant. The CA mortar layer stress decreases as the temperature increases. Therefore, it is necessary to consider temperature changes when conducting dynamic analysis.
  • The increase in vehicle speed and fastener stiffness exacerbates the dynamic responses of the CRTS I ballastless track. The vertical acceleration of the vehicle body rises rapidly, which may affect the passenger’s travel experience. The vertical stress in the CA mortar layer rises in a linear manner. The entire system will thus face more intense vibrations. The compression of the CA mortar layer and the growth rate of the vertical stress also rise continuously. Due to the impact of the train load, the CA mortar edge of experiences the highest stress and is most likely to be damaged.
In the near future, we will conduct experiments related to the damage of CA mortar layers, such as viscoelastic responses after freeze–thaw cycles. We will use this model to analyze the impact of damage on the entire system. Finally, we will compare the calculated results with relevant high-speed railway standards to provide maintenance recommendations.

Author Contributions

Conceptualization, Y.C., B.W. and L.Y.; methodology, Y.C.; software, Y.C.; writing—original draft, Y.C.; supervision, X.S.; writing—reviewing and editing, L.Y. and X.S. resources, B.W. and X.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 12302265; the Natural Science Foundation of Jiangsu Province, grant number BK20230469; and the Basic Science (Natural Science) Research Projects in Higher Education Institutions in Jiangsu Province, grant number 23KJB130008.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

εtotal strain of the viscoelastic model
εestrain of the elastic body
εvestrain of the viscoelastic body
σtotal stress of the viscoelastic model
E, η, αparameters of the fractional Maxwell model
E1, E2, λparameters of the Zener model
T(s)transfer function in the LC domain, where s is the complex variable
J(t)creep compliance
G(t)relaxation modulus
G0, k, gi, τiparameters of the generalized Maxwell model
Ttemperature
Mv, Mb, Mwmass of the vehicle, bogie, and wheelset
Jv, Jbpitch rotational inertia of the vehicle and bogie
Zv, Zb, Zwvertical displacement of the vehicle, bogie, and wheelset
βv, βbpitch nodal displacement of the vehicle and bogie
lv, lbhalf of the length of the vehicle and bogie
Kvb, Kbwstiffness of the primary suspension and the secondary suspension
Cvb, Cbwdamping of the primary suspension and the secondary suspension
Frw,kwheel–rail force of the k-th wheelset
q(t)vertical wheel–rail force
Gwheel–rail contact constant
Z(t)compression amount between the wheel and rail
Rradius of the wheel
F(t)tangential wheel–rail force
Sv(ϕ)American power spectral density of track irregularity
j, Av, ϕ, ϕcparameters of the American power spectral density of track irregularity

References

  1. Zeng, X.; Xie, Y.; Deng, D.; Wang, P.; Qu, F. A study of the dynamic mechanical properties of CRTS I type CA mortar. Constr. Build. Mater. 2016, 112, 93–99. [Google Scholar] [CrossRef]
  2. Gautier, P.-E. Slab track: Review of existing systems and optimization potentials including very high speed. Constr. Build. Mater. 2015, 92, 9–15. [Google Scholar] [CrossRef]
  3. Wang, J.; Chen, Y.; Fan, X.; Li, J. Effects of strain rate and confining pressure on compressive behavior of cement asphalt mortar. Mater. Des. (1980–2015) 2015, 65, 772–779. [Google Scholar] [CrossRef]
  4. Zhao, G.; Zhao, L.; Liu, Y. Incompatible thermal deformation of interlayers and corresponding damage mechanism of high-speed railway track structure. High-Speed Railw. 2023, 1, 37–46. [Google Scholar] [CrossRef]
  5. Fang, L.; Huang, J.; Chen, Z.; Zhang, B.; Xu, S.; Lu, J. Frequency Dependency and Constitutive Model of Dynamical Mechanical Properties of Cement Asphalt Binders. J. Mater. Civ. Eng. 2024, 36, 04024129. [Google Scholar] [CrossRef]
  6. Xie, Y.; Fu, Q.; Long, G.; Zheng, K.; Song, H. Creep properties of cement and asphalt mortar. Constr. Build. Mater. 2014, 70, 9–16. [Google Scholar] [CrossRef]
  7. Fu, Q.; Xie, Y.; Long, G.; Meng, F.; Song, H. Temperature sensitivity and model of stress relaxation properties of cement and asphalt mortar. Constr. Build. Mater. 2015, 84, 1–11. [Google Scholar] [CrossRef]
  8. Wang, J.; Wu, X.; Fan, X.; Chen, Y. Stress–strain model of cement asphalt mortar subjected to temperature and loading rate. Constr. Build. Mater. 2016, 111, 164–174. [Google Scholar] [CrossRef]
  9. Zhu, H.; Zeng, X.; Lan, X.; Song, H.; Long, G.; Xie, Y. Effects of High Temperature on Creep Properties of Cement and Emulsified Asphalt Mortar. J. Mater. Civ. Eng. 2022, 34, 04022128. [Google Scholar] [CrossRef]
  10. Yuan, Q.; Liu, W.; Pan, Y.; Deng, D.; Liu, Z. Characterization of Cement Asphalt Mortar for Slab Track by Dynamic Mechanical Thermoanalysis. J. Mater. Civ. Eng. 2016, 28, 04015154. [Google Scholar] [CrossRef]
  11. You, Q.; Wei, B.; Ma, J.; Sesay, T.; Wan, H. Interpreting the creep behavior of asphalt mortar at high temperature through experimental and numerical methods. Constr. Build. Mater. 2020, 258, 120317. [Google Scholar] [CrossRef]
  12. Xu, Y.; Xu, Z.; Guo, Y.; Jia, H.; Huang, X.; Wen, Y. Mathematical modeling and test verification of viscoelastic materials considering microstructures and ambient temperature influence. Mech. Adv. Mater. Struct. 2022, 29, 7063–7074. [Google Scholar] [CrossRef]
  13. Zhao, L.; Dai, X. Dynamic Response Analysis of Slab Ballast Track Considering Viscoelastic Characteristics of CA Mortar. J. Railw. Sci. Eng. 2018, 15, 8. (In Chinese) [Google Scholar]
  14. Liu, D.; Liu, Y.; Ren, J.; Yang, R.; Liu, X. Contact Loss beneath Track Slab Caused by Deteriorated Cement Emulsified Asphalt Mortar: Dynamic Characteristics of Vehicle-Slab Track System and Prototype Experiment. Math. Probl. Eng. 2016, 2016, 3073784. [Google Scholar] [CrossRef]
  15. Ren, J.; Li, H.; Cai, X.; Deng, S.; Wang, J.; Du, W. Viscoelastic deformation behavior of cement and emulsified asphalt mortar in China railway track system I prefabricated slab track. J. Zhejiang Univ.-Sci. A 2020, 21, 304–316. [Google Scholar] [CrossRef]
  16. Deng, S.; Ren, J.; Wei, K.; Ye, W.; Du, W.; Zhang, K. Fatigue damage evolution analysis of the CA mortar of ballastless tracks via damage mechanics-finite element full-couple method. Constr. Build. Mater. 2021, 295, 123679. [Google Scholar] [CrossRef]
  17. Chen, H.; Li, W.; Jiang, Y.; Xiao, L. Fatigue life prediction for CA mortar in CRTS II railway slab track subjected to combined thermal action and vehicle load by mesoscale numerical modelling. Constr. Build. Mater. 2024, 437, 136987. [Google Scholar] [CrossRef]
  18. Zhai, W. Vehicle–Track Coupled Dynamics Models; Springer: Berlin/Heidelberg, Germany, 2020. [Google Scholar]
  19. Wu, B.; Wen, Z.; Wu, T.; Jin, X. Analysis on thermal effect on high-speed wheel/rail adhesion under interfacial contamination using a three-dimensional model with surface roughness. Wear 2016, 366–367, 95–104. [Google Scholar] [CrossRef]
  20. Wu, B.; Yang, Y.; Xiao, G. A transient three-dimensional wheel-rail adhesion model under wet condition considering starvation and surface roughness. Wear 2024, 540–541, 205263. [Google Scholar] [CrossRef]
  21. Hajikarimi, P.; Nejad, F.M. Applications of Viscoelasticity: Bituminous Materials Characterization and Modeling; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
  22. Zhao, R.; Jing, F.; Wang, R.; Cai, J.; Zhang, J.; Wang, Q.; Xie, H. Influence of oligomer content on viscosity and dynamic mechanical properties of epoxy asphalt binders. Constr. Build. Mater. 2022, 338, 127524. [Google Scholar] [CrossRef]
  23. Cai, S.M.; Chen, Y.M.; Liu, Q.X. Development and validation of fractional constitutive models for viscoelastic-plastic creep in time-dependent materials: Rapid experimental data fitting. Appl. Math. Model. 2024, 132, 645–678. [Google Scholar] [CrossRef]
  24. Su, X.; Chen, Y.; Li, J.; Wu, B. Viscoelastic study of cement and emulsified asphalt mortar under temperature variations based on a novel variable-order fractional Maxwell model. Constr. Build. Mater. 2024, 451, 138828. [Google Scholar] [CrossRef]
  25. Podlubny, I. An introduction to fractional derivatives, fractional differential equations, to methods of their solution and some of their applications. Math. Sci. Eng. 1999, 198, 0924.34008. [Google Scholar]
  26. Su, X.; Yao, D.; Xu, W. A new method for formulating linear viscoelastic models. Int. J. Eng. Sci. 2020, 156, 103375. [Google Scholar] [CrossRef]
  27. Su, X.; Yao, D.; Xu, W. Processing of viscoelastic data via a generalized fractional model. Int. J. Eng. Sci. 2021, 161, 103465. [Google Scholar] [CrossRef]
  28. Wang, L. Study on the Viscoelastic and Thermomechanical Properties of CA Mortar; Dalian Jiaotong University: Dalian, China, 2023. [Google Scholar]
  29. Li, J.H.; Farquharson, C.G.; Hu, X.Y. Three effective inverse Laplace transform algorithms for computing time-domain electromagnetic responses. Geophysics 2016, 81, E113–E128. [Google Scholar] [CrossRef]
  30. Chen, M.; Sun, Y.; Zhu, S.; Zhai, W. Dynamic performance comparison of different types of ballastless tracks using vehicle-track-subgrade coupled dynamics model. Eng. Struct. 2021, 249, 113390. [Google Scholar] [CrossRef]
  31. Li, L.; Nimbalkar, S.; Zhong, R. Finite element model of ballasted railway with infinite boundaries considering effects of moving train loads and Rayleigh waves. Soil Dyn. Earthq. Eng. 2018, 114, 147–153. [Google Scholar] [CrossRef]
  32. Yu, Y.; Zhao, L.; Zhou, C. Vertical Coupling Mechanism between the Sitting Human Body and the Vehicle-track System. J. Mech. Eng. 2022, 58, 290–300. [Google Scholar] [CrossRef]
  33. Nan, H. Analysis of Dynamic Characteristics and Parameter Study of CRTSII Type Slab Ballast Track on Roadbed; Lanzhou Jiaotong University: Lanzhou, China, 2012. [Google Scholar]
  34. Zhai, W.; Wei, K.; Song, X.; Shao, M. Experimental investigation into ground vibrations induced by very high speed trains on a non-ballasted track. Soil Dyn. Earthq. Eng. 2015, 72, 24–36. [Google Scholar] [CrossRef]
  35. Wang, M.Z.; Cai, C.B.; Zhu, S.Y.; Zhai, W.M. Experimental study on dynamic performance of typical nonballasted track systems using a full-scale test rig. Proc. Inst. Mech. Eng. Part F-J. Rail Rapid Transit 2017, 231, 470–481. [Google Scholar] [CrossRef]
  36. Xu, H. CRTS I Type Plate Rail CA Mortar Dynamic Mechanical Properties Test and Theoretical Research; Southwest Jiaotong University: Chengdu, China, 2015. [Google Scholar]
  37. Zhang, S. Beijing-Tianjin Intercity High-Speed Railway System Commissioning Technology; China Railway Press: Beijing, China, 2008. [Google Scholar]
  38. Zhang, X.; Chai, Y.; Liu, J. Experimental study on the effect of temperature on the mechanical properties of cement-asphalt mortar. Bull. Chin. Ceram. Soc. 2015, 34, 538–543. (In Chinese) [Google Scholar] [CrossRef]
Figure 1. CRTS I ballastless track.
Figure 1. CRTS I ballastless track.
Lubricants 13 00058 g001
Figure 2. The FM model.
Figure 2. The FM model.
Lubricants 13 00058 g002
Figure 3. Zener model.
Figure 3. Zener model.
Lubricants 13 00058 g003
Figure 4. Fitting curves of the creep compliance.
Figure 4. Fitting curves of the creep compliance.
Lubricants 13 00058 g004
Figure 5. Finite element model for the compressive creep experiment with CA mortar.
Figure 5. Finite element model for the compressive creep experiment with CA mortar.
Lubricants 13 00058 g005
Figure 6. Simulation results of the creep experiment.
Figure 6. Simulation results of the creep experiment.
Lubricants 13 00058 g006
Figure 7. Vertical vehicle–track–subgrade system model [30].
Figure 7. Vertical vehicle–track–subgrade system model [30].
Lubricants 13 00058 g007
Figure 8. Vertical track irregularity.
Figure 8. Vertical track irregularity.
Lubricants 13 00058 g008
Figure 9. The VTS model.
Figure 9. The VTS model.
Lubricants 13 00058 g009
Figure 10. Track subsystem simulation results. (a) Rail displacement; (b) slab displacement; (c) fastener force; (d) CA mortar stress.
Figure 10. Track subsystem simulation results. (a) Rail displacement; (b) slab displacement; (c) fastener force; (d) CA mortar stress.
Lubricants 13 00058 g010
Figure 11. Comparison of the simulation results with the relevant literature. (a) Slab displacement and rail displacement; (b) fastener force and CA mortar stress.
Figure 11. Comparison of the simulation results with the relevant literature. (a) Slab displacement and rail displacement; (b) fastener force and CA mortar stress.
Lubricants 13 00058 g011
Figure 12. Comparison of time history curves of dynamic responses of the track subsystem. (a) Rail displacement; (b) slab displacement; (c) CA mortar displacement; (d) compressive displacement of CA mortar; (e) CA mortar stress; (f) base displacement.
Figure 12. Comparison of time history curves of dynamic responses of the track subsystem. (a) Rail displacement; (b) slab displacement; (c) CA mortar displacement; (d) compressive displacement of CA mortar; (e) CA mortar stress; (f) base displacement.
Lubricants 13 00058 g012aLubricants 13 00058 g012b
Figure 13. Comparison of the maximum simulation outcomes. (a) Rail and slab displacement; (b) CA mortar and concrete base displacement; (c) compressive displacement of CA mortar; (d) CA mortar stress.
Figure 13. Comparison of the maximum simulation outcomes. (a) Rail and slab displacement; (b) CA mortar and concrete base displacement; (c) compressive displacement of CA mortar; (d) CA mortar stress.
Lubricants 13 00058 g013
Figure 14. Comparison of the maximum dynamic response results of the vehicle subsystem. (a) Vertical acceleration of the vehicle body; (b) wheel–rail force.
Figure 14. Comparison of the maximum dynamic response results of the vehicle subsystem. (a) Vertical acceleration of the vehicle body; (b) wheel–rail force.
Lubricants 13 00058 g014
Figure 15. Growth rate of dynamic responses at different temperatures.
Figure 15. Growth rate of dynamic responses at different temperatures.
Lubricants 13 00058 g015
Figure 16. Comparison of the maximum dynamic response results of the track subsystem. (a) Slab displacement; (b) CA mortar displacement; (c) compression displacement of CA mortar; (d) CA mortar stress.
Figure 16. Comparison of the maximum dynamic response results of the track subsystem. (a) Slab displacement; (b) CA mortar displacement; (c) compression displacement of CA mortar; (d) CA mortar stress.
Lubricants 13 00058 g016
Figure 17. Comparison of the maximum dynamic response results of the track subsystem. (a) Rail displacement; (b) CA mortar displacement; (c) CA mortar acceleration; (d) CA mortar stress.
Figure 17. Comparison of the maximum dynamic response results of the track subsystem. (a) Rail displacement; (b) CA mortar displacement; (c) CA mortar acceleration; (d) CA mortar stress.
Lubricants 13 00058 g017
Figure 18. Position diagram for the CA mortar layer.
Figure 18. Position diagram for the CA mortar layer.
Lubricants 13 00058 g018
Figure 19. Comparison of the maximum dynamic response results of CA mortar. (a) CA mortar displacement; (b) CA mortar stress.
Figure 19. Comparison of the maximum dynamic response results of CA mortar. (a) CA mortar displacement; (b) CA mortar stress.
Lubricants 13 00058 g019
Figure 20. CA mortar layer compressive stress cloud diagram.
Figure 20. CA mortar layer compressive stress cloud diagram.
Lubricants 13 00058 g020
Table 1. Transfer functions of basic mechanical components.
Table 1. Transfer functions of basic mechanical components.
ComponentTransfer Function T(s)
Hooke’s springE
Newtonian dashpotλs
Abel dashpotηsα
where λ is the viscosity coefficient of the Newtonian dashpot, and s is the complex number in the Laplace domain.
Table 2. Fitting parameters of the creep compliance.
Table 2. Fitting parameters of the creep compliance.
T/°CFM ModelZener Model
E/Mpaη/Mpa·sααR2E1/MpaE2/Mpaλ/Mpa·sR2
−2046.477268.7080.2530.97734.19145.280245.1910.970
027.892477.3580.3200.99523.45342.088399.2710.977
2023.132146.7510.2190.99217.20136.594264.6570.962
4010.984140.7540.1900.95710.07741.962103.7630.907
Table 3. Fitting parameters of the Prony series.
Table 3. Fitting parameters of the Prony series.
No.τi/sgi
T/°C
−2002040
11 × 10−26.160 × 10−21.662 × 10−26.395 × 10−24.072 × 10−2
21 × 10−14.530 × 10−21.721 × 10−23.912 × 10−21.942 × 10−2
31 × 1007.066 × 10−23.402 × 10−25.758 × 10−22.814 × 10−2
41 × 1011.060 × 10−16.541 × 10−28.264 × 10−24.082 × 10−2
51 × 1021.381 × 10−11.184 × 10−11.078 × 10−15.697 × 10−2
61 × 1031.600 × 10−11.800 × 10−11.314 × 10−17.752 × 10−2
71 × 1041.210 × 10−11.991 × 10−11.223 × 10−19.925 × 10−2
G0/MPa46.56327.93123.15611.036
Table 4. Symbols in vehicle subsystem equations and their meanings.
Table 4. Symbols in vehicle subsystem equations and their meanings.
SymbolComponentMeaning
Mv, JvVehicleMass and pitch rotational inertia
Zv, βvVertical displacement and pitch nodal displacement
lvHalf of the length
Mb, JbBogieMass and pitch rotational inertia
Zb, βbVertical displacement and pitch nodal displacement
lbHalf of the length
MwWheelsetMass
Kvb, CvbPrimary suspensionStiffness and damping
Kbw, CbwSecondary suspensionStiffness and damping
Frw,kThe k-th wheelsetWheel–rail force
Table 5. USA vertical track irregularity parameters.
Table 5. USA vertical track irregularity parameters.
Track GradeAv/(cm2·m/rad)ϕc/(rad/m)
11.21070.8245
21.01810.8245
30.68160.8245
40.53760.8245
50.20590.8245
60.03390.8245
Table 6. Vehicle and fastener parameters.
Table 6. Vehicle and fastener parameters.
ParameterValue
Vehicle body mass/t42.400
Bogie mass/t3.400
Wheelset mass/t2.200
Nodding moment of inertia of the vehicle body/(kg·m2)2.74 × 106
Nodding moment of inertia of the bogie/(kg·m2)7.2 × 103
Primary suspension stiffness/(N/m)1.04 × 107
Primary suspension damping/(N·s/m)5 × 104
Secondary suspension stiffness/(N/m)4 × 105
Secondary suspension damping/(N·s/m)5 × 104
Fastener stiffness/(N/m)5 × 107
Table 7. Track and subgrade parameters.
Table 7. Track and subgrade parameters.
StructuralDensity/(kg/m3)Elastic Modulus/MPaPoisson’s Ratio
Rail78502.06 × 1050.3
Slab25003.65 × 1040.2
Lug boss24403.4 × 1040.2
Base25003.6 × 1040.2
Subgrade surface22502.5 × 1020.27
Subgrade bottom21301.6 × 1020.32
Subgrade body20001.2 × 1020.35
CA mortar1800——0.45
Table 8. Comparison of the simulation results with relevant literatures.
Table 8. Comparison of the simulation results with relevant literatures.
ParameterResult in This PaperResults in the Literature [36]Measured Value [36]Measured Value [37]
Rail dis/mm1.3921.5370.3–0.880.66–1.32
Slab dis/mm0.3310.3670.27–0.390.16–0.37
CA mortar stress/KPa46.583125.50125.63–38.48≤45
Fastener force/kN52.41360.45014.4–65.824.4–46.3
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, Y.; Wu, B.; Yao, L.; Su, X. Dynamic Response Analysis of Ballastless Tracks Considering the Temperature-Dependent Viscoelasticity of Cement-Emulsified Asphalt Mortar Based on a Vehicle–Track–Subgrade Coupled Model. Lubricants 2025, 13, 58. https://doi.org/10.3390/lubricants13020058

AMA Style

Chen Y, Wu B, Yao L, Su X. Dynamic Response Analysis of Ballastless Tracks Considering the Temperature-Dependent Viscoelasticity of Cement-Emulsified Asphalt Mortar Based on a Vehicle–Track–Subgrade Coupled Model. Lubricants. 2025; 13(2):58. https://doi.org/10.3390/lubricants13020058

Chicago/Turabian Style

Chen, Yunqing, Bing Wu, Linquan Yao, and Xianglong Su. 2025. "Dynamic Response Analysis of Ballastless Tracks Considering the Temperature-Dependent Viscoelasticity of Cement-Emulsified Asphalt Mortar Based on a Vehicle–Track–Subgrade Coupled Model" Lubricants 13, no. 2: 58. https://doi.org/10.3390/lubricants13020058

APA Style

Chen, Y., Wu, B., Yao, L., & Su, X. (2025). Dynamic Response Analysis of Ballastless Tracks Considering the Temperature-Dependent Viscoelasticity of Cement-Emulsified Asphalt Mortar Based on a Vehicle–Track–Subgrade Coupled Model. Lubricants, 13(2), 58. https://doi.org/10.3390/lubricants13020058

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop