Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Transportation Geotechnics 17 (2018) 178–191

Contents lists available at ScienceDirect

Transportation Geotechnics
journal homepage: www.elsevier.com/locate/trgeo

The stiffening of soft soils on railway lines T


a b,⁎ a b c
K. Dong , D.P. Connolly , O. Laghrouche , P.K. Woodward , P. Alves Costa
a
Heriot Watt University, EH14 4AS, UK
b
University of Leeds, LS2 9JT, UK
c
University of Porto, 4099-002 Porto, Portugal

ARTICLE INFO ABSTRACT

Keywords: Railway tracks experience elevated rail deflections when the supporting soil is soft and/or the train speed is
Railway ground improvement greater than approximately 50% of the wave propagation velocity in the track-soil system (i.e. the critical ve-
Railroad track stiffness locity). Such vibrations are undesirable, so soil replacement or soil improvement of the natural soil (or alter-
Railway critical velocity natively mini-piles or lime-cement treatment) is often used to increase track-ground stiffness prior to line
Rail-track dynamics thin-layer method (TLM)
construction. Although areas of existing soft subgrade might be easily identified on a potential new rail route, it
Subgrade soil replacement
Stiffening soft soil
is challenging to determine the type and depth of ground remediation required. Therefore, major cost savings
can be made by optimising ground replacement/improvement strategies.
This paper presents a numerical railway model, designed for the dynamic analysis of track-ground vibrations
induced by high speed rail lines. The model simulates the ground using a thin-layer finite element formulation
capable of calculating 3D stresses and strains within the soil during train vehicle passage. The railroad track is
modelled using a multi-layered formulation which permits wave propagation in the longitudinal direction, and is
coupled with the soil model in the frequency-wavenumber domain. The model is validated using a combination
of experimental railway field data, published numerical data and a commercial finite element package. It is
shown to predict track and ground behaviour accurately for a range of train speeds.
The railway simulation model is computationally efficient and able to quickly assess dynamic, multi-layered
soil response in the presence of ballast and slab track structures. Therefore it is well-suited to analysing the effect
of different soil replacement strategies on dynamic track behaviour, which is particularly important when close
to critical speed. To show this, three soil-embankment examples are used to compare the effect of different
combinations of stiffness improvement (stiffness magnitude and remediation depths up to 5 m) on track beha-
viour. It is found that improvement strategies must be carefully chosen depending upon the track type and
existing subgrade layering configuration. Under certain circumstances, soil improvement can have a negligible
effect, or possibly even result in elevated track vibration, which may increase long-term settlement. However,
large benefits are possible, and if detailed analysis is performed, it is possible to minimise soil improvement
depth with respect to construction cost.

Introduction degradation. Therefore it is important to have a methodology capable


of assessing remediation measures for both loading cases, ensuring
Background track deflections meet desired tolerances.
The relationship between increases in track deflection and train
The global railway network is undergoing rapid expansion, with speed is well documented [51] and the speed at which maximum track
train vehicles becoming faster, heavier, and having less headway be- displacement occurs due to a moving load is known as the ‘critical
tween successive passages. High axle freight loads are problematic on velocity’. Although critical velocity effects (i.e. very high rail dis-
soft soil sites because they cause excessive track deflection. Similarly, placements) have been observed on numerous lines around the world,
faster trains are problematic because when the train speed exceeds the most well documented case was at Ledsgard in Sweden [37,32]. At
approximately 50% of the natural ground wave speed, increases in this site there were low stiffness soil layers resulting in a low site critical
track deflection occur. If deflections are high, it is a safety concern and velocity. Therefore high track deflections were observed, particularly as
results in more frequent maintenance due to accelerated track train speed increased.


Corresponding author.
E-mail address: d.connolly@leeds.ac.uk (D.P. Connolly).

https://doi.org/10.1016/j.trgeo.2018.09.004
Received 6 June 2018; Received in revised form 24 August 2018; Accepted 10 September 2018
Available online 15 September 2018
2214-3912/ © 2018 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/BY/4.0/).
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Ensuring soil stiffness’ before the construction of a new line is periodic modes computed and the solution transferred to the space
constructed is challenging because soils are typically multi-layered and domain via the Floquet transform.
the dynamic behaviour of the track-soil system depends on multiple Alternatively, to further reduce computational requirements, 2.5D
variables such as train speed. Therefore, on a new line there are 2 models can be used approximate a 3D domain, and only require a si-
challenging scenarios which may be encountered after a critical velo- milar number of computations as a 2D problem. Yang et al. [50] pro-
city scoping study: posed this type of finite element solution, with an infinite element
absorbing boundary solution. Alternatively [3] used an equivalent
1. A low stiffness soil configuration which will likely require re- linear 2.5D FEM-IEM approach to study the effects of soil non-linearity
mediation (e.g. train speed > 70% of critical velocity) on the dynamic response of high-speed lines. Bian et al. [12,11] also
2. A semi low stiffness soil configuration for which the need for re- combined a 2.5D finite element model with one-quarter car model to
mediation is unclear (e.g. train speed between 50 and 70% of critical study the vibrations caused by vertical track irregularities. Alter-
velocity) natively, François et al. [21] and Galvín et al. [23] presented efficient
approaches for coupling both 2.5D FEM and 2.5D BEM, where the ir-
For both of these cases it is desirable to know how ground re- regular domain was modelled using 2.5D FEM and the ground modelled
mediation will affect track behaviour. Two types of remediation often using 2.5D BEM. Alves Costa et al. [5] also proposed a numerical
used on new lines are soil replacement and soil improvement [47]. scheme, based on the coupling between 2.5D FEM-BEM, to obtain the
These are potentially expensive and thus to minimise cost, analysis is track-ground response. The result was found to agree well with field
required to determine the optimal depth and stiffness of the remediated data from a test site in Portugal.
ground. When the system is considered invariant in the horizontal direction,
The first step in analysing the potential for critical velocity effects to the dynamic response of the ground can be simulated using a combi-
occur is to determine the site critical speed. For a homogenous soil this nation of analytical equations, combined with finite elements for the
is straightforward, however in reality soils are multi-layered. Therefore non-invariant direction (i.e., the vertical direction). This approach is
an approach for solely calculating the critical velocity, rather than the known as the Thin-Layer Method (TLM), where the medium is dis-
track response was proposed by Sheng et al. [43], Alves Costa et al. [2], cretised into thin layers in just a single (vertical) direction. These ele-
Bian et al. [10], Mezher et al. [38]. These approaches relied on the ments are similar to traditional 1D finite elements, however the hor-
analysis of the wave dispersion relationships in the track-ground system izontal directions are solved analytically, thus allowing the solution to
and found to give fast and reliable estimates of the critical velocity for be computed in a fraction of the time required for a full 3D model. The
layered soils. development of the generalised TLM is described by Kausel [28], Kausel
In an attempt to simulate track dynamics at critical velocity, early and Roesset [31], Kausel [30,29]. It has been used for railway problems
researchers such as Lamb [36], Fryba [22] used analytical expressions by Bian and Chen [13] to obtain the semi-analytical solution for three-
to model the problem as a moving point load placed on an elastic half- dimensional layered ground responses due to harmonic moving loads.
space. Later, Krylov [35] and Dieterman and Metrikine [19] used a Building on this work, Alves Costa et al. [3], Alves Costa [4], Bian et al.
moving load to represent a train wheel, analytical expressions for the [10] also combined thin-layered elements with a 2.5D finite element
track structure and Green’s functions for the soil. method to predict train induced vibrations.
Expanding upon this to simulate more complex layered soil cases, This work builds upon the work of these previous researchers and
Kaynia et al. [32] proposed a semi-analytical model for the prediction combines a TLM ground model with analytical modelling of the track
of ground response. An embankment was modelled using an infinite structure. This results in a method that requires low computational
viscoelastic beam and Green’s function used to calculate the ground effort to calculate track and ground responses due to high speed trains.
stiffness matrix. Further, Sheng et al. [43] and Thompson [48] proposed The model is validated numerically and experimentally, and then used
a semi-analytical approach in which the track was modelled analyti- to investigate the effect of soil replacement/improvement below
cally in the frequency domain and the soil was modelled using an in- railway lines.
tegral transformation method. A similar approach was also followed by
Alves Costa et al. [1] and Alves Costa [4]. The model was found to have Model description
agreement with field data collected at Ledsgard, Sweden.
Although analytical approaches are computationally efficient thus The model is a semi-analytical approach to compute vertical track
allowing for sensitivity analyses, they typically rely on a variety of displacements and three dimensional soil stresses/strains. It comprises
assumptions thus making them valid for only a subset of real-world of two main components: (1) an analytical track model, and (2) a semi-
situations [26]. Therefore, to analyse the dynamics of more complex analytical thin-layer method ground model. Both sub-models are solved
track structures, three dimensional (3D) models have been proposed. in the frequency domain and coupling is performed in the wavenumber-
These include Hall [25] who developed a 3D finite element track-soil frequency domain, using relaxed boundary conditions between the
model using ABAQUS, which was subject to moving point loads. track and soil. This means that only the vertical components are used
Building upon this, Connolly et al. [17,16], Shih et al. [44,45], Powrie for coupling because the 1D track model and 3D soil model have a
et al. [42] and Varandas et al. [49] also suggested using 3D models. different number of degrees of freedom [32,39]. The advantage of this
Further, El Kacimi et al. [27] proposed a Fortran based 3D FE code in approach is that the deep-wave propagation generated near critical
the time domain, while Galvín and Domínguez [24] proposed a fre- velocity can be simulated both accurately and efficiently.
quency domain solution. Kouroussis et al. [33] also suggested using
sub-modelling, where a multibody vehicle model was considered and Track model
combined with a 3D ABAQUS ground model surrounded by infinite
elements. Auersch [8] and O’Brien & Rizos [40] coupled a multi-body The track model is modelled in the wavenumber-frequency domain
vehicle system with a 3D FEM-BEM model to simulate the vibrations as previously described in Alves Costa [4], Alves Costa et al. [2],
induced by high speed rails. Mezher et al. [38]. Different formulations are used to describe the
A challenge however with fully 3D numerical modelling is that it is ballasted and concrete slab tracks respectively, both of which assume
computationally demanding. Alternatively, to reduce analysis run the track to be infinite in the direction of train passage.
times, researchers have proposed using the periodic nature of railway Fig. 1 illustrates the ballasted track model. It is composed of rail,
tracks to reduce the number of degrees of freedom [14,15,20,7,6]. To railpad, sleeper and ballast components, with the properties of discrete
do so, the track-soil can be divided into thin 3D finite element slices, the elements (e.g. sleepers and railpads) distributed along its length. Eq. (1)

179
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

2
keq (k x , ) = + G
uzz (k x , k y , z = 0, ) Ctg dk y (3)
where Ctg is a scaling factor for the track-ground coupling. For the case
of a ballasted track, coupling is achieved through compatibility of
displacements at the track centre point on the track-ground boundary:
sin(k y b)
Ctg =
ky b (4)

sin(k y b)2
Ctg =
Fig. 1. Ballasted track model structure.
(k y b)2 (5)
Whereas for the slab track, coupling is achieved through compat-
describes the formulation which permits 1D wave propagation, where: ibility of the average displacements across the track-ground boundary:
k x is the Fourier images of coordinates x ; mr and ms are the mass of rail For both track cases, relaxed boundary conditions are assumed, i.e.,
per metre and the equivalent distributed mass of the sleepers respec- the compatibility and equilibrium conditions are only respected in the
tively; kp is the complex stiffness of the railpad, i.e, including damping vertical direction. This is considered valid because vertical excitation is
(kp = kp (1 + i cp) ), where indicates a complex number; Cp is the dominant during train passage.
ballast compressional wave speed; Eb is the ballasts Young’s modulus;
hb is the ballast height; 2b is the track width; ur , us and ubb are the rail, Soil model
sleeper, and ballast (bottom) displacements respectively ( represents
frequency domain response); P is the vertical force acting on the rail. Thin-layer method formulation
The thin-layer method (TLM) is a semi-analytical approach com-
EIr k x4 + kp 2m
r kp 0 monly used to solve wave propagation problems in 3D unbounded
2 Eb b
2m 2 Eb b domains (e.g. soils). It relies on discretizing the vertical plane into a
kp kp +
tan
hb
Cp
s
sin
hb
Cp
finite number of layers, each thin with respect to wavelength, meaning
Cp Cp
vertical displacements can be assumed to vary linearly across each
0
2 Eb b 2 Eb b
+ keq layer. Regarding the horizontal directions, the analytical wave equation
sin
hb
Cp tan
hb
Cp
Cp is used to rapidly calculate harmonic displacements without suffering
Cp
from the element aspect-ratio restriction of the finite element method.
u r (k x , ) P (k x , ) The thin layer formulation is undertaken in the frequency-wave-
us (k x , ) = 0 number domain, with the response computed for each individual fre-
ubb (k x , ) 0 (1) quency under steady-state harmonic conditions. It allows the soil to be
discretized along the z-direction with the infinite and homogeneous
Fig. 2 illustrates the layout of the slab track model, which unlike the property along the horizontal plane, as shown in Fig. 3, in which n
ballasted track, considers the bending stiffness of the concrete slab. The represents the total number of vertically discretised soil layers. To en-
governing equations for the slab track are shown in Eq. (2), where: EIsl sure the propagation of stresses/strains inside the soil domain is mod-
is the bending stiffness of the slab; msl is the mass of the slab per meter elled correctly, three-node quadratic thin-layer elements are used
and usl is the vertical slab displacement in the frequency domain. (Fig. 4). Also, to ensure the thickness of each layer is small with respect
to the wavelength, all thin layers meet the criteria outlined in Eq. (6),
EIr k x4 + kp 2m
r kp u r (k x , ) where kmax is the maximum wavenumber under consideration.
kp EIsl k x4 + kp 2m
sl + keq usl (k x , )
wavelength 2
h= =
P (k x , ) 8 8k max (6)
=
0 (2) The displacement field inside each thin layer is approximated by
For both track types, track-soil coupling is achieved using a complex means of the interpolation function (shape function), i.e., u = NU ,
equivalent stiffness (keq ) defined in the wavenumber-frequency domain, where U is a vector containing the nodal displacements. Also, N = N( )
as shown in Eq. (3). Higher train speeds typically result in greater wave is the shape function matrix of the form, N = [N1 IN2 IN3 I] where Na is
propagation across the track-soil interface. Therefore the accuracy of the shape function and I the 3 × 3 identity matrix and represents the
coupling across this interface is vital for reliable simulation. In this local coordinates. For a three noded thin-layer element, the shape
equation, uzz
G
is the Green’s function describing the vertical displace- functions are:
ment of the surface of the ground, while k x and k y are the Fourier
images of coordinates x and y respectively. For both tracks, coupling is
achieved by taking into account the equilibrium of loads and the
compatibility of displacements along the track-ground interface [46]:

Fig. 3. Schematic drawing of the procedure of the vertical discretization and


Fig. 2. Slab track model structure. domain transform in the thin-layer method.

180
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Fig. 4. Discretization into thin layers with quadratic interpolation.

1 2 1 h h
N1 ( ) =
2 2 (7) [K5] = [B2]T [D][B3 ]|J | d + [B3 ]T [D][B2 ]|J | d
0 0 (18)
N2 ( ) = 1 2
(8)
Also,|J | is the determinant of the Jacobian matrix and D is the
1 2 1 constitutive matrix, in which elastic and isotropic material properties
N3 ( ) = +
2 2 (9) are used in this paper [11].
This classic thin-layer formulation gives the response of a medium
Next, strains and stresses inside each layer are approximated as
with a fixed rigid base, however an infinitely deep soil is required for
{ } = [B]{U} and { } = [D]{ } = [D][B]{U} , where [B] is the strain ma-
the majority of critical velocity problems. To overcome this, a variety of
trix and [B] = [ B1 B 2 B3 ] and
absorbing boundary formulations have been proposed [30,9]. In this
ik x Na 0 0 work, Thompson-Haskell matrices are formed in a similar manner to the
0 ik y Na 0 previous derivation and appended to those above. The size of each
Na matrix is 3(n + 1) × 3(n + 1) to account for the displacements in ver-
0 0
z tical, horizontal and longitudinal directions at the half-space interface.
[Ba] = ik N ik N 0
y a x a

0 Na
ik y Na Equivalent stiffness formulation
z
Na After the equivalent stiffness is computed in the frequency-wave-
0 ik x Na
z (10) number domain using Eq. (3), the system of equations in Eqs. (1) and
(2) are then solved to obtain the displacements of the track compo-
where k x and k y are the Fourier images of the x and y directions re-
nents.
spectively.
Then, scaled Green’s function R (k x , k y , ) can be represented as:
Then, rewriting Ba as: [Ba]=[Ba]1 + ik x [Ba]2 + ik y [Ba]3
R (k x , k y , ) = (L (k x , ) Ctg ) g (k
x, k y, ) (19)
0 0 0
0 0 0 Na 0 0 0 0 0 where g (k
x , k y , ) is the displacement/strain/stress Green’s function;
0 0 Na 0 0 0 0 Na 0 L (k x , ) Ctg represents the load function, in which L (k x , ) is computed
z
0 0 0 0 0 0 by multiplication of the equivalent stiffness and lower track displace-
[Ba ]1 = 0 0 0 [Ba ]2 = [Ba ]3 =
0 Na 0 Na 0 0
0 Na
0 ment. Also, Ctg is given in Eqs. (4) and (5).
0 0 0 0 0 Na
z Inverting the scaled Green’s function in the transversal and long-
Na 0 0 Na 0 0 0
z
0 0 itudinal directions using the inverse Fourier transform gives the ground
responses in the time-space domain:
(11)
The wave equation for the thin-layer system is simplified and ex- 1
R (x , y , t ) = R (k x , k y , ) e i ( t kx x ky y) dk
x dk y d
pressed in the frequency-wavenumber domain as ([K] 2 [M]) U = P , (2 )3 (20)
where U and P are displacements and tractions, respectively, and M is
Finally, it is worth noting that because the TLM model is computed
the corresponding mass matrix. K is the global stiffness matrix:
in the wavenumber-frequency domain, the moving load effect is taken
[K] = [K 0] + ik x [K1] + ik y [K2] + k x2 [K3] + k y2 [K 4] + k x k y [K5] (12) into account using the shift property of the Fourier transform that al-
lows frequency to be related to wavenumber, i.e., = k x c , where
where: is the excitation frequency (set to 0 Hz in this work), k x is the wa-
h venumber in the direction of moving load, and c is the moving load
[K 0] = [B1]T [D][B1]|J| d velocity.
0 (13)
h h Model validation
[K1] = [B1]T [D][B2 ]|J| d [B 2 ]T [D][B1]|J| d
0 0 (14) The model is validated against three test cases which are a combi-
nation of numerical and experimental results. Firstly, its ability to
h h
model soil improvement below a railway track is verified. Next its
[K2] = [B1]T [D][B3 ]|J | d [B3 ]T [D][B1]|J| d
0 0 (15) output is compared against numerical results from published literature,
and finally it is validated against experimental results from a railway
h line in Portugal [5,18].
[K3] = [B2]T [D][B2 ]|J | d
0 (16)
Soil replacement validation
h
[K 4] = [B3]T [D][B3]|J | d One advantage of using the TLM for soil modelling is that it is
0 (17) computationally efficient compared to 3D FE analysis. However, one

181
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Fig. 5. Left: Layout of soil replacement geometry, Right: Cut through ABAQUSs model showing track and soil response during receptance test.

Fig. 6. Comparison between simplified and detailed soil replacement geometry (Left: static rail receptance excitation, Right: moving load excitation time history).

limitation compared to 3D approaches is that all horizontal layers are


infinitely long/wide. This is acceptable for many problems because the
soil layers below and near the track can be assumed to be horizontal
and infinite. However, when considering soil replacement/improve-
ment, to minimise cost, remediation is typically only performed directly
below the track. The TLM cannot capture this localised improvement
and instead assumes the improvement is infinite rather than just below
the track.
To determine the effect of this assumption on the ability of the
proposed model to simulate soil replacement, 3D FE tests are performed
using ABAQUS. The 3D track consists of rail, railpad, slab and sub-
ballast components. The track is fully coupled to a 3D soil model with x,
y, z dimensions: 50 * 32 * 12 m; which utilises infinite element ab-
sorbing boundaries in the free-field to reduce artificial reflections. Two
soils, with constant density, 2000 kg/m3, are considered and shown in
Fig. 7. Schematic representation of track-ground interaction model.
Fig. 5:

Table 1
Soil properties used in the simulation for train velocity of 70 km/h and 200 km/h.
Soil layer Layer thickness (m) Density (kg/m3) Young’s modulus (MPa) Poisson ratio Damping

c = 70 c ≥ 185 c = 70 c ≥ 185 c = 70 c ≥ 185

Crust 1.1 1500 23 19 0.49 0.49 0.04 0.063


Organic clay 3 1260 6 4 0.5 0.5 0.02 0.058
Clay 1 4.5 1475 19 16 0.5 0.5 0.05 0.098
Clay 2 6 1475 33 32 0.5 0.5 0.05 0.064
Half-space – 1475 44 44 0.5 0.5 0.05 0.06

182
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Fig. 8. Train loads applied in the numerical simulation.

Table 2 compares the track displacement time history for a train running close
Velocity and direction of the compared simulation cases. to the critical velocity (550 km/h) for both cases. It is seen that again
Case 1 Case 2 Case 3 the response is similar, with a small discrepancy in the peak value.
Therefore, considering the validation is for the most challenging test
Velocity (km/h) 70 km/h 185 km/h 200 km/h case (i.e. case of a vertical replacement barrier) it can be concluded
Direction Southbound Northbound Southbound
that the proposed model is capable of accurately modelling the effects
of soil replacement.
1. In real-life, when considering soil replacement, it is likely that due
to the practicalities of construction the sides of the replacement zone Numerical validation
would be sloped (dashed grey lines in Fig. 5). However, for vali-
dation the improved area is considered to have vertical sides (solid A numerical comparison is made against an alternative, peer-re-
grey lines in Fig. 5). This is the worst-case scenario and thus most viewed and validated track-ground railway model [32]. Like all models
likely to challenge the limitations of the model. The Young’s mod- though, the peer-reviewed approach was originally found to have dis-
ulus of the improved soil is 150 MPa, while the Young’s modulus of crepancies between its results and field data [32]. Therefore, it is in-
the non-remediated soil is set to 45 MPa. Considering that soil im- tended for a qualitative comparison.
provement can be undertaken to an arbitrary depth, and that it is the Fig. 7 illustrates the modelling scenario, for which a combined
vertical improvement edges likely to introduce errors, the soil im- track/embankment structure rests on a multi-layered soil. Also, the soil
provement is considered for the full model depth. is composed of 5 layers, with the lowest layer being infinitely deep, as
2. A fully homogenous soil with properties: 150 MPa. This represents described in Table 1. However, it should be noted that to account for
the TLM approximation of post-soil improvement where the im- non-linear soil behaviour (i.e. stiffness degradation with strain), the soil
provement was performed to the infinite horizontal layers properties in the previous [32] study are manually changed to account
for this. Therefore, to ensure consistency between validations, the same
To compare the results generated by both models, track receptance ‘degraded’ properties are used for comparison (i.e. the track/embank-
is considered. Fig. 6 (left) shows the results for all frequencies are ment has a bending rigidity EI of 80 MN/m2 when the train speed is
similar. A small discrepancy is seen at low frequency which is gov- either 185 km/h or 200 km/h, and 200 MN/m2 when the train speed is
erned by the static stiffness of the overall track-ground model, how- 70 km/h). It is worth noting that the bending stiffness is updated by
ever the difference is negligible. This is because wave energy first changing the Young’s modulus E , instead of changing the dimensions of
propagates from the track into the soil directly below the track (i.e. the embankment. Train loading is chosen to replicate an X2000 train
where remediation is located). Then, when it spreads from this region, running in two directions (Northbound and Southbound), at speeds of
for the case of the remediated soil it moves through a stiff to soft soil 70, 185 and 200 km/h (Fig. 8) (see Table 2).
interface thus not generating significant reflection. Similarly, in the Fig. 9 shows the rail displacement time history comparison between
homogenous TLM assumption, it also does not experience a reflection, the published result and the solution calculated using the proposed
thus making the approximations similar. Further, Fig. 6 (right) model. In this example, the velocity is lower (70 km/h) than the critical

Fig. 9. Track displacement for Southbound train at 70 km/h (case 1) in time domain (Left) and frequency domain (Right).

183
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Fig. 10. Track displacement for Northbound train at 185 km/h (case 2) in time domain (Left) and frequency domain (Right).

Fig. 11. Track displacement for Southbound train at 200 km/h (case 3) in time domain (Left) and frequency domain (Right).

speed, meaning the response is relatively quasi-static. This is re- frequencies is partly due to having to approximate the published time
presentative of a soft-soil site, where although significant wave pro- histories because raw time history data was unavailable.
pagation is not occurring (i.e. the train speed is not close to the critical Finally, Fig. 11 shows the comparison between the proposed model
velocity), track displacements are still high. It is seen that the new and the published results at a speed of 200 km/h. Again the speed is
model is a strong match to the published data. close to the critical velocity and the dynamic effects are dominant.
Fig. 10 shows the second validation, at the higher speed of 185 km/ However, compared to the speed of 185 km/h, the displacements as-
h in the Northbound direction. On this occasion the train speed is closer sociated with the leading and trailing wheels have reduced, while the
to the critical velocity meaning dynamic wave propagation is sig- response due to the central wheels has increased. Again, the model
nificantly more prominent than at 70 km/h. This results in amplified predicted this accurately. Therefore is it concluded that the model is
displacements, in both the upward and downward directions, particu- accurate and any discrepancies between models are as likely to be from
larly for the leading and trailing wheels. Again the model is able to the published result as from the proposed model. Also, it should be
accurately predict track response. The discrepancy at higher stated that the track-ground coupling in the model presented by Kaynia

Fig. 12. Geometry and mechanical properties of the Alfa-Pendular train.

184
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Fig. 13. Rail displacement due to the train passage in both time content (Left) and in frequency content (Right).

et al. [32] is slightly different from the presented approach, since hy- frequencies and their harmonics are accurately simulated. Any small
brid time-frequency domain is adopted in the former. discrepancies at high frequencies are likely to have been due to wheel-
rail irregularities, or to any other dynamic excitation mechanisms,
Experimental validation which were ignored during prediction, since only quasi-static excitation
is taken into account.
In addition to numerical validation, an experimental validation is Fig. 14 (Left) shows the sleeper velocity during train passage and
performed against data from a railway line in western Portugal, near Fig. 14 (Right) shows the corresponding frequency content. Again the
the town of Carregado. At this site, a comprehensive series of tests were agreement is strong for both cases.
used to assess the soil characteristics, track properties and the system
response due to the passage of Alfa-Pendular trains [5]. Soil replacement strategies
The train is composed of 6 vehicles, with the geometry and axle
loads shown in Fig. 12. Regarding the geometric settings of the model, On railway lines, areas of low stiffness ground can be remediated
there are 10 soil layers supported by bedrock, and ballast and sub- using a variety of techniques, including soil replacement and soil im-
ballast lay on top of the uppermost soil layer. The height of the ballast is provement. These techniques are expensive, particularly as depth of
0.57 m, the Young’s modulus 97 MPa, the Poisson ratio 0.12 and the remediation increases, meaning it is desirable to optimise/minimise
density 1590 kg/m3. The same parameters but for the sub-ballast are improvement depth. This is challenging to determine because at high
0.55 m, 212 MPa, 0.2 and 1910 kg/m3, respectively. One difference train speeds, the low frequency wave energy generated penetrates deep
between the track described in Eq. (1) and that at Carregado, is the into the underlying soil. This can be solved using the model because its
presence of a sub-ballast layer below the ballast. To account for this thin-layer formulation makes it well suited for analysing the effect of
change, the analytical track equation is modified as shown in Eq. (21), soil replacement and soil improvement.
where Esb , Cpsb , hsb , usb are the Young’s modulus, compression wave To show the capabilities of the model, three soil test cases are un-
speed, thickness and displacement of the sub-ballast, respectively. dertaken to determine the effect of improvement depth and stiffness for
both ballasted and slab tracks. They are chosen to represent scenarios
EIk x4 + kp 2
mr kp 0 0 that could be encountered at soft soil sites and have the properties
outlined in Table 3:
2 Eb b 2 Eb b
kp kp + 0
tg hb Cpb sin hb Cpb
Cpb Cpb

2m 1. A homogenous low stiffness soil


2. A low stiffness soil overlying a stiffer soil
s
2 Eb b 2 Eb b 2 Esb b
0 3. A low stiffness soil sandwiched below a stiff embankment and above
sin hb Cpb tg hb Cpb sin hsb Cpsb
Cpb Cpb Cpsb a stiffer underlying soil
2 Esb b
+
tg hsb Cpsb Regarding soil improvement, for each of the three cases, the soft soil
Cpsb
is improved to a maximum depth of 5 m, considering the effect of five
2 Esb b 2 Esb b
0 0 + keq cumulative 1 m thick improvements. All improvements are considered
as a homogeneous horizontal layer and the track displacement is ana-
sin hsb Cpsb tg hsb Cpsb
Cpsb Cpsb

ur (k x , )
lysed for each. Further, for case 1, three different stiffness’ of im-
P (k x , ) provement are considered. The train loading is a single 18 tonne axle
us (k x , ) 0
×
ub(k x , )
= load moving at 330 km/h, which is considered as the maximum op-
0
0 erational speed for new high speed rail lines in UK, for all three study
usb(k x , ) (21)
cases. The ballasted and slab track properties are shown in Table 4.
Fig. 13 (Left) compares the simulated rail displacement and the field
measurement in the time domain. Further, Fig. 13 (Right) shows the Soil case 1
corresponding frequency content comparison. It is seen that there is a
strong match between the predicted and measured values. Regarding Soil test case 1 consists of a homogenous, infinitely deep soil with a
the time history, the correlation is strong for both the magnitude and stiffness of 45 MPa. Five different depths of soil improvement from the
timing of response. Regarding the frequency content, the key dominant soil–track interface are considered (1, 2, 3, 4, 5 m), and for each, three

185
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Fig. 14. Vertical sleeper velocity due to the train passage in both time content (Left) and in frequency content (Right).

in Table 5. As expected, all displacements for the ballasted track are


Table 3
Soil properties. greater than those for the concrete slab. Considering the ballasted track,
all depths of improvement result in reduced displacements, however
Soil Layer Young’s Poisson’s ratio Density Damping
adding additional improvement at large depth below the ground has
thickness (m) modulus (MPa) (kg/m3)
less of an effect than at the surface (e.g. adding 1 m of improvement to
1 [∞] [45] 0.35 1800 0.03 the natural soil provides a greater percentage of reduction compared to
2 [5 ∞] [45 120] 0.35 1800 0.03 adding 1 m at a depth of 4 m). This is because the Rayleigh wave energy
3 [2 5 ∞] [200 45 120] 0.35 1800 0.03 is confined close to the soil surface.
Considering the slab track, the performance benefits are less clear,
with 1 m of improvement (at 150 and 200 MPa stiffness’s) resulting in
Table 4
Track properties (ballasted track and slab track). increased track deflections. This can be explained by analysing the re-
lationship between train speed and rail deflections, (i.e. the dynamic
Ballasted track Slab track
amplification factor). Fig. 16 shows this relationship for the cases of no
Rail EIr (Nm2) 1.26 × 107 1.26 × 107
soil improvement and 1 m of improvement. It is seen that the train
mr (kg/m) 120 120 speed is marginally past the critical velocity when no improvement is
Railpad kp (N/m) 5.5 × 10 8 5.5 × 10 8 performed. However, improving the ground stiffness shifts the critical
cp (Ns/m) 2.5 × 105 2.5 × 105 velocity to a higher speed meaning that the 1 m improvement cases are
Sleepers ms (kg/m) 490 – located closer to the peak, resulting in greater displacement amplifi-
Ballast/slab h (m) 0.35 0.35 cation. Other than these cases, track displacements are lower than the
E (MPa) 150 3 × 10 4 non-improved case for all improvement depths and stiffness’s. Again,
2b (m) 2.5 2.5 the relationship between depth of improvement and rail displacements
(kg/m3) 1600 2500
is non-linear.
Regarding the effect of the stiffness of the replacement soil, as ex-
pected, stiffer soils result in lower track deflections. For example, for
the ballasted 1 m replacement case, the rail displacements are 2.71 mm
different magnitudes of stiffness improvement are tested (150, 200,
and 2.35 mm when the replacement stiffness is 150 and 250 MPa re-
250 MPa).
spectively. Comparing this to the 2 m case, 2 m of 150 MPa soil has a
Fig. 15 shows the effect on rail displacement due to these changes
rail displacement of 1.97 mm, meaning the use of 2 m of softer fill
for both ballasted and slab tracks, with the results further summarised

Fig. 15. Maximum vertical rail displacement for Soil 1. (Left) Ballasted track; (Right) Slab track; Dashed black line represents the original value.

186
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Table 5 time histories for low speed is regular but oscillations occur as it ap-
Rail displacement improvement in percentage terms for Soil 1 at a train speed proaches the critical velocity. Comparing the maximum stress values at
of 330 km/h. different speeds, the x and y stress components increase by 423% and
Improvement 150 MPa 200 MPa 250 MPa 448% respectively, while the xz shear stress increases by 371%. In
depth comparison, although the maximum z stress component remains the
Ballasted Slab Ballasted Slab Ballasted Slab largest, it only increases by 124% due to the speed change. Therefore at
low speed the response is dominated by the vertical response, while at
1m 16.3% −4.0% 22.8% −1.1% 27.3% 1.5%
2m 38.6% 10.9% 44.6% 16.8% 48.3% 21.1% critical speed the entire stress field becomes important.
3m 48.1% 23.1% 53.4% 29.5% 56.7% 33.9%
4m 52.9% 30.9% 58.0% 37.3% 61.1% 41.6%
5m 55.7% 35.8% 60.6% 42.1% 63.6% 46.3% Soil case 2

Soil test case 2 consists of a 5 m deep soft layer (45 MPa) overlying a
provides better performance than 1 m of stiffer fill. Alternatively stiffer layer (120 MPa), which is homogenous, and infinitely deep. Five
though, at greater depths the opposite is true, for example with 4 m of different depths of soil improvement from the soil–track interface are
stiffer fill providing lower deflections than 5 m of softer fill. Also, for the considered (1, 2, 3, 4, 5 m), where 5 m means all of the soft 45 MPa
ballasted track, the displacement reduction between 150, 200 and layer is replaced. For each case, the soft layer is replaced with a soil of
250 MPa is relative constant for all depths of improvement. However, stiffness of 250 MPa, and the effect on rail displacement is analysed.
for the slab track this not true, with the percentage benefit between Fig. 18 shows the effect of soil replacement for both ballasted and slab
150 MPa and 250 MPa at 1 m, lower than that for the 5 m case. track cases, with the percentage terms shown in Table 6. Again the
Therefore, considering the non-linear relationships between improved concrete slab track has lower deflections due to its elevated bending
soil stiffness and improvement depth, soil remediation must be under- stiffness, however in contrast to soil case 1, all depths of improvement
taken carefully to ensure displacement tolerances are met and cost is result in reduced track deflections. For the ballasted track the re-
minimised. lationship between improvement depth and rail displacement is highly
Additionally, Fig. 17 shows a stress time history comparison at the non-linear. When 1 m of improvement is performed, displacements re-
surface of the soil below the central line of the ballasted track, for duce by 49.1%, however the reductions are only 64.4, 69.4, 71.8 and
speeds 20 m/s and 92 m/s. It is observed that the shape of the stress 73.2% for 2, 3, 4, 5 m replacements respectively. Therefore for this

Fig. 16. Maximum vertical rail displacement vs train speed, for Soil 1 with 1 m of soil replacement. (Left) Ballasted track; (Right) Slab track; Dashed black line
represents train speed.

Fig. 17. Stress time history comparison at the surface for ballasted track. (Left) c = 20 m/s (Right) c = 92 m/s.

187
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Fig. 18. Maximum vertical rail displacement for Soil 2. (Left) Ballasted track; (Right) Slab track; Dashed black line represents the original value.

embankment overlying the soft-stiff stratum from soil case 2 (Fig. 21).
Table 6 The embankment is modelled in a simplified manner, i.e., considering it
Rail displacement improvement in percentage for Soil 2 at a train speed of
as an additional layer, infinitive in the transversal direction. To simu-
330 km/h.
late the presence of an embankment, it is modelled as an additional soil
Improvement depth 250 MPa layer, which has been shown to be a reasonable approximation in Alves
Costa et al. [3]. This results in a 5 m thick soft layer (45 MPa) sand-
Ballasted Slab
wiched between two stiffer layers. The upper embankment layer has a
1m 49.1% 19.3% stiffness of 200 MPa and the lower ground layer is infinitely deep with a
2m 64.4% 36.1% stiffness of 120 MPa. Again, improvement depths between 1 and 5 m are
3m 69.4% 45.7%
considered.
4m 71.8% 50.9%
5m 73.2% 53.9%
Fig. 22 and Table 7 show the relationship between soil replacement
depth and track displacement. It is observed that the stiff embankment
causes a significant reduction in rail displacement compared to Fig. 18.
case, depending upon the desired deflection levels, it may be most cost This is because the dispersion characteristics are shifted the critical
effective to remediate only the top 1 m of soil. For the slab track the velocity to a higher value, due to the presence of the embankment.
relationship is also non-linear, however less pronounced. Again, in- Similar to soil case 2, all improvement depths result in reduced track
creasing improvement depth reduces deflections, however the change displacements, however the relationship is no longer relatively linear.
in performance benefit reduces as the improvement depth approaches Instead, only a small reduction is achieved by improving 2 m compared
5 m. For dynamic amplification curves, please see Fig. 20. to 1 m, while improving 3 m compared to 2 m results in a large re-
Fig. 19 shows the effect of soil replacement on displacements. It is duction. This is true for both track types, but more pronounced for the
seen that the unmodified soil gives rise to significant wave propagation, ballasted case. Therefore, depending on the desired rail deflection tol-
with ‘Mach cone’ type contours present. In comparison, for the fully erances, it might be most efficient to replace the top 3 m rather than
remediated soil stratum, the response is more symmetric in both hor- either 2 or 4 m. It should also be noted that although the rail dis-
izontal directions. placements for this particular track-soil combination may not have re-
quire remediation in practise, the finding of non-linearity in soil im-
provement depth is also applicable to other track-soil stiffness
Soil case 3 combinations.
In addition to rail displacements, Fig. 23 shows the resulting max-
Soil test case 3 is chosen to determine the effect of an embankment imum vertical stresses within the embankment (2 m) and soil to a depth
[34,41] overlying a soft soil site. Therefore it consists of a 2 m high

Fig. 19. Rail displacement comparisons for Soil 2. (Left) Original soil; (Right) 5 m depth of improvement.

188
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Fig. 20. Maximum vertical rail displacement vs train speed, for Soil 2 with different depths of soil replacement strategies. (Left) Ballasted track; (Right) Slab track.

Table 7
Rail displacement improvement in percentage for Soil 3 at 300 km/h.
Improvement depth 250 MPa

Ballasted Slab

1m 4.0% 2.9%
2m 7.3% 5.5%
3m 21.8% 20.8%
4m 28.5% 29.1%
5m 32.2% 33.9%

lower at 7 m compared to the softer soils, the upper 2 m of the stiffer


soil experiences slightly higher stress levels.

Discussion

When planning, designing and constructing a new railway line it is


important to assess the track-soil dynamics along the route. If the op-
Fig. 21. Schematic diagram of the slab track used in the simulation overlaying erational train speed is greater than 50% of the combined track-soil
on the soil. wave speed then there is the potential for dynamic effects to cause
excessive track deflections. To mitigate this risk, at the planning stage, a
scoping exercise can be performed to approximate the absolute critical
velocity. This can be done for example using dispersion curve analysis
of 5 m below the embankment. It is seen that despite the soil replace-
based upon historical geotechnical data or surface wave testing data. If
ment having a significant impact of rail deflections, the effect on stress
the train speed is calculated to be greater than 50% of the critical ve-
levels is less defined. This is particularly true for the ballasted track,
locity then the TLM model can be used to perform a more detailed
however for the slab track, although the stresses for stiffer soils are
assessment (i.e. calculation of track-ground DAF curves). Finally, if soil

Fig. 22. Maximum downward rail displacement for Soil 3. (Left) Ballasted track; (Right) Slab track; Dashed black line represents the original value.

189
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

Fig. 23. Vertical stress vs Depth for Soil case 3. (Left) Ballasted track; (Right) Slab track.

replacement is deemed necessary then model can be combined with a Competitividade e Internacionalização (POCI) and by national funds
costing tool to optimise the depth of remediation required, with respect (PIDDAC) through FCT/MCTES.
to price.
Lastly, it should be noted that the aforementioned analysis assumes References
linear soil behaviour. As train speeds approach the critical velocity,
strain levels within the soil can move to the high strain range [3]. This [1] Alves Costa P, Calcada R, et al. A 2.5D finite element model for simulation of un-
results in a non-linear relationship between strain and stiffness. The bounded domains under dynamic loading. In: Numerical methods in geotechnical
engineering; 2010. p. 397–404.
thin-layer method approach used within the model is well suited to the [2] Alves Costa P, et al. Critical speed of railway tracks. Detailed and simplified ap-
simulation of these effects. proaches. Available at: Transp Geotech 2015;2:30–46. https://doi.org/10.1016/j.
trgeo.2014.09.003.
[3] Alves Costa P, Calçada R, et al. Influence of soil non-linearity on the dynamic re-
Conclusion sponse of high-speed railway tracks. Soil Dyn Earthq Eng 2010;30(4):221–35.
[4] Alves Costa P. Vibrações Do Sistema Via-Maciço Induzidas Por Tráfego Ferroviário.
Modelação Numérica E Validação Experimental; 2011.
Soil replacement or improvement is commonly used on new railway [5] Alves Costa P, Calçada R, Silva Cardoso A. Track-ground vibrations induced by
lines to increase track stiffness, however is expensive. Therefore a railway traffic: In-situ measurements and validation of a 2.5D FEM-BEM model. Soil
model was developed capable of rapidly assessing the effect of different Dyn Earthq Eng 2012;32(1):111–28.
[6] Arlaud E, Costa D’Aguiar S, Balmes E. Receptance of railway tracks at low fre-
soil improvement depths and stiffness’ on track behaviour. The model
quency: numerical and experimental approaches. Available at: Transp Geotech
can therefore be used to optimise soil replacement with respect to 2016;9:1–16. https://doi.org/10.1016/j.trgeo.2016.06.003.
construction cost. The modelling methodology comprised of a thin- [7] Arlaud E, Costa D’Aguiar S, Balmes E. Validation of a reduced model of railway
track allowing long 3D dynamic calculation of train-track interaction. In: Computer
layer method ground model, coupled in the wavenumber-frequency
methods and recent advances in geomechanics – proceedings of the 14th int. con-
domain to an analytical track model. It was capable of simulating the ference of international association for computer methods and recent advances in
deep wave propagation that occurs when approaching critical velocity. geomechanics, IACMAG 2014, (September); 2015. p. 1193–1198.
Validation was performed using a combination of field experimental [8] Auersch L. The excitation of ground vibration by rail traffic: theory of vehicle-track-
soil interaction and measurements on high-speed lines. J Sound Vib
data and published numerical results, and the model found to have 2005;284(1–2):103–32.
strong accuracy. To show the model capabilities, three earthwork si- [9] Barbosa JM de O, Park J, Kausel E. Perfectly matched layers in the thin layer
tuations were considered: a homogenous soft soil, a soft soil over a stiff method. In: Computer methods in applied mechanics and engineering, 217–220
(October); 2012. p. 262–274. Available at: https://doi.org/10.1016/j.cma.2011.12.
soil, and a stiff embankment resting on a soft soil over a stiff soil. It was 006.
found that due to the complex nature of the 3D wave propagation and [10] Bian X, et al. Numerical analysis of soil vibrations due to trains moving at critical
its dependency on multiple variables, the relationship between im- speed. Available at: Acta Geotech 2016;11(2):281–94. https://doi.org/10.1007/
s11440-014-0323-2.
provement depth and rail displacement can be highly non-linear. [11] Bian X, et al. A 2.5D finite element approach for predicting ground vibrations
Therefore soil improvement must be designed carefully to achieve the generated by vertical track irregularities. J Zhejiang Univ-Sci A
desired track performance. 2011;12(12):885–94 Available at: http://link.springer.com/10.1631/
jzus.A11GT012.
[12] Bian X, Chen Y, Hu T. Numerical simulation of high-speed train induced ground
Acknowledgements vibrations using 2.5D finite element approach. Sci China Ser G: Phys Mech Astron
2008;51(6):632–50.
[13] Bian X, Chen Y. An explicit time domain solution for ground stratum response to
Model development was a collaboration between the University of harmonic moving load. Acta Mech Sin/Lixue Xuebao 2006;22(5):469–78.
Porto, Heriot-Watt University and the University of Leeds. The authors [14] Chebli H, Othman R, et al. 3D periodic BE-FE model for various transportation
structures interacting with soil. Comput Geotech 2008;35(1):22–32.
would like to thank the Leverhulme Trust (UK), the University of Porto,
[15] Chebli H, Clouteau D, Schmitt L. Dynamic response of high-speed ballasted railway
Heriot Watt University and the University of Leeds. Without their tracks: 3D periodic model and in situ measurements. Soil Dyn Earthq Eng
support, this research would not have been possible. In addition, the 2008;28(2):118–31.
participation of last author was financially supported by: Project POCI- [16] Connolly D, Giannopoulos A, Fan W, et al. Optimising low acoustic impedance
back-fill material wave barrier dimensions to shield structures from ground borne
01-0145-FEDER-007457 – CONSTRUCT – Institute of R&D In Structures high speed rail vibrations. Constr Build Mater 2013;44:557–64.
and Construction funded by FEDER funds through COMPETE2020 – [17] Connolly D, Giannopoulos A, Forde MC. Numerical modelling of ground borne vi-
Programa Operacional Competitividade e Internacionalização (POCI) – brations from high speed rail lines on embankments. Available at: Soil Dyn Earthq
Eng 2013;46:13–9. https://doi.org/10.1016/j.soildyn.2012.12.003.
and by national funds through FCT – Fundação para a Ciência e a [18] Correia dos Santos N, et al. Experimental analysis of track-ground vibrations on a
Tecnologia; Project POCI-01-0145-FEDER-029577 – funded by FEDER stretch of the Portuguese railway network. Available at: Soil Dyn Earthq Eng
funds through COMPETE2020 – Programa Operacional 2016;90:358–80. https://doi.org/10.1016/j.soildyn.2016.09.003.

190
K. Dong et al. Transportation Geotechnics 17 (2018) 178–191

[19] Dieterman HA, Metrikine AV. The equivalent stiffness of a half-space interacting 1995;44(2):149–64.
with a beam. Critical velocities of moving load along the beam. Eur J Mech A/Solids [36] Lamb H. On the propagation of tremors over the surface of an elastic solid. Philos
1996;15:67–90. Trans R Soc Lond 1904;203:1–42.
[20] Ferreira PA, López-Pita A. Numerical modelling of high speed train/track system for [37] Madshus C, Kaynia AM. High-speed railway lines on soft ground: dynamic beha-
the reduction of vibration levels and maintenance needs of railway tracks. Available viour at critical train speed Available at: J Sound Vib 2000;231(3):689–701http://
at: Constr Build Mater 2015;79:14–21. https://doi.org/10.1016/j.conbuildmat. linkinghub.elsevier.com/retrieve/pii/S0022460X99926470.
2014.12.124. [38] Mezher SB, et al. Railway critical velocity – analytical prediction and analysis.
[21] François S, et al. A 2.5D coupled FE-BE methodology for the dynamic interaction Transp Geotech 2016;6(October):84–96.
between longitudinally invariant structures and a layered halfspace. Available at: [39] Montalbán L, et al. Analysis of vibrations generated by rail corrugation in a high
Comput Methods Appl Mech Eng 2010;199(23–24):1536–48. https://doi.org/10. speed track, 2015;2015: 12–16.
1016/j.cma.2010.01.001. [40] O’Brien J, Rizos DC. A 3D BEM-FEM methodology for simulation of high speed train
[22] Fryba L. Vibration of soilds and structures under moving loads. Netherlands: induced vibrations. Soil Dyn Earthq Eng 2005;25(4):289–301.
Springer; 1972. [41] Oliver B, et al. The effect of embankment on high speed rail ground vibrations.
[23] Galvín P, et al. A 2.5D coupled FE-BE methodology for the prediction of railway Available at: Int J Rail Transp 2016;4(4):229–46. https://doi.org/10.1080/
induced vibrations. Soil Dyn Earthq Eng 2010;30:1500–12. 23248378.2016.1220844.
[24] Galvín P, Domínguez J. Analysis of ground motion due to moving surface loads [42] Powrie W, Yang LA, Clayton CRI. Stress changes in the ground below ballasted
induced by high-speed trains. Eng Anal Boundary Elem 2007;31(11):931–41. railway track during train passage. Proc Inst Mech Eng, Part F: J Rail Rapid Transit
[25] Hall L. Simulations and analyses of train-induced ground vibrations in finite ele- 2007;221(2):247–62.
ment models. Soil Dyn Earthquake Eng 2003;23(5):403–13. [43] Sheng X, Jones CJC, Thompson DJ. A comparison of a theoretical model for quasi-
[26] Hendry M, Hughes DA, Barbour L. Track displacement and energy loss in a railway statically and dynamically induced environmental vibration from trains with
embankment Available at: Proc Inst Civ Eng – Geotech Eng measurements. J Sound Vib 2003;267(3):621–35.
2010;163(1):3–12http://www.icevirtuallibrary.com/doi/10.1680/geng.2010.163. [44] Shih JY, Thompson DJ, Zervos A. The effect of boundary conditions, model size and
1.3. damping models in the finite element modelling of a moving load on a track/ground
[27] El Kacimi A, et al. Time domain 3D finite element modelling of train-induced vi- system. Available at: Soil Dyn Earthq Eng 2016;89:12–27. https://doi.org/10.1016/
bration at high speed. Available at: Comput Struct 2013;118:66–73. https://doi. j.soildyn.2016.07.004.
org/10.1016/j.compstruc.2012.07.011. [45] Shih JY, Thompson DJ, Zervos A. The influence of soil nonlinear properties on the
[28] Kausel, E. An explicit solution for the Green functions for dynamic loads in layered track/ground vibration induced by trains running on soft ground. Available at:
media; 1981. Transp Geotech 2017;11:1–16. https://doi.org/10.1016/j.trgeo.2017.03.001.
[29] Kausel E. Thin-layer method: formulation in the time domain. Int J Numer Meth [46] Steenbergen MJMM, Metrikine AV. The effect of the interface conditions on the
Eng 1994;37(6):927–41. dynamic response of a beam on a half-space to a moving load. Eur J Mech, A/Solids
[30] Kausel E. Wave propagation in anisotropic layered media. Int J Numer Meth Eng 2007;26(1):33–54.
1986;23(8):1567–78. [47] Steenbergen MJMM, Metrikine AV, Esveld C. Assessment of design parameters of a
[31] Kausel E, Roesset JM. Stiffness matrices for layered soils. Bull Seismol Soc Am slab track railway system from a dynamic viewpoint. J Sound Vib
1981;71(6):1743–61. 2007;306(1–2):361–71.
[32] Kaynia AM, Madshus C, Zackrisson P. Ground vibration from high-speed trains: [48] Thompson D. Railway noise and vibration: mechanisms, modelling and means of
Prediction and countermeasure. J Geotech Geoenviron Eng 2000;126(Igvc):531–7. control. Elsevier; 2008.
[33] Kouroussis G, et al. Discrete modelling of vertical track-soil coupling for vehicle- [49] Varandas JN, et al. A numerical study on the stress changes in the ballast due to
track dynamics. Available at: Soil Dyn Earthq Eng 2011;31(12):1711–23. https:// train passages. Available at: Procedia Eng 2016;143(Ictg):1169–76. https://doi.
doi.org/10.1016/j.soildyn.2011.07.007. org/10.1016/j.proeng.2016.06.127.
[34] Kouroussis G, et al. Railway cuttings and embankments: experimental and numer- [50] Yang YB, Hung HH, Chang DW. Train-induced wave propagation in layered soils
ical studies of ground vibration. Available at: Sci Total Environ using finite/infinite element simulation. Soil Dyn Earthq Eng 2003;23(4):263–78.
2016;557–558:110–22. https://doi.org/10.1016/j.scitotenv.2016.03.016. [51] Zhai W, He Z, Song X. Prediction of high-speed train induced ground vibration
[35] Krylov VV. Generation of ground vibrations by superfast trains. Appl Acoust based on train-track-ground system model. Earthq Eng Eng Vibr 2010;9(4):545–54.

191

You might also like