You are on page 1of 22

Vehicle System Dynamics

International Journal of Vehicle Mechanics and Mobility

ISSN: 0042-3114 (Print) 1744-5159 (Online) Journal homepage: https://www.tandfonline.com/loi/nvsd20

Influence of some vehicle and track parameters on


the environmental vibrations induced by railway
traffic

G. Kouroussis , O. Verlinden & C. Conti

To cite this article: G. Kouroussis , O. Verlinden & C. Conti (2012) Influence of some vehicle
and track parameters on the environmental vibrations induced by railway traffic, Vehicle System
Dynamics, 50:4, 619-639, DOI: 10.1080/00423114.2011.610897

To link to this article: https://doi.org/10.1080/00423114.2011.610897

Published online: 09 Sep 2011.

Submit your article to this journal

Article views: 759

View related articles

Citing articles: 12 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=nvsd20
Vehicle System Dynamics
Vol. 50, No. 4, April 2012, 619–639

Influence of some vehicle and track parameters on the


environmental vibrations induced by railway traffic
G. Kouroussis*, O. Verlinden and C. Conti
Faculty of Engineering, Department of Theoretical Mechanics, Dynamics and Vibrations,
Université de Mons – UMONS, Place du Parc 20, B-7000 Mons, Belgium

(Received 10 March 2011; final version received 30 July 2011 )

A study is performed on the influence of some typical railway vehicle and track parameters on the
level of ground vibrations induced in the neighbourhood. The results are obtained from a previously
validated simulation framework considering in a first step the vehicle/track subsystem and, in a second
step, the response of the soil to the forces resulting from the first analysis. The vehicle is reduced to
a simple vertical 3-dof model, corresponding to the superposition of the wheelset, the bogie and the
car body. The rail is modelled as a succession of beam elements elastically supported by the sleepers,
lying themselves on a flexible foundation representing the ballast and the subgrade. The connection
between the wheels and the rails is realised through a non-linear Hertzian contact. The soil motion
is obtained from a finite/infinite element model. The investigated vehicle parameters are its type
(urban, high speed, freight, etc.) and its speed. For the track, the rail flexural stiffness, the railpad
stiffness, the spacing between sleepers and the rail and sleeper masses are considered. In all cases, the
parameter value range is defined from a bibliographic browsing. At the end, the paper proposes a table
summarising the influence of each studied parameter on three indicators: the vehicle acceleration, the
rail velocity and the soil velocity. It namely turns out that the vehicle has a serious influence on the
vibration level and should be considered in prediction models.

Keywords: railway systems; vehicle–track interaction; vibration control; track models; multibody
systems; finite elements

1. Introduction

Over the last 20 years the railway has been the subject of much criticisms in terms of vibrations.
The development of new lines, especially for the high-speed network, is often felt by residents
as high vibratory nuisances even though the rail network modernisation has the aim to improve
the dwellers’ quality of life. In the case of urban tramways, the distance between the railway
track and the surrounding buildings is decreased and the evaluation of buildings damage and
dwellers’ discomfort becomes necessary.
The development of prediction models has been the subject of intensive research. However,
the problem is complex as it involves three main subsystems: the vehicle, the track and the
soil, which should be integrated in the model with a sufficient accuracy. Simple analytical

*Corresponding author. Email: georges.kouroussis@umons.ac.be

ISSN 0042-3114 print/ISSN 1744-5159 online


© 2012 Taylor & Francis
http://dx.doi.org/10.1080/00423114.2011.610897
http://www.tandfonline.com
620 G. Kouroussis et al.

models were firstly proposed, establishing a close relationship between the vehicle axle loads
and the soil surface motion, but their application is limited to specific conditions and cases.
The perfect example is without any doubt the Krylov model [1], where the vehicle, considered
as a sequence of constant axle loads, rides on a flexible beam lying on Winkler’s foundation.
The rail deflection is combined with the sleepers distribution, to determine the forces acting
on the soil surface. Whatever the interesting results brought by this model, it is obvious that
it does not consider the contribution of the vehicle dynamics.
In railway traffic, the vehicle speed and weight are systematically pointed up as the dominant
parameters when ground vibrations are concerned. According to the literature review, the
vehicle speed is an important parameter which contributes significantly to the ground vibration
level, as it induces a spread of the rail deflection. For example, Picoux and Le Houédec [2]
have modelled the vehicle as vertical and lateral rectangular loads, related to the effect of axle
weight and to the centrifugal force in curve passing, respectively. Their results underline the
particular case of high speed, compared to the Rayleigh wave speed. The same literature shows
that the vehicle dynamics and its effect in the soil-transmitted vibrations are rarely underlined
although vehicle suspensions have most probably an important role, especially when the
vehicle moves on a discrete irregularity. Recently, Kouroussis et al. [3] have validated their
numerical approach in the case of urban railway traffic and have demonstrated the benefit of
modifying the motor wheel vertical stiffness on the ground level: a reduction of 70% was
observed, without changing the axle loads, when the vehicle moves on a local defect.
Some sensitivity studies are also found in literature, depending on whether the track or the
soil is analysed. In their parametric study, Yang et al. [4] emphasise various soil effects, but
from a simple model for the vehicle (set of harmonic loads) and the track (beam uniformly
supported by a Winkler’s foundation). The effect of the track configuration has been treated
by Metrikine et al. considering the rail continuously supported [5] or periodically supported
by the sleepers [6], and applied in the case of high-speed vehicles, demonstrating that the
vertical deflection of the rail is identical in both models, under some conditions. Anti-vibration
solutions directly supported by the track are well-known examples of vibration reduction by
modifying or adding local stiffnesses on the track. Wang et al. have studied the efficiency of
rubber-modified concrete material through a bi-dimensional model [7] when a harmonic load
is applied on the track–soil subsystem. Vogiatzis and Chaikali [8] have described numerically
the benefit of the fasterner-less Prefarail technology for the tramway in Athens. Wang et al. [9]
have analysed the vibratory responses of rail foundations utilising different ballast materials.
They conclude that the rubber-modified asphalt concrete is a satisfactory solution for filtering
the ground-borne vibrations. All these examples bring interesting clues on the influence of
the track configuration. Although the soil configuration was largely studied, the track is rarely
analysed in detail. Railpads and ballast are often included in the track flexibility through a
Winkler’s foundation. The sleeper influence (mass and spacing) is rarely taken into account,
if so the sleeper is uniformly distributed along the track. Grassie et al. [10] have shown that
these models are deficient from 50 Hz, a discretely supported Euler–Bernoulli beam being
better adapted for the high frequencies. A Timoshenko beam can also be used, as it takes into
account the shear deformation of the rail, but the Euler–Bernoulli theory provides acceptable
results for the response to vertical forces at frequencies less than about 500 Hz, and gives
informations about the so-called pinned–pinned resonance mode.
The assumption of harmonic load is a simple approach to analyse the benefit of anti-vibration
solutions. With this approach, it is possible to determine the sensitivity of the soil vibratory
level to the vehicle and track components and consequently to identify the most interesting
ones. The main limitation of the frequency approach is that the global vibration level also
depends on the frequency content of the source. The latter corresponds to the rail/wheel
contact forces and actually derives from the vehicle/track interaction.
Vehicle System Dynamics 621

The main objective of this paper is to present a comprehensive study, considering some
vehicle and track parameters whose influence has been poorly studied so far in literature. The
results are yielded by the numerical prediction model recently developed and validated by
Kouroussis et al. [3,11]. The paper first presents the main characteristics and assumptions of
the model. The envisaged vehicle and track parameters, and their range, established according
to a literature survey, are then outlined. The investigated vehicle parameters are the vehicle
type and the speed. Only a quarter-vehicle model has been retained. Although very simple,
this model has the advantage of avoiding any particular geometry of the vehicle and is suf-
ficient for the considered study. Concerning the track, the rail flexural stiffness, the railpad
stiffness, the spacing between sleepers and the rail and sleeper masses are analysed. The vari-
ous configurations are then compared, by means of three indicators classically used to assess
the human or structural risk due to vibration exposure. They are associated to the onboard
vertical acceleration, the track deflection and the soil velocity. The last section summarises
the parametric study, emphasising among other things the effect of the vehicle and the track
dynamics. The main finding is finally given.

2. Numerical model and basic assumptions

The proposed numerical model works in two successive stages. The first one implements
the simulation of the vehicle/track subsystem. The vibration being generated not only from
the repetitive nature of the track but also from local or distributed irregularities on the rail
surface, is analysed. The history of ground forces yielded by this first step are then applied
to the tridimensional finite/infinite element (FIEM) model of the soil. Both analyses are
performed in the time domain. It has been shown that the assumption of decoupling between
the soil and the vehicle/track subsystems is valid as far as the soil is sufficiently rigid with
respect to the ballast. Full details about the track modelling and the decoupling validity are
provided in [3,12]. The proposed prediction model has been validated against experimental
measurements for some typical railway applications, namely the Brussels tramway [3,13] and
the Thalys and Eurostar high-speed trains [11].
Only a half-bogie model is used in this study, considering the wheelset, the bogie frame
and the car body parts as rigid bodies (with the masses mw , mb and mc , respectively, Figure 1).
Primary and secondary suspensions are represented by linear springs ki and dampers di (i = 1
or 2) and the wheel/rail contact is induced by the Frail/wheel force (Hertz’s contact law). The

v0
mc
d2 k2
mb
d1 k1
mw
Frail/wheel
L

dp kp
−Fwheel/rail m
db kb
df kf

Figure 1. Basic model for the parametric study.


622 G. Kouroussis et al.

equations governing the motion of the vehicle take the form:


⎡ ⎤⎧ ⎫ ⎡ ⎤⎧ ⎫
mc 0 0 ⎨ q̈c ⎬ d2 −d2 0 ⎨ q̇c ⎬
⎣0 mb 0 ⎦ q̈b + ⎣−d2 d1 + d2 −d1 ⎦ q̇b
⎩ ⎭ ⎩ ⎭
0 0 mw q̈w 0 −d1 d1 q̇w
⎡ ⎤⎧ ⎫ ⎧ ⎫
k2 −k2 0 ⎨ qc ⎬ ⎨ −mc g ⎬
+ ⎣−k2 k1 + k2 −k1 ⎦ qb = −mb g . (1)
⎩ ⎭ ⎩ ⎭
0 −k1 k1 qw Frail/wheel − mw g

As the vehicle is reduced to the vertical plane, a single rail is used for the track behaviour.
Euler–Bernoulli beam elements are used to represent the rail (Young’s modulus Er , cross
section Ar , density ρr and geometrical moment of inertia Ir ), discretely supported by lumped
masses m representing the half-sleepers, with a regular spacing L. Railpad (subscript p) and
ballast (b) are also described by linear springs and dampers. The ballast is connected to
the foundation by a viscoelastic layer (kf , df ) corresponding to the direct soil impedance.
The problem is solved in the time domain, allowing to consider moreover the non-linear
effect of the wheel–rail contact. The track irregularity is taken into account directly in the
vehicle/track simulation using a stochastic approach based on data collected by Garg an
Dukkipati [14].
The soil simulation is performed using a 3-D finite element model, in which infinite elements
are added at the border of the region of interest. Time domain simulation is more appropriate to
simulate the transient response of the soil: a recent paper [15] has shown that a compact finite–
infinite element model gives accurate solutions in time analysis so that tridimensional cases
can be considered on usual computers. The calculation time can also be limited with the help of
an explicit integration scheme coupled with parallel computer programming [16]. The detailed
description of the soil modelling is reported in [15], represented here by a homogeneous half-
space (density ρ, Young’s modulus E, Poisson’s ratio ν and viscous damping ratio β). Soil
spring kf and damping df can be derived from Wolf’s formulas [17] referring to a homogeneous
half-space (defined by its Young’s modulus E, its density ρ and its Poisson’s ratio ν)
a 0.75 
Eb
kf = 1.55 + 0.8 , (2)
(1 − ν 2 ) b

2ab(1 + ν)ρ
df = 1.48 kf (3)
πE

for a load applied on a rectangular surface with 2a and 2b dimensions (b < a).

3. Outline of the study

3.1. Reference case

Table 1 presents the reference parameters of the complete model, based on the Brussels tram
and site characteristics [29]. To cover a realistic range of vehicle and track parameters values,
a survey of literature has been made. The choice of this site is caused by the large knowledge
of the vehicle and the track behaviour, obtained from tram constructors or collected thanks to
in situ measurements (track and soil parameters). Values are adapted to be coherent with the
quarter-car model. Reference vehicle speed is assumed to be equal to 100 km/h.
Vehicle System Dynamics 623

Table 1. Reference and studied-case parameters.

Component Parameter Reference value Other values

Train Car body mass ( 41 ) mc 4000 kg 6000 kg [18], 12,420 kg [19],


13,000 kg [20], 17,500 kg [21]
Bogie frame mass ( 21 ) mb 900 kg 800 kg [21], 1400 kg [20],
5000 kg [18,19]
Wheelset mass mw 320 kg 1200 kg [18,19], 1400 kg [21],
2027 kg [20]
k1 6 MN/m 1.12 MN/m [22], 2 MN/m [18],
Primary suspension 2.28 MN/m [21]
d1 6 kN s/m 12 kN s/m [22], 15 kN s/m [18],
23.3 kN s/m [21]
k2 0.5 MN/m 0.6 MN/m [20], 4 MN/m [18]
Secondary suspension d2 30 kN s/m 4 kN s/m [20], 30 kN s/m [18]
Speed v0 100 km/h From 50 km/h to 300 km/h
Track (EB50T) Section Ar 63.8 cm2 
Geometrical moment of inertia Ir 1987 cm4 rail UIC 60, UIC 50 and others
Young’s modulus Er 210 GPa
Density ρr 7850 kg/m3
Track irregularity h(x) class 2 [14] Fixed
Railpad Stiffness kp 90 MN/m 6 MN/m [23], 60 MN/m [18],
100 MN/m [24], 180 MN/m [19],
237 MN/m [25], 280 MN/m [10]
Damping dp 30 kN s/m 5 kN s/m [26], 28 kN s/m [19],
52 kN s/m [18], 63 kN s/m [10]
Sleeper Mass m 90.84 kg 52 kg [19], 110 kg [10], 125 kg [22],
200 kg [25], 290 kg [27]
Spacing L 0.72 m 0.60 m [25], 0.698 m [10]
Ballast Stiffness kb 32 MN/m 12 MN/m [18], 40 MN/m [28],
120 MN/m [19], 180 MN/m [10],
190 MN/m [22]
Damping db 52 kN s/m 12 kN s/m [18], 50 kN s/m [28],
82 kN s/m [10], 100 kN s/m [22]
Soil Young’s modulus E 146 MPa 
Density ρ 2000 kg/m3 No variation in this study
Poisson’s ratio ν 0.28
Damping β 0.0003 s

3.2. Analysed parameters

With a global model for the vehicle and the track, the parametric study will be as complete
as possible. The results of this sensitivity analysis are presented in the following sections,
classified in terms of vehicle and rail parameters, as presented in Figure 2.
Some components and their values are more detailed, for example, the railpad and ballast
stiffnesses. Several comments can be made regarding the proposed values:

• For the vehicle, stability and comfort studies were performed by constructors and have
allowed to understand vehicle dynamics, implying that masses’ proportion and suspensions
parameters vary in a lesser range. An analysis based on the train type is more pertinent.
• A multitude of rail profiles exist, depending on the country, and will also be investigated.
• The sensitivity of rail irregularity was studied in [12], showing the great influence of these
parameters on the ground vibrations. In this study, the track quality is assumed to be bad
(class 2 according to Garg’s and Dukkipati’s analytical characterisation [14]).
• In the reference case, the soil is considered as homogeneous and viscoelastic.
624 G. Kouroussis et al.

vehicle track

rail railpad sleeper sleeper ballast


type speed mass
type stiffness spacing stiffness

Figure 2. Analysed vehicle and track parameters.

3.3. Criteria of comparison

Indicators are used as yardsticks to compare the influence of the vehicle and track parameters
on the soil surface. In order to have an overall view of the vehicle/track dynamics problem, the
vertical car body accelerations and the rail deflections will be compared, as these parameters
are related to the passenger’s discomfort and the rail stability, respectively. As the soil vertical
velocity is significant and is systematically used in literature to describe the ground vibrations,
the vertical direction will be the only one presented afterwards, for the purpose of concision.
Two indicators have been retained for the soil:
• The peak particular velocity (PPV) defined as the maximum absolute amplitude of the
velocity time signal [30]
PPV = max |v(t)|, (4)
• the DIN weighted vibrations severity KBF described by the maximum of the time-averaged
signal
 
1 t
KBF (t) = KB2 (ξ ) e−(t−ξ )/τ dξ (τ = 0.125 s) (5)
τ 0
of the weighted velocity signal KB(t) obtained by flowing the original velocity signal
through the high-pass filter
1
HKB (f ) =  , (6)
1 + (5.6/f )2
according to [31].

4. Influence of the vehicle parameters

4.1. The vehicle type

Various vehicles have been studied, with their own characteristics, as presented in Table 2.
Five vehicles are analysed, ordered against the increasing axle load. Vehicle 1 is the reference
vehicle, that is to say, the model of the tram T2000 of Brussels. Vehicle 2 is based on the
German ICE, which has more flexible primary suspension but more damped, compared to
the reference vehicle. Vehicle 3 presents higher axle load due to the important weight of the
bogie while the car body mass is comparable. Vehicle 4 is derived from the Thalys HST with
important car body mass. Finally, vehicle 5 is typical of freight train. The axle loads vary from
Vehicle System Dynamics 625

Table 2. Data used in the simulation, for various studied vehicles.

Vehicle 1 Vehicle 2 Vehicle 3 Vehicle 4 Vehicle 5


Parameters (Brussels tram) (German ICE) (Corail) (Thalys HST) (Freight train)

mc 4000 kg 3750 kg 6000 kg 13,000 kg 17,500 kg


mb 900 kg 1250 kg 5000 kg 1400 kg 800 kg
mw 320 kg 500 kg 1200 kg 2027 kg 1400 kg
k1 6 MN/m 0.72 MN/m 2 MN/m 1.15 MN/m 22.8 MN/m
k2 0.5 MN/m 1.8 MN/m 4 MN/m 0.6 MN/m 4 MN/m
d1 6 kN s/m 40 kN s/m 15 kN s/m 2.5 kN s/m 2.33 kN s/m
d2 30 kN s/m 30 kN s/m 30 kN s/m 4 kN s/m 60 kN s/m
Car body mode 1.7 Hz 2.2 Hz 2.6 Hz 0.9 Hz 2.2 Hz
Bogie mode 18.2 Hz 4.3 Hz 7.1 Hz 7.3 Hz 38.7 Hz
Axle load 52 kN 55 kN 100 kN 164 kN 197 kN

(a) 1
Brussels tram
German ICE
Acceleration (m/s )

Corail
2

0.5
Thalys HST
Freight train
0

−0.5

−1
0 0.2 0.4 0.6 0.8 1
Time (s)
Time analysis
(b)
0
10
Brussels tram
German ICE
−1
10 Corail
Amplitude (m/s2)

Thalys HST
Freight train
−2
10

−3
10

−4
10
0 20 40 60 80 100
Frequency (Hz)

Third-octave band analysis

Figure 3. Sensitivity of the car body acceleration to the vehicle type.

52 kN (for the tram) to 197 kN (for the freight train). Car body and bogie bounce modes can
be easily calculated and are lower than 10 Hz (around 2 and 8 Hz, respectively), except for the
bogie modes of the tram and the freight train, at 18.2 and 38.7 Hz, respectively.
Figures 3 and 4 show the simulated results, on vehicle and track sides respectively. It turns
v
out that the Thalys vehicle provides less level to the car body (rms value az,rms of 0.14 m/s2 )
than the other vehicles, which at a first glance may appear coherent as the vehicle suspensions
are designed to bring a high comfort (Figure 3). The high levels are obtained for the Corail
v
(az,rms = 0.37 m/s2 ) and freight trains (az,rms
v
= 0.32 m/s2 ). For the last vehicle, the two bounce
modes appear in the results. For the track (Figure 4), the axle load is clearly the main parameter
626 G. Kouroussis et al.

(a) −3
x 10
0.5

Displacement (m)
−0.5

−1

−1.5 Brussels tram


−2 German ICE
Corail
−2.5 Thalys HST
Freight train
−3
0 0.1 0.2 0.3 0.4 0.5
Time (s)
Time analysis
(b) −2
10
Brussels tram
German ICE
−4
10 Corail
Amplitude (m)

Thalys HST
Freight train
−6
10

−8
10

−10
10
0 20 40 60 80 100
Frequency (Hz)
Spectral analysis

Figure 4. Sensitivity of the rail deflection to the vehicle type (deflection at mid-span of sleeper).

which imposes the rail deflection: Thalys and freight trains give the highest levels. The spectral
content shows that, for high axle load, a resonance appears around 40 Hz.
For the soil, the level diminishes, of course, with the distance from the track. The Thalys vehi-
cle implies the highest ground vibrations, in terms of PPV and KBF,max indicators (Figure 5),
in contrast to the freight train, the heaviest but generating reasonable ground vibrations. In
addition to these indicators, the time velocity signals are added to the figure, showing the
presence of a high frequency component in the Thalys and Corail train results, missing in the
freight train case. This frequency corresponds to the sleepers excitation mechanism frequency
defined by
v0
fs = , (7)
L
equal to 38.6 Hz with v0 being the vehicle speed and L the sleeper spacing. For the freight train,
this frequency is very close to the bogie bounce mode and the vehicle plays the unexpected
role of vibration absorber. The rail vibrations, induced by the passing of the vehicle at a
speed v0 is attenuated directly by the vehicle itself, whose bogie mode is identical to this
passing frequency. Figure 4 shows this effect: the peak response around 40 Hz comes down
to two peaks with lower height (the difference is more visible if the track structure is not
damped). Considering the KBF,max indicator, the Corail and freight train levels converge to the
same values (Figure 5(c)). Due to the structural damping, the high frequencies are strongly
attenuated by the soil. It turns out that vibrations induced by railway vehicles depend not only
on the axle load but also, in some cases, on the dynamic characteristics of the vehicle itself.
Vehicle System Dynamics 627

(a) Brussels tram German ICE Corail Thalys HST Freight train
2
Velocity (mm/s)

Velocity (mm/s)

Velocity (mm/s)

Velocity (mm/s)

Velocity (mm/s)
1 1 2 2

−1 −1 −2 −2
−2
0 1.5 0 1.5 0 1.5 0 1.5 0 1.5
Time (s) Time (s) Time (s) Time (s) Time (s)

Time history at 2 m from the track

(b)
Brussels tram
German ICE
Corail
Thalys HST
PPV (mm/s)

0
10 Freight train

−1
10
2 4 6 8 10 12 14
Distance from the track (m)
PPV
(c)
Brussels tram
0
10 German ICE
Corail
KBF,max (mm/s)

Thalys HST
Freight train

−1
10

2 4 6 8 10 12 14
Distance from the track (m)
KBF,max

Figure 5. Sensitivity of the PPV and KBF,max indicators to the vehicle type.

4.2. The vehicle speed

The speed will vary from 50 to 300 km/h for the reference vehicle, corresponding to a realistic
railway operating scale, from a tram speed to a HST speed (Table 3). The studied high speeds
are clearly out of the operating speed range of the reference vehicle but the purpose of this
analysis is to compare independently the vibration levels of the vehicle, the track and the soil.
An additional case will be studied where the vehicle speed is close to 600 km/h (167 m/s),
related to the particular case of the soil super-critical speed, with respect to Rayleigh wave
speed cR (equal, in this case, to 156 m/s). The Mach number, defined by
v0
MR = (8)
cR
is here useful in classifying the studied speed in sub- or super-Rayleigh cases. The
correspondence is illustrated in Table 3.
Next figures present the results for the vehicle–track subsystem. For vehicles (Figure 6),
the maximum onboard acceleration level increases with speed, in a proportional manner. The
628 G. Kouroussis et al.

Table 3. Various analysed speeds and the corresponding Mach number.

Speed 50 km/h 100 km/h 150 km/h 200 km/h 300 km/h 600 km/h

MR 0.09 0.18 0.27 0.36 0.53 1.07

(a) 3

2
Acceleration (m/s )
2

0 50 km/h
100 km/h
−1 150 km/h
200 km/h
−2 300 km/h
600 km/h
−3
0 0.2 0.4 0.6 0.8 1
Time (s)
Time analysis
(b) 0
10
50 km/h
100 km/h
−1
10 150 km/h
Amplitude (m/s2)

200 km/h
300 km/h
−2
10 600 km/h

−3
10

−4
10
0 20 40 60 80 100
Frequency (Hz)
Third-octave band analysis

Figure 6. Sensitivity of the car body acceleration to the vehicle speed.

rms acceleration has a typical value of


v
az,rms = 0.09, 0.14, 0.20, 0.24, 0.30, 0.62 m/s2

for the studied speeds, respectively. For the track, the rail deflection does not present variation
in term of maximum level but the shape clearly changes, approaching to an impulse when the
speed is very high (Figure 7(a)). This phenomenon appears also in the frequency content where
the speed extends the frequency band but diminishes its low frequency spectrum (Figure 7(b)).
Note that the track deflection symmetry vanishes with high speed (transition from quasi-static
deflection to dynamic deflection). The same observations can be made for the soil surface
forces.
A marked difference appears for the maximum level in the soil results (Figure 8) where
a linear evolution of the level appears with the vehicle speed, until 300 km/h (an additional
vehicle speed of 500 km/h is simulated and added in the figure). For the super-Rayleigh speed
(v0 = 600 km/h), one can observe a sub-Rayleigh-to-super-Rayleigh transition in the vibra-
tory levels and in the vibration shapes as well. This phenomenon was observed experimentally
by Madshus and Kaynia [32] where the Swedish HST X–2000 has run on a soft soil site and
Vehicle System Dynamics 629

−4
(a) x 10
2

Displacement [m]
−2
50 km/h
−4 100 km/h
150 km/h
−6 200 km/h
300 km/h
600 km/h
−8
0 0.1 0.2 0.3 0.4 0.5
Time [s]

(b) Time analysis


−4
10
50 km/h
100 km/h
150 km/h
Amplitude [m]

−6
10 200 km/h
300 km/h
600 km/h

−8
10

−10
10
0 20 40 60 80 100
Frequency [Hz]
Spectral analysis

Figure 7. Sensitivity of the rail deflection to the vehicle speed (deflection at mid-span of sleeper).

induced large ground deformations, when the vehicle speed was close to the wave velocities
in the soil.
Figure 9 presents the difference of soil motion between the reference and the super-Rayleigh
speed where the generated wavefronts are completely different. This phenomenon is similar to
the aeroacoustic sub-critical phenomenon and its supersonic bang. Nevertheless the running
speed is super-critical in terms of Rayleigh waves but not for body waves: it stays in the
subsonic case (v0 < cS , with cS the shear waves velocity), and at the limit of the transonic
case (cS < v0 < cP , with cP being the compression wave velocity) as the shear wave velocity
cS = 169 m/s is larger than the vehicle speed. It can also be emphasised that the wavefront is
not straight but curved, essentially due to soil structural attenuation. The wavefront angle, the
so-called Mach angle, can be calculated by

1 cR
αM = arcsin = arcsin (v0 ≥ cR ), (9)
MR v0

and is equal to 69.2◦ in this case, in agreement with the predicted results.
The present analysis covers a large speed range, which is not the case in practice (e.g.
commercial speed of trams is usually around 50–100 km/h). As a complement to the Figure 8,
the sensibility of the vehicle speed is again studied, according to each vehicle type previously
studied in each specific realistic range of speed. Figure 10 displays the PPV level at 5 and
10 m from the track. The increase with the speed is observed for all the vehicles, except for
the Thalys HST results in Figure 10(a) where the passing of v0 = 100 to 150 km/h induces
630 G. Kouroussis et al.

(a) 50 km/h 100 km/h 150 km/h 300 km/h


200 km/h 600 km/h
Velocity (mm/s)

Velocity (mm/s)
Velocity (mm/s)

Velocity (mm/s)
Velocity (mm/s)

Velocity (mm/s)
1 2 15
1 1 1

−1 −1
−1 −1 −2 −15
0 3 0 1.5 0 1 0 0.75 0 0.5 0 0.25
Time (s) Time (s) Time (s) Time (s) Time (s) Time (s)

Time history at 2 m from the track


(b) 20

15
PPV (mm/s)

2m

10
increasing distance

5 5m

10 m
15 m
0
0 100 200 300 400 500 600
Vehicle speed (km/h)
PPV
(c) 5

4
(mm/s)

2m
3
F,max

2 increasing distance
KB

1 5m

10 m
15 m
0
0 100 200 300 400 500 600
Vehicle speed (km/h)
KB F,max

Figure 8. Sensitivity of the PPV and KBF,max indicators to the vehicle speed, for various distances from the track.

Figure 9. Comparison of the ground waves propagation (vertical component vz ) at sub- and super-Rayleigh regime.

a decrease of vibratory level (at 150 km/h, the sleepers excitation mechanism frequency fs is
equal to 57.9 Hz and is more attenuated by the soil damping effect at small distances from the
track). Excluding this local case, these results show that the vehicle speed strongly contributes
to the vibratory level and with the same importance for each studied vehicle.
Vehicle System Dynamics 631

(a) PPV (mm/s) (b) PPV (mm/s)


0 1 2 3 4 0 0.5 1 1.5 2

Freight train

Thalys HST Thalys HST

Corail Corail
50 km/h 50 km/h
German ICE German ICE
100 km/h 100 km/h
150 km/h 150 km/h
Brussels tram 200 km/h Brussels tram 200 km/h
300 km/h 300 km/h

At 5 m from the track At 10 m from the track

Figure 10. Influence of the peak particle velocity to the vehicle speed and the vehicle type.

5. Influence of the track parameters

The flexibility of the track is often referred to as the most significant parameter for ground
vibrations. The track supports the vehicle, takes up the axle loads and dispatches these forces
through the ballast. The maximum track deflection and its shape are expected to determine
the level of vibrations. In our model, three track modes can be identified with the reference
values of Table 1:

• a first resonance frequency where the rail and sleepers vibrate in phase (at 88 Hz for the
reference values) that we will call T 1 mode,
• a second mode (T 2 mode) where rail and sleepers vibrate out of phase (at 341 Hz),
• the third mode, called pinned-pinned resonance (P–P mode), where the rail vibrates with a
wavelength equal to two sleeper bays (at 875 Hz).

These modes can influence thereafter the rail deflection and its spectrum, as illustrated in
Figure 11 where the three modes are observable.
For the vehicle, the conclusion is not obvious. A preliminary analysis is made here in order to
emphasise the whole effect of the track flexibility on the vehicle dynamics. Figure 12 compares
the vehicle motion considering a flexible track (reference values of Table 1) or a rigid track.
The results, both for car body acceleration and for bogie acceleration, are very close, showing
that considering a flexible track does not modify the vehicle dynamics, in the presented case. It

−7
10
Above a sleeper
At mid−span of sleeper
Receptance (m/N)

−8
10 P−P mode

−9
10 T1 mode T2 mode

−10
10
0 200 400 600 800 1000
Frequency (Hz)

Figure 11. Track receptance for the reference case.


632 G. Kouroussis et al.

(a) 0.5
Rigid track
Flexible track

Acceleration (m/s2)
0

−0.5
0 0.2 0.4 0.6 0.8 1
Time (s)
Car body acceleration
(b)
4
Rigid track
Flexible track
Acceleration (m/s2)

−2

−4
0 0.2 0.4 0.6 0.8 1
Time (s)
Bogie frame acceleration

Figure 12. Impact of the track flexibility to the vehicle motion.

certainly explains that the simulation of a complete process taking into account the track and the
soil is not yet usually performed by the vehicle constructors, who reasonably consider the track
and, therefore, the soil as rigid in the studied cases. It is however important to analyse the effect
of each component on the ground vibrations. The influence of the rail type, the railpad and bal-
last stiffnesses, and the sleepers (spacing and masses) is summarised in the following sections.

5.1. Track flexibility

The flexibility (per unit of length) of the track derives from the rail, the railpad and the ballast
flexibilities, and the sleepers spacing. For the sensitivity analysis, the following parameters
are investigated around their usual range, also obtained from literature:

• Various rail types, ordered by the mass per unit of length mr and the flexural stiffness Er Ir
are analysed (Table 4). In all cases, the rail is assumed to be built of standard steel (Young’s
modulus: 210 GPa and density: 7850 kg/m3 ).
• As the railpad role is to absorb a part of rail vibrations and to allow the travel of rail without
damaging the sleeper, various railpads have been studied, essentially on the basis of their
stiffness and their damping to a lesser extent. Table 5 presents various simulation cases, from
very soft railpads to stiff railpads. The case of kp = 1 GN/m corresponds to the situation of
a railpad that has lost its elastic behaviour.
• The sleeper spacing is limited: it is usually of 0.60 m but, for the low-loaded tracks or tracks
consisting of continuous welded rail, it grows up to 0.72 m. Two cases will be investigated,
as described in Table 6.
Vehicle System Dynamics 633

Table 4. Data for the various studied rails (profile and mechanical properties).

Moment of inertia Section Flexural stiffness Mass mr


Rail type Ir (cm4 ) Ar (cm2 ) Er Ir (MNm2 ) (kg/m)

SMR 29a 624 37.94 1.31 29.8


UIC 50 1940 63.92 4.07 50.2
EB50T 1988 63.82 4.17 50.0
UIC 54 2127 69.34 4.47 54.4
BS 113 Ab 2349 71.83 4.93 56.4
UIC 60 3055 76.87 6.42 60.3
c
NP4aM 3535 79.45 7.42 62.4
a
British rail (from [10]).
b
German rail (very light, from [20]).
c
Grooved rail (used in embedded track structures, from a Belgian site).

Table 5. Various railpad data used in the simulations.

Case 1 (reference) Case 2 (very soft) Case 3 Case 4 Case 5 (stiff)

kp 90 MN/m 6 MN/m 60 MN/m 180 MN/m 1 GN/m


dp 30 kN s/m 5 kN s/m 52 kN s/m 28 kN s/m 63 kN s/m
fT 1 88 Hz 57 Hz 86 Hz 89 Hz 89 Hz
ξT 1 41% 18% 40% 42% 44%
fT 2 341 Hz 135 Hz 285 Hz 479 Hz 1124 Hz
ξT 2 45% 59% 85% 30% 25%

Table 6. Investigated sleepers spacing.

Case 1 (reference) Case 2


Spacing L 0.72 m 0.60 m

fT 1 88 Hz 92 Hz
ξT 1 41% 43%
fT 2 341 Hz 395 Hz
ξT 2 45% 45%
fP−P 875 Hz 1260 Hz

Table 7. Various ballast data used in the simulations.

Parameters Case 1 (nominal) Case 2 (soft) Case 3 Case 4 Case 5 (stiff)

kb 32 MN/m 12 MN/m 40 MN/m 120 MN/m 190 MN/m


db 52 kN s/m 12 kN s/m 40 kN s/m 40 kN s/m 100 kN s/m
fT 1 88 Hz 58 Hz 94 Hz 128 Hz 143 Hz
ξT 1 41% 18% 31% 20% 26%
fT 2 341 Hz 341 Hz 346 Hz 361 Hz 361 Hz
ξT 2 45% 38% 43% 42% 51%

• The railpad, ballast stiffness and damping will be studied, considering five cases from soft
to stiff ballast (Table 7).
Tables 5–7 also present the track modes. Ballast and railpad stiffnesses have the greatest
influence on the T 1 and T 2 track modes, due to their large range, but in a different way:
eigenfrequencies of T 1 and T 2 modes change in function of their stiffnesses but each element
affects essentially one mode (T 2 mode for the railpad and T 1 mode for the ballast). The
634 G. Kouroussis et al.

−4
(a) x 10
2

Displacement (m)
UIC 50
−2
UIC 54
UIC 60
−4 EB50T
BS 113 A
−6 SMR 29
NP4aM
−8
0 0.1 0.2 0.3 0.4 0.5
Time (s)
To the rail type
(b) −4
x 10
2

0
Displacement (m)

−2

−4
case 1 (reference)
−6 case 2 (very soft)
case 3
−8 case 4
case 5 (stiff)
−10
0 0.1 0.2 0.3 0.4 0.5
Time (s)
To the ballast

Figure 13. Sensitivity of the rail deflection (time analysis at mid-span of sleeper).

sleeper spacing not only plays a role on these track modes but also changes the track P–P
mode. For the rail types, small variations in the T 1 and T 2 modes can be observed. However,
the P–P modes present important variation from 1048 Hz for the heaviest to 633 Hz for the
light rail.
From all the performed simulations, focusing on these parameters, some general observa-
tions can be made:

• For the track (Figure 13(a) and (b)), it is evident that, for any component, the higher the
flexibility, the larger the rail deflection.
• The ground vibrations diminish with a flexible track (Figure 14). A compromise is also
necessary between rail deflection limitation and ground vibrations. As the ground vibrations
are essentially dominated by the low frequencies up to 100 Hz, T 2 and P–P modes do not
affect them.

For the sleepers, the possibility of spacing variation being limited, the two cases are comparable
(see Figure 17 in the next subsection).
Anti-vibration solutions can be studied from this stage: an interesting solution, proposed by
various companies specialised in vibration isolation systems, is to stagger standard railpads
with soft pads in order to force a compromise between soil vibrations attenuation and rail
deflection. Figure 15 gives typical results, considering nominal type 1 (90 MN/m) alternatively
with type 2 railpads (6 MN/m). Compared with a track equipped with soft pads (type 2), the
rail deflection is effectively diminished with the use of staggered railpads. At the ground level
(Figure 16), the beneficial effect is also emphasised at few meters from the track (here 5 m).
Vehicle System Dynamics 635

(a) 100
case 1 (reference)
case 2 (soft pad)
case 3
case 4

PPV (mm/s)
case 5 (stiff pad)

−1
10

2 4 6 8 10 12 14
Distance from the track (m)
PPV
(b)
case 1 (reference)
case 2 (soft pad)
case 3
KBF,max (mm/s)

case 4
−1
10 case 5 (stiff pad)

−2
10
2 4 6 8 10 12 14
Distance from the track (m)
KB F, max

Figure 14. Sensitivity of the soil vibrations indicators to the railpad flexibility.

x 10-3

0
Vertical deflection (m)

-0.5 deflexion maximale a 5.6 10–4

deflexion maximale a 7.9 10–4


nominal railpads
-1
soft railpads
staggered railpads

-1.5 deflexion maximale a 1.56 10–3

0.72 m

Distance from the track

Figure 15. Rail deflection in the case of staggered railpads.

A significant reduction of ground vibrations (10–15% compared to the nominal railpad case)
is obtained with staggered railpads, without notably increasing the rail deflection (40%). With
soft pads, the maximum rail deflection is nearly triple as the one obtained with nominal railpads
for a gain of 40% on the ground vibration level.
636 G. Kouroussis et al.

(a) 0
10
nominal railpads
soft railpads
staggered railpads

PPV (mm/s)

−1
10

2 4 6 8 10 12 14
Distance from the track (m)

PPV
(b)
nominal railpads
soft railpads
staggered railpads
KBF,max (mm/s)

−1
10

−2
10
2 4 6 8 10 12 14
Distance from the track (m)

KB F, max

Figure 16. Benefit of the staggered railpads on the PPV and KBF,max indicators.

5.2. Rail and sleeper mass

The rail and the sleeper masses define the second set of parameters set that has been investi-
gated. The rail profile inevitably defines its mass (Table 4). Sleepers have for a long time been
made of wood, but concrete is now widely used (in mono-bloc or bi-bloc configuration). Steel
sleepers are currently used as well, although far less from wood or concrete ties due to their
low damping. Table 8 summarises the characteristics of the sleepers considered in this study.
No variation is observed with the sleeper mass, in rail deflection as well as in ground
vibrations (Figure 17). The same observation can be made for the rail mass. Neither the
track mass nor the rail mass do affect the maximum track deflection and therefore the ground
vibrations. Although the mass and the flexural stiffness of the rail are bounded, it can be
observed that the rail mass does not influence the track deflection (Figure 13(a)) but only the
flexural stiffness intervenes. This can be explained by the low vehicle speed considered in this
study, compared to the critical velocity of the track

4Er Ir
vcrit = 4 2 2 . (10)
(ρr Ar + m /L )(L/kb + L/kp + L/ks )
2 2

The track mass contributes to the dynamic deflection and on its effect on the ground, only
when the vehicle speed is close to the critical velocity vcrit of the track. For typical values of
the track parameters, vcrit = 377 m/s (235 m/s in the most unfavourable case), clearly out of
the commercial speed range.
Vehicle System Dynamics 637

(a) 0
10
case 1 (m = 90.84 kg ; L = 0.72 m)
case 2 (m = 52 kg ; L = 0.72 m)
case 3 (m = 200 kg ; L = 0.72 m)
case 4 (m = 290 kg ; L = 0.72 m)

PPV (mm/s)
case 5 (m = 90.84 kg ; L = 0.60 m)

−1
10

2 4 6 8 10 12 14
Distance from the track (m)
PPV
(b)
case 1 (m = 90.84 kg ; L = 0.72 m)
case 2 (m = 52 kg ; L = 0.72 m)
case 3 (m = 200 kg ; L = 0.72 m)
KBF,max (mm/s)

case 4 (m = 290 kg ; L = 0.72 m)


−1
10 case 5 (m = 90.84 kg ; L = 0.60 m)

−2
10
2 4 6 8 10 12 14
Distance from the track (m)
KB F,max

Figure 17. Sensitivity of the PPV and KBF,max indicators to the sleepers (mass and spacing).

Table 8. Investigated sleepers mass.

Case 1 (reference) Case 2 (of wood) Case 3 (concrete bi-bloc) Case 4 (concrete mono-bloc)

Mass m (kg) 90.84 52 200 290


fT 1 (Hz) 88 100 68 59
ξT 1 % 41 45 33 29
fT 2 (Hz) 341 395 296 282
ξT 2 % 45 60 34 31

5.3. Track irregularity

The rail irregularity could be analysed but the same conclusions will be brought as the ones
published in a recent article [12]: the irregularity vertical profile plays an important role
(dynamic deflection) in the transmitted ground vibrations, amplifying the axle load effect
(quasi-static deflection). Moreover, the vehicle speed strongly influences the effect of track
irregularity on the ground vibrations level and the vehicle dynamics intervenes in the forefront.

6. Summary and conclusion

A parametric study has been presented with global emphasis on the vehicle and track parame-
ters. Table 9 encapsulates the dependence of three indicators (the vehicle vertical acceleration
638 G. Kouroussis et al.

Table 9. Sensibility analysis summary.

Indicators
Sub-system Parameters Range azv uzrail vzsoil

Train Axle load N 50–200 kN   


Vehicle speed v0 50–600 km/h  ∼ 
Track Rail flexural stiffness Er Ir 1.3–7.4 MNm2 –  
Rail mass ρr Ar 30–62 kg/m – – –
Railpad stiffness kp 6–103 MN/m –  
Sleepers mass m 50–290 kg/m – – –
Sleepers spacing L 0.6–0.72 m – ↑ ↑
Ballast stiffness kb 12–190 MN/m –  

Notes:  , high increasing when the parameter value increases; ↑ , light increasing when the parameter value increases;

, poor sensibility of the parameter; ∼ , poor sensibility of the parameter but the signal shape changes; –, no sensibility
of the parameter; ↓ , light decreasing when the parameter value increases;  , high decreasing when the parameter
value increases.

azv , the dynamic rail deflection uzrail , and the soil vibrations velocity vzsoil ) on the main vehi-
cle and track parameters. Some of these results agree with observations found in literature,
especially for the track parameters influence.
For the vehicle parameters, this analysis demonstrates that the axle load is not the only
vehicle parameter which influences the ground vibrations since the vehicle dynamics modifies
the forces transmitted to the track and therefore to the ground. The vehicle speed also has a
strong influence on the result: the special effect of super-critical speed is portrayed.
The track parameters (rail type, railpad, ballast, sleepers) also have an effect on the ground
vibrations but their value ranges are limited for practical reasons. The track flexibility is
the primary means to control the ground vibrations when the variation range only considers
values encountered in practice whereas the track mass (rail and sleepers) does not influence
for commercial vehicle speeds. Anti-vibration solutions, as the staggered railpads, were also
studied in this paper, showing their benefit without loss of sight of their influence on the other
sub-systems (rail deflection limit).

References

[1] V.V. Krylov, Effect of track properties on ground vibrations generated by high-speed trains, Acustica-acta
Acustica 84 (1998), pp. 78–90.
[2] B. Picoux and D.L. Houédec, Propagation d’ondes à la surface d’un massif soumis à un effort latéral et vertical
du à un trafic ferroviaire, 15ème Congrès Français de Mécanique Association Française de Mécanique, Nancy,
France, 3–7 September 2001, pp. 364–369.
[3] G. Kouroussis, O. Verlinden, and C. Conti, On the interest of integrating vehicle dynamics for the ground
propagation of vibrations: The case of urban railway traffic, Veh. Syst. Dyn. 48 (2010), pp. 1553–1571.
[4] Y.B. Yang, H.H. Hung, and D.W. Chang, Train-induced wave propagation in layered soils using finite/infinite
element simulation, Soil Dyn. Earthquake Eng. 23 (2003), pp. 263–278.
[5] A.V. Metrikine, A.V. Vostrukhov, and A.C.W.M. Vrouwenvelder, Drag experienced by a high-speed train due
to excitation of ground vibrations, Int. J. Solid Struct. 38 (2001), pp. 8851–8868.
[6] A.V. Vostroukhov and A.V. Metrikine, Periodically supported beam on a visco-elastic layer as a model for
dynamic analysis of a high-speed railway track, Int. J. Solids Struct. 40 (2003), pp. 5723–5752.
[7] J. Wang and X. Zeng, Numerical simulations of vibration attenuation of high-speed train foundations with
varied trackbed underlayment materials, J. Vibr. Control 10 (2004), pp. 1123–1136.
[8] C. Vogiatzis and S. Chaikali, Vibration and ground borne noise criteria and mitigation measures at Athens
tramway, 11th International Congress on Sound and Vibration, St. Petersburg, Russia, 5–8 July 2004.
[9] J. Wang, X. Zeng, and D.A. Gasparini, Dynamic response of high-speed rail foundations using linear hysteretic
damping and frequency domain substructuring, Soil Dyn. Earthquake Eng. 28 (2008), pp. 258–276.
[10] S.L. Grassie, R.W. Gregory, D. Harrison, and K.L. Johnson, The dynamic response of railway track to high
frequency vertical excitation, J. Mech. Eng. Sci. 24 (1982), pp. 77–90.
Vehicle System Dynamics 639

[11] G. Kouroussis, O. Verlinden, and C. Conti, Free field vibrations caused by high-speed lines: Measurement and
time domain simulation, Soil Dyn. Earthquake Eng. 31 (2011), pp. 692–707.
[12] G. Kouroussis, O. Verlinden, and C. Conti, A two–step time simulation of ground vibrations induced by the
railway traffic, Proc. IMechE, Part C: J. Mech. Eng. Sci. 2011. doi:10.1177/0954406211414483.
[13] G. Kouroussis, O. Verlinden, and C. Conti, Contribution of vehicle/track dynamics to the ground vibrations
induced by the Brussels tramway, ISMA2010 International Conference on Noise and Vibration Engineering,
P. Sas, ed., Leuven, Belgium, 20–22 September 2010, pp. 3489–3502.
[14] V.K. Garg and R.V. Dukkipati, Dynamics of Railway Vehicle Systems, Academic Press, Toronto, Canada, 1984.
[15] G. Kouroussis, O. Verlinden, and C. Conti, Ground propagation of vibrations from railway vehicles using a
finite/infinite-element model of the soil, Proc. IMechE, Part F: J. Rail Rapid Transit 223 (2009), pp. 405–413.
[16] G. Kouroussis, O. Verlinden, and C. Conti, On the infinite element modelling of soil for the prediction of railway
vibrations, Third International Congress Design and Modelling of Mechanical Systems, Hammamet, Tunisia,
16–18 March 2009.
[17] J.P. Wolf, Foundation Vibration Analysis Using Simple Physical Models, Prentice-Hall, NJ, 1994.
[18] T.H. Young and C.Y. Li, Vertical vibration analysis of vehicle/imperfect track systems, Veh. Syst. Dyn. 40
(2003), pp. 329–349.
[19] W.M. Zhai and H. True, Vehicle-track dynamics on a ramp and on the bridge: Simulation and measurements,
Veh. Syst. Dyn. 33 (Suppl.) (1999), pp. 604–615.
[20] C. Esveld Modern Railway Track, 2nd MRT Production, Zaltbommel, The Netherlands, 2001.
[21] J. Kisilowski and K. Knothe (eds.), Advanced Railway Vehicle System Dynamics, Wydawnictwa Naukowo–
Techniczne, Warsaw, Poland, 1991.
[22] C. Andersson and J. Oscarsson, Dynamic train/track interaction including state-dependent track properties and
flexible vehicle components, Veh. Syst. Dyn. 33 (Suppl.) (1999), pp. 47–58.
[23] C. Dochy, G. Verheulpen, and P. Vanhonacker, Vibrations dans l’environnement lors du passage du tram dans
le tunnel de Laeken: Situation actuelle (2004) par rapport à la situation après installation des voies (1994),
Tunnels et Ouvrages Souterrains 191 (2005), pp. 271–275.
[24] G. Degrande, Free field vibration measurements during the passage of a Thalys high-speed train, Tech. report,
Katholieke Universiteit te Leuven, 2000.
[25] P. Chatterjee, G. Degrande, S. Jacobs, J. Charlier, P. Bouvet, and D. Brassenx, Experimental results of free
field and structural vibrations due to underground railway traffic, 10th International Congress on Sound and
Vibration, Stockholm, Sweden, 7–10 July 2003.
[26] J. Kogut, G. Lombaert, S. François, G. Degrande, W. Haegeman, and L. Karl, High speed train induced vibra-
tions: In situ measurements and numerical modelling, 10th International Congress on Sound and Vibration,
Stockholm, Sweden, 7–10 July 2003.
[27] K. Knothe and Y. Wu, Receptance behaviour of railway track and subgrade, Arch. Appl. Mech. 68 (1998), pp.
457–470.
[28] J. Kogut, G. Degrande, and W. Haegeman, Free field vibrations due to the passage of an IC train and a Thalys
HST on the high speed track L2 Brussels–Köln, 6th National Congress on Theoretical and Applied Mechanic,
National Committee for Theoretical and Applied Mechanics, Ghent, Belgium, 26–27 May 2003.
[29] B. de Saedeleer, S. Bilon, S. Datoussaïd, and C. Conti, Vibrations induced by urban railway vehicles — Modeling
of the Vehicle/Track system, Proceeding of the “transport and environment” study days of the BSMEE, Mons,
Belgium, 28–29 May 1998.
[30] Deutsches Institut für Normung, DIN 4150-3: Structural vibrations — Part 3: Effects of vibration on structures
(1999), 12 pp.
[31] Deutsches Institut für Normung, DIN 4150-2: Structural vibrations — Part 2: Human exposure to vibration in
buildings (1999), 21 pp.
[32] C. Madshus and A.M. Kaynia, High-speed railway lines on soft ground: Dynamic behaviour at critical train
speed, J. Sound Vibr. 231 (2000), pp. 689–701.

You might also like