Vous êtes sur la page 1sur 16

ARTICLE IN PRESS

Soil Dynamics and Earthquake Engineering 30 (2010) 8297

Contents lists available at ScienceDirect

Soil Dynamics and Earthquake Engineering


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

Numerical modelling of vibrations from a Thalys high speed train in the Groene Hart tunnel
S. Gupta, H. Van den Berghe, G. Lombaert, G. Degrande
Department of Civil Engineering, K.U.Leuven, Kasteelpark Arenberg 40, B-3001 Leuven, Belgium

a r t i c l e in f o
Article history: Received 17 November 2008 Received in revised form 18 September 2009 Accepted 23 September 2009 Keywords: High speed train induced vibrations Groene Hart tunnel Coupled periodic nite elementboundary element model Dynamic vehicletrack interaction Dynamic tracktunnelsoil interaction

a b s t r a c t
This paper presents a numerical study of vibrations due to a Thalys high speed train in the Groene Hart tunnel, which is part of the high speed link South between Amsterdam and Antwerp and the worlds largest bored tunnel. A coupled periodic nite elementboundary element model is used to predict the free eld response due to the passage of a Thalys high speed train in the Groene Hart tunnel. A subdomain formulation is used, where the track and the tunnel are modelled using a nite element method, while the soil is modelled as a layered half space using a boundary element method. The tunnel and the soil are assumed to be invariant in the longitudinal direction, but modelled as a periodic structure using the Floquet transformation. A general analytical formulation to compute the response of three-dimensional periodic media excited by moving loads is adopted. The Groene Hart area is marshy and completely saturated. The top soil consists of layers of peat and clay with a very low density and shear wave velocity. The numerical model allows to understand the effect of these soft layers on vibration levels, resulting in an amplication of the horizontal response and a large contribution of the quasi-static forces at high train speeds. Vibration levels are assessed using the Dutch SBR guideline. It is concluded that the operation of high speed railway trafc in the Groene Hart tunnel is not expected to cause serious vibration problems. & 2009 Elsevier Ltd. All rights reserved.

1. Introduction The Groene Hart is a relatively thinly populated area in the Netherlands enclosed by large cities of the Randstad, a very densely populated area accommodating nearly half of the Dutch population. It is a rural area that offers many green spaces with agriculture, nature and recreation as primary activities. It is threatened by expansion of the big cities around it and other land use which may destroy its unique landscape. With further expansion and advances in the Randstad, the concern over the preservation of the Groene Hart region has increased. The high speed link South (HSL-Zuid) between Amsterdam and Antwerp aims to improve connectivity between important economic centers of Europe. It crosses the provinces of Northand South-Holland and passes through the Groene Hart area. It was decided to bore a tunnel to conserve the grasslands of the Groene Hart and to leave the landscape undisturbed. A 7 km long tunnel was therefore constructed between Leiderdorp and Hazerswoude-Dorp. The tunnel has been bored using the Aurora tunnel boring machine, that has been manufactured for the purpose of boring a tunnel with an external diameter of 14.5 m.

Corresponding author. Tel.: + 32 16 32 16 67; fax: +32 16 32 19 88.

E-mail address: geert.degrande@bwk.kuleuven.be (G. Degrande). 0267-7261/$ - see front matter & 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.soildyn.2009.09.004

This makes the Groene Hart tunnel the largest bored tunnel in the world. The tunnel consists of a single tube with a length of over 7 km. Including the access ramps, the total length is about 8.5 km. One of the requirements for the design of the high speed link including the Groene Hart tunnel has been to comply with the Dutch SBR (Building Research Foundation) guideline for measuring and assessing vibrations [1]. In the present paper, the problem of high speed train (HST) induced vibrations in the Groene Hart area is examined by assessing computed free eld vibration levels using the SBR guideline. The environmental impact of vibrations due to railways has recently gained a lot of importance on account of increasing public sensitivity to noise and vibration and stricter legislation to achieve sustainable development. Great efforts have been made in the last two decades to develop numerical models for prediction of ground-borne vibrations from railways. As far as deterministic modelling is concerned, analytical and various numerical methods such as nite element, boundary element and integral transform methods have been employed to model the coupled tracktunnelsoil system. The problem of vibrations from high speed railway trafc has been studied by a large number of researchers. Most studies have focused on solving the traintracksoil interaction problem, which allows to compute the vibrations from surface railways. Sheng et al. [2,3] have modelled the tracksoil interaction by

ARTICLE IN PRESS
S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297 83

means of a model with an innite layered beam on top of a layered half space and have coupled a train model to this track model [4,5]. Auersch [6] has coupled a nite element model for a nite part of the track to a boundary element model for the soil. This model is used for the calculation of the track compliance in the solution of the vehicletrack interaction problem. Metrikine et al. [7] have studied the stability of a moving train bogie, modelled as a two degree of freedom system, which is coupled to a beam of innite length for the track and a homogeneous half space model for the soil, following an approach proposed by Metrikine and Popp [8]. Lombaert and Degrande [9] and Lombaert et al. [10] apply this methodology in a boundary element formulation to predict vibrations induced by railway trafc [10,11]. The model has been validated by means of experiments that have been performed at the occasion of the homologation tests of the new [10,11]. Most HST track on the line L2 between Brussels and Koln of these advanced numerical models assume the geometry to be invariant in the longitudinal direction and use the Fourier transform to efciently formulate the problem in the frequency wavenumber domain. Similar models with varying degree of sophistication have also been developed for predicting vibrations due to underground railway trafc. Recently developed models such as the semianalytical pipe-in-pipe model [1214], the model based on coupling of the nite element method and integral transform methods [15] and the coupled nite elementboundary element models [1619] account for three-dimensional dynamic interaction between the track, the tunnel and the soil. These models exploit the invariance or periodicity of the system in the longitudinal direction using the Fourier or the Floquet transform to formulate the problem in the frequencywavenumber domain. Sheng et al. [20] have referred to this approach as the wavenumber nite elementboundary element method. They have used this methodology to compare the transfer functions of two alternative tunnel designs, a single double-track tunnel and a pair of single-track tunnels. Gupta et al. [21] have used the coupled periodic nite elementboundary element approach for predicting vibrations due to subway trafc in Beijing. The present paper demonstrates the applicability of this model in predicting the vibrations from high speed trains running in tunnels. The paper is organized in the following manner. Section 2 reviews the characteristics of the soil, the tunnel, the track and the train. Section 3 outlines the modelling of the Groene Hart tunnel using the coupled periodic nite elementboundary element model and explains the computation of the transfer functions and the axle loads. Section 4 presents the response of the tunnel and

the free eld due to the passage of a Thalys HST in the Groene Hart tunnel. Finally, in Section 5, the predicted vibration levels are compared to the SBR guideline to evaluate human exposure to vibrations.

2. System characteristics 2.1. Characteristics of the soil The Groene Hart is a wet area with lot of marshes, polders and reclaimed lakes. The soil is essentially peaty and completely saturated. The geotechnical prole along the Groene Hart tunnel is shown in Fig. 1a. The top layers mainly consist of peat and clay, while the underlying half space is sand. A shallow peat layer of thickness 3.7 m is present on top of the clay layer up to a depth of about 10 m. Under the clay layer is a thin layer of peat. Under this second peat layer are the medium and coarse sand layers. Between PK 24 and PK 29 (project kilometers), the tunnel is mainly embedded in a layer of sand. The top part of the tunnel lies in a medium sand, while the bottom part of the tunnel lies in the coarse sand. The shear wave velocity of the soil has been determined through seismic cone penetration tests (SCPT) [22]. These tests have been performed at three locations S1, S2 and S3 along the tunnel (Fig. 1a). The soil density is obtained from classical soil mechanics tests [22]. For the numerical calculations, the shear wave velocity prole is determined by averaging the shear wave velocity obtained from the SCPT tests on location S3 (Fig. 1b). The depth of the tunnel at this location is 26.25 m with respect to the free surface. Four layers on top of a half space are considered (Fig. 1b), whose properties are summarized in Table 1. The soil is modelled as a horizontally layered half space, consisting of saturated soil layers, that are modelled using Biots theory as equivalent mono-phasic media or frozen mixtures with a shear
Table 1 Dynamic soil characteristics of the Groene Hart area at location S3. Layer 1 2 3 4 5 Thickness (m) 3.7 7.0 8.3 9.3 1 Cs0 (m/s) 50 75 180 240 260 Cp0 (m/s) 1761 1719 1685 1715 1726

r (kg=m3 )
1107 1500 1970 1970 1970

n0 ()
0.4996 0.4990 0.4942 0.4900 0.4884

b ()
0.025 0.025 0.025 0.025 0.025

0 1.69

PEAT

10 Depth NAP [m]

CLAY

20
MEDIUM SAND

27.94 30
COARSE SAND

40 0 100 200 300 400 500 Cs [m/s]

Fig. 1. (a) Geotechnical prole along the Groene Hart tunnel and (b) shear wave velocity prole obtained from the SCPT tests on location S3 (gray line) and used for the computations (black line).

ARTICLE IN PRESS
84 S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297

wave velocity Cs0, a longitudinal wave velocity Cp0 (and a corresponding high value of Poissons ratio n0 ) and a mixture density r [2325]. As no experimental data about the hysteretic material damping ratio bs and bp in shear and volumetric deformation in each layer are available, a uniform value b bs bp 0:025 is assumed for all layers. 2.2. Characteristics of the tunnel The Groene Hart tunnel has been bored using a slurry shield tunneling method at a depth of about 26 m. Per section, ten prefabricated concrete sections were placed to form the tunnel lining and joined with bolts in the circumferential and longitudinal direction. On the tunnel oor a technical gallery was set up for easy execution of the project and maintenance work. The space between the gallery and the tunnel is lled with a mixture of sand and cement for stabilization. This is covered by a concrete oor with a thickness of 0.27 m, on which the track is installed. The tunnel hosts two tracks, which are separated by a partition wall of thickness 0.45 m. This partition wall has been provided in view of safety and easy evacuation. Fig. 2 shows the cross section of the tunnel. The tunnel has an internal radius ri 6:65 m and an outer radius re 7:25 m. Additional dimensions are shown on Fig. 2. The concrete tunnel lining has a Youngs modulus Et 35 000 MPa, a Poissons ratio nt 0:20, a density rt 2500 kg=m3 and a hysteretic material damping ratio bt 0:02. The mixture of the sand and cement has a Youngs modulus Esc 24 000 MPa, a Poissons ratio nsc 0:20 and a density rsc 1850 kg=m3. 2.3. Characteristics of the track Two tracks are installed in the tunnel, according to a track system that is commonly used on high speed lines in Germany and the Netherlands. It is a ballastless track that uses concrete biblock sleepers with a lattice truss, which makes it easier to install and ensures an accurate gauge. The tracks are installed at 3.5 m from the center of the tunnel. A geotextile layer is placed between the track and the substructure. The layer is xed on the concrete oor of the tunnel with dowels. Its main purpose is to increase the friction between the track and the tunnel bed to avoid any possible separation. The stiffness of the geotextile layer is Kg 2:23 109 N=m3 , while the damping in the layer is Cg 2:31 104 N s=m3 . The reinforced concrete slab is 2.8 m wide and 0.24 m thick. Prefabricated

concrete bi-bloc sleepers with dimensions 0:645 m 0:292 m 0:016 m are cast into the slab. Elastic pads, base plates, and rail pads discretely support the UIC 60 rails on sleepers at an interval d 0:65 m. The base plate pads have a stiffness kbp 30 106 N=m and a damping cbp 8 103 N s=m. These soft pads provide some degree of vibration isolation. The mass of the base plates is 12 kg. The rail pads are stiff and have a stiffness krp 1 109 N=m and a damping crp 7:77 104 N s=m [26]. The track unevenness has been measured with a NSTO measurement train on the part of the high speed line South that includes the Groene Hart tunnel. Fig. 3 compares the one-third octave band spectrum of the measured unevenness with the track classes 1 and 6 of the Federal Railroad Administration (FRA). FRA track classes have been derived on the basis of extensive measurements on the US railway network; six classes of track quality are dened, class 1 being the poorest and class 6 the best [27]. Superimposed on this graph are the TSI + [28] and ISO 3095:2005 [29] limits. The unevenness measured on the present HST track is higher than FRA track class 6 at shorter wavelengths or higher frequencies, which can cause greater wheel-rail rolling noise. For a train speed of v 300 km=h, wavelengths between 83 cm and 83 m are important for predicting ground-borne vibrations in the frequency range between 1 and 100 Hz.

2.4. Characteristics of the train The high speed train is an articulated Thalys train with two power cars and eight carriages; the total length of the train is equal to 200.18 m. The power cars are supported by two bogies and have four axles. The carriages next to the power cars share one bogie with the neighboring carriage, while the six other carriages share both bogies with neighboring carriages. The total number of bogies equals 13 and, consequently, the number of axles on the train is 26. The carriage length Lt , the distance Lb between bogies, the axle distance La , the total axle mass Mt , the sprung axle mass Ms and the unsprung axle mass Mu of all carriages are summarized in Table 2.

60 Unevenness [dB ref 1 m] 40 20 ISO limit 0 TSI limit 20 2048 512 128 32 8 Wavelength [cm] 2

Fig. 3. One-third octave band spectra of the measured unevenness along the high speed line South (solid black line) and FRA track classes 1 (dashed gray line) and 6 (solid gray line). Superimposed on the graph are the TSI+ (dashed black line) and ISO 3095:2005 limits (dash-dotted black line).

Table 2 Geometrical and mass characteristics of the Thalys HST. Lt (m) Power cars Side carriages Central carriages 22.15 21.84 18.70 Lb (m) 14.00 18.70 18.70 La (m) 3.00 3.00 3.00 Mt (kg) 17000 17000 17000 Ms (kg) 14937 14937 14937 Mu (kg) 2027 2027 2027

Fig. 2. Cross section of the Groene Hart tunnel.

ARTICLE IN PRESS
S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297 85

3. Numerical modelling of the Groene Hart tunnel A coupled periodic nite elementboundary element model [16,17] is used to model the response of the Groene Hart tunnel during the passage of a Thalys high speed train at a speed of 300 km/h. In the following subsections, the methodology used to derive in the frequency domain the response of periodic media due to moving loads, based on the transfer functions formulated in the frequencywavenumber domain and the frequency content of the axle loads, will be briey recapitulated. The reader is referred to complementary literature [16,17,30] for a more detailed description of the methodology. 3.1. Response of periodic media due to moving loads The coupled tracktunnelsoil system is subjected to vertical loads moving along the longitudinal direction ey . In the xed frame of reference, the distribution of na vertical axle loads on the coupled tracktunnelsoil system is written as the summation of the product of Dirac functions that determine the time dependent position xk t fxk0 ; yk0 vt ; zk0 gT and the time history gk t of the k-th axle load:

known. The computation of the transfer functions and the axle loads is presented in the following subsections. 3.2. Transfer functions of the tracktunnelsoil system A coupled periodic nite elementboundary element model is used to solve the dynamic tunnelsoil interaction problem and to compute the transfer functions in the frequencywavenumber domain. A subdomain formulation is employed where the nite element method is used to model the track and the tunnel, while the boundary element method is used to model the soil as a horizontally layered halfspace. As the reference cell of the tunnel ~ t x ~ ; ky ; o are decomposed is bounded, the tunnel displacements u ~ t x ~ ; ky that are periodic functions of on a basis of tunnel modes W ~ s x ~ ; ky ; o are the second kind [16,17]. The soil displacements u written as the superposition of waves that are radiated by the ~ t x ~ ; ky into the soil. A weak variational formulatunnel modes W tion of the problem results in the following system of equations in the frequencywavenumber domain [16,17,32]: Kt ky o2 Mt ky Ks ky ; oat ky ; o Ft ky ; o 3

rbx; t

na X k1

dx xk0 dy yk0 vt dz zk0 gk tez

with yk0 the initial position of the k-th axle that moves with the train speed v along the y-axis and ez the vertical unit vector. A periodic structure can be analyzed using the Floquet transform [16,17] by restricting the problem domain O to a single periodic unit ~ . If the spatial period is L, then the position x of or reference cell O ~ nLey , any point in the problem domain is decomposed as x x ~ is the position in the reference cell and n is the cell number. where x The response of periodic domains subjected to the moving loads is given as follows in the frequency domain [30,31]: na Z 1 1 X ~ nLey ; o ^ o ky vexpiky nL yk0 ^ i x g u 2p k 1 1 k Z L=2 0 ~ ~ x ~ 0 dky ~ 0 h expiky y ; ky ; o dy 2 zi ~ ; x
L=2

where ky ky 2mp=L such that ky A p=L; p=L. The transfer 0 ~ ~ x ; ky ; o of the coupled tracktunnelsoil system in function h zi ~ ; x the frequencywavenumber domain is the Floquet transform of the 0 ~ ^ x nLey ; o in the frequencyspatial domain transfer function h zi ~ ; x ~ in the direction ei due to a and is dened as the displacement at x ~ 0 in the direction ez . Expression (2) unit load applied at x demonstrates that the response to moving loads can be calculated 0 ~ ~ x ^ k o are ; ky ; o and the axle loads g if the transfer functions h zi ~ ; x

where Kt ky and Mt ky are the projection of the nite element stiffness and mass matrix of the tunnels reference cell on the ~ t x ~ ; ky . Ks ky ; o is the dynamic stiffness matrix tunnel modes W of the soil and is computed using the periodic boundary element formulation based on GreenFloquet functions dened on the periodic structure with period L [16,17,32]. Ft ky ; o is the projection of the force vector in the reference cell on the tunnel ~ t x ~ ; ky . Finally, at ky ; o are the modal coordinates that modes W determine the contribution of the modes in the kinematical basis ~ t x ~ ; ky . W In the present analysis, the Groene Hart tunnel is modelled as a periodic structure with period L 0:65 m, corresponding to the sleeper spacing ds . It is worth to mention that, in the low frequency range of interest, the tunnel could also have been modelled as an equivalent invariant structure using a two-and-ahalf dimensional method. The periodic approach, however, has the advantage that existing three-dimensional boundary element technology for layered media can be reused, since the Green Floquet functions have the same singularities as the threedimensional Greens functions [16,17]. A two-and-a-half dimensional approach would have necessitated the analysis of the singularities of the two-and-a-half dimensional Greens functions, which has now been avoided. A nite element model of the reference cell is made using ANSYS. The circular concrete lining of the tunnel and the tunnel invert consisting of sand/cement stabilization are modelled using 8-node brick elements, including incompatible bending modes (Fig. 4). The size of the nite elements is governed by the

Mode 5 at 5.9 Hz.

Mode 9 at 29.75 Hz.

Mode 17 at 75.03 Hz.

Fig. 4. (a,b) Two in-plane modes and (c) the rst out-of-plane mode of the generic cell of the Groene Hart tunnel.

ARTICLE IN PRESS
86 S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297

boundary element mesh along the tunnelsoil interface, so that ideally a minimum of N 8 elements are used per shear min wavelength ls Cs =fmax , with Cs 240 m=s the shear wave velocity in the soil layer around the apex of the tunnel and fmax 100 Hz the maximum excitation frequency. This results in a min recommended element length le ls =N 0:3 m. In the present model, two elements are used in the longitudinal direction (le 0:325 m), while only 36 elements are used to model the tunnel lining in the circumferential direction (le 0:7625 m). Thus, the model is expected to be accurate up to 40 Hz, while some discretization errors may occur at higher frequencies. A ner mesh is not used because of the long computation times involved. The partition wall and the concrete oor that rests on the sand/

Fig. 5. Real part of the vertical displacement of the coupled tracktunnelsoil system due to a harmonic excitation at (a) 10 Hz and (b) 40 Hz on the rails.

cement stabilization are also modelled with 8-node brick elements. The partition wall is connected to the tunnel wall with a hinge connection; there is no connection between the concrete oor and the tunnel wall. ~ t x ~ ; ky ; o in As the tunnel is bounded, the displacement eld u ~ t x ~ ; ky the tunnel can be decomposed on a basis of functions W [16,17]. This kinematic basis is selected by performing an eigenvalue analysis of the tunnels reference cell. A complex eigenvalue problem is formed at each wavenumber after imposing the periodicity condition of the second kind on the reference cell in order to comply with the denition of the Floquet transform [16,17,30]. In the present analysis, the eigenvalue problem is solved for only a limited number M of lower modes at selected wavenumbers ky1 ; ky2 ; . . . ; kyq in the range 0; p=L. Fig. 4 shows two in-plane and the rst out-of-plane exible modes of the generic cell computed at zero wavenumber. The eigenvectors computed at selected wavenumbers are collected in a matrix and a singular value decomposition is applied to obtain the kinematic basis. For the problem under consideration, a convergence study has demonstrated that a kinematic basis consisting of 80 vectors guarantees good modelling accuracy [25]. The track is incorporated in the tunnel using the Craig Bampton substructuring technique. The tunnel hosts two tracks, but only the right track is considered in the present analysis, as it is expected that the presence of the left track does not affect the dynamic traintracktunnelsoil interaction. A nite element model of the track is developed according to the track properties described in Section 2.3. The slab is connected to the tunnel oor by discrete supports, which represent the stiffness of the

Fig. 6. Observation points (a) on the slab track, (b) on the tunnel, and (c) in the free eld.

ARTICLE IN PRESS
S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297 87

geotextile layer. The concrete slab is modelled using plate elements, while the bi-bloc sleepers are represented using mass elements and joined to the slab using rigid connections. The base plates are modelled using mass elements, while EulerBernoulli beam elements are used for the rails. The base plate pads and the rail pads are represented by spring elements that account for the stiffness and the damping of the resilient material. The rail pads and base plates discretely support the rail at an interval ds 0:65 m, resulting in a periodic track model. Eigenvalue analysis of the track on a rigid base demonstrates that resonance of the rail and the base plate on the base plate pads occurs at 106.54 Hz; 25 track modes on a rigid base have been found sufcient to obtain convergence. The response of the tunnel in the frequencywavenumber domain is obtained through solution of the coupled system of Eq. (3). Once the displacements and stresses along the tunnelsoil interface are known, the representation formula of elastodynamics is applied to compute the free eld response around the tunnel. Transfer functions are nally obtained by evaluating the

Table 3 Coordinates of the observation points on the track, on the tunnel, and in the free eld. Points TR1 TR2 TR3 TR4 T1 T2 T3 A B C D E F Position Rail Base plate Sleeper Slab Tunnel oor Tunnel wall Tunnel apex Free eld Free eld Free eld Free eld Free eld Free eld x (m) 2.635 2.635 2.635 2.537 3.385 6.650 0 0 14.5 43.5 0 14.5 43.5 y (m) 0 0 0 0 0 0 0 0 0 0 0 0 0 z (m) 29.516 29.536 29.672 29.692 29.797 26.250 19.600 0 0 0 10.7 10.7 10.7

inverse Floquet transformation from the wavenumber to the spatial domain [16,17,32]. As an example, Fig. 5 shows the vertical component of the transfer function of the coupled tracktunnel soil system for a unit harmonic load at excitation frequencies of 10 and 40 Hz; a load of 0.5 N is applied on each rail, which is equivalent to a unit load on the system. Apart from the response of the tunnel, the response of the soil has been computed in a limited number of output points in the reference cell, allowing to visualize the response on two horizontal and one vertical plane after evaluation of the inverse Floquet transformation. The wave propagation pattern in the soil is asymmetric, as the harmonic load is applied on the right track, which is eccentrically positioned with respect to the center of the tunnel. The presence of the tunnel signicantly inuences wave propagation in the soil, particularly at 40 Hz, where the wavelengths are shorter than the diameter of the tunnel. Computing the response on a denser grid of points in the reference cell would allow to visualize the transfer functions in more detail in the different soil layers, but requires substantial computational effort. The transfer functions are subsequently discussed in more detail for a limited number of observations points on the slab track, in the tunnel and in the free eld, as indicated on Fig. 6; the coordinates of these points are given in Table 3. Fig. 7a shows the track receptance in the frequency range 1 200 Hz for the track on a rigid base. Superimposed on the same graph is the rail receptance in the frequency range 1100 Hz, computed with the coupled nite elementboundary element model, accounting for the exibility of the tunnelsoil system; the latter does not substantially affect the response of the track. The rst peak in the response appears at the rst natural frequency of the track at 106.54 Hz. Fig. 7bd show the transfer function at the base plate, the sleeper, and the tunnel apex. These transfer functions do not exhibit any peak in the frequency range 1 100 Hz. The maximum reduction in the response is from the base plate to the sleeper, which is due to the exible base plate pads. The reduction in response from the rail to the base plate is not large as stiff rail pads are used. Superimposed on Fig. 7d is the

Displacement [dB ref 1m/N]

140

Displacement [dB ref 1m/N]

150

160

170

180

190

200 0 50 100 150 Frequency [Hz] 200

210 0 20 40 60 80 Frequency [Hz] 100

Displacement [dB ref 1m/N]

150

Displacement [dB ref 1m/N]

200

170

220

190

240

210 0 20 40 60 80 Frequency [Hz] 100

260 0 20 40 60 80 Frequency [Hz] 100

Fig. 7. (a) Track receptance for the track on a rigid base (black line) and for the track in the tunnel (gray line) and transfer functions at (b) the base plate (TR2), (c) the sleeper (TR3), and (d) the tunnel apex (T3). Superimposed on (d) is the transfer function at the tunnel apex for the tunnel without the track (gray line).

ARTICLE IN PRESS
88 S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297

transfer function at the tunnel apex for the case of the tunnel without the track, i.e. when the load is applied directly on the concrete oor. At low frequencies, the track does not signicantly inuence the response at the apex, which can also be explained by the fact that the transmittance at low frequencies is equal to one. Figs. 8 and 9 show the horizontal and vertical component of the transfer function at the observation points A, B, C, D, E, and F in the free eld, for a harmonic load applied on the rails. The

points A, B, and C are located on the free surface, while points D, E, and F are located at a depth of 10.7 m, at the interface of the soft clay layer and the relatively stiffer sand layer (Fig. 6c). Comparison of the response in the points D, E, and F with the response in the points A, B, and C allows to appreciate if and how vibration components are amplied through the soft clay and peat layers. Considering rst the horizontal response (Fig. 8), it can be observed that the response in the points A and D directly above

Displacement [dB ref 1m/N]

Displacement [dB ref 1m/N]

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

Displacement [dB ref 1m/N]

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

Displacement [dB ref 1m/N]

Displacement [dB ref 1m/N]

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

Displacement [dB ref 1m/N]

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

Fig. 8. Horizontal component of the transfer function at the points (a) A, (b) B, (c) C, (d) D, (e) E, and (f) F in the free eld. Superimposed on the graphs is the transfer function in the free eld for the tunnel without the track (gray line).

Displacement [dB ref 1m/N]

Displacement [dB ref 1m/N]

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

Displacement [dB ref 1m/N]

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

Displacement [dB ref 1m/N]

Displacement [dB ref 1m/N]

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

Displacement [dB ref 1m/N]

200 220 240 260 280 0 20 40 60 80 Frequency [Hz] 100

Fig. 9. Vertical component of the transfer function at the points (a) A, (b) B, (c) C,(d) D, (e) E, and (f) F in the free eld. Superimposed on the graphs is the transfer function in the free eld for the tunnel without the track (gray line).

ARTICLE IN PRESS
S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297 89

the tunnel is non-zero as the harmonic load is applied on both rails of the track, which is eccentrically positioned with respect to the center of the tunnel. The response in the free eld is characterized by an oscillating behavior due to interference of compression and shear or Rayleigh waves. The horizontal response in the point B at the surface is clearly higher than the response in the point E at the bottom of the softer layers; a similar observation can be made, but to a less extent, when comparing the response in the points F and C, while no amplication is observed between points D and A. This amplication is related to vertically propagating shear waves in the soft top layers, as will be explained in a following paragraph. Superimposed on Fig. 8 are the transfer functions computed without the presence of the track when the harmonic load is directly applied on the tunnel invert. The track has an inuence from 40 Hz onwards. The response of the tracktunnelsoil system is higher than that of the tunnelsoil system due to the resonance of the track at 106 Hz. The vertical component of the transfer function (Fig. 9) is generally lower than the horizontal component and no pronounced amplication through the soft layers is observed. The previous observations can be explained by visualizing the wave eld around the tunnel upon harmonic excitation on the track. It has been mentioned, however, that the computation of the transfer functions of the coupled tracktunnelsoil system in

a large number of receivers in the reference cell is expensive. Although it has also been recognized that the tunnel importantly inuences the transfer functions, especially at higher frequencies, the amplication of waves in the soil can also be studied by considering the Greens functions of the layered soil (without tunnel). These Greens functions are computed with the ElastoDynamics Toolbox (EDT) [33], applying a vertical unit harmonic load at a depth of 29.5 m, corresponding to the depth of the concrete oor of the tunnel where the load is applied in the previous cases. The soil has the same properties as given in Table 1. Fig. 10 shows the horizontal and vertical component, as well as the modulus, of the Greens function in the layered halfspace for a harmonic point load applied at a depth of 29.5 m, at an excitation frequency of 20, 40, and 80 Hz. For the horizontal component, it can clearly be observed how shear waves propagate upwards in the stiffer sand layer(s) in a shear window marked by a zone that makes an angle from almost 303 to 603 with the horizontal axis; at the interfaces with the softer materials (clay and peat layers), the shear waves are refracted and propagate upwards at a steeper angle within the softer material, while the amplitude increases. These gures clearly illustrate how the response is amplied by upward propagating shear waves in the soft layers. Depending on the excitation frequency, this phenomenon is more pronounced in

Fig. 10. Horizontal component (left), vertical component (middle) and modulus (right) of the Greens function in a layered halfspace for a harmonic point load applied at a depth of 29.5 m, corresponding to the depth of the tunnel invert, at a frequency of (a) 20 Hz, (b) 40 Hz, and (c) 80 Hz.

ARTICLE IN PRESS
90 S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297

an area along the surface at a certain distance from the vertical axis of symmetry; at larger distances along the surface, waves are attenuated due to geometrical and material damping in the soil. The vertical component is maximum in the sand layer(s) at a depth corresponding to the point of application of the load and attenuates with increasing lateral distance. No amplication of the vertical component occurs in the soft clay and peat layers, where the vertical component is low. The modulus of the transfer function in the supercial layers is therefore governed by the horizontal component and by the upward propagation of shear waves, as can clearly be observed in Fig. 10. 3.3. Axle loads The static and dynamic component of the axle loads are accounted for in the following. The static component is due to the weight of the train, as summarized in Table 2. The resulting static axle load is equal to wk 170 kN. Dynamic axle loads are related to parametric and unevenness excitation. Parametric excitation occurs at harmonics of the sleeper passage frequency fs v=ds 128:2 Hz, with ds 0:65 m the sleeper distance and v 300 km=h the speed of the train. These frequencies lie outside the frequency range of interest for ground-borne vibration (180 Hz). Due to the Doppler effect, the rst harmonic of the axle load contributes to the response at a xed point in the free eld in a frequency range fs =1 v=C ; fs =1 v=C , with C the wave speed in the ground. The waves radiated behind the load can have much lower frequencies than fs and lie into the frequency range of interest. Higher frequencies are not considered for the prediction of ground-borne vibrations as they are signicantly attenuated in the free eld due to material damping in the soil; they can be

important, however, for the prediction of re-radiated noise in buildings. The parametric excitation depends on the variation in the track receptance along the sleeper bay. Fig. 11 shows the track compliance over the sleeper and at mid-span between sleepers. Since the rail and the base plate assembly are supported by soft base plate pads, the variation in the track receptance is small. Therefore, parametric excitation is not signicant for the present analysis. Only dynamic forces due to the wheel/rail unevenness are considered. ^ d o due to wheel/rail unevenness can be The dynamic forces g calculated by considering the following vehicletrack interaction equation [10,21]: ^ t og ^ v o C ^ d o u ^ w=r o C 4

^ t o is the ^ v o is the vehicle compliance matrix, C where C compliance matrix of the track in the moving frame of reference, ^ w=r o is the unevenness represented in the frequency and u domain [10]. This approach is valid only for invariant tracks, but can be conveniently used for the present periodic analysis as the variation in the track receptance along the sleeper bay is negligible. Superimposed on Fig. 11 is the diagonal element ^ t o of the track compliance matrix in the moving frame of C 11 reference for a train speed of 300 km/h. The resonance frequency near 106 Hz shifts to lower frequencies due to the Doppler effect, ^ t o is not much but, even at a high train speed, the element C 11 different from the corresponding element in the xed frame of ^ t o and reference. Fig. 12 shows the off-diagonal elements C 12 t ^ C 21 o of the track compliance matrix in the moving and the xed frame of reference. These elements are clearly inuenced by the train speed. In the following, the track receptance in the moving frame of reference is therefore used to compute the wheel/rail interaction forces.

Displacement [dB ref 1m/N]

140 RMS force [dB ref 1 N] 0 50 100 150 Frequency [Hz] 200

100

160

80

180

60

200

40 100

101 102 Frequency [Hz]

103

Fig. 11. The track receptance over the sleeper (dashed gray line) and at mid-span between sleepers (solid gray line) for a train speed of 0 km/h. Superimposed on the graph is the track receptance over the sleeper for a train speed of 300 km/h (black line).

Fig. 13. One-third octave band RMS spectrum of the contact force at the front axle of the Thalys high speed train computed with the measured unevenness (black line) and FRA track class 6 (gray line) for a train speed of 300 km/h.

Displacement [dB ref 1m/N]

140

Displacement [dB ref 1m/N]

140

160

160

180

180

200 0 50 100 150 Frequency [Hz] 200

200 0 50 100 150 Frequency [Hz] 200

^ t o and (b) C ^ t o of the track compliance matrix for a train speed of 0 km/h (gray line) and 300 km/h (black line). Fig. 12. The elements (a) C 12 21

ARTICLE IN PRESS
S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297 91

At frequencies of more than a few Hertz, the vehicles primary and secondary suspension isolate the body and the bogie from the wheelset [34]. The unsprung masses therefore are the only components that affect the vertical dynamic loads. Each axle is modelled as a mass Mu and the vehicle compliance matrix is equal ^ v o diagf1=Mu o2 g of order 26. to the diagonal matrix C The measured unevenness uw=r y is transformed to the wavenumber domain and expressed in the frequency domain. Eq. (4) is subsequently used to calculate the dynamic wheeltrack interaction forces. Fig. 13 shows the one-third octave band RMS spectrum of the traintrack interaction force at the front axle of the Thalys HST. The frequency content of this force exhibits a clear maximum near the wheeltrack resonance frequency of about 42 Hz. Superimposed on the same graph is also the dynamic force computed using the FRA track class 6. The comparison between the contact force for the measured unevenness and the FRA track class 6 shows that the unevenness on the track corresponds to the quality of the FRA track class 6. For further analysis, the dynamic forces due to the measured unevenness are used.

4. Response due to a passage of a Thalys HST in the tunnel The static and dynamic component of the axle loads are used to compute the response in the tunnel and in the free eld. The frequency content of the response is calculated by means of Eq. (2), which uses the transfer functions in the frequencywavenumber domain and the frequency content of the axle loads. The time history is obtained by evaluating the inverse Fourier transformation.

Fig. 14 show the time history, the linear frequency spectrum and the one-third octave band RMS spectrum of the vertical velocity on the rail (TR1). The passage of every axle can clearly be identied from the time history that shows a nearly uniform response for every axle (Fig. 14a). Although the power cars and the adjacent carriages have a different axle composition than the central carriages, a modulation of the frequency content is observed with peaks at the fundamental bogie passage frequency fb v=Lb 4:46 Hz of the central carriages and its higher harmonics, modulated at the axle passage frequency fa v=La 27:78 Hz [35] (Fig. 14b). The contribution of the quasi-static forces is also shown (Fig. 14b and c). It can be observed that the quasi-static response is vital for the response of the track in the low frequency range up to 30 Hz [10]. The dominant frequency content at the frequencies above 30 Hz is due to rail unevenness. Unlike the maxima near the traintrack resonance frequency in the dynamic force (Fig. 13), the maxima in the response are spread over a wider frequency range around the traintrack resonance frequency due to the Doppler effect. Fig. 15 shows the time history, the linear frequency spectrum and the one-third octave band RMS spectrum of the vertical velocity in the tunnel on the concrete oor (T1). The contribution of each axle can be distinguished in the time history, but is less clear than in the response on the rail. The contribution of quasistatic forces has diminished and shifted to lower frequencies. The dominant frequency content above 20 Hz is due to the dynamic component of the axle loads. Figs. 1619 show the time history and the one-third octave band spectrum of the horizontal and the vertical velocity in the free eld at the points A, B, and C along the surface, and the points

0.4 Velocity [m/s/Hz] Velocity [m/s] 0.2 0 0.2 0.4 3

0.05 0.04 0.03 0.02 0.01 0 2 1 0 Time [s] 1 2 3 0 20 40 60 80 100 RMS velocity [m/s]

101 102 103 104 105 1 2 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

Frequency [Hz]

Fig. 14. (a) Time history, (b) linear frequency spectrum and (c) one-third octave band RMS spectrum of the vertical velocity on the rail (TR1) for a passage of the Thalys HST at a speed of 300 km/h. Superimposed on graphs (b) and (c) is the contribution of quasi-static forces (gray line).

x 103 2 1 0 1 2 3 Velocity [m/s/Hz] 6

x 104 103 4 RMS velocity [m/s] 0 20 40 60 80 100 104 105 106 107 1 2 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

Velocity [m/s]

0 2 1 0 Time [s] 1 2 3

Frequency [Hz]

Fig. 15. (a) Time history, (b) linear frequency spectrum and (c) one-third octave band RMS spectrum of the vertical velocity on the concrete oor (T1) for a passage of the Thalys HST at a speed of 300 km/h. Superimposed on graphs (b) and (c) is the contribution of quasi-static forces (gray line).

ARTICLE IN PRESS
92 S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297

x 104 6 4
Velocity [m/s] Velocity [m/s]

x 104 6 4 2 0 2 4 6 3 2 1 0 Time [s] x 104 6 4


Velocity [m/s] Velocity [m/s]

x 104 6 4 2 0 2 4 6 3 2 1 0 Time [s] x 104 6 4


Velocity [m/s]

2 0 2 4 6 3 2 1 0 Time [s] x 104 6 4 1 2 3

Velocity [m/s]

2 0 2 4 6 3 2 1 0 Time [s] 1 2 3

2 0 2 4 6 3 2 1 0 Time [s] 1 2 3

2 0 2 4 6 3 2 1 0 Time [s] 1 2 3

Fig. 16. Time history of the horizontal velocity at points (a) A, (b) B, (c) C, (d) D, (e) E, and (f) F in the free eld for a passage of the Thalys HST at a speed of 300 km/h.

103 RMS velocity [m/s] RMS velocity [m/s] 104 105 106 107

103 104 105 106 107 RMS velocity [m/s] 1 2 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

103 104 105 106 107

4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

103 RMS velocity [m/s] 104 105 106 107 RMS velocity [m/s]

103 104 105 106 107 RMS velocity [m/s] 1 2 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

103 104 105 106 107

4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

Fig. 17. One-third octave band spectrum of the horizontal velocity at points (a) A, (b) B, (c) C, (d) D, (e) E, and (f) F in the free eld for a passage of the Thalys HST at a speed of 300 km/h. Superimposed on the graphs is the contribution of quasi-static forces (gray line).

D, E, and F at a depth of 10.7 m along the interface between the soft clay layer and the stiffer sand layer. At the point E, which is relatively close to the tunnel, the horizontal velocity (Fig. 16e) is larger than the vertical velocity (Fig. 18e), which is due to the dominant contribution of shear waves to the horizontal response;

the vertical response is lower due to the low compressibility of the medium. The horizontal velocity is higher on the surface (points B and C, Fig. 16b and c) than at depth (points E and F, Fig. 16e and f) due to amplication of vertically polarized shear waves in the soft soil layers. Such an amplication is not observed on the vertical

ARTICLE IN PRESS
S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297 93

x 104 6 4 Velocity [m/s] Velocity [m/s] 2 0 2 4 6 3 2 1 0 Time [s] x 104 6 4 Velocity [m/s] 2 0 2 4 6 3 2 1 0 Time [s] 1 2 3 Velocity [m/s] 6 4 2 0 2 4 6 1 2 3 6 4 2 0 2 4 6

x 104 6 4 Velocity [m/s] 3 2 1 0 Time [s] x 104 6 4 Velocity [m/s] 2 0 2 4 1 2 3 2 0 2 4 6

x 104

0 Time [s]

x 104

0 Time [s]

6 3

0 Time [s]

Fig. 18. Time history of the vertical velocity at points (a) A, (b) B, (c) C, (d) D, (e) E, and (f) F in the free eld for a passage of the Thalys HST at a speed of 300 km/h.

103 RMS velocity [m/s] RMS velocity [m/s] 104 105 106 107 1 2 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

103 104 105 106 107 1 2 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz] RMS velocity [m/s]

103 104 105 106 107 1 2 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

103 RMS velocity [m/s] 104 105 106 107 1 2 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz] RMS velocity [m/s]

103 104 105 106 107 1 2 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz] RMS velocity [m/s]

103 104 105 106 107 1 2 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

Fig. 19. One-third octave band spectrum of the vertical velocity at points (a) A, (b) B, (c) C, (d) D, (e) E, and (f) F in the free eld for a passage of the Thalys HST at a speed of 300 km/h. Superimposed on the graphs is the contribution of quasi-static forces (gray line).

component, which is lower at the surface (Fig. 18ac) than at depth (Fig. 18df), due to the attenuation of waves as well as the reection of waves at layer interfaces. Similar observations have been made in the discussion of the transfer functions (Fig. 10). These observations explain why the horizontal peak particle

velocity (PPV) along the surface is much higher than the vertical PPV; in the point B, for example, the horizontal and vertical PPV are equal to 0.60 and 0.11 mm/s, respectively. The frequency content shifts towards lower frequencies for increasing distance from the track as higher frequency

ARTICLE IN PRESS
94 S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297

components are attenuated due to material damping in the soil (Figs. 17 and 19). The contribution of the dynamic axle loads dominates the response at frequencies above 8 Hz. As the train speed is high, the contribution of quasi-static forces can also be identied in the free eld response, especially on the vertical component; it has shifted to lower frequencies, as the response in the time domain becomes broader. Metrikine et al. [36] have shown that, for a tunnel embedded in soft soil, the critical speed is usually less than the minimum wave velocity in the soil and depends on the depth of the tunnel. As in the present analysis the shear wave velocity of the soil around the tunnel is much higher than the train speed, the motion is not in the super-critical regime. The prominent contribution of quasi-static forces in the vertical response can be explained from the transfer functions in the free eld. In Fig. 9, a difference of more than 20 dB can be observed between the transfer function at low frequencies below 10 Hz and the transfer functions at higher frequencies around the wheel track resonance frequency of 42 Hz. This large difference is also reected in the response to moving loads, where a large magnitude of the quasi-static response is observed at low frequencies in comparison to the dynamic response around 42 Hz. In order to better understand the high quasi-static response in the free eld, the response at a lower train speed of 100 km/h is compared with the response at a higher speed of 300 km/h. Fig. 20 shows a contour plot of the logarithm of the vertical displacement on the free surface at the point B in the frequencywavenumber domain. The dominant response is conned to low wavenumbers. The extent of this region is larger for softer soils with low wave velocities. Superimposed on the graph are the load velocity lines (ky o=v) for both train speeds. This relation implies that the

phase velocity Cy o=ky of all waves radiated by the constant load must be equal to the velocity v of the load. For the lower train speed, the load velocity line lies in the region of higher wavenumbers, where the response is strongly attenuated by the tunnelsoil system. For the higher train speed, the load velocity line crosses the region with high response, causing signicant amplication of the quasi-static response. This is also illustrated in Fig. 21, where the vertical velocity in the point B due to quasistatic excitation at the two considered train speeds (100 and 300 km/h) is compared. At the low train speed, the displacement eld generated by the load consist of strongly attenuated waves, while, at the high train speed, waves are excited in the structure which can transport energy to distant points. A difference of more than a factor of 10 is observed in the vertical velocity due to quasistatic excitation at a train speed of 100 and 300 km/h. Such an amplication is not expected in the case of moving dynamic loads, as for harmonically varying loads, waves are always radiated, irrespective of the train speed.

5. Evaluation of human exposure to vibrations In order to assess the possibility of annoyance to people due to the passage of a Thalys high speed train in the Groene Hart tunnel, the Dutch SBR guideline part B [1] is applied. This guideline is used for evaluating human exposure to vibrations in buildings. In the present analysis, however, only free eld vibrations are considered. The SBR guideline [1] is very similar to the German DIN 4150 norm [37] and based on the assessment of the maximum Vmax of the running effective velocity veff t and of the vibration level Vper over an evaluation period (day, evening and night), dened as v u n u1 X 5 v2 Vper t n i 1 eff ;max;30;i Herein, n is the number of 30 s intervals in the evaluation period and veff ;max;30;i is the maximum of the running effective velocity veff t in the i-th interval of 30 s. Values of veff ;max;30;i;i lower than 0.1 are set equal to zero. The maximum vibration level Vmax and the vibration level Vper over an evaluation period are compared to limit values A1 , A2 and A3 . If the maximum vibration level Vmax is less than the limit value A1 , then the standard is satised. If the maximum vibration level Vmax is greater than the limit value A2 , then the standard is not satised. If the maximum vibration level Vmax lies between the limit values A1 and A2 , then the vibration level Vper over the evaluation period should be calculated. The standard is not

2 Wavenumber [rad/m] 1.5 1 0.5 0 20 40 60 80 Frequency [Hz] 100

25 30 35 40 45

~ z ky ; o on the free surface at Fig. 20. Logarithm of the vertical displacement u point B in the frequencywavenumber domain. Superimposed on the graph are the load velocity lines (ky o=v) for a train speed of 100 km/h (dashed line) and 300 km/h (solid line).

1 Velocity [m/s/Hz] 0.8 0.6 0.4 0.2 0

x 104 104 RMS velocity [m/s] 105 106 107 108 2 4 6 8 10 1 2 Frequency [Hz] 4 8 16 31.5 63 125 250 Onethird octave band center frequency [Hz]

Fig. 21. (a) Linear frequency spectrum and (b) one-third RMS spectrum of the vertical velocity at point B due to the quasi-static excitation for a passage of the Thalys train at a speed of 100 km/h (gray line) and 300 km/h (black line).

ARTICLE IN PRESS
S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297 95

satised if Vper is greater than A3 . The threshold values for different building classications and for each evaluation period are summarized in Table 4. These limit values are for repeated occurrence of vibrations due to railways [1]. Figs. 22 and 23 show the running effective horizontal and vertical velocity on the free surface at points A, B and C. Superimposed on these graphs are the limit values A1 and A2 for repeated occurrence of vibrations in residential buildings during day time, assuming that high speed train trafc through the Groene Hart tunnel will mainly be concentrated during the day. The maximum values Vmax for the vertical velocity at points A, B and C are equal to 0.0369, 0.0287 and 0.0119, respectively, and lower than the threshold value A1 0:10; the SBR guideline is satised for the vertical component along the surface. The maximum values Vmax for the horizontal velocity at points A, B and C are equal to 0.2352, 0.3621 and 0.1148, respectively. These values exceed the threshold value A1 0:10, but do not exceed the upper threshold value A2 0:40. A further analysis based on the vibration level Vper over the evaluation period is needed to check
Table 4 Limit values for Vmax and Vper for repeated occurrence of vibrations due to railways according to the SBR guideline part B [1]. Type of facility A1 A2 A3 A1 Night 0.10 0.10 0.15 0.15 0.10 0.20 0.20 0.60 0.60 0.10 0.05 0.05 0.07 0.07 A2 A3

Day and evening Hospitals Residential buildings University and ofce buildings Meeting rooms Critical work spaces 0.10 0.10 0.15 0.15 0.10 0.40 0.40 0.60 0.60 0.10 0.05 0.05 0.07 0.07

whether the standard is satised or not. Since the exact number of train passages is not known, an inverse procedure is followed to determine the maximum number of train passages that can be allowed such that the SBR guideline is not violated. For the sake of simplicity, it is assumed that only one train passage occurs per interval of 30 s. It has been agreed by the authorities that the SBR guideline should be met in buildings located at a lateral distance of 50 m from the Groene Hart tunnel. The farthest observation point considered in this paper is point C, which is located at a lateral distance of 43.5 m from the tunnel. Using the limiting value A3 0:05 for residential buildings and the maximum vibration level veff ;max;30;i 0:1148 at the point C, the number of train passages that can be allowed during the day is equal to 273. Likewise for the evening time, the maximum number of train passages that can be allowed is 91. The present frequency of the Thalys train in the Groene Hart area is less than the number of passages that could be allowed and, therefore, the SBR guideline is not expected to be violated. If the observation point B at a distance of 14.5 m from the tunnel is chosen, then the maximum number of train passages that can be allowed during the day and evening is calculated as 27 and 9, respectively. These numbers may be critical for buildings located closer to the tunnel as the SBR guideline is violated if the present frequency of the Thalys high speed train in the region is slightly increased. The analysis presented in this paper is based on free eld vibration levels. A more rigorous analysis can be carried out by calculating the response in buildings to check whether the guideline is satised or not. Due to presence of soft soil layers, the buildings in the region are likely to be founded on pile foundations. The vibration levels on the foundation and in the building can be determined by solving a dynamic soilstructure interaction problem [38]. The vibrations at the foundation level

Running effective velocity

Running effective velocity

0.4

A2

0.4

A2

Running effective velocity

0.6

0.6

0.6 A2

0.4

0.2

A1

0.2

A1

0.2

A1

0 3

0 1 Time [s]

0 3

0 1 Time [s]

0 3

0 1 Time [s]

Fig. 22. The running effective horizontal velocity veff t on the free surface at points (a) A, (b) B and (c) C. Superimposed on the graphs are the limits A1 and A2 for repeated occurrence of vibrations in residential buildings due to railways (day time) according to the SBR guideline part B (gray lines).

Running effective velocity

0.4

A2

Running effective velocity

0.4

A2

Running effective velocity

0.6

0.6

0.6 A2

0.4

0.2

A1

0.2

A1

0.2

A1

0 3

0 1 Time [s]

0 3

0 1 Time [s]

0 3

0 1 Time [s]

Fig. 23. The running effective vertical velocity veff t on the free surface at points (a) A, (b) B and (c) C. Superimposed on the graphs are the limits A1 and A2 for repeated occurrence of vibrations in residential buildings due to railways (day time) according to the SBR guideline part B (gray lines).

ARTICLE IN PRESS
96 S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297

depend on the relative stiffness between the foundation (piles) and the soil and are expected to be less than the free eld vibrations as the soil is very soft compared to the foundation. The response in the building may be amplied due to resonances of oors and other structural elements. Nevertheless, it can be concluded as per the SBR guideline applied at a lateral distance of 43.5 m from the tunnel, that the operation of Thalys high speed trains in the Groene Hart area is not expected to be a matter of concern for human exposure to vibrations.

FoundationFlanders (FWO-Vlaanderen). The nancial support of IWT and FWO-Vlaanderen is gratefully acknowledged. The authors are grateful to Dr. H. Stuit and Dr. W. Gardien from Movares and to Dr. P. Holscher and Dr. V. Hopman from Deltares for providing useful information and input data for performing this analysis.

References
[1] Stichting Bouwresearch. SBR deel B: Hinder voor personen in gebouwen door trillingen: meeten beoordelingsrichtlijn; 2002. [2] Sheng X, Jones CJC, Petyt M. Ground vibration generated by a harmonic load acting on a railway track. Journal of Sound and Vibration 1999;225(1):328. [3] Sheng X, Jones CJC, Petyt M. Ground vibration generated by a load moving along a railway track. Journal of Sound and Vibration 1999;228(1):12956. [4] Sheng X, Jones CJC, Thompson DJ. A comparison of a theoretical model for quasi-statically and dynamically induced environmental vibration from trains with measurements. Journal of Sound and Vibration 2003;267(3):62135. [5] Sheng X, Jones CJC, Thompson DJ. A theoretical model for ground vibration from trains generated by vertical track irregularities. Journal of Sound and Vibration 2004;272(35):93765. [6] Auersch L. The excitation of ground vibration by rail trafc: theory of vehicle tracksoil interaction and measurements on high-speed lines. Journal of Sound and Vibration 2005;284(12):10332. [7] Metrikine AV, Verichev SN, Blauwendraad J. Stability of a two-mass oscillator moving on a beam supported by a visco-elastic half-space. International Journal of Solids and Structures 2005;42:1187207. [8] Metrikine AV, Popp K. Steady-state vibrations of an elastic beam on a viscoelastic layer under moving load. Archive of Applied Mechanics 2000;70: 399408. [9] Lombaert G, Degrande G, Clouteau D. Numerical modelling of free eld trafc induced vibrations. Soil Dynamics and Earthquake Engineering 2000;19(7):47388. [10] Lombaert G, Degrande G, Kogut J, Franc - ois S. The experimental validation of a numerical model for the prediction of railway induced vibrations. Journal of Sound and Vibration 2006;297(35):51235. [11] Lombaert G, Degrande G. Ground-borne vibration due to static and dynamic axle loads of InterCity and high speed trains. Journal of Sound and Vibration 2009;319(35):103666. [12] Forrest JA, Hunt HEM. A three-dimensional tunnel model for calculation of train-induced ground vibration. Journal of Sound and Vibration 2006;294:678705. [13] Forrest JA, Hunt HEM. Ground vibration generated by trains in underground tunnels. Journal of Sound and Vibration 2006;294:70636. [14] Hussein MFM, Hunt HEM. A numerical model for calculating vibration from a railway tunnel embedded in a full-space. Journal of Sound and Vibration 2007;305:40131. [15] Muller K, Grundmann H, Lenz S. Nonlinear interaction between a moving vehicle and a plate elastically mounted on a tunnel. Journal of Sound and Vibration 2008;310:55886. [16] Clouteau D, Arnst M, Al-Hussaini TM, Degrande G. Freeeld vibrations due to dynamic loading on a tunnel embedded in a stratied medium. Journal of Sound and Vibration 2005;283(12):17399. [17] Degrande G, Schevenels M, Chatterjee P, Van de Velde W, Holscher P, Hopman V, Wang A, Dadkah N. Vibrations due to a test train at variable speeds in a deep bored tunnel embedded in London clay. Journal of Sound and Vibration 2006;293(35):62644. [18] Sheng X, Jones CJC, Thompson DJ. Prediction of ground vibration from trains using the wavenumber nite and boundary element methods. Journal of Sound and Vibration 2006;293:57586. [19] Andersen L, Jones CJC. Coupled boundary and nite element analysis of vibration from railway tunnelsa comparison of two- and three-dimensional models. Journal of Sound and Vibration 2006;293:61125. [20] Sheng X, Jones CJC, Thompson DJ. Modelling ground vibrations from railways using wavenumber nite- and boundary-element methods. Proceedings of the Royal Society AMathematical Physical and Engineering Sciences 2005;461:204370. [21] Gupta S, Liu W, Degrande G, Lombaert G, Liu W. Prediction of vibrations induced by underground railway trafc in Beijing. Journal of Sound and Vibration 2008;310:60830. [22] Hopman V, de Feijter JW. Grondonderzoek HSL, Groene Hart tunnel, seismische sonderingen. Report CO-380840/671, GeoDelft; July 2000. [23] Biot MA. Theory of propagation of elastic waves in a uid-saturated porous solid. I. Low-frequency range. Journal of the Acoustical Society of America 1956;28(2):16878. [24] Schevenels M, Degrande G, Lombaert G. The inuence of the depth of the ground water table on free eld road trafc induced vibrations. International Journal for Numerical and Analytical Methods in Geomechanics 2004;28(5):395419. [25] Van den Berghe H. Trillingen ten gevolge van HST verkeer in de Groene Harttunnel. Masters thesis, Department of Civil Engineering, K.U.Leuven; 2008.

6. Conclusion In this paper, a coupled periodic nite elementboundary element model has been used to model the Groene Hart tunnel and predict the response due to the passage of a Thalys high speed train. The shear wave velocity of the soil has been determined from SCPT tests, while the density has been obtained from classical soil mechanics tests. The soil is completely saturated and consists of layers of peat, clay and sand. For numerical computations, a layered halfspace with four layers on top of a halfspace is considered. The top layers consisting of peat and clay are very soft. This results in an increase of the horizontal response on the surface due to amplication of vertically polarized shear waves, a phenomenon that is explained by analyzing the Greens functions of the layered halfspace. The tunnel geometry is complex as it consists of the tunnel invert, a technical gallery and a partition wall; it is conveniently modelled using the nite element method. The track is efciently incorporated in the tunnel using the CraigBampton substructuring method. The response due to the passage of a Thalys high speed train in the tunnel is computed. The response on the track is due to the quasi-static as well as the dynamic axle loads. At frequencies above 8 Hz, the free eld response is dominated by the dynamic axle loads. The contribution of the quasi-static axle loads is also present but limited to frequencies below 8 Hz. The amplitude of the quasi-static response increases with the train speed. Moreover, the frequency content spreads to higher frequencies as the train speed increases. The horizontal response on the free surface is higher than the vertical response as it is amplied due to the presence of soft top layers. The SBR guideline has been applied to evaluate the human response to vibrations from underground railways. Vibration levels are calculated at three observation points on the free surface and compared to limit values. The vertical component of the velocity complies with the SBR guideline as it is lower that the lower threshold value in the SBR guideline. For the horizontal component, the vibration level Vmax lies between the lower and upper threshold. Based on the maximum vibration intensity value Vmax at a lateral distance of 43.5 m from the tunnel on the free surface, the maximum number of train passages that are possible during the day and in the evening without violating the SBR guideline is evaluated as 273 and 91, respectively. These numbers suggest that the operation of the Thalys high speed train in the Groene Hart region will not cause serious vibration problems.

Acknowledgments The results presented in this paper have been obtained within the frame of the SBO project IWT 03175 Structural damage due to dynamic excitation: a multi-disciplinary approach, funded by IWT. The third author is a Postdoctoral fellow of the Research

ARTICLE IN PRESS
S. Gupta et al. / Soil Dynamics and Earthquake Engineering 30 (2010) 8297 97

[26] N.N. Report reference B0-IVW-050020642, Movares, Geotechnical Engineering and Environment Group; April 2005. [27] Hamid A, Yang TL. Analytical description of track-geometry vibrations. Transportation Research Record 1981;838:1926. [28] Diehl RJ, Holm H. Roughness measurements-have the necessities changed?. Journal of Sound and Vibration 2006;293:77783. [29] International Organization for Standardization. ISO 3095:2005: railway applicationsacousticsmeasurement of noise emitted by railbound vehicles; 2005. [30] Gupta S. Numerical modelling of subway induced vibrations. PhD thesis, Department of Civil Engineering, K.U.Leuven; 2008. [31] Chebli H, Othman R, Clouteau D. Response of periodic structures due to canique 2006;334:34752. moving loads. Comptes Rendus Me [32] Clouteau D, Elhabre ML, Aubry D. Periodic BEM and FEM-BEM coupling: application to seismic behaviour of very long structures. Computational Mechanics 2000;25:56777.

[33] Schevenels M, Franc - ois S, Degrande G. EDT: an elasto dynamics toolbox FOR MATLAB. Computers & Geosciences 2009;35(8):17524. [34] Knothe K, Grassie SL. Modelling of railway track and vehicle/track interaction at high frequencies. Vehicle Systems Dynamics 1993;22: 209262. [35] Degrande G, Lombaert G. An efcient formulation of Krylovs prediction model for train induced vibrations based on the dynamic reciprocity theorem. Journal of the Acoustical Society of America 2001;110(3):137990. [36] Metrikine AV, Vrouwenvelder T. Surface ground vibration due to a moving train in a tunnel: two-dimensional model. Journal of Sound and Vibration 2000;234(1):4366. Normung. DIN 4150 Teil 2: Erschutterungen [37] Deutsches Institut fur im Bauwesen, Einwirkungen auf Menschen in Gebauden; 1999. [38] Fiala P, Degrande G, Augusztinovicz F. Numerical modelling of ground borne noise and vibration in buildings due to surface rail trafc. Journal of Sound and Vibration 2007;301(35):71838.

Vous aimerez peut-être aussi