You are on page 1of 24

Acta Mech 235, 1395–1418 (2024)

https://doi.org/10.1007/s00707-023-03605-3

O R I G I NA L PA P E R

Paul König · Christoph Adam

A model considering the longitudinal track–bridge


interaction in ballasted railway bridges subjected
to high-speed trains

Received: 3 February 2023 / Revised: 29 April 2023 / Accepted: 7 May 2023 / Published online: 27 May 2023
© The Author(s) 2023

Abstract In this paper, a dynamic interaction model of the coupled system of railway bridge, foundation,
subsoil, ballast, track, and high-speed train is presented, with special emphasis on the longitudinal interaction
between the track and the bridge structure, taking into account the flexibility of the ballast. After a description
of the model of this interaction system, the equations of motion are given separately for each subsystem. The
discretization of the bedded rails is performed by two different approaches. In the first approach, the deflection
of the rails is expanded into the eigenfunctions of a finitely long bedded beam representing the rails. In the
second, simplified approach, the track response is represented by a superposition of the static deflection of the
infinitely long bedded beam due to a concentrated load. The coupling of the bridge structure with the track is
achieved by a component mode synthesis technique, which in the first approach leads to a representation of
the equations of motion in state-space. A discrete substructure technique is used to couple this subsystem with
the train model. The two presented strategies are verified by comparison with results of a finite element model
of this interaction system. Several application examples reveal the influence of the horizontal track–bridge
interaction and other modeling parameters on the dynamic bridge response.

1 Introduction

The constantly growing demand for efficient transportation has led to a steady expansion of high-speed rail
networks over the last two decades. This development makes dynamic evaluation necessary for more accurate
prediction of the behavior of railway bridges. Excessive accelerations of the bridge deck in response to the
passing train may lead to instability of the track ballast and consequently to misalignment of the rails, requiring
shortening of maintenance intervals or reduction of train speeds. Since excessive vibrations in combination
with misaligned rails can also affect the running safety of the train and, in the worst case, cause the train
to derail, limits for the maximum acceleration of the bridge deck have been introduced in various design
specifications. Predicting the dynamic response of railway bridges has therefore become an important task for
the engineer when designing railway bridges.
Although the prediction of the dynamic response is of undisputed importance and a variety of modeling
strategies have been developed to date, discrepancies between prediction and measurement are still observed
[4,20]. The reason for these discrepancies is the frequent uncertainties in mechanical modeling. While the
first mechanical models summarized in the books of Frýba [8,9] were already able to take into account the

P. König (B), C. Adam


Universität Innsbruck, Unit of Applied Mechanics, Technikerstr. 13, 6020 Innsbruck, Austria

P. König
E-mail: paul.koenig@uibk.ac.at
C. Adam
E-mail: christoph.adam@uibk.ac.at
1396 P. König, C. Adam

resonance excitation of the bridge structure caused by the regular axle spacing, the models often idealized
the train as moving individual loads and also neglected the interaction with the substructure of the track as
well as the subsoil under the foundation of the bridge. Early methods considering the dynamic interaction of
bridge and train by modeling the train with so-called mass-spring-damper (MSD) systems can be found in
Yang et al. [33], with simple two-dimensional and more sophisticated three-dimensional formulations. A more
recent publication Stoura and Dimitrakopoulos [29] presented an approach to decoupling the vehicle-bridge
interaction problem.
The large influence of the soil–structure interaction, frequently discussed by various authors [10,23,31,34],
represents an additional uncertainty factor that is often ignored by standard approaches. In an attempt to
accurately represent the subsurface, discretizing the half-space with finite elements (FE) or boundary elements
[10,23,31] often leads to a high computational cost. Therefore, recent contributions [12–14] have developed
simplified interaction models that idealize the subsoil beneath the foundation as spring-dashpot elements.
The mechanical models to represent the ballast superstructure in practical application mainly use the vertical
interaction properties of the track. These models, which mostly focus on two-dimensional descriptions, vary
greatly in their level of detail, ranging from a beam representing the rails connected to the ground or bridge
by a single intermediate layer [3,13,14,16,21], to models with two intermediate layers consisting of springs
and dashpots representing the rail pad and ballast between the rail beam and the ground or bridge. Track
models that additionally represent the substructure as a viscoelastic layer can be also found in the literature
[19,35]. In addition to the vertical interaction, the consideration of the longitudinal interaction between the
track and the bridge has been shown to have a significant impact on the dynamic behavior of the structure
[2,17,22,30], affecting both the damping and the natural frequencies of the coupled system. In these studies,
the longitudinal stiffness of the ballast is usually modeled with linear-elastic or linear-elastic ideally plastic
springs, often referring to the admissible horizontal stiffness and shear resistance specified in EN1991-2 [6].
While the assumption of linear or bi-linear material behavior is convenient for numerical approaches, recent
experimental investigations [26,28] provide important insights into the strictly nonlinear stiffness properties,
which strongly depend on the displacement amplitude, as well as into the frequency-dependent damping
mechanisms of the ballast.
With the aim to partially close the gap between modeling approaches with vertically oriented interaction
models considering train, track, bridge and subsoil and modeling approaches considering longitudinal track–
bridge interaction, this paper extends the vertically oriented models of König et al.[14] and König et al.
[13] with longitudinal ballast interaction. Here, the involved train, track, and bridge-soil subsystems are first
described individually, focusing on the longitudinal track–bridge interaction. The equations of motion of the
track and bridge-soil subsystems are then coupled using component mode synthesis (CMS) [3,14]. The paper
focuses on the modal expansion of the viscoelastically bedded track based on König et al. [14]. Alternatively,
a simplified but computationally more efficient approach based on a Rayleigh–Ritz approximation of the
deflection of the stand-alone track subsystem proposed in König et al. [13] is also presented. By applying a
discrete substructure technique (DST) to the resulting equations of the track–bridge system and the train MSD
system, the full description of the coupled system is derived for both modeling approaches.
The verification of the proposed modal expansion-based approach (Approach 1) is achieved by comparing
the time history results obtained with the presented method with the results of a corresponding FE model.
On the basis of a numerical study, insights into the influence of track–bridge interactions in the longitudinal
direction are provided. Furthermore, the second approach (Approach 2) based on the simplifying Rayleigh-Ritz
approximation of the rail deformation is compared with Approach 1 and the applicability of both methods is
demonstrated.

2 Mechanical model and fundamental equations

2.1 Modeling strategy

The planar mechanical model considered here, representing the dynamically interacting subsystems of the train
moving at constant speed v, the track, and the single-span bridge resting on the subsoil, is shown in Fig. 1.
The conventional train, consisting of Nc individual cars, each consisting of a car body, two bogies, and four
axles, is represented by an MSD system.
Both the slender bridge structure and the rails are modeled as Euler-Bernoulli beams. The beams are
assumed to be rigid in the axial direction, prohibiting longitudinal deformations. The origin x = 0 of the axis
coordinate x is located at the left end of the bridge beam, which has the length L b . Hinged supports are placed
A model considering the longitudinal track–bridge interaction... 1397

Fig. 1 Planar MSD model of the j-th vehicle crossing the track resting on the bridge with viscoelastic supports (modified from
König et al. [14])

at both ends of the (in reality infinitely long) rail beam of length L r for Approach 1. However, the length
L 0 of the beam section before and after the bridge structure must be chosen sufficiently large to prevent the
influence of these “artificial” boundary conditions on the dynamic system response [14]. For the Rayleigh-Ritz
approximation of the infinite rail beam used in Approach 2, these “artificial” boundary conditions are not
required. Regardless of the modeling approach chosen, the influence of the train on the interacting system is
considered only for the finite section of length L r between points A and D.
The stiffness and damping of the ballast are accounted for by continuous vertical and horizontal viscoelastic
bedding of the track. Experiments have shown that these assumptions are appropriate for approximating the
bridge-track system response [17]. In addition, continuous vertical damping is added to the bridge beam to
account for energy dissipation in the ballast due to the absolute motion of the bridge and track as proposed in
Stollwitzer et al. [28].
The lumped mass at the two beam ends representing the bridge structure corresponds to the foundation
mass and a proportional soil mass. The spring-damper elements below the lumped mass model the properties
of the soil, which is in reality a half-space. The spring and damping parameters can be derived, for example,
with the cone model of Wolf [32], requiring the soil density ρs , the constrained modulus E s , the Poisson’s
ratio of the soil ν and foundation base area A0 . Further details on the modeling strategy for the mechanical
subsystems as well as a description of the involved model parameters are provided in the subsequent sections.

2.2 Track–bridge–soil subsystem

2.2.1 Ballast

The interaction between the rails and the bridge structure is primarily determined by the ballast bed. The ballast
bed is flexible and has a major impact on the damping behavior of the system. In the model, these properties
1398 P. König, C. Adam

Fig. 2 Detail of the coupled rail and bridge beam system (left). Interaction model based on the longitudinal and vertical ballast
parameters (center). Longitudinal displacement difference due to rotation of the cross sections and corresponding shear traction
(right)

Fig. 3 Free-body diagram of the rail and bridge-soil subsystems

are represented by viscoelastic elements. Figure 2 shows a detail of the layerwise setup of the coupled bridge-
track structure, the corresponding viscoelastic model of the ballast between bridge beam and rail beam, and a
deformed element according to the subsequently described modeling assumptions. The vertical flexibility and
damping properties of the ballast is accounted for by continuously distributed vertical spring-damper elements
with parameters kz and cz between the bridge and the rail subsystem. With wb (x, t) denoting the vertical
displacement of the bridge and with the rail deflection wr (x, t), the vertical deformation of the ballast layer
w(x, t) becomes
w(x, t) = wb (x, 0, L b ) − wr (1)
The window function (x, 0, L b ) = H (x) − H (x − L b ), which is composed of the two Heaviside step
functions H (x) and H (x − L b ), defines the range of interaction between the track and the bridge of span
length L b . The corresponding distributed vertical interaction force between the bridge beam and the rail beam

F(x, t) = kz w + cz ẇ (2)


is depicted in Fig. 3, which shows the free-body diagram of the rail and the bridge-soil subsystems.
Experiments have shown that there is also in horizontal direction a relative displacement between the rails
and the bridge structure [17,28], which builds up over the ballast height. Since the horizontal displacement is
primarily concentrated at the base of the sleepers, it is assumed that a shear plane forms at the lower edge of
the sleepers [17]. Additionally, it is also assumed that the rotation of the cross-section of the rail beam is also
A model considering the longitudinal track–bridge interaction... 1399

imposed on the ballast layer between the rails and the shear plane, see Fig. 2. Accordingly, the rotation of the
cross-section of the bridge beam is imposed on the ballast layer below the shear plane [17,28]. Thus, in the
model conception, the overall horizontal relative displacement u is formed in the shear plane and composed
of
u(x, t) = u b (x, t) + u r (x, t) (3)
where the displacement component u b is related to the cross-sectional rotation ϕb = −wb,x of the bridge
beam and the displacement component u r to the cross-sectional rotation ϕr = −wr,x of the rail beam as
follows: see Fig. 2,

u b (x, t) = −ϕb (x, t)rb = wb,x (x, t)rb , 0 ≤ x ≤ Lb


(4)
u r (x, t) = −ϕr (x, t)rr = wr,x (x, t)rr , −L 0 ≤ x ≤ L b + L 0
In this relation, rb denotes the distance from the axis of the bridge beam to the base of the sleepers (i.e., the shear
plane) and rr the distance from the base of the sleeper to the axis of the rail beam, see Fig. 2. The resistance to
the horizontal relative displacement between the bridge structure and the rails and the corresponding damping
effect is represented by continuously distributed horizontal springs and dampers with parameters kx and cx at
the height of the shear plane, as shown in Fig. 2. Therefore, the shear traction T in the shear plane depicted in
Fig. 2 is composed of the components Tb and Tr ,
T (x, t) = Tb (x, t) + Tr (x, t) (5)
with

Tb (x, t) = kx u b (x, t) + cx u̇ b (x, t) , 0 ≤ x ≤ L b


(6)
Tr (x, t) = kx u r (x, t) + cx u̇ r (x, t) , − L 0 ≤ x ≤ L b + L 0
Stollwitzer et al. [28] reports that, in addition to the ballast related damping effects due to the ballast
deformation in horizontal and vertical direction, there is also a damping mechanism due to the absolute
displacement of the ballast. This is accounted for in the model by a distributed viscous bedding of the bridge
structure with the damping parameter cob , see Fig. 1. The corresponding distributed damping force
Fob (x, t) = −cob ẇb (x, t) (7)
is also depicted in Fig. 2. The mass of the ballast and the sleeper is added to the mass of the bridge beam.
However, in Sect. 5.4 an example is considered where the sleeper mass is added to the mass of the rail beam,
discussing the possible effects of the additional mass moving together with the rail beam.

2.2.2 Rails

The characteristic parameters of the Euler-Bernoulli beam of length L r = 2L 0 + L b representing the rail
subsystem are given by the effective bending stiffness E Ir of the two rails, their mass per unit length ρ Ar ,
the lateral and longitudinal bedding stiffness coefficients kz and kx , and the lateral and longitudinal bedding
damping coefficients cz and cx . Based on the modeling assumptions, the equation of motion of the rail beam
shown in Fig. 3 can be derived as follows:

ρ Ar ẅr (x, t) + E Ir wr,x x x x (x, t) = f r (x, t) + F(x, t) + (Tb,x (x, t) + Tr,x (x, t))rr (8)
It is noted that the shear traction due to the relative horizontal deformation of the ballast enters the above
equation in the form (Tb,x (x, t) + Tr,x (x, t))rr , since it is applied at the distance rr from the beam axis. The
interaction forces resulting from the contact of train wheels (axles) with the rails read as [14]
( j)

Nc 
na
( j)  ( j)  ( j) ( j)
f r (x, t) = Fk (t)δ x − xk (t) (t, tAk , tDk ) (9)
j=1 k=1

( j) ( j) ( j)
where xk = vt −lk is the axle position of the k-th axle of the j-th vehicle at time t. The variable lk denotes
the initial distance of the k-th axle of the j-th vehicle from x = −L 0 , see Fig. 3. The position of the k-th
( j) ( j)  ( j) 
axle load of the j-th vehicle Fk at xk is defined by the Dirac-delta function δ x − xk (t) . The variable
1400 P. König, C. Adam

( j) ( j) ( j) ( j)
(t, tAk , tDk ) = H (t − tAk ) − H (t − tDk ) denotes the time range in which the respective axle load crosses
( j) ( j)
the track between point A and point D. The variables tAk and tDk denote the time instances at which the axle
load arrives at point A and leaves the rail beam at point D (points A, B, C, D are indicated in Figs 1 and 2).
The boundary conditions representing the hinged supports at points A and D, added for Approach 1, read
as

wr (x = −L 0 ) = wr (x = L b + L 0 ) = 0
wr,x x (x = −L 0 ) = wr,x x (x = L b + L 0 ) = 0 (10)

2.2.3 Bridge, foundation and subsoil

With the constant flexural rigidity E Ib , mass per unit length ρ Ab and the span L b , the equation of motion of
the bridge beam subjected to the vertical interaction force F(x, t) and the horizontal shear traction T (x, t) at
distance rb from the beam axis (see Fig. 2) is derived as

ρ Ab ẅb (x, t) + E Ib wb,x x x x (x, t) = −F(x, t) + T,x (x, t)rb − Fob (x, t) , 0 ≤ x ≤ L b (11)

The lumped mass and the spring-damper element at both boundaries, which represent the soil properties,
yield the following boundary conditions [12]

(x = 0) : m b ẅb (0, t) + cb ẇb (0, t) + kb wb (0, t) + E Ib wb,x x x (0, t) = 0,


wb,x x (0, t) = 0
(12)
(x = L b ) : m b ẅb (L b , t) + cb ẇb (L b , t) + kb wb (L b , t) − E Ib wb,x x x (L b , t) = 0,
wb,x x (L b , t) = 0

2.3 Train subsystem

As an example, the planar model of a conventional train described in detail in Salcher and Adam [24] is
considered. In this model, each of the j = 1, ..., Nc vehicles has ten degrees of freedom (DOFs). As shown
in Fig. 1, these vehicles consist of a car body (whose variables are dentoed by the subscript“p”), two bogies
(subscript“s”) and four axles (subscript“a”). Herein, four vertical translational DOFs describe the axle dis-
( j) ( j)
placements (u al , l = 1,..., 4), two vertical translational DOFs represent the displacements of the bogies (u s1 ,
( j) ( j)
u s2 ) and one vertical translational DOF is used for the displacement of the car body, u p . The rotational
( j) ( j)
DOFs considered in the present model involve the rotations of the bogies (ϕs1 , ϕs2 ) and the rotation of the
( j) ( j) ( j) ( j)
car body, ϕp . The corresponding masses are m a , m s and m p . The moments of inertia are denoted by
( j) ( j)
Is and Ip . Furthermore, the parameters of the spring-damper elements between the axles and boogies are
( j) ( j)
the spring stiffness ka and the damping parameter ca . Accordingly, stiffness and damping parameters of
( j) ( j)
elements connecting the car body and the bogies are denoted as ks and cs , respectively. The geometrical
( j) ( j)
parameters defining the positions of the spring-damper elements are h a and h s . The equations of motion of
the complete train [24]

Mc üc + Cc u̇c + Kc uc = Fc (13)


result from adding the equations of motion of all vehicles. Consequently, the mass, damping and stiffness matri-
ces Mc , Cc , Kc and the vector of degrees of freedom uc are composed of the corresponding matrices/vectors
( j) ( j) ( j) ( j)
of the individual vehicles Mc , Cc , Kc and uc , j = 1, ..., Nc , as
   
Mc = diag Mc(1) , ..., Mc( j) , ..., Mc(Nc ) , Cc = diag C(1) c , ..., C ( j)
c , ..., C(Nc )
c
   T (14)
Kc = diag Kc(1) , ..., Kc( j) , ..., Kc(Nc ) , uc = uc(1) , ..., uc( j) , ..., uc(Nc )
A model considering the longitudinal track–bridge interaction... 1401

The vector of interaction forces Fc between the train axles and the rails is also composed of the individual
( j)
vectors of interaction forces Fc of all vehicles,
 T
Fc = Fc(1) , ..., Fc( j) , ..., Fc(Nc ) (15)
The system matrices, vector of degrees of freedom and interaction force vector for the j-th vehicle can be
found in Salcher and Adam [24]; Hirzinger et al. [12].
The state-space representation of Eq. (13), needed for the coupling procedure based on König et al. [14],
is given by (e.g., Hirzinger et al. [12])
Ac ḋc + Bc dc = fc (16)
with        
Cc Mc Kc 0 Fc uc
Ac = , Bc = , fc = , dc = (17)
Mc 0 0 −Mc 0 u̇c

3 Coupling of the subsystems

In the following, two different strategies are used for coupling the subsystems. In the first approach, hereafter
referred to as Approach 1, the modally expanded rail and bridge subsystems are coupled with a CMS procedure
as described in König et al.[14]. The second approach, hereafter referred to as Approach 2, involves a simplified
description of track deflection based on a Rayleigh-Ritz approximation, similar to König et al. [13]. Since these
coupling strategies are already described in detail in the aforementioned publications König et al. [14] and
König et al. [13], the procedure will be described only briefly, pointing out the main differences with these
references.

3.1 Modal series expansion of the bridge-soil response

Both strategies require the approximation of bridge displacement wb (x, t) by a complex modal series expansion
in Nb modes in the form of

Nb
(m) (m)

Nb
(m) (m)
wb (x, t) ≈ yb (t)b (x) + ¯ (x)
ȳb (t) (18)
b
m=1 m=1

(m) (m)
where yb (t) is the m-th modal coordinate of the non-classically damped bridge-soil subsystem and b (x)
(m) ¯ (m) (x) represent the correspond-
denotes the corresponding complex eigenfunction. The variables ȳb (t) and  b
ing complex conjugate counterparts. For the derivation of the underlying modal properties of the stand-alone
bridge-soil subsystem based on its homogeneous equation of motion, given by

ρ Ab ẅb + cob ẇb + E Ib wb,x x x x = 0 (19)


it is referred to König et al. [14]. The only notable difference to König et al. [14] is the complex natural
(m)
frequency sb ,
(m) 4
λb E Ib
−cob ± 2 − 4ρ A
cob b L 4b
(m)
sb = (20)
2ρ Ab
which includes the damping parameter cob capturing the damping effect due to the absolute motion of the
(m)
ballast, also present in Eq. (19). The parameter λb corresponds to the m-th eigenvalue [14]. The generalized
(m)
modal mass Mb , derived from the orthogonality relations of the non-classically damped bridge-soil system
reads as [7,15]

(m) 1 (m) (m) (m) (m)


Mb = (m)
cob Jb0 + cb Pb + ρ Ab Jb0 + m b Pb
2sb
1402 P. König, C. Adam

⎛ ⎞
 
1⎜ 1 (m) (m) (m) (m) ⎟
= ⎝ E Ib Jb2 + kb Pb + ρ Ab Jb0 + m b Pb ⎠ (21)
2 (m) 2
−sb

with

(m)  Lb (m) 2 (m)  Lb (m) 2


Jb0 = 0 b (x) dx , Jb2 = 0 b,x x (x) dx
(m) (m) 2 (m) 2
Pb = b (0) + b (L b ) (22)

(m) (m)
The normalizing constants ab and bb of the associated orthogonality relations are related to the generalized
modal mass in the form of [7,15]
(m) (m) (m) (m) (m) 2 (m)
ab = 2sb Mb , bb = −2 sb Mb (23)

3.2 Rail response

As described in Biondi et al. [3], it is useful to separate the rail deflection wr (x, t) into two fractions,

wr (x, t) = wr(b) (x, t) + wr(f) (x, t) (24)


(b)
The first contribution to the rail deflection wr (x, t) corresponds to the rail response due to interaction forces
resulting from the ballast deformation from the bridge displacement wb (x, t). It represents therefore the
coupling term between the rail and the bridge-soil subsystems. According to Biondi et al. [3] and König et al.
(b)
[14], wr (x, t) is approximated as


Nb
(m)

Nb
(m)
wr(b) (x, t) ≈ yb (t) (m)
r (x) + ȳb (t) ¯ r(m) (x) (25)
m=1 m=1

(m) (m) (m)


where the modal coordinates yb (t) and ȳb (t) are multiplied by a corresponding shape function r (x)
and its complex conjugate ¯ r(m) (x), respectively. This shape function is the solution of an ordinary differential
equation resembling the quasi-static version of the equation of motion Eq. (8) [3],
(m) (m) (m)
r,x x x x (x) − rr kx r,x x (x) + kz r (x)
2
E Ir
(m) (26)
= rbrr kx b,x x (x) + kz b (x) (x, 0, L b )
The solution is found numerically, taking into account the boundary conditions of the rail beam (see Eq. (10)).
(f)
The second part wr (x, t) to the response is the deflection of the stand-alone track in response to the axle
(f)
loads. Depending on the coupling strategy, wr (x, t) is approximated differently, as discussed below.

3.2.1 Modal series representation of the rail response for Approach 1


(f)
As shown in König et al. [14], the response wr (x, t) of the stand-alone rail bream can be approximated by
the complex modal series expansion in Nr modes,


Nr 
Nr
wr(f) (x, t) ≈ yr(n) (t)(n)
r (x) +
¯ (n)
ȳr(n) (t) r (x) (27)
n=1 n=1

where (n)
r (x) and  ¯ (n)
r (x) represent the n-th complex eigenfunction and its complex conjugate counterpart,
(n) (n)
and yr (t) and ȳr (t) are the n-th modal coordinate and its complex conjugate counterpart, respectively, of
A model considering the longitudinal track–bridge interaction... 1403

the rail beam. Since the rail beam is simply supported (see Fig. 1) with constant bedding stiffness and damping
values, the general form of the n-th eigenfunction given in König et al. [14] simplifies to a pure sine function,
(n)
λr (x + L 0 )
(n)
r (x) = C
(n)
sin , λ(n)
r = nπ (28)
Lr
(n)
where λr is the eigenvalue of the n-th mode and C (n) is an arbitrary scaling factor. The eigenfunctions are
(n) (n)
real-valued because the eigenvalue λr is also real-valued. However, the natural frequency sr is complex
(n)
valued, and related to the corresponding eigenvalue λr as
 
(n)  (n) 2  (n) (n) 
(n)
− cz + X c ± cz + X c − 4ρ Ar kz + X k + X EI
sr = (29)
2ρ Ar
with
(n) (n) (n)
(λr )2 rr2 cx (n) (λr )2 rr2 kx (n) (λr )4 E Ir
X c(n) = , X k = , X EI = (30)
L 2r L 2r L 4r
The orthogonality conditions for the eigenfunctions yield the n-th generalized modal mass Mr(n) as [7,13,15]
1 (n) (n) (n)
Mr(n) = cz Jr0 + rr2 cx Jr1 + ρ Ar Jr0
2sr(n)
    (31)
1 1 (n) (n) (n) (n)
= − 2
kz Jr0 + rr kx Jr1 + E Ir Jr2 + ρ Ar Jr0
2
2 (n)
sr
with
(n)  L b +L 0 (n) 2 (n)  L b +L 0 (n) 2
Jr0 = −L 0 r (x) dx , Jr1 = −L 0 r,x (x) dx
(n)  L b +L 0 (n) 2
Jr2 = −L 0 r,x x (x) dx (32)
(n) (n)
The scaling constants ar and br can again be expressed in terms of the generalized modal mass [7,15],
2
ar(n) = 2sr(n) Mr(n) , br(n) = −2 sr(n) Mr(n) (33)

3.2.2 Rayleigh-Ritz approximation of the rail response for Approach 2


(f)
In the alternative solution strategy, the deflection wr (x, t) is approximated by a Rayleig-Ritz approach, where
the static displacement of the infinitely long, elastically bedded beam due to a concentrated load serves as a
shape function [13],

Na
wr(f) (x, t) ≈ ϕr∗ (x − x p (t))yr p (t) (34)
p=1
Here, x p (t) refers to the position of the p-th axle of the Na axles of the whole train ( p = 1, ..., Na ), yr p (t)
denotes the vertical displacement at the p-th axle and ϕr∗ (x − x p (t)) is the shape function corresponding to
the static deflection centered around the p-th axle. With the local spacial coordinate x̂ p = x − x p (t), the static
deflection of the vertically bedded rail beam due to a concentrated load reads as [11,13]

ϕr∗ (x̂ p ) = e−β x̂ p (sin β x̂ p + cos β x̂ p ) , x̂ p ≥ 0


kz
β= 4
(35)
4E Ir
As can be seen, with these symmetric shape functions the influence of the longitudinal bedding of the
rail beam with stiffness kx is not taken into account. Therefore, this approximation cannot be considered as
a full replacement of Approach 1, but it represents a very efficient alternative if appropriately verified in the
parameter space of interest.
1404 P. König, C. Adam

3.3 Component mode synthesis in Approach 1

Application of CMS on the rearranged equations of motion of the bridge-soil and track substructures (Eq. (11)
and (8)),      
ρ Ab 0 ẅb (x, t) E Ib 0 wb,x x x x (x, t)
+
0 ρ Ar ẅr (x, t) 0 E Ir wr,x x x x (x, t)
     
kf −kf wb (x, t) cf + cob −cf ẇb (x, t)
+ +
−kf kf wr (x, t) −cf cf ẇr (x, t)
 2    2   (36)
rb kx rb rr kx wb,x x (x, t) rb cx rbrr cx ẇb,x x (x, t)
− −
rbrr kx rr2 kx wr,x x (x, t) rbrr cx rr2 cx ẇr,x x (x, t)
 
0
=
fr
as described in König et al. [14] ultimately leads to the equations of motion of the coupled track–bridge–soil
subsystem in terms of modal coordinates in state-space [14],
AB ḣB + BB hB = fB (37)
Herein the vectors hB and fB denote the vector of modal coordinates and the force vector, respectively. These
vectors are specified in König et al. [14]. The system matrices AB and BB are given by
   
Ab + Mb Sb + Cb Mbr Sr + Cbr Bb + Kb Kbr
AB = , BB = (38)
Mrb Sb + Crb Ar Krb Br
and are composed of individual sub-matrices as shown in König et al. [14]. The sub-matrices are listed in
Appendix A.
Similar to the coupling procedure in state-space used in König et al. [14], a representation closer related
to the formulation used in König et al. [13] can be derived in the form of
MB ḧB + CB ḣB + KB hB = fB (39)
with the system matrices given by
   
0 Mbr Bb + Kb Kbr
MB = , KB = ,
Mrb 0 Krb Br
 
Ab + Cb + Mb Sb Cbr
CB = (40)
Crb Ar

It is emphasized that both formulations of Eq. (37) and (39) can be used interchangeably. However, the
present implementation of Approach 1 is based on the formulation of Eq. (37) and the subsequent coupling
procedure of König et al. [14], while Eq. (39) has been added only to illustrate the similarities and differences
with Approach 2, which is described below.

3.4 Component mode synthesis in Approach 2

In contrast to the state-space representation of the coupled track–bridge–soil subsystem equations in Approach
1 (Eq. (37)), in Approach 2 the coupled system equations are provided in standard representation as [13]

M h¨ B (t) + 
 B (t) h˙ B (t) + K
CB (t) B (t)
hB (t) = 
fB (t) (41)
 B (t), 
The system matrices M CB (t) and KB (t) are composed of sub-matrices
   
 br (t) 
M B (t) = 0 M
, B (t) = Bb + Kb Kbr (t) ,
K
 rb (t) Mr
M rb (t)
K Kr
 
A + Cb + Mb Sb Cbr (t) 

CB (t) = b (42)
Crb (t) Cr
A model considering the longitudinal track–bridge interaction... 1405

Fig. 4 Corresponding assumption coupling axle and rail displacements (modified from Hirzinger et al. [12])

given in Appendixes A and B. The vector  hB (t) comprises both the modal coordinates of the bridge beam and
the generalized coordinates of the rail beam (see König et al. [13]). The force vector  fB (t) is given in König
(f)
et al. [13]. The difference in the approximation of the rail deflection component wr in the two approaches
leads to differences in the sub-matrices in Eq. (38) (or Eq. 40) and Eq. (42), in the coupling matrices between
the rail and the bridge (subscripts “br”and “rb”), and in the matrices involving the time-dependent coordinates
of the rail (subscript “r”). For these subsystem matrices, please refer to Appendix B.

4 Coupled equations of the complete system

The coupling of the train with the track–bridge–soil subsystem is based on the assumption of a rigid contact
between wheel and rail according to the so-called corresponding assumption [36].
As shown in Fig. 4, also an irregularity profile function Iirr (x) can be added to the track deflection, to
account for random track irregularities in vertical direction, i.e., deviations from the perfectly straight and
smooth track.

4.1 Coupled equations of motion in Approach 1

In analogy to König et al. [14], a DST is applied to condense the equations of motion Eq. (13) of the train
subsystem and the equations of motion Eq. (37) of the track–bridge–soil subsystem to a coupled set of equations
of motion, with the result [14]
A(t)ẋ(t) + B(t)x(t) = f(t) (43)
where A(t) and B(t) are the time-dependent system matrices of the fully coupled system of train, track, bridge
and subsoil. The vectors x(t) and f(t) denote the vector of system coordinates and the force vector, respectively.
Due to the coupling procedure based on the corresponding assumption, the DOFs related to the displacements
of the train axles are condensed into the modal coordinates of the rail beam. Thus, the vector x(t) is composed
of the modal coordinates of the bridge-soil subsystem, the modal coordinates of the rail beam, and the DOFs
of the train related to the displacements and rotations of the car bodies and bogies of the individual vehicles.
For a more detailed description of the coupling procedure, reference is made here to König et al. [14]. This set
of equations is solved numerically by application of the Runge–Kutta method.

4.2 Coupled equations of motion in Approach 2

Application of DST as described in König et al. [13] to the equations of motion Eq. (13) of the train subsystem
and the equations of motion Eq. (41) of the track–bridge–rail subsystem according to Approach 2 finally yields
the single set of coupled equations
¨ + C(t)
M(t)
x(t) x(t) x(t) = 
˙ + K(t) f(t) (44)
where M(t), C(t) and K(t) denote the time-dependent mass, damping and stiffness matrices of the coupled
x(t) and 
system. The vectors  f(t) are the vectors of the system coordinates and forces of Approach 2. Similar
1406 P. König, C. Adam

Table 1 Parameters of the bridge beam [13] and the rail beam [21]

Bridge parameters Value Rail parameters Value Unit


Lb 17.5 L0 30 m
E Ib 1.356 · 1010 E Ir 12.831 · 106 Nm2
ρ Ab 7830 ρ Ar 120.73 kg/m

Fig. 5 Two degree of freedom vehicle and irregularity profile superimposed to the rail surface

to Approach 1, the coupling procedure of Approach 2 condenses the axle DOFs of the train into DOFs of the
rail beam. However, the difference is that the rail DOFs are now represented by the time-dependent coordinates
of the Rayleigh-Ritz approximation instead of modal coordinates. To solve Eq. (44), the Newmark-β method
has proven to be efficient.

5 Computations

5.1 Verification

Considering the additional approximations introduced in Approach 2, the verification of the presented modeling
strategies is divided into two parts. In this section, a rigorous verification of Approach 1 is performed by
comparing the results of the developed MATLAB [18] code with the outcomes of finite element (FE) analyses
using the software suite ABAQUS [1]. Approach 2 is verified in the subsequent section by comparison with
representative results from Approach 1.
The bridge considered in all computational examples is the same steel bridge as considered in König et
al. [13]. The parameters of the bridge beam as well as the parameters of the rail beam consisting of two 60E1
(UIC60) rails are listed in Table 1.
For the vertical stiffness and damping coefficients of the ballast, the fixed values kz = 104 · 106 N/m2 and
cz = 50 kNs/m2 are chosen, as given in Yang et al. [33]. The eccentricities for the longitudinal interaction
model are rb = 0.7 m for the bridge beam and rr = 0.3 m for the rail beam. The base values of the horizontal
ballast stiffness and damping coefficients are kx = 10.4 · 106 N/m2 and cx = 50 kNs/m2 , but are also varied to
verify the model for a wider range of parameters. Model verification involving the soil–structure interaction
of the bridge and underlying soil have already been performed in König et al. [14] and König et al. [13].
Since the focus is on verifying the track–bridge interaction model in the longitudinal direction, the influence
of underlying soil is therefore omitted and the bridge is considered as a beam on rigid supports.
The described track–bridge system is crossed by a simple MSD system representing the interaction model
with respect to one axle of the ICE 3 train model moving at a constant speed of v = 70 m/s. This interaction
model, shown in Fig. 5, is characterized by the parameters given in Table 2. Figure 5 also shows the vertical
irregularity profile superimposed to the perfectly straight rail geometry. This irregularity profile was randomly
generated to represent a track of poor quality, based on the methodology presented in Claus and Schiehlen [5].
The FE model represents an equivalent interaction system using the same input parameters and is visualized
in Fig. 6. Both, the bridge and the rail beam are modeled using Euler-Bernoulli beam elements (B23) with a
uniform mesh-length of 6.125 cm. The continuous vertical bedding is idealized by discrete spring and dashpot
elements at each node of the beam elements. The eccentricities for the longitudinal interaction are modeled
A model considering the longitudinal track–bridge interaction... 1407

Table 2 Parameters of two degree of freedom vehicle model [14,19], for the variables see Fig. 5

Parameter Value Unit


ms 15125 kg
ma 1800 kg
ka 1.764 · 106 N/m
ca 4.800 · 104 Ns/m

Fig. 6 Schematic of FE model with used element types

with rigid massless beam elements (RB2D2), equally spaced at a distance of 6.125 cm, and connected with
horizontally oriented spring and dashpot elements representing the longitudinal bedding. This configuration
is shown in Fig. 6, where it should be noted that the longitudinal offset has been added for illustration purposes
only and the respective element nodes of the rail and bridge beams are situated at the same positions in the
longitudinal direction. Hence, in the actual FE model, the nodes of the rigid beam elements of the rail beam
and bridge beam, which are connected by longitudinal spring and damper elements, are situated at the same
positions inside the shear zone. Furthermore, the two DOF system is modeled with rigid body lumped masses
(Mass) connected by a spring and a damper element, respectively. In the FE model, rigid contact between
the moving axle and the rail beam was implemented using ABAQUS tube-to-tube contact elements (ITT21).
The deviation of the rail geometry from a perfectly straight line is accounted for by offsetting the rail beam
nodes perpendicular to the straight beam axis by the amount of the irregularity profile of Fig. 5. To keep Fig. 6
relatively simple, this deviation of the rail nodes in the vertical direction is not illustrated. It is also noted that
the relatively small element size of 6.125 cm was used to sufficiently approximate the irregularity profile along
with the continuous bedding. In the verification examples, a constant time increment of t = 10−4 s is used.

5.2 Variation of the horizontal stiffness

To verify the prediction of the system response, several computations were performed with different horizontal
stiffness coefficients. In the computations with the proposed approach (Approach 1), the bridge displacement
is approximated by Nb = 6 modes. The deflection of the stand-alone track is approximated by Nr = 100
modes, emphasizing the need to include such a large number of track modes to adequately predict the system
response (see also König et al. [14] and König et al. [13]). In Fig. 7, the bridge response at the quarter point
(Fig. 7a, b) and the center point (Fig. 7c, d), predicted by the semi-analytical approach (“Approach 1; ref.
sol.”), is compared with the results of the FE model (“FEM; ref. sol.”). The dashed vertical lines indicate the
time instants at which the MSD system arrives at the left bridge support (tB ) and leaves the bridge at the right
support (tC ). As can be seen, the bridge displacements computed with the proposed approach perfectly match
those of the FE model, see Fig. 7a, b). This excellent agreement is also observed for the accelerations (Fig. 7c,
d).
1408 P. König, C. Adam

Fig. 7 Approach 1 versus FE analysis. Two different values of horizontal ballast stiffness kx . a Bridge displacement and b
acceleration at x = L b /4. Second row: c Displacement and d acceleration at x = L b /2

To further verify the modeling approach, another set of computations was performed in which the horizontal
ballast stiffness was set to ten times its original value (10kx = 104 · 106 N/m2 ). Comparing the respective lines
representing the results of the proposed approach and those of the FE model, again very good agreement is
observed in the bridge displacements and accelerations. The comparison of the two sets of computations with
different horizontal stiffness values clearly illustrates the stiffening effect related to the longitudinal interaction
in the reduced displacement amplitude as well as in the reduced period of the free vibration response after the
MSD system has left the bridge.

5.3 Variation of the damping

The verification of the response prediction taking into account the damping due to the longitudinal track–bridge
interaction is done in the same manner as for the horizontal stiffness. The results of the computations with
the base value parameter set already shown in Fig. 7 are shown again in Fig. 8 as a reference. Since the track
irregularities were found to have a greater impact on the bridge response at the quarter point, Fig. 8 shows results
at this position only. The base value of horizontal damping was also multiplied by ten (10cx = 500 kNs/m2 ).
Again, very good agreement is observed of the predicted bridge displacement (Fig. 8a) and acceleration
(Fig. 8b), further verifying the proposed approach.
Next, the consideration of the additional damping parameter cob , which is the last ballast-related damping
mechanism reported in Stollwitzer et al. [28], is verified. While this parameter was omitted in the above
computations, Fig. 9 also shows results where this damping parameter is set to cob = 8.7 kNs/m2 (“with cob ”),
again showing excellent agreement between the proposed Approach 1 and the FE model.
From the verification examples presented here, it can be concluded that the contribution of the additional
damping mechanisms in the horizontal and vertical directions has also been successfully integrated into the
semi-analytical modeling approach.
A model considering the longitudinal track–bridge interaction... 1409

Fig. 8 Approach 1 versus FE analysis. Two different values of horizontal ballast damping cx . a Bridge displacement and b
acceleration at x = L b /4

Fig. 9 Approach 1 versus FE analysis. Consideration of ballast damping cob . a bridge displacement and b acceleration at x = L b /4

5.4 Application problems

Considering the large uncertainties in the ballast parameters in combination with the highly nonlinear properties
reported in the literature, a thorough parametric study is presented for the application example, where the
influence of the model parameters is further discussed. For this purpose, the dynamic response of the example
bridge, which was also used for verification, is further analyzed. The base parameters of the bridge and track
are the same as in Sect. 5.1. However, taking into account structural damping of the bridge, the modal damping
coefficient of ζ̃b(m) = 0.8125% according to [6] for a steel bridge of the given length is assigned to each mode of
(m)
the rigidly supported bridge. This is achieved by modulating the complex natural frequency sb as described
in König et al. [14]. In contrast to the simple MSD system used for the verification in 5.1, a more thorough
representation of the Nc = 8 vehicles of the ICE 3 train model is considered. Each vehicle is represented by
the same ten DOF system with parameters (compare Fig. 1), which include the masses (m p , m s , m a ), moments
of inertia (Ip , Is ), stiffness coefficients (ks , ka ), damping coefficients (cs , ca ), and dimensions (h s , h a ), given in
König et al. [14] and Nguyen et al. [19]. All analyses discussed below involve individual time history analyses
of train crossings at various train speeds from 10 to 90 m/s at uniform intervals of v = 0.5 m/s. The bridge
response is evaluated at 101 uniformly distributed points along the track, and representative maximum absolute
bridge displacements (max|wb |) and maximum absolute bridge accelerations (max|ẅb |) are computed for each
considered train speed. These response quantity are then presented in the form of response spectra in Figs 10
to 14 for a variety of input parameter configurations.

5.4.1 Variation of horizontal ballast parameters

As in the verification examples, the horizontal stiffness and damping parameters kx and cx are varied. The
parameter variation is performed in the range of 0.5 to 10 times the respective base value. In this section, the
1410 P. König, C. Adam

Fig. 10 Approach 1. Different values of horizontal ballast stiffness kx . Response spectra. a Maximum absolute displacement and
b maximum absolute acceleration

influence of track irregularities is not considered. Response spectra of maximum absolute bridge deflection
(max |wb |) and maximum absolute bridge acceleration (max |ẅb |) resulting from variation of kx are depicted
in Fig. 10. Here, the results derived from computations with the base values of the input parameters are
represented by the black lines and the blue lines represent results with different horizontal stiffness. In Fig. 10
(a), the increase in the system stiffness of the complete coupled system associated with the increase in kx is
clearly visible, both in the decrease in vertical displacements and in the shift of response peaks toward higher
train speeds.
The peaks in the response spectra apparent in this figure are related to critical train speeds, at which the
system is excited to resonance by the crossing train. One reason of resonance is the uniform axle spacing of the
individual vehicles of the train crossing the bridge at constant speed. The corresponding so-called second-order
(m)
resonance speeds vi are defined as follows for a bridge without a track [33]:
(m)
(m) dc f b
vi = , m = 1, .., Nb , i = 1, ... (45)
i
For the considered ICE 3 train model, the regular spacing is dc = 24.775 m. Estimating the natural frequencies
of the track–bridge model with the natural frequencies of the stand-alone simply supported bridge beam [9],
 2
(m) m π E Ib
f b,rigid = , m = 1, 2, ... (46)
Lb 2 ρ Ab
reveals that the most pronounced response peak in Fig. 10 is related to the first mode (m = 1) at the resonant
speed of v3(1) = 58.61 m/s. Therefore, the shift in response peaks with increasing horizontal ballast stiffness kx is
due to the increased natural frequencies of the coupled track–bridge system compared to the natural frequencies
of the stand-alone bridge from Eq. (46). In both cases, the displacement response quantity max |wb | and the
acceleration response quantity max |ẅb |, a decrease in the overall response is observed. However, this effect
is less pronounced in the acceleration response.
Similar to the results with variation of the horizontal stiffness in Fig. 10, the results with variation of the
damping parameter cx are shown in Fig. 11. Here, the decrease in resonance peaks with increasing horizontal
ballast damping is clearly visible, while, as expected, no significant changes of the respective resonance speeds
are observed.

5.4.2 Effect of track irregularities

In view of the results of the previous section derived from the variation of the horizontal ballast parameters, track
irregularities are considered in the following computations in the form of the irregularity profile Iirr (x) depicted
in Fig. 5. Since the influence of ballast parameters is likely to be even more pronounced when considering the
imperfect track geometry, different values of kx and cx are also considered here. However, computations are only
performed using the base values of kx and cx and ten times these base values, respectively. The corresponding
A model considering the longitudinal track–bridge interaction... 1411

Fig. 11 Approach 1. Different values of horizontal ballast damping cx . Response spectra. a Maximum absolute displacement and
b maximum absolute acceleration

Fig. 12 Approach 1 versus Approach 2. Track irregularities considered. Two different values of the horizontal ballast stiffness
kx . Response spectra. a Maximum absolute displacement and b maximum absolute acceleration

results are shown in Fig. 12 for the different values of kx . In addition to the results derived with Approach
1, Fig. 12 also contains results derived with the simplified Approach 2 to also verify the latter method for
cases where the influence of the longitudinal track–bridge interaction is most pronounced. Examining Fig. 12
(a), effects of the horizontal stiffness parameter are similar to Fig. 10 (a) and only a small effect of the track
irregularities on the maximum absolute displacement max |wb | is visible when comparing the corresponding
lines in Figs 10 (a) and 12 (a).
However, inspection of Fig. 12 (b) shows the significant influence of track irregularities on the acceleration
response quantity max |ẅb |. Moreover, the tendency observed in Fig. 10 (b) for the acceleration response peaks
to decrease with increasing horizontal ballast stiffness is present only for response peaks below v = 50 m/s, and
the most prominent response peak increases rather dramatically when a high ballast stiffness of kx is considered.
Apart from that, an overall good agreement between the results of Approach 1 and Approach 2 is observed
in Fig. 12. However, at higher train speeds, minor deviations between the two approaches are observed. This
may be attributed to the simplifications of the Rayleigh-Ritz approximation introduced in Approach 2, as well
as to a larger influence of the track model on the bridge response at higher speeds when track irregularities are
considered. This slight disagreement in results was also observed for the purely vertically oriented model in
König et al. [13].
It is also evident from Fig. 12 (b) that, compared to Fig. 10 (b), an increased influence of the longitudinal
stiffness parameter on the computed bridge accelerations is observed when additional track irregularities are
considered. However, the bridge displacements appear to be affected by the longitudinal stiffness parameter
to a similar extent with or without track irregularities, see Figs 10 (a) and 12 (a). This may be due to the fact
that the vertical bridge accelerations are generally more sensitive to amplifications due to track irregularities
than the vertical displacements (see, e.g., Salcher and Adam [25]). Together with the additional amplified
1412 P. König, C. Adam

Fig. 13 Approach 1 versus Approach 2. Track irregularities considered. Two different values of the horizontal ballast damping
cx . Response spectra. a Maximum absolute displacement and b maximum absolute acceleration

relative movement between bridge and track, this is likely the cause of the overall greater influence of the track
parameters on the acceleration response of the bridge in the presence of track irregularities.
For completeness, Fig. 13 includes results of analyses with ten times the base value of the longitudinal ballast
damping cx , where the effect of the additional damping becomes visible again. For all examples in Figs 12
and 13, Approach 2 agrees quite well with the values derived with Approach 1, justifying the simplifications
introduced in Approach 2.

5.4.3 Influence of soil–structure interaction

The last example comprises the full interaction model considering the soil–structure interaction based on
the same soil and foundation parameters also used in König et al. [13]. These are the constrained modulus
E s = 2.5 · 108 N/m2 of the soil, the soil density ρs = 2300 kg, Poisson’s ratio ν = 0.28, base area of
the foundation A0 = 40 m2 and the foundation mass m b = 2.5 · 105 kg. Application of the cone model of
Wolf Wolf and Deeks [32] yields the following equivalent subsoil coefficients, i.e., the stiffness coefficient
kb = 1.514 · 109 N/m and the damping coefficient cb = 3.033 · 107 Ns/m [13].
Considering the frequency- and displacement-dependent characteristics of the longitudinal ballast prop-
erties reported from experiments [26,28], the problem of appropriate selection of equivalent linear model
parameters kx and cx remains. For the damping parameter, it is easy to argue that a minimum value should
be chosen to avoid unrealistically high damping, which subsequently underestimates the predicted vibration
response of the bridge. Such a value is also recommended in Stollwitzer et al. [26] and given as cx = 62 kNs/m2 .
However, the selection of a suitable value for kx is not as straight forward. In addition to the relative displace-
ment amplitude within the assumed shear plane in the ballast and the excitation frequency, the loading by the
axle loads of the train and the ballast condition (frozen or unfrozen) also strongly influence the longitudinal
stiffness of the ballast. Therefore, the computational example presented here covers only the case of unfrozen
ballast, where the longitudinal stiffness parameter is estimated based on measurements on a loaded track. For
this case, Stollwitzer et al. [26] proposes the following regression formula for the estimation of kx ,
kx = −9.03 ln (u 0 ) + 23.92 [kN/mm/m] , u 0 in mm (47)
where u 0 denotes the relative displacement amplitude in longitudinal direction. Additional regression formulas
for unloaded and frozen states are also specified in [26]. Since u 0 is not constant over the length of the bridge,
an equivalent displacement amplitude u 0eq must be used. In the same context, such an equivalent displacement
amplitude is given in Stollwitzer and Fink [27] as

π (rr + rb )2
u 0eq = wb0 eb + (48)
Lb 2
based on the fundamental mode shape in form of a sine half wave and strain energy considerations. Herein wb0
denotes the bridge deflection at the center and eb the additional support eccentricity of the bridge. Since the
displacement amplitude varies quite significantly for different values of kx , the deflection at the center is simply
A model considering the longitudinal track–bridge interaction... 1413

Fig. 14 Response spectra—full interaction model. a maximum absolute deflection and b maximum absolute acceleration

estimated by taking the average of the peak values of the maximum absolute deflection of the two response
spectra shown in Fig. 10 with the lowest and highest horizontal stiffness, resulting in w0eq ≈ 6 · 10−3 m.
Since no additional support eccentricity eb is considered in the present model, the equivalent displacement
amplitude is obtained as u 0eq ≈ 0.76 · 10−3 m. Consequently, substituting u 0eq into Eq. (47) leads to an
estimated longitudinal stiffness coefficient kx = 26.40 · 106 N/m2 .
In addition to the vertical and horizontal ballast coefficients, ballast-related damping due to absolute
movement of the ballast can be accounted for in the form of the damping parameter cob . Based on a regression
for the damping coefficient cob , which is also frequency dependent [27], and the first natural frequency of the
(1)
stand-alone rigidly supported bridge f b,rigid = 7.10 Hz, the damping parameter is estimated as cob = 8.7 · 103
Ns/m2 .
Given the detailed interaction model and the damping mechanisms involved due to soil and ballast model,
(m)
applying the full structural damping of ζ̃b = 0.8125% could result in an uncertain response prediction
because the system damping is overestimated. Therefore, only the span-independent amount of structural
damping, given in EN1991-2 [6] as ζ̃b(m) = 0.5%, is assigned to the modes of the bridge beam.
Figure 14 shows the bridge response spectra of the full model, including the track irregularity profile. Here,
the solid lines represent the results without and with considering the damping parameter cob (“wo. cob ”and“w.
cob ,” respectively). To facilitate visual comparison of these results with the previous results for the rigidly
supported bridge, Fig. 14 (a) shows the maximum absolute value of the bridge displacement relative to the
straight bridge axis max |wb,rel | defined as [13]
max |wb,rel | = max |wb (x, t) − (wb (L b , t) − wb (0, t))/L b | (49)

In the underlying model of these two spectral representations, as in all previous computations, the sleeper
mass is included in the bridge mass per unit length ρ Ab . However, considering all the modeling details of the
presented ballast interaction model, the obvious problem of the mechanical model idealizing the entire track
as a continuous system must also be addressed. Although the presented model does not include a detailed
description of the discrete support of the rails on the sleepers, it can still be argued that, at least in the case of a
relatively stiff rail-sleeper connection, the sleeper mass can contribute significantly to the dynamic response of
the rail beam. Therefore, the results of two additional computations are included in Fig. 14 that add a distributed
mass per unit length δρ Ar = 500 the kg/m to the rail beam instead to the bridge (“with sleeper mass”).
In Fig. 14, the results denoted by “wo. cob ; without sleeper mass,” were obtained using a similar horizontal
damping parameter as in the reference solutions (“ref. sol.”) of Figs 12 and 13, without applying the additional
damping cob and sleeper mass δρ Ar . Therefore, a comparison of these results is possible, giving an indication
of the influence of the additionally considered soil–structure interaction on the results of Fig. 14. Comparing
the response peaks of these results for both the displacement in Fig. 14 (a) and the acceleration in Fig. 14 (b)
with the response peaks of the reference solution of Figs 12 and 13, significantly lower values are observed
when the soil–structure interaction is considered. This indicates an increased system damping associated with
the soil–structure interaction.
Moreover, the inclusion of the additional damping parameter cob also contributes significantly to the
system damping, which is evident from the reduced response peaks. The consideration of the sleeper mass
1414 P. König, C. Adam

as a distributed mass for the rail girder has no significant effect on the relative bridge deflection. However,
as the train speed increases, the bridge acceleration tends to be greater when the sleeper mass is added to the
rail beam. Considering the observation that track irregularities significantly affect the bridge acceleration at
train speeds above 55 m/s (see Figs 12 and 13), it can be concluded that the increase in bridge acceleration in
Fig. 14, results from the increased effect of rail vibration on the bridge due to the additional mass of the rail
beam combined with the additional excitation mechanism due to track irregularities.

6 Summary and conclusions

In more complex mechanical models of track–bridge interaction systems, the properties of the rails, rail
pads, sleepers, and ballast are accounted for by multiple layers of vertically oriented continuously distributed
spring-damper elements. However, recent experimental studies have shown that also the longitudinal ballast
interaction may have a significant impact on the system behavior. Effects such as an increase in system stiffness
and additional damping have been reported. Therefore, in the present contribution, modeling approaches were
developed to address the complex nature of longitudinal track–bridge interaction in response prediction of
high-speed railway bridges. These approaches involve discretization of the track in two different forms. While
the first approach uses a modal series expansion of the rail deflection, the second approach adopts a Rayleigh-
Ritz approximation of this response quantity. The coupling of the track and bridge-soil subsystems is achieved
by variations of the component mode synthesis. The application of a discrete substructuring technique to the
track–bridge–soil subsystem and the mass-spring-damper model of the high-speed train leads to the equation
of motion of the entire interaction system. The resulting models allow the analysis of several interaction
effects that affect the prediction of the dynamic system response. In addition to the longitudinal track–bridge
interaction, these include the vertical track–bridge interaction, the train-track interaction, and the soil–structure
interaction of the bridge.
The two modeling approaches were developed with the aim of providing an accurate description of the
model considering the longitudinal track–bridge interaction and an efficient computation. Based on the theory
and the implementation of the two approaches, it can be concluded that Approach 1 provides a more accurate
description of the model taking into account the longitudinal interaction, but at the cost of higher computational
times due to the large number of rail modes to be considered. In contrast, Approach 2 uses a simplified
description of the track, resulting in a much lower computational cost, but with the possible loss of accuracy.
However, for the parameters considered in this publication, Approach 2 provided reliable results for the
predicted bridge response.
Using the models based on the modeling approaches presented, the results of a variety of response analyses
were shown, providing insight into the influence of longitudinal bridge-track interaction effects on the system
response. In addition, the response of the model with realistic parameter configurations was shown, including
the soil–structure interaction of an example bridge. From these results, the following conclusions can be drawn
for the train-bridge interaction systems considered, taking into account the modeling assumptions:
• The longitudinal track–bridge interaction leads to stiffening and damping effects, which become visible in
the predicted displacement and acceleration response of the bridge.
• The dynamic response prediction is strongly affected by the presence of track irregularities, which in turn
strongly depends on the longitudinal interaction parameters.
• The additional damping effect of the ballast, related to the absolute motion of the ballast, leads to a
significant reduction of the displacement and acceleration peaks at resonant speeds.
• Consideration of sleeper mass in the rail beam leads to prediction of higher accelerations when track
irregularities are present.
• Soil–structure interaction has a strong influence on the predicted dynamic response, i.e., the response peaks
due to the additional damping are reduced.
The presented models provide improved simulation tools for more realistic estimation of the dynamic response
of the dynamic train-bridge interaction system, which can be used to conduct extensive parameter studies in
the future.

Acknowledgements The computational results presented have been achieved (in part) using the HPC infrastructure LEO of the
University of Innsbruck.
A model considering the longitudinal track–bridge interaction... 1415

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use,
sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original
author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other
third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit
line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted
by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To
view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Funding Open access funding provided by University of Innsbruck and Medical University of Innsbruck. The financial support
of parts of the research presented in this paper by the Tyrolean Science Fund (TWF, project number 346016) is gratefully
acknowledged.

Declarations

Competing interests The authors have no relevant financial or non-financial interests to disclose.

Appendix A Sub-matrices - Approach 1

This Appendix specifies system matrix components in the equations of motion of the coupled track–bridge–soil
subsystem (Eqs (37) and (38)) according to Approach 1. The diagonal matrices Sb and Sr containing complex
natural frequencies of the bridge beam and the rail beam read as [14]
 
(1) (2) (N ) (1) (2) (N )
Sb = diag sb , sb , ..., sb b , s̄b , s̄b , ..., s̄b b ,
  (A1)
Sr = diag sr(1) , sr(2) , ..., sr(Nr ) , s̄r(1) , s̄r(2) , ..., s̄r(Nr )

The matrices Ab and Bb containing the normalizing constants for the modal equations of the bridge beam,
specified in Eq. (23), read as [14]
 
Ab = diag ab(1) , ab(2) , ..., ab(Nb ) , āb(1) , āb(2) , ..., āb(Nb ) ,
  (A2)
(1) (2) (N ) (1) (2) (N )
Bb = diag bb , bb , ..., bb b , b̄b , b̄b , ..., b̄b b

The matrices Ar and Br containing the normalizing constants for the modal equations of the rail beam, given
in Eq. (33), read as [14]
 
Ar = diag ar(1) , ar(2) , ..., ar(Nr ) , ār(1) , ār(2) , ..., ār(Nr ) ,
  (A3)
Br = diag br(1) , br(2) , ..., br(Nr ) , b̄r(1) , b̄r(2) , ..., b̄r(Nr )

The matrices Mb , Cb and Kb differ from those specified in König et al. [14], because in the present
model the longitudinal stiffness and damping of the ballast are considered. For the present model, they are
derived as
 L b +L 0
Mb = ρ Ar  r  Tr dx,
−L 0
 Lb  Lb
Cb = cz b Tb dx − b  Tr dx
0 0
 Lb  L b +L 0 
−  r Tb dx + r
T
r dx
0 −L 0
  Lb  Lb
−cx rb2 b Tb,x x dx + rb rr b  Tr,x x dx (A4)
0 0
 Lb  L b +L 0 
+rbrr  r Tb,x x dx + rr2  r  Tr,x x dx
0 −L 0
1416 P. König, C. Adam

 Lb  Lb 
Kb = kz b Tb dx − b  Tr dx
0 0
  Lb  Lb 
−kx rb2 b Tb,x x dx + rb rr b  Tr,x x dx
0 0

with the vectors of eigenfunctions of the bridge beam b and the vector of shape functions r [14],
 T
(1) (2) (N ) ¯ (1) ¯ (2) ¯ (Nb ) ,
b = b , b , ..., b b ,  b , b , ..., b
 T (A5)
(Nb ) ¯ (1) ¯ (2)
r =
(1) (2)
r , r , ..., r , r , r , ..., ¯ r(Nb )

Note that for Kb in Eq. (A4) the identity given by Eq. (26) has been considered. With the vector of eigen-
functions of the rail beam r given by [14]
 T
(Nr ) ¯ (1) ¯ (2) ¯
r = (1)
r ,  (2)
r , ...,  r ,  r ,  r , ...,  (Nr )
r , (A6)

the sub-matrices resulting from coupling of the bridge-soil and track subsystems have the following form,
 L b +L 0
Mbr =Mrb T
= ρ Ar  r Tr dx,
−L 0
 L b +L 0  Lb 
Cbr =cz  r Tr dx − b Tr dx
−L 0 0
  L b +L 0  Lb 
− cx rr2  r Tr,x x dx + rb rr b Tr,x x dx ,
−L 0 0
 L b +L 0  Lb 
Crb =cz r  Tr dx − r Tb dx
−L 0 0
  L b +L 0  Lb 
− cx rr2 r  Tr,x x dx + rb rr r Tb,x x dx ,
−L 0 0
 L b +L 0  Lb  (A7)
Kbr =kz  r Tr dx − b Tr dx
−L 0 0
  L b +L 0  Lb 
− kx rr2  r Tr,x x dx + rb rr b Tr,x x dx
−L 0 0
 L b +L 0
+ E Ir  r Tr,x x x x dx
−L 0

=kx rbrr − b (L b )r,x (L b ) + b (0)r,x (0)

+ b,x (L b )r (L b ) − b,x (0)r (0) ,

Krb =0

Here Krb = 0 follows directly from Eq. (26). The expression for Kbr was simplified by integrating by parts
for terms containing spatial derivatives, together with the identity of Eq. (26).

Appendix B Sub-matrices - Approach 2

In the following, the sub-matrices in the equation of motion of Approach 2 (Eqs (41) and (42)) are specified.
The sub-matrices related to the modal coordinates of the bridge (subscript “b”) are the same as for Approach
1 and can be taken directly from Appendix A. However, the matrix components related to the coupling of the
A model considering the longitudinal track–bridge interaction... 1417

substructures of the rail beam and bridge-soil beam differ from Approach 1. With the definition of the vector
of rail shape functions ∗r ,
 T
∗r = ϕr∗ (x − x1 (t)), ϕr∗ (x − x2 (t)), ..., ϕr∗ (x − x Na (t)) (B8)
the coupling matrices of Eq. (42), denoted by the subscripts “br” and “rb”, are obtained. Simply replacing
r by ∗r in the expressions of Eq. (A7) results in the equivalent representations of the sub-matrices M  br (t),
 rb (t), 
M Cbr (t),  br (t) and K
Crb (t), K rb (t). Here the time dependency of the matrices is a direct result of the
time dependency of Eq. (B8).
The matrix K br (t) can be simplified in the same manner as Kbr in Eq. (A7), i.e., It can be shown that

Kbr (t) ≈kx rbrr − b (L b )∗r,x (L b ) + b (0)∗r,x (0)
 (B9)
+ b,x (L b )∗r (L b ) − b,x (0)∗r (0) ,

if L 0 is sufficiently large. A similar simplification can also be found in König et al. [13].
The remaining sub-matrices of Eqs (41) and (42) related to the time-dependent coordinates of the rail beam
(subscript“r”) can be found using the same procedure as in König et al. [13]. Herein the dynamic response of
the rail due to an individual axle load F p (t) is considered at first. In this regard, the equation of motion of the
stand-alone rail beam approximated by the p-th Rayleigh-Ritz shape function can be derived from Eq. (8) as
∗ ∗ ∗ ∗
E Ir ϕr,x x x x ( x̂ p )yr p (t) + ρ Ar ϕr ( x̂ p ) ÿr p (t) + cz ϕr ( x̂ p ) ẏr p (t) + kz ϕr ( x̂ p )yr p (t)
∗ ∗
  (B10)
− rr2 kx ϕr,x x yr p (t) − rr cx ϕr,x x ẏr p (t) = F p (t)δ x̂ p (t, tA p , tD p )
2

As a shape function for the Rayleigh-Ritz approximation of the rail beam deflection, the deflection of the
infinitely long elastically bedded beam due to a single load is employed (Eq. (35)). However, also in Approach
2 the influence of the train on the interacting system is only considered for a finite section of the rail beam.
Choosing the considered section of the track in same interval as in Approach 1, −L 0 ≤ x p ≤ L b + L 0 , the
time window in which the p-th axle crosses the track between points A and D is defined by (t, tA p , tD p )
[13]. As shown in König et al. [13], the equation of motion of a single degree of freedom (SDOF) oscillator,
equivalent to the bedded rail beam under point load, can be found by pre-multiplying Eq. (B10) by ϕr∗ (x̂ p ) and
integrating from −∞ to ∞, resulting in
   
3 3 2
ρ Ar ÿr p + cz + βrr cx ẏr p +
2
kz + βrr kx yr p = F p (t)(t, tA p , tD p )
2
(B11)
2β 2β β
As in König et al. [13], the Na SDOF systems that capture the response of the rail beam to the individual axle
forces are considered to be decoupled because the deflection shape around the axle load is very isolated. This
consideration, together with Eq. (B10), leads to the diagonal matrices Mr , Cr , Kr , which represent the Na
decoupled SDOF oscillators,
   
3 3 3 3
Mr = diag ρ Ar , ..., ρ Ar , Cr = diag cf + βrr2 cx , ..., cf + βrr2 cx ,
2β 2β 2β 2β
  (B12)
2 2
Kr = diag kf + βrr2 kx , ..., kf + βrr2 kx
β β

References

1. ABAQUS: ABAQUS (2016). Providence, RI, United States (2015)


2. Battini, J.M., Ülker-Kaustell, M.: A simple finite element to consider the non-linear influence of the ballast on vibrations of
railway bridges. Eng. Struct. 33(9), 2597–2602 (2011). https://doi.org/10.1016/j.engstruct.2011.05.005
3. Biondi, B., Muscolino, G., Sofi, A.: A substructure approach for the dynamic analysis of train-track-bridge system. Comput.
Struct. 83(28–30), 2271–2281 (2005). https://doi.org/10.1016/j.compstruc.2005.03.036
4. Chordá-Monsonís, J., Romero, A., Moliner, E., Galvín, P., Martínez-Rodrigo, M.: Ballast shear effects on the dynamic
response of railway bridges. Eng. Struct. 272, 114957 (2022). https://doi.org/10.1016/j.engstruct.2022.114957
5. Claus, H., Schiehlen, W.: Modeling and simulation of railway bogie structural vibrations. Veh. Syst. Dyn. 29(sup1), 538–552
(1998). https://doi.org/10.1080/00423119808969585
1418 P. König, C. Adam

6. EN1991-2. Eurocode 1: Actions on structures—Part 2: traffic loads on bridges (consolidated version) 2012
7. Foss, K.A.: Coordinates which uncouple the equations of motion of damped linear dynamic systems. J. Appl. Mech. 25,
361–364 (1958)
8. Frýba, L.: Dynamics of railway bridges. Thomas Telford Publishing, London (1996)
9. Frýba, L.: Vibration of Solids and Structures Under Moving Loads. Springer, Berlin (1999)
10. Galvín, P., Domínguez, J.: High-speed train-induced ground motion and interaction with structures. J. Sound Vib. 307(3),
755–777 (2007). https://doi.org/10.1016/j.jsv.2007.07.017
11. Hetényi, M.: Beams on Elastic Foundation: Theory with Applications in the Fields of Civil and Mechanical Engineering.
Scientific Series. University of Michigan Press, University of Michigan Ann Arbor, Mich (1946)
12. Hirzinger, B., Adam, C., Salcher, P.: Dynamic response of a non-classically damped beam with general boundary conditions
subjected to a moving mass-spring-damper system. Int. J. Mech. Sci. 185, 105877 (2020). https://doi.org/10.1016/j.ijmecsci.
2020.105877
13. König, P., Salcher, P., Adam, C.: An efficient model for the dynamic vehicle-track-bridge-soil interaction system. Eng. Struct.
253, 113769 (2022). https://doi.org/10.1016/j.engstruct.2021.113769
14. König, P., Salcher, P., Adam, C., Hirzinger, B.: Dynamic analysis of railway bridges exposed to high-speed trains considering
the vehicle–track–bridge–soil interaction. Acta Mech. 232(11), 4583–4608 (2021). https://doi.org/10.1007/s00707-021-
03079-1
15. Krenk, S.: Complex modes and frequencies in damped structural vibrations. J. Sound Vib. 270(4–5), 981–996 (2004). https://
doi.org/10.1016/S0022-460X(03)00768-5
16. Lei, X., Zhang, B.: 11. Influence of track stiffness distribution on vehicle and track interactions in track transition. In:
Proceedings of the Institution of Mechanical Engineers, Part F: Journal of Rail and Rapid Transit 224, 592–604 (2010).
https://doi.org/10.1243/09544097JRRT318
17. Mähr, T.: Theoretical and experimental investigations on the dynamic behaviour of railway bridges with ballast superstructure
under moving loads. Ph. D. thesis, TU Wien, 2009
18. MATLAB: MATLAB (R2020a). Natick, Massachusetts (2020)
19. Nguyen, K., Goicolea, J.M., Galbadón, F.: Comparison of dynamic effects of high-speed traffic load on ballasted track using
a simplified two-dimensional and full three-dimensional model. In: Proceedings of the Institution of Mechanical Engineers,
Part F: Journal of Rail and Rapid Transit 228(2), 128–142 (2012). https://doi.org/10.1177/0954409712465710
20. Rebelo, C., Simões da Silva, L., Rigueiro, C., Pircher, M.: Dynamic behaviour of twin single-span ballasted railway viaducts—
field measurements and modal identification. Eng. Struct. 30(9), 2460–2469 (2008). https://doi.org/10.1016/j.engstruct.2008.
01.023
21. Rigueiro, C., Rebelo, C., Simões da Silva, L.: Influence of ballast models in the dynamic response of railway viaducts. J.
Sound Vib. 329(15), 3030–3040 (2010). https://doi.org/10.1016/j.jsv.2010.02.002
22. Rocha, J., Henriques, A.A., Calçada, R.: Probabilistic safety assessment of a short span high-speed railway bridge. Eng.
Struct. 71, 99–111 (2014). https://doi.org/10.1016/j.engstruct.2014.04.018
23. Romero, A., Solís, M., Domínguez, J., Galvín, P.: Soil–structure interaction in resonant railway bridges. Soil Dyn. Earthq.
Eng. 47, 108–116 (2013). https://doi.org/10.1016/j.soildyn.2012.07.014
24. Salcher, P., Adam, C.: Modeling of dynamic train-bridge interaction in high-speed railways. Acta Mech. 226(8), 2473–2495
(2015). https://doi.org/10.1007/s00707-015-1314-6
25. Salcher, P., Adam, C.: Estimating exceedance probabilities of railway bridge vibrations in the presence of random rail
irregularities. Int. J. Struct. Stab. Dyn. 20, 2041005 (2020). https://doi.org/10.1142/S0219455420410059
26. Stollwitzer, A., Bettinelli, L., Fink, J.: The longitudinal track-bridge interaction of ballasted track in railway bridges: experi-
mental determination of dynamic stiffness and damping characteristics. Eng. Struct. 274, 115115 (2023). https://doi.org/10.
1016/j.engstruct.2022.115115
27. Stollwitzer, A., Fink, J.: Die rechnerische Bestimmung der Dämpfung von Stahl-Eisenbahnbrücken - Teil 2: Verifizierung
anhand von Bestandsbrücken. Stahlbau 90(6), 449–462 (2021). https://doi.org/10.1002/stab.202100013
28. Stollwitzer, A., Fink,J., Malik,T.: Influence of the ballasted track on the dynamic behaviour of steel railway bridges. ce/papers
4(2-4): 2013–2020. https://doi.org/10.1002/cepa.1516 2021
29. Stoura, C.D., Dimitrakopoulos, E.G.: A modified bridge system method to characterize and decouple vehicle-bridge inter-
action. Acta Mech. 231(9), 3825–3845 (2020). https://doi.org/10.1007/s00707-020-02699-3
30. Ticona Melo, L., Malveiro, J., Ribeiro, D., Calçada, R., Bittencourt, T.: Dynamic analysis of the train-bridge system con-
sidering the non-linear behaviour of the track-deck interface. Eng. Struct. 220, 110980 (2020). https://doi.org/10.1016/j.
engstruct.2020.110980
31. Ülker-Kaustell, M., Karoumi, R., Pacoste, C.: Simplified analysis of the dynamic soil–structure interaction of a portal frame
railway bridge. Eng. Struct. 32(11), 3692–3698 (2010). https://doi.org/10.1016/j.engstruct.2010.08.013
32. Wolf, J.P., Deeks, A.: Foundation Vibration Analysis: A Strength of Materials Approach. Butterworth-Heinemann, Oxford
(2004)
33. Yang, Y.B., Yau, J., Yao, Z., Wu, Y.: Vehicle-bridge interaction dynamics: with applications to high-speed railways. World
Scientific, Singapore (2004)
34. Zangeneh, A., Svedholm, C., Andersson, A., Pacoste, C., Karoumi, R.: Identification of soil-structure interaction effect in a
portal frame railway bridge through full-scale dynamic testing. Eng. Struct. 159, 299–309 (2018). https://doi.org/10.1016/j.
engstruct.2018.01.014
35. Zhai, W., Wang, K., Cai, C.: Fundamentals of vehicle-track coupled dynamics. Veh. Syst. Dyn. 47, 1349–1376 (2009). https://
doi.org/10.1080/00423110802621561
36. Zhang, N., Xia, H., Guo, W.W., De Roeck, G.: A vehicle-bridge linear interaction model and its validation. Int. J. Struct.
Stab. Dyn. 10(02), 335–361 (2010). https://doi.org/10.1142/S0219455410003464

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional
affiliations.

You might also like