You are on page 1of 16

Engineering Structures 211 (2020) 110430

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Dynamic effects on a train-bridge system caused by stochastically generated T


turbulent wind fields
P.A. Montenegroa, , D. Barbosaa, H. Carvalhob, R. Calçadaa

a
CONSTRUCT – LESE, Faculty of Engineering, University of Porto, Rua Dr. Roberto Frias s/n, 4200–465 Porto, Portugal
b
Department of Structural Engineering, Federal University of Minas Gerais, Av. Pres. Antônio Carlos, 6627 - Pampulha, 31270-901 Belo Horizonte, Brazil

ARTICLE INFO ABSTRACT

Keywords: The present paper presents a methodology for evaluating the possibility of derailment of trains running over
Stochastic wind velocity field bridges subjected to crosswinds. The dynamic model consists of a train–bridge interaction coupling system
Turbulent wind that takes into account the contact interface between wheel and rail from which the contact forces arise. These
Train-bridge interaction contact forces are the inputs for the safety criteria that are later used to evaluate derailment. The wind
Running safety
fluctuating component is accurately considered through a stochastic process that generates correlated wind
Railway bridges
High-speed
velocity fields, which are later converted into wind loads that are applied to both the bridge and the train.
Before evaluating the running safety, the dynamic response of the train–bridge system is assessed to under-
stand the influence of the train speed and wind velocity on it. Then, the running safety is analyzed based on
three criteria, namely the Nadal, Prud’homme and unloading indexes, being the influence of the bridge lateral
flexibility in the derailment risk also discussed. Finally, the results obtained with the stochastic wind model are
compared with those obtained with a simplified discrete gust model proposed by the European normative. The
maximum wind velocities, obtained through the stochastic and simplified formulations, for which no derail-
ment criteria allowance is exceeded are computed and a comparative critical analysis between the models is
presented.

1. Introduction using couples train–bridge–wind models can be found in the literature.


Xu et al. [7] proposed a framework to analyze the response of a long
Bridges are fundamental lifelines in railway engineering, especially suspension bridge subjected to turbulent winds and the passage of
with regard to high–speed railways (HSR), where the necessity to en- high–speed trains. However, in this study, the safety evaluation was
sure smother tracks with large curve radius led to an increase in the purely assessed based on the bridge response, since the train was tra-
number of viaducts and bridges. Some countries in Asia, such as China, velling inside a close deck and, consequently, it was not directly sub-
Japan and Taiwan, have a highly developed HSR network in which jected to the wind load. Later, Xu et al. [8] assessed the running safety
some of the lines have more than 75% of viaducts [1–3]. This fact led to and passenger comfort of a train crossing a cable–stayed bridge at a
an increase in the probability of a train being over a bridge during the single speed, not allowing to draw general conclusions regarding the
occurrence of natural hazards, such as earthquakes or crosswinds. Re- topic. Li et al. [9] presented an analytical model for analyzing the dy-
garding the latter, some of these bridges are situated in regions with namics of a wind–vehicle–bridge system, including the effects of the
deep valleys, where the crosswinds tend to be extremely severe to the static and buffeting displacements of the deck on the vehicle response
train’s safety. Moreover, the accident risks and their consequences tend and the vehicle effects on the aerodynamic loads applied to the bridge.
to increase due to high speeds [4,5], while the development of in- Again, a limited number of scenarios were analyzed, namely one with
creasingly lighter trains may worsen the situation in terms of over- no wind and another with a wind mean velocity of 25 m/s, both with a
turning derailments caused by crosswinds [6]. Therefore, the determi- single train speed of 250 km/h. Xia et al. [10] and Guo et al. [11]
nation of threshold values for the wind velocity above which the train performed a running safety analysis of two existing suspension bridges
should stop is a topic of the utmost importance and that deserves to be in China, namely the Tianxingzhou and the Tsing Ma bridges, respec-
thoroughly studied. tively. Both conclude that the dynamic response of the bridge, espe-
Few studies regarding the train running safety against crosswinds cially the lateral one, is mainly influenced by the wind load, being the


Corresponding author.
E-mail addresses: paires@fe.up.pt (P.A. Montenegro), up201405298@fe.up.pt (D. Barbosa), hermes@dees.ufmg.br (H. Carvalho), ruiabc@fe.up.pt (R. Calçada).

https://doi.org/10.1016/j.engstruct.2020.110430
Received 11 September 2019; Received in revised form 23 January 2020; Accepted 24 February 2020
Available online 14 March 2020
0141-0296/ © 2020 Elsevier Ltd. All rights reserved.
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

influence of the train speed substantially lower. While the first also While the former represents the wind field in a simplified way with
concluded that the locomotive car is less susceptible to overturning single gust fixed in space, the latter is more accurate and realistic, since
derailment due to its larger weight, the latter concluded that for a mean it considers the spatial and temporal variability of the wind field and
wind velocity of 30 m/s the running safety is no longer guaranteed in the corresponding correlations between the generated profiles. Hence,
the Tsing Ma bridge and the traffic should be closed. Xu and Zhai the second approach is considered in the present work to study the
[12,13] analyzed the couple effects between randomly generated winds response of the train-bridge system and the running safety against
and irregularities and compared their influence in the vehicle’s re- crosswinds.
sponse. While the carbody response is mainly governed by the cross-
wind effects, the wheel–rail interaction and track vibration is more
2.1. Correlation between wind velocity fields
affected by the track irregularities.
More recently, some studies regarding the influence of wind bar-
Although the wind velocity field is fully three–dimensional and,
riers were also published. Zhang et al. [14] studied the effects of wind
theoretically, should be represented through a multidimensional sto-
barriers with different heights and levels of porosity, concluding that
chastic process, it is accepted for engineering practice to simplify this
the barriers should have porosities ranging between 10% and 20% in
process and turn it into three independent one–dimensional multi-
order to be effective. According to the authors, and for certain train
variate stochastic processes. Based on this, the cross–spectral density
speeds, the critical wind velocities increased from 12.5 m/s up to
matrix S(ωmf) that correlates the wind spectra of each of the n gen-
32.5 m/s due to the presence of barriers. Guo et al. [15] also con-
eration points is given by [20]
ducted studies regarding this topic on two different bridge solutions
belonging to the wind prone HSR line of Lanzhou–Urumqi. They S11 ( mf ) S12 ( mf ) S1n ( mf )
conclude that for wind velocities up to 35 m/s the critical train speeds S21 ( mf ) S22 ( mf ) S2n ( mf )
on the two bridges were very similar. However, for higher levels of S( mf ) =
wind, the derailment indexes drop considerably in a trough–beam
Sn1 ( mf ) Sn2 ( mf ) Snn ( mf ) (1)
bridge when compared to a box–beam bridge, leading to the conclu-
sion that the latter has a better performance against crosswinds. where each element Sjm(ωmf) is the cross-spectra between generation
Olmos and Astiz [6] also conducted numerical simulations on a high points j and m defined as
pier viaduct belonging to the Spanish HSR network and analyzed the
influence of wind barriers and tuned mass dampers (TMDs) in the Sjm ( mf ) = Sjj ( mf ) Smm ( mf ) jm (djm , mf ), j, m = 1, 2, , n; j m
train’s running safety. The TMDs installed at the top of the piers sig-
(2)
nificantly decreased the bridge lateral response and, consequently,
increased the train safety. Moreover, the combined installation of and ωmf is a double indexed frequency dependent on the frequency
TMDs and wind barriers improved the threshold wind velocity when increment Δω expressed as
the train crossed the viaduct at 340 km/h from 25 m/s to 37.5 m/s, m
leading to the conclusion that both countermeasures can have a po- mf = (f 1) + , f = 1, 2, ,N
(3)
n
sitive effect when installed together.
All the aforementioned publications simulate the wind velocity field Assuming that the spectra do not vary along the longitudinal di-
with a stochastic process that takes into account the correlation and rection yields
coherence between the different wind time–histories in each generation
point. However, no comparisons with other simplified wind models
S11 ( mf ) = S22 ( mf ) = =Snn ( mf ) = S( mf ) (4)
have been performed so far, including those proposed by the norms. which allows the following simplification of Eq. (2)
The present work aims to fill this gap, by comparing the results ob-
tained with a stochastic wind model with those obtained with the Sjm ( mf ) = S( mf ) jm (djm , mf ) (5)
simplified Chinese Hat discrete gust model recommended by the
European norm EN 14067–6 [16], presented in a previous publication where djm is the distance between the generation points j and m and Γjm
[17]. Such comparison allowed a discussion, which is presented at the (djm, ωmf) is the Davenport’s coherence function [21] given by
end of the paper, regarding the advantages and disadvantages of each mf djm
model. Moreover, the present study also draws important conclusions jm (djm , mf ) =e 2 U¯z
(6)
regarding the influence of the lateral flexibility of the bridge structure
(columns and deck) in the train’s running safety. where Ūz is the mean wind velocity at height z and λ is the wind cor-
The case study of this paper consists of the Volga River bridge, relation factor. By considering that the n generation points are uni-
which belongs to the future HSR line connecting the Russian cities of formly spaced along the bridge deck with a distance d, Γjm(djm, ωmf) may
Moscow and Kazan. A train–bridge interaction tool previously devel- be redefined as [20]
oped by the authors [18] and successfully used for the evaluation of |j m |
mf d
train running safety in other situations [17,19] is adopted to assess 2 U¯z = C|j m|
jm (djm , mf ) = e
wheel–rail contact forces that serve as input data to analyze the de- (7)
railment through criteria existing in the literature.

2. Stochastic process for simulating correlated wind velocity


2.2. Stochastic generation of wind velocity fields
fields
The lower triangular matrix H(ωmf) that arises from the Cholesky’s
To simulate the wind loads acting on bridges and trains it is es- decomposing of the cross–spectral density matrix S(ωmf) defined in Eq.
sential to correctly define the wind velocity fields that affect these two (1) can be expressed as
structural systems. Two methodologies can be adopted to define these
fields: (1) the discrete gust approach, which is adopted in the current H( mf ) = S( mf ) G( mf ) (8)
European norm EN 14067-6 [16] and briefly exposed later in Section 5,
and (2) the stochastic approach, which consists of generating artificial where the coefficient matrix G(ωmf) may be defined as (see Eqs. (1) and
wind velocity profiles using power spectral density (PSD) functions. (7))

2
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

1 0 the air density and Vw,b(t) is the resultant wind velocity given by
C 1 C2
C2 C 1 C2 1 C2 Vw, b (t ) = [U¯z + u (t )]2 + w (t ) 2 (14)
G( mf ) =
C3 C2 1 C2 C 1 C2 1 C2 leading to an incidence angle α of

Cn 1 Cn 2 1 C 2 Cn 3 1 C2 Cn 4 1 C2 1 C2 1 w (t )
(t ) = tan
(9) U¯z + u (t ) (15)
Finally, the fluctuating component of the wind velocity time–- The moment Mm,b acting on the deck arises from the vertical eccen-
histories in each generation point j in the horizontal uj(t) and vertical wj tricity ez between the deck gravity center, Cg, and the pressure center of
(t) directions given by [20] the lateral windward area, Cp, and from the horizontal eccentricity ey
j N taken as Bb/4 as stated in EN 1991-1-4 [23] and shown in Fig. 1.
u j (t ) = 2 Su ( mf ) Gjm ( mf )cos( mf t+ mf ) In the present work, the drag and lift aerodynamic coefficients have
m=1 f =1 (10) been calculated according to the procedure defined in Section 8.3 from
j N
EN 1991-1-4 [23].
wj (t ) = 2 Sw ( mf ) Gjm ( mf )cos( mf t+ mf )
m=1 f =1 (11) 2.4.2. Loads on railway vehicles
The wind loads acting on the vehicle are calculated by a procedure
where, Su(ωmf) and Sw(ωmf) are the horizontal and vertical wind
similar to that used in the bridge. Considering that the main stream
spectra (see Eq. (4)), N is the number of considered wind frequencies,
flow acts perpendicular to the direction of the vehicle movement, the
mf is a set of independent random phase angles uniformly distributed
resultant wind velocity Vw,v(t) is given by the same expression defined
in the range [0, 2π] and Gjm(ωmf) is an element of matrix G(ωmf) defined
in Eq. (14). Therefore, the relative wind velocity to the vehicle Vr(t) can
in Eq. (9).
be defined as (see Fig. 2)

2.3. Simplification in the generation of wind velocity fields Vr (t ) = Vv2 + Vw2 , v = Vv2 + [U¯z + u (t )]2 + w (t ) 2 (16)

In order to improve the computational time of the wind velocity while the yaw angle β is given by
fields, the simplification proposed by Cao et al. [20], which consists of
Vw, v [U¯z + u (t )]2 + w (t ) 2
neglecting the cosines of Eqs. (10) and (11) whose values of Gjm(ωmf) (t ) = tan 1 = tan 1
are less than 10−3, has been adopted in the present work. This sim- Vv Vv (17)
plification is justified by the fact that the values of Gjm(ωmf) rapidly
decrease with the increase of the distance djm between points j and m. where Vv is the vehicle’s running speed.
Finally, the forces acting at the center of gravity Cg of the carbody of
each car of the train, namely the drag, Fd,v(t), and lift, Fl,v(t), forces and
2.4. Wind loads computation and application
the rolling moment Mm,v(t) are depicted in Fig. 2 and can be expressed
by the following equations
2.4.1. Loads on bridges
The wind loads acting on bridges are divided into the steady–state 1
Fd, v (t ) = Vr (t ) 2Cd, v ( , ) Av
and buffeting components. While the former is induced by the mean 2 (18)
wind velocity, the latter is caused by the wind fluctuating flow in the
1
horizontal u(t) and vertical directions w(t). Hence, and according to the Fl, v (t ) = Vr (t )2Cl, v ( , ) Av
2 (19)
quasi–steady theory, the drag, Fd,b, and lift, Fl,b, wind forces per unit
length of the deck (see Fig. 1) can be expressed as [22] 1
Mm, v (t ) = Vr (t )2Cm, v ( , ) Av Hv Fd, v dgc
1 2 (20)
Fd, b (t ) = Vw, b (t ) 2Cd, b ( ) Hb
2 (12) where Cd,v(α,β), Cl,v(α,β) and Cm,v(α,β) are the aerodynamic drag, lift
1 and moment coefficients of the vehicle, respectively, that depend on the
Fl, b (t ) = Vw, b (t )2Cl, b ( ) Bb incidence and yaw angles, and Av and Hv are the windward exposed
2 (13)
area and height of the vehicle, respectively. In the present work, and
in which Cd,b(α) and Cl,b(α) are the drag and lift wind force aerodynamic due to the similarity of the train used in this study, the aerodynamic
coefficients of the bridge dependent on the angle of incidence α, Hb and coefficients given by the Annex C of EN 14067-6 [16] for the ICE3 train
Bb are the height and width of the deck exposed area, respectively, ρ is have been adopted for all the calculations. In this normative, the rolling
moment aerodynamic coefficient refers to the moment along the long-
Bb itudinal axis centered with the track and situated at the top of the rails.
Therefore, since the moment has to be applied at the center of gravity of
the carbody, it is necessary to remove the portion of the moment that
Mm,b
Cg Fd ,b arises from the eccentricity dgc between this point and the top of the
ez rails (see Fig. 2).
Cp Uz +u(t) Due to the movement of the train, the wind loads at each instant
Hb

have to be spatially interpolated, since the vehicle position may not


w(t) coincide with the position of a generation point. Moreover, the loads
Vw,b (t)
have to be also temporally interpolated, since the instants of the dy-
namic analysis do not necessarily have to coincide with the instants of
z the generated wind velocity fields defined in Section 2.
Fl,b y
ey x
3. Train-bridge-wind interaction system

Fig. 1. Resulting wind velocity and wind loads acting on the bridge. The dynamic train-bridge interaction coupling system is solved by a

3
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

Cp Uz +u(t) y
Train running direction
w(t) Vv
Cg Fd,v Vw,v (t) z x
dcg Fl,v Mm,v z
Vw,v (t) Vr (t)
y x
(a) (b)
Fig. 2. Relative wind velocity and wind loads acting on the vehicle: (a) front view and (b) plan view.

direct method, based on the Lagrange multiplier method, which avoids and current Newton iteration, respectively.
an iterative procedure to ensure the coupling between the two systems. Regarding the wheel–rail contact model, the nonlinear Hertz con-
Additional constraint equations that relate the displacements of the tact theory [27] is used to analyze the normal contact problem, in
train with the nodal displacements of the bridge are added to the which the normal contact force Fn (see Fig. 3) between the wheel and
governing equilibrium equations, forming a single system that can be rail is given by [28]
directly and efficiently solved. However, when the contact non- 3
linearities are taken into account an iterative algorithm has to be added Fn = Kh d 2 (22)
to the formulation to solve the nonlinear equations [24,25]. The
where d is the penetration and Kh is a generalized stiffness coefficient
iterative schemes most widely used for the solution of nonlinear finite
that depends on the material properties of the bodies in contact, such as
element equations, and that are also used in the present algorithm, are
the Young modulus and the Poisson ratio, and the curvatures of the
based on the Newton method [26].
surfaces at the contact point. As for the tangential creep forces in the
Later, Montenegro et al. [18] extended the formulation to take into
longitudinal, Fξ, and lateral, Fη, directions (see Fig. 3), these are pre-
account the geometry of the wheel and rail and the behavior of the
calculated and stored in a lookup table, based on the USETAB algorithm
contact interface. Based on this work, the train–bridge coupling can be
[29], to be later interpolated during the dynamic analysis as a function
mathematically expressed as
of the creepages and the semi–axes ratio of the contact ellipse (see [18]
for details about how the table was built).
¯ D
K ¯ aF i + 1 (at + t , i, Xt + t , i)

¯ 0
= The formulation of the train-bridge interaction numerical tool used
H Xi + 1 r¯ (21)
in the present work, referred hereinafter as VSI (Vehicle–Structure
Interaction Analysis), has been programmed in MATLAB® [30] and
where K̄ is the effective stiffness matrix of the train-bridge system, D̄
presented in detail in [18]. In this tool, both the bridge and the train are
and H̄ are transformation matrices that relate, respectively, the contact
modeled using the finite element method (FEM) software ANSYS® [31],
forces in the local coordinate system (target element coordinate system
being their structural matrices imported by MATLAB® [30], where all
xte,yte,zte defined by the rails, as shown in Fig. 3) with the nodal forces
the dynamic analysis is performed.
in the global coordinate system and the displacements of the nodes of
the structure (target elements defined by the rails) in the global co-
ordinate system with the displacements of the contact element points 4. Dynamic analysis of the train–bridge system subjected to
defined in the local coordinate system (see [7] for details). The vectors stochastically generated winds
Δa and ΔX are the incremental nodal displacements and contact forces,
respectively, is the residual force vector and r̄ the vector with the 4.1. Application case
irregularities that may exist in the contact interface. Superscript t + Δt
refers to the current time step, while i and i + 1 indicates the previous The case study addressed in the present work consists of the Volga
River Bridge, near the Russian city of Kazan (see Fig. 4), which belongs
to the future HSR link between Moscow and Yekaterinburg. Winds with
gust up to 130 km/h are expected to occur in the region, meaning that
the running safety of trains against crosswinds is an issue of the utmost
importance, especially when they cross the bridge at the operating
speed of 350 km/h.
The structural solution consists of a prestressed concrete bridge with

Fig. 3. Normal and tangential contact forces in the contact interface between Fig. 4. Location of the Volga River bridge in the Moscow-Yekaterinburg HSR
wheel and rail. link [32].

4
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

(a)

(b) (c)
Fig. 5. Structural drawings of the Volga River HSR bridge: (a) elevation view, (b) cross–sections and (c) middle column.

four continuous spans of 98 + 2×190 + 98 m, totalizing 576 m, as represented in Fig. 7, including an illustration of the numerical model
shown in Fig. 5. The deck is formed by a cast-in-place single-cell box developed in ANSYS [31]. The main components of the car, namely
girder with 6200 mm width and height ranging from 5138 mm at the carbody, bogies and wheelsets, are modeled using beam finite elements,
middle of each span to 12639 mm over the main piers (see Fig. 5b). The while spring–dampers are used to simulate the suspensions that connect
middle pier is monolithically connected to the deck (see Fig. 5c) with those same components. In Fig. 7 k, c, m and I represents stiffness,
variable cross-section along its height, while the remaining piers are damping, concentrated mass and rotary inertia, respectively, while the
connected through bearing supports. The top part of the cross–section subscripts cb, b and w denote carbody, bogie and wheelset, respectively.
consists of a 13800 mm wide platform that supports two ballasted To better understand the train dynamic behavior, the numerical mode
tracks. shapes and the corresponding frequencies of the vehicle are presented
in Appendix B. However, as already mentioned in [17], the vehicle’s
4.2. Bridge and train model manufacturer does not allow publishing the train properties due to
confidential matters.
4.2.1. Volga River high-speed railway bridge
As mentioned before in Section 3, the structural models have to be 4.3. Sources of excitation
previously modeled with a commercial FEM software in order to obtain
the stiffness, mass and damping matrices to be later imported to MA- 4.3.1. Wind load
TLAB® [30], where the dynamic analyses are carried out. In the present In the present work, the Kaimal’s Su [35] and Lumley and Panofsky’s
work, the FEM software used to model the bridge is ANSYS [31], as Sw [36] spectra (see Eqs. (10) and (11)) have been adopted to generate
shown in Fig. 6. The main structural components, namely the deck and the wind flows in the horizontal and vertical directions, respectively.
the piers, are modeled using beam finite elements, while the track These spectra are expressed as
components, such as ballast, pads and fasteners are simulated using u2 200fz
spring–dampers whose properties were obtained in the literature (see Su ( ) =
(1 + 50fz )5/3 (23)
Appendix A). In order to take into account the mass of the non–-
structural components, which is mainly composed by the ballast, mass u2 3.36fz
Sw ( ) =
point elements are also included in the model. Moreover, the connec- (1 + 10fz )5/3 (24)
tions between the deck, which is modeled along its gravity center, and
the track are defined through rigid link elements (see Fig. 6.b). Finally, where ω is the frequency, fz is the normalized frequency dependent on
an extension of the track is modeled in both bridge’s extremities to the height of the generation point relative to the ground z given by
efficiently simulate the transition between the bridge and the adjacent z
fz =
structures (see Fig. 6c). 2 U¯z (25)
The shapes of the main global modes and respective natural fre-
quencies may be consulted in Appendix A. Since the dynamic vehicle and u is the friction velocity of the wind flow given by
structure interaction analysis is solved through a direct integration al- kŪz
u =
gorithm (see Section 3), the damping matrix is defined through the
Rayleigh proportional matrix [33]. Therefore, a damping ratio of 1%
ln ( )
z
z0 (26)
(proposed by EN 1991-2 [34] for concrete bridges) is set for the first In which k is the Von Karman constant equal to 0.4 [22] and z0 is
vertical mode (f1v = 0.530 Hz) and third lateral mode (f3l = 1.339 Hz) the roughness length.
in order to ensure that that the main global modes are correctly In the present work, the Volga River bridge deck is located at a
damped. height z = 48 m, while the roughness length z0 is 0.05, which corre-
sponds to the standard terrain category II of EN 1991-1-4 [23]. More-
4.2.2. High-speed railway vehicle over, the wind field is composed of 33 generation points equally spaced
Due to lack of information regarding the Russian high–speed train, by 18 m along the 576 m length of the deck. In order to guarantee that
the vehicle used in this work consists of a Japanese Shinaknsen train the train enters the bridge after recovering from the transient effects, 10
whose properties were known. The train model is developed in ANSYS more generation points are considered before the first point of the
[31] and consists of ten independent cars with 25 m length each and bridge.
with an average static axle load of 110 kN. The dynamic model is For exemplification purposes, and to visually observe the coherence

5
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

Z
X
Y

Z
(a)
Z
Y X
X
Y

Deck girder
Deck girder Column
Rail Rail
Sleeper Sleeper
Rigid link Rigid link
Mass point Mass point

(b) (c)
Fig. 6. Numerical model of the Volga River HSR bridge: (a) overview, (b) connection between deck and track and (c) track extension at the bridge’s extremity.

between the wind velocity time–histories generated in different points, 1 2


c
2
SG, C ( ) = AG, C
the horizontal component of the wind velocity field of four points is l2 ( 2
r + 2)( 2
c + 2)( 2
s + 2)
(28)
plotted in Fig. 8. In this figure, points A and B are 1 m apart, while
points C and D are located 18 m and 180 m away, respectively, from where Ω is the spatial frequency, the subscripts A, V, G and C indicate
point A. For the generation of this wind field, and for all others simu- the alignment, elevation level, gauge and cross level irregularities, l is
lated in the present study, an upper frequency ωup = 6 Hz and a fre- half of the gauge, Ωc, Ωr and Ωs are constant factors given in [40] and A
quency increment Δω = 0.002 Hz are adopted, totalizing N = 3000 are the irregularity scale factors ( AA = AG = 2 × 10 6 rad. m and
frequencies. Furthermore, the correlation factor λ from Davenport’s AV = AC = 4 × 10 6 rad. m ). These factors have been chosen to obtain
coherence function presented in Eq. (6) can assume different values the alert limit level of irregularities defined in EN13848–5 [41]. Finally,
according to the literature (7 in [37], 9 in [38], 10 in [20] or 16 in the irregularity profile function r(x) is given by
[39]). In the present work, the value proposed by Cao et al. [20] of N 1
λ = 10 is adopted. As expected, the coherence between points A and B, r (x) = 2 An cos( nx + n)
which are very close to each other, is very high, but it starts to decrease n=0 (29)
when the distance between the points becomes higher. where n are the random phase angles uniformly distributed in the
Finally, based on the wind velocity fields generated according to the range [0, 2π] and Ωn are a series of N spatial discrete frequencies de-
aforementioned parameters, the wind loads applied to both the bridge fined in the interval [Ω0, Ωup] with increments ΔΩ, in which Ω0 and Ωup
and the train can be computed through the procedure previously de- are the minimum and maximum frequencies considered. For ex-
scribed in Section 2.4. emplification purposes, Fig. 9 presents the PSD function for the eleva-
tion irregularity and a 200 m stretch of the vertical and lateral devia-
tions of the left rail.
4.3.2. Track irregularities
The track irregularities considered in the present work are artifi-
cially generated according to the procedure described by Claus and 4.4. Analysis parameters
Schiehlen [40]. According to [40], and based on several measurements
of track irregularities, the PSD functions S used to generate the profiles The calculations in the present work are carried out for train speeds
can be standardized and expressed as between 140 km/h and 420 km/h and for 10 min mean wind velocities
at height z = 48 m (hereinafter referred to as Ūz ) up to 30 m/s. The
1 2 turbulent intensity Iv associated to this scenario is 14.56%, which is
c
SA, V ( ) = AA, V much smaller than the 24.46% assumed in the normative Chinese Hat
2 ( 2
r + 2)( 2
c + 2)
(27)
discrete gust model proposed in [16] and investigated in [17]. Due to

6
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

(a) (b)

(c) (d)
Fig. 7. Dynamic model of the railway vehicle: (a) lateral view, (b) front view, (c) ANSYS model overview and (d) detail of the bogie.

the significant differences between the discrete gust and stochastic problem, a constant time step Δt is used in all calculations, while the
formulations, the running safety critical train speeds obtained with both following parameters for the HHT implicit method have been adopted:
approaches are compared and discussed. α = –0.1, β = 0.3025 and γ = 0.6. According to Hughes [42], values of
In terms of the parameters related to the solution of the dynamic α in the interval [–1/3, 0] ensure second–order accuracy and

30 30
Point A Point B Point C Point A Point B Point D

25 25
Uz + u(t)

Uz + u(t)

20 20

15 15

10 10
0 30 60 90 120 150 0 30 60 90 120 150
Time (s) Time (s)
(a) (b)
Fig. 8. Horizontal component of the wind velocity field at different points for Ūz = 20 m/s : (a) points A, B and C and (b) points A, B and D.

7
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

(a) (b)
Fig. 9. Rail irregularities: (a) elevation PSD function and (b) deviations from the left rail.

unconditional stability to the integration algorithm. excessive vertical accelerations on the deck. Therefore, it is obvious that
Finally, regarding the running safety analysis presented in Section this criterion does not cover the type of scenarios that are being eval-
4.6, the possibility of derailment is evaluated based on three criteria, uated in this study, thus making it necessary to evaluate them through
namely Nadal, Prud’homme and unloading. If the maximum value of at train–bridge interaction models and using proper derailment criteria
least one of the criteria exceeds the limit allowance during the time that based on wheel–rail contact forces. Finally, it can also be seen that the
the train is still over the bridge, the circulation is considered to be vertical response is practically not affected by the wind speed, since its
unsafe for that specific combination of wind and train speeds. main direction is lateral.
The displacement response of the bridge at the midspan of the third
4.5. Dynamic response of the system span when the train is running at 300 km/h subjected to three different
wind loads is plotted in Fig. 11. This speed corresponds to the higher
4.5.1. Bridge response critical speed according to Fig. 10 within the operating limit of this
In order to assess the possible resonances that may occur in the high–speed line, which is 350 km/h as mentioned in [17]. It is clear
bridge due to the passage of the train, the maximum vertical accel- that the lateral response is strongly influenced by the wind forces, while
erations evaluated at the same aforementioned point (midspan of the the vertical one is mainly affected by the passage of the train. In fact,
third span) for train speeds ranging from 140 km/h to 420 km/h are the vertical response of the bridge tends to be slightly lower for higher
depicted in Fig. 10. As expected, since the bridge is continuous, it is not wind velocities due to the fact that the lifting force acting on the train
significantly prone to resonance effects. Note that the level of vertical alleviates its weight and, consequently, the dynamic effect caused by it.
accelerations is still very small when compared to the limit of 3.5 m/s2 Similar conclusions can be drawn from the accelerations evaluated at
imposed by EN 1990-Annex A2 [43] (the maximum vertical accelera- the same point of the bridge and depicted in Fig. 12. The passage of the
tion is 0.14 m/s2). This limit is defined to indirectly control the running train over the third span is clearly noticed, not only from the vertical
safety through the bridge response, since for this level of accelerations response, but also in the lateral one, where higher frequencies appear in
the ballast may start to experience instabilities that may lead to de- the response caused by the lateral forces transmitted by the train to the
railment. However, in the present study, as will be seen in Section 4.6, bridge through the wheel–rail interface.
the running safety is put at risk mainly due to the presence of strong
crosswinds that may lead to vehicle overturning and not due to
4.5.2. Train response
Fig. 13 shows the lateral and vertical accelerations in the carbody of
0.15 the first car for the same scenarios previously described in Section 4.5.1.
Uz = 0 m/s
Vertical acceleration (m.s-2)

Uz = 10 m/s The wind load has a strong influence on both the lateral and vertical
0.12 Uz = 20 m/s responses due to the drag and lift forces that are directly applied to the
train. Moreover, it is clear that the track irregularities impose higher
frequency vibrations to the vehicle, since, even without wind, it is pos-
0.09 sible to observe important fluctuations in the acceleration response on
both lateral and vertical directions. These accelerations, however, are
0.06 already substantially smaller than those registered in the wheelsets and
bogies due to the filtering capacity of the train suspensions.

0.03
4.6. Safety assessment against crosswinds

0 4.6.1. Derailment criteria


140 180 220 260 300 340 380 420 In the present work, the train running safety against crosswinds is
Vv (km/h) assessed based on the Nadal, Prud’homme and unloading criteria. The
first is defined as the ratio between lateral Y and vertical Q contact
Fig. 10. Maximum vertical accelerations at the midspan of the third span for forces. The Prud’homme criterion aims to limit the total lateral contact
different train and mean wind speeds. force in one wheelset ws Y to avoid panel shift. Finally, the unloading

8
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

25 10
Uz = 0 m/s

Vertical displacement (mm)


Lateral displacement (mm) 20 Uz = 10 m/s 5
Uz = 20 m/s
15 0

10 -5

5 -10
Uz = 0 m/s
0 -15 Uz = 10 m/s
Uz = 20 m/s
-5 -20
0 2 4 6 8 10 0 2 4 6 8 10
Time (s) Time (s)
(a) (b)
Fig. 11. Bridge displacement response at the midspan of the third span for Vv = 300 km/h: (a) lateral and (b) vertical directions.

criterion controls the excessive lift of the two wheels i and j of one side Finally, it is important to notice that the train’s stability is not re-
of the bogie by comparing the static vertical load Q0 with the corre- lated to any specific critical location in the bridge, since the maximum
sponding dynamic force. Table 1 shows the dimensionless indexes ξ of values of the derailment factors are related to the instants in which the
each criterion, the corresponding allowances and the filter types that wind speed, and consequently the wind load, is higher. Therefore, de-
have to be applied to the indexes’ time-histories. References for the pending on the random generated stochastic wind field used in a spe-
allowances and filter types are also included in the table. cific simulation, the maximum values of the derailment factors may
occur in different positions of the bridge.

4.6.2. Running safety evaluation 4.6.3. Influence of the bridge lateral flexibility in the safety evaluation
The distribution of the maximum safety indexes obtained for each One question that may arise from the running safety evaluation
combination of train and wind speeds are depicted in Fig. 14 (Ūz refers against crosswinds presented in the previous Section 4.6.2 consists of
to the 10 min mean wind velocity at height z = 48 m). The first main evaluating the influence of the lateral flexibility of the bridge in the
conclusion that may be drawn from these graphics is that the train can train derailment risk. To answer this question, the following four sce-
safely cross the bridge at any speed if the wind mean velocity does not narios illustrated in Fig. 15 have been investigated:
exceed 15 m/s. Even for wind loads characterized by mean velocities of
20 m/s, the safety is put at risk only for very high train speeds (higher (i) Scenario A: both the bridge and the train are subjected to a wind
than 350 km/h), based on the Prud’homme criteria. However, this field characterized by a mean speed of 20 m/s (scenario evaluated
problem becomes significantly more important when the mean wind in Section 4.6.2).
velocity reaches 25 m/s, since the unloading criterion allowance is (ii) Scenario B: no wind loads are applied to the system (scenario
exceeded for train speeds lower than 300 km/h, which is still a common evaluated in Section 4.6.2).
travelling speed in a typical HSR line. Hence, for this particular case, (iii) Scenario C: only the train is subjected to a wind load corresponding
the introduction of countermeasures, such as the installation of a to a mean speed of 20 m/s (hypothetical scenario).
monitoring system to send information to the train to reduce its speed (iv) Scenario D: only the bridge is subjected to a wind load corre-
when a wind velocity threshold is reached or the introduction of wind sponding to a mean speed of 20 m/s (hypothetical scenario).
barriers, could be a viable option to decrease the risk of derailment.

0.2 0.1
Uz = 0 m/s
Vertical acceleration (m.s-2)
Lateral acceleration (m.s-2)

Uz = 10 m/s
Uz = 20 m/s
0.1 0.05

0 0

-0.1 -0.05 Uz = 0 m/s


Uz = 10 m/s
Uz = 20 m/s
-0.2 -0.1
0 2 4 6 8 10 0 2 4 6 8 10
Time (s) Time (s)
(a) (b)
Fig. 12. Bridge acceleration response at the midspan of the third span for Vv = 300 km/h: (a) lateral and (b) vertical directions.

9
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

1 1
Uz = 0 m/s Uz = 0 m/s

Vertical acceleration (m.s-2)


Lateral acceleration (m.s-2) Uz = 10 m/s
Uz = 20 m/s
Uz = 10 m/s
Uz = 20 m/s
0.5 0.5

0 0

-0.5 -0.5

-1 -1
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Time (s) Time (s)
(a) (b)
Fig. 13. Accelerations in the carbody of the first car for Vv = 300 km/h: (a) lateral and (b) vertical directions.

Table 1
Safety criteria used in the present study to assess the train running safety.
Criterion Criterion index Allowances Filter type

Nadal =
Y 0.8 [44] Low–pass filter with a cut–off frequency of 20 Hz using a 4th order filter after performing
N Q
a sliding mean based on a window size of 2.0 m [45].
Prud’homme ws Y 1.0 [44]
P = 2Q0
10 + [kN]
3
Unloading Qi + Qj 0.9 [16] Low–pass filter with a cut–off frequency of 2 Hz using a 4th order Butterworth filter [16].
U =1
2Q 0

Based on these four scenarios, the maximum Prud’homme and un- those obtained with the normative Chinese Hat wind model proposed in
loading indexes for each analyzed running speed are depicted in EN 14067-6 [16] and shown in the authors’ previous publication [17].
Fig. 16. Moreover, the time–histories of these both criteria for the four In this latter model, the turbulent component of the wind is modelled
analyzed scenarios and a train running speed of Vv = 300 km/h are also through an ideal single gust evaluated for a fixed wind scenario, con-
plotted in Fig. 17. It can be observed that both safety indexes obtained sidering a reference height of z = 4 m and a roughness length of
in the scenarios A and C are very similar, meaning that the derailment z0 = 0.07 m, which leads to a turbulence intensity of 24.46%. Details
risk is mostly influenced by the wind that is acting directly on the train about the formulation of the wind velocity profile through the Chinese
and not so much by the lateral response of the bridge when subjected to Hat model can be found in the Annex I of EN 14067-6 [16].
crosswinds. For the scenarios A and C, the maximum Prud’homme The Volga River bridge has been previously subjected to a running
index when the train is running at Vv = 300 km/h is, respectively, safety study presented in [17] using the normative Chinese Hat discrete
0.799 and 0.766, while the unloading index reaches 0.677 and 0.657. gust model. In that study, it was assumed that the gust peak occurs at
This conclusion is emphasized when the scenarios B and D are com- the midspan of the third span of the bridge in order to cause maximum
pared since the wind that is acting only on the bridge in the latter lateral displacements to the deck. Hence, in order to confront the sto-
scenario does not significantly affect the safety indexes obtained in the chastic and discrete wind models, the results obtained in the present
former one, in which no wind loads are applied to the system. paper are compared with those obtained in [17].
The aforementioned results and conclusions, although cannot be The distribution of the maximum Prud’homme and unloading safety
extrapolated to other types of bridges without performing the re- indexes obtained in the present study (stochastic wind model) and in
spective calculations, may be seen as a first step to facilitate future [17] for the combination of trains speeds ranging from 140 km/h to
analyses of this type, since in some cases the bridge model can be sig- 420 km/h with three levels of wind velocity are plotted in Fig. 18 (the
nificantly simplified (considered rigid, for example) due to its low in- subscripts st and dg indicates stochastic and discrete gust models, re-
fluence in the train safety. However, in situations where the running spectively, while Ūz refers to the 10 min mean wind velocity at height
safety evaluation is not the only topic to be addressed, for example, z = 48 m for both models). It can be observed that the Chinese Hat
when the bridge response has to be assessed, such simplifications model is more conservative for all the scenarios, leading to higher
cannot be undertaken. Hence, further studies should be carried out to safety index values. Such results are due to the fact that the discrete
analyze the flexibility limits of the structure for which the aforemen- gust model proposed in [16] is not suitable to be used in bridges since it
tioned conclusions remain valid. imposes a height above the terrain of z = 4 m with an associated tur-
bulence intensity of 24.46%. The decks of the bridges, however, may be
5. Comparison between stochastic and normative discrete gust located at higher elevations, resulting in much lower turbulence in-
wind models tensity and, consequently, in a lower ratio between peak and mean
wind velocity. In the present case, since the deck is located at z = 48 m,
The present section aims to compare the critical wind velocities the turbulence intensity is only 14.56%, leading to lower peak wind
obtained with the stochastic wind model adopted in this paper with velocities using the stochastic model. Thus, for the same mean wind

10
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

(a) (b)

(c)
Fig. 14. Maximum safety indexes for several wind speeds: (a) Nadal, (b) Prud’homme and (c) unloading criteria.

velocity Ūz at height z = 48 m, the discrete gust model leads to higher available in the literature that compares both models in terms of safety
peak wind velocities and, generally, to more unfavorable safety in- indexes.
dexes. For the particular case of the Volga River bridge studied in this
Fig. 19 presents the running safety curve that indicates the critical work, in which the maximum mean wind velocity registered is
wind velocities that the vehicle can withstand, for each analyzed run- Ūz = 20 m/s [17], the running safety is guaranteed for train speeds up
ning speed, before exceeding any of the safety criteria allowances. The to 200 km/h when considering the discrete gust wind model, while
degree of conservativeness of the Chinese Hat gust model is once again when using the stochastic model the train can safely cross the bridge to
very notorious, leading to critical mean wind velocities 5 to 6 m/s lower speeds up to 380 km/h.
than those obtained with the stochastic model, which may significantly
condition the operation. Thus, although more complex to implement, 6. Conclusions
the models based on stochastic wind velocity fields are an interesting
option when investigating the derailment of trains running over The present article presents a framework for evaluating the stability
bridges. These conclusions are of the utmost importance to the current of trains moving over a bridge subjected to crosswinds. The analyses
state of the art on this field since there is currently no information have been performed using a train–bridge interaction tool programmed

Fig. 15. Scenarios for assessing the influence of the bridge flexibility in the train running safety.

11
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

(a) (b)
Fig. 16. Maximum safety indexes for scenarios A to D: (a) Prud’homme and (b) unloading criteria.

1 1
Scenario A Scenario C Scenario A Scenario C
Scenario B Scenario D Scenario B Scenario D
0.8 0.8
P

U
Prudhomme factor

Unloading factor

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Time (s) Time (s)
(a) (b)
Fig. 17. Time-history of the safety indexes for scenarios A to D for Vv = 300 km/h: (a) Prud’homme and (b) unloading criteria.

(a) (b)
Fig. 18. Comparison of the maximum safety indexes for several wind velocities obtained with the stochastic (present study) and discrete gust [17] models: (a)
Prud’homme and (b) unloading criteria.

12
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

Fig. 19. Running safety curve obtained with the stochastic (present study) and Chinese Hat [17] wind models.

in MATLAB that imports the structural matrices of the train and bridge obtained in the discrete gust model is higher than that obtained with
models previously developed in the commercial finite element method the stochastic formulation, leading to an unrealistic overestimation of
software ANSYS. The proposed approach simulates the wind velocity the critical wind velocities. Such conclusions are extremely important
field through independent one–dimensional multivariate stochastic to the current state of the art of this field since this information is
processes based on cross-spectral matrices that correlate the wind currently not available in the literature. Hence, although more complex
spectra in each generation point. and computationally less attractive when compared to the discrete gust
The proposed framework has been tested in a bridge that belongs to models, the stochastic generation of wind velocity fields leads to more
the future Russian HSR network. The lateral response of the bridge is accurate and realistic results, making them a fundamental tool to in-
strongly influenced by the wind loads, while the vertical one is mainly vestigate the train running safety in the presence of crosswinds. The
affected by the train passage. A slight reduction in the displacement’s discrete gust models, however, may be used in a first estimation of the
vertical response of the bridge has been detected due to the effect of the critical wind velocities in the design phase, but further studies should
lifting force acting on the train that alleviates its weight. be carry out to improve them in order to obtain more realistic results
Regarding the running safety analysis, the results indicate that for when analyzing the running safety of trains moving over bridges.
wind loads characterized by mean wind velocities up to 15 m/s the
train can always safely cross the bridge. However, for wind velocities of Credit authorship contribution statement
20 m/s, the Prud’homme factor starts to exceed the allowance value for
very high speeds (higher than 350 km/h), while the limits of the re- P.A. Montenegro: Conceptualization, Methodology, Software,
maining factors are exceeded only for wind velocities of 25 m/s but for Validation, Investigation, Writing - original draft, Writing - review &
typical travelling speeds in HSR networks of about 300 km/h. The in- editing. D. Barbosa: Methodology, Software, Visualization. H.
stallation of countermeasures, such as wind barriers, could be a solution Carvalho: Methodology, Investigation, Writing - review & editing. R.
to overcome this problem. Moreover, the results show that, for this Calçada: Visualization, Supervision.
particular bridge, the lateral flexibility of the structure has an almost
insignificant influence in the train’s running safety, since the safety Declaration of Competing Interest
indexes remain almost the same for the scenarios with and without
wind load applied to the bridge. Therefore, the derailment risk is mostly The authors declare that they have no known competing financial
influenced by the wind that is acting directly on the train and not so interests or personal relationships that could have appeared to influ-
much by the lateral response of the bridge. Such conclusions, however, ence the work reported in this paper.
should not be extrapolated to other types of bridges with different
lateral flexibilities without performing further studies that analyze its Acknowledgements
limits of validity.
Important conclusions regarding the comparison between the sto- This work was financially supported by: Base Funding - UIDB/
chastic and the discrete gust wind model proposed by the European 04708/2020 and Programmatic Funding - UIDP/04708/2020 of the
normative are also drawn in the present paper. In the single discrete CONSTRUCT - Instituto de I&D em Estruturas e Construções ‑ funded by
gust model, the turbulent component of the wind is simply considered national funds through the FCT/MCTES (PIDDAC). The authors would
as an ideal single gust related to a reference height above the ground of also like to express their gratitude to the FiberBridge – Fatigue
4 m, which is much lower than the height of a typical deck. Such as- strengthening and assessment of railway metallic bridges using fiber-
sumption led to conservative results, with critical wind velocities 5 to reinforced polymers (POCI-01-0145-FEDER-030103) by FEDER funds
6 m/s lower than those obtained with the stochastic formulation, since through COMPETE2020 (POCI) and by national funds (PIDDAC)
the turbulence intensity associated with a mean wind velocity at a through the Portuguese Science Foundation (FCT/MCTES). Finally, the
height z = 4 m above the ground is much higher than that at a height authors would also like to thank Eng. Alexander Bolkovoy and Eng.
z = 48 m (height of the deck of the Volga River bridge). As a con- Ivan Chebykin from OJSC Institute Gyprostroymost in Moscow, Russia,
sequence, and for the same scenario, the maximum wind velocity for providing the design data relative to the Volga River HSR bridge.

13
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

Appendix A. – Mechanical and dynamic properties of the Volga River bridge

Table A.1 presents the main mechanical properties of the numerical model of the Volga River high–speed railway bridge.
Fig. A.1 presents the numerical mode shapes and the natural frequencies of the first three vertical and lateral modes of vibration of the Volga
River high–speed railway bridge.

Table A1
Mechanical properties of the numerical model of the Volga River Bridge.
Designation Value Unit Reference

Modulus of elasticity of the deck concrete 39.5 GPa Designer


Modulus of elasticity of the pier concrete 34.5 GPa Designer
Density of concrete 2500 kg/m3 Designer
Non–structural mass 14,500 kg/m Designer
Ballast's longitudinal stiffness 30 MN/m/m [46]
Ballast's transversal stiffness 7.5 MN/m/m [47]
Ballast's vertical stiffness 100 MN/m/m [47]
Ballast's damping (3 directions) 50 kN.s/m/m [48]
Fastener's longitudinal stiffness 20 MN/m [49]
Fastener's transversal stiffness 20 MN/m [49]
Fastener's vertical stiffness 500 MN/m [50]
Fastener's rotational stiffness 45 kN.m/rad [47]
Fastener's longitudinal damping 50 kN.s/m [49]
Fastener's transversal damping 50 kN.s/m [49]
Fastener's vertical damping 200 kN.s/m [50]

Fig. A1. Numerical mode shapes and natural frequencies of the Volga River Bridge.

14
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

Appendix B. – Dynamic properties of the Shinkansen high–speed train

Fig. B.1 presents the numerical mode shapes and the natural frequencies of the Shinkansen car adopted in the present study. The carbody and
bogie are denoted by CB and B, respectively.

(a) CB: 1st rolling - 0.58 Hz (b) CB: bouncing - 0.86 Hz (c) CB: 2nd rolling - 0.87 Hz

(d) CB: yawing - 0.96 Hz (e) CB: pitching - 1.04 Hz (f) B: bouncing - 6.61 Hz

(g) B: pitching - 6.91 Hz (h) B: rolling - 9.34 Hz (i) B: yawing - 26.90 Hz


Fig. B1. Numerical mode shapes and natural frequencies of the Shinkansen car.

References compstruc.2008.04.007.
[11] Guo WW, Xia H, Youlin X. Running safety analysis of a train on the Tsing Ma Bridge
under turbulent winds. Earthquake Eng Eng Vibrat 2010;9(3):307–18. https://doi.
[1] Dai G, Hu N, Liu W. The recent improvement of high-speed railway bridges in org/10.1007/s11803-010-0015-3.
China. 34th IABSE Symposium for Bridge and Structural Engineering: Large [12] Xu L, Zhai W. A model for vehicle–track random interactions on effects of cross-
Structures and Infrastructures for Environmentally Constrained and Urbanised winds and track irregularities. Veh Syst Dyn 2019;57(3):444–69. https://doi.org/
Areas, Venice, Italy. 2010. 10.1080/00423114.2018.1469775.
[2] Ishibashi T. Shinkansen structures in Japan. Porto, Portugal: Workshop on Bridges [13] Xu L, Zhai W. Cross wind effects on vehicle–track interactions: a methodology for
for High-Speed Railways; 2004. dynamic model construction. J Comput Nonlinear Dyn 2019;14(3):1–11. https://
[3] Kao TC, Lin CK. Taiwan high speed rail & its impact to regional development. 4th doi.org/10.1115/1.4042142.
CECAR - Civil Engineering Conference in the Asian Region, Taipei, Taiwan. 2007. [14] Zhang T, Xia H, Guo WW. Analysis on running safety of train on bridge with wind
[4] Castillo E, Calviño A, Grande Z, Sánchez-Cambronero S, Gallego I, Rivas A, et al. A barriers subjected to cross wind. Wind Struct 2013;17(2):203–25. https://doi.org/
Markovian-Bayesian network for risk analysis of high speed and conventional 10.12989/was.2013.17.2.203.
railway lines integrating human errors. Comput-Aided Civ Infrastruct Eng [15] Guo WW, Xia H, Karoumi R, Zhang T, Li X. Aerodynamic effect of wind barriers and
2016;31(3):193–218. https://doi.org/10.1111/mice.12153. running safety of trains on high-speed railway bridges under cross winds. Wind
[5] Andersson A, O’Connor A, Karoumi R. Passive and adaptive damping systems for Struct 2015;20(2):213–36. https://doi.org/10.12989/was.2015.20.2.213.
vibration mitigation and increased fatigue service life of a tied arch railway bridge. [16] EN 14067-6. Railway applications - Aerodynamics - Part 6: Requirements and test
Comput-Aided Civ Infrastruct Eng 2015;30(9):748–57. https://doi.org/10.1111/ procedures for cross wind assessment, European Committee for Standarization
mice.12116. (CEN), Brussels, 2010.
[6] Olmos JM, Astiz MA. Improvement of the lateral dynamic response of a high pier [17] Montenegro PA, Carvalho H, Calçada R, Bolkovoy A, Chebykin I. Stability of a train
viaduct under turbulent wind during the high-speed train travel. Eng Struct running over the Volga River high speed railway bridge during crosswinds. Struct
2018;165:368–85. https://doi.org/10.1016/j.engstruct.2018.03.054. Infrastruct Eng 2019:1–17. https://doi.org/10.1080/15732479.2019.1684956 (in
[7] Xu YL, Xia H, Yan QS. Dynamic response of suspension bridge to high wind and press).
running train. J Bridge Eng 2003;8(1):46–55. https://doi.org/10.1061/(ASCE) [18] Montenegro PA, Neves SGM, Calçada R, Tanabe M, Sogabe M. Wheel-rail contact
1084-0702(2003) 8:1(46). formulation for analyzing the lateral train-structure dynamic interaction. Comput
[8] Xu YL, Xia H, Zhang N. Vibration of coupled train and cable-stayed bridge systems Struct 2015;152:200–14. https://doi.org/10.1016/j.compstruc.2015.01.004.
in cross winds. Eng Struct 2004;26(10):1389–406. https://doi.org/10.1016/j. [19] Montenegro PA, Calçada R, Vila-Pouca N, Tanabe M. Running safety assessment of
engstruct.2004.05.005. trains moving over bridges subjected to moderate earthquakes. Earthquake Eng
[9] Li Y, Qiang S, Liao H, Xu YL. Dynamics of wind–rail vehicle–bridge systems. J Wind Struct Dyn 2016;45:483–504. https://doi.org/10.1002/eqe.2673.
Eng Ind Aerodyn 2005;93(6):483–507. https://doi.org/10.1016/j.jweia.2005.04. [20] Cao Y, Xiang H, Zhou Y. Simulation of stochastic wind velocity field on long-span
001. bridges. J Eng Mech 2000;126(1):1–6. https://doi.org/10.1061/(ASCE)0733-
[10] Xia H, Guo WW, Zhang N, Sun GJ. Dynamic analysis of a train–bridge system under 9399(2000) 126:1(1).
wind action. Comput Struct 2008;86(19–20):1845–55. https://doi.org/10.1016/j. [21] Davenport AG. The dependence of wind load upon meteorological parameters.

15
P.A. Montenegro, et al. Engineering Structures 211 (2020) 110430

Toronto, Canada: University of Toronto Press; 1968. 007.


[22] Holmes JD. Wind Loading of Structures. 3rd ed. New York, USA: Spon Press - Taylor [38] Strömmen E. Theory of Bridge Aerodynamics. 1st ed. Berlin, Germany: Springer;
& Francis Group; 2001. 2006.
[23] EN 1991-1-4. Eurocode 1: Actions on structures - Part 1-4: General actions - Wind [39] Simiu E, Scanlan RH. Wind eEffects on Structures. 3rd ed. New York, USA: John
actions, European Committee for Standarization (CEN), Brussels, 2010. Wiley & Sons; 1996.
[24] Montenegro PA, Neves SGM, Azevedo AFM, Calçada R. A nonlinear vehicle-struc- [40] Claus H, Schiehlen W. Modeling and simulation of railway bogie structural vibra-
ture interaction methodology with wheel-rail detachment and reattachment, tions. Veh Syst Dyn 1998;29(1):538–52. https://doi.org/10.1080/
COMPDYN 2013 - 4th ECCOMAS Thematic Conference on Computational Methods 00423119808969585.
in Structural Dynamics and Earthquake Engineering, Kos, Greece, 2013. [41] EN 13848-5. Railway applications - Track - Track geometry quality - Part 5:
[25] Neves SGM, Montenegro PA, Azevedo AFM. Calçada R, A direct method for ana- Geometric quality assessment, European Committee for Standarization (CEN),
lyzing the nonlinear vehicle–structure interaction. Eng Struct 2014;69:83–9. Brussels, 2005.
https://doi.org/10.1016/j.engstruct.2014.02.027. [42] Hughes TJR. The finite element method: Linear static and dynamic finite element
[26] Bathe KJ. Finite element procedures. Upper Saddle River, NJ: Prentice-Hall; 1996. analysis. New York City, NY, USA: Dover Publications; 2000.
[27] Hertz H. Ueber die Berührung fester elastischer Körper [On the contact of elastic [43] EN 1990-Annex A2. Eurocode 0: Basis of structural design - Annex 2: Application for
solids]. J für die reine und angew Mathematik 1882;92:156–71. bridges (normative), European Committee for Standarization (CEN), Brussels, 2001.
[28] Shabana A, Zaazaa KE, Sugiyama H. Railroad vehicle dynamics: A computational [44] TSI. Technical specification for interoperability relating to the infrastructure sub-
approach. Boca Raton, USA: CRC Press; 2008. system of the trans-European high-speed rail system, Official Journal of the
[29] Kalker JJ. Book of tables for the Hertzian creep-force law, 2nd Mini Conference on European Union, Brussels, 2002.
Contact Mechanics and Wear of Wheel/Rail Systems. Hungary: Budapest; 1996. [45] EN 14363. Railway applications - Testing and Simulation for the acceptance of
[30] MATLAB®. Release R2018a, The MathWorks Inc., Natick, MA, USA, 2018. running characteristics of railway vehicles - Running Behaviour and stationary
[31] ANSYS®. Academic Research, Release 19.2, ANSYS Inc., Canonsburg, PA, USA, tests, European Committee for Standarization (CEN), Brussels, 2016.
2018. [46] UIC. 774-3-R. Track/bridge interaction - Recommendations for calculation. 2nd Ed.
[32] Institute Giprostroymost, Accessed November 2017. http://www. Paris: International Union of Railways (UIC); 2001.
giprostroymost.ru/. [47] ERRI D 202/RP 11. Improved knowledge of forces in CWR track (including
[33] Clough RW, Penzien J. Dynamics of structures. Third Ed. New York City, NY, USA: switches): Parametric study and sensivity analysis of CWERRI, European Rail
McGraw-Hill Inc; 2003. Research Institute, Utrecht, 1999.
[34] EN 1991-2. Eurocode 1: Actions on structures - Part 2: Traffic loads on bridges, [48] Wu YS, Yang YB. Steady-state response and riding comfort of trains moving over a
European Committee for Standarization (CEN), Brussels, 2003. series of simply supported bridges. Eng Struct 2003;25(2):251–65. https://doi.org/
[35] Kaimal JC, Wyngaard JC, Izumi Y, Coté OR. Spectral characteristics of surface-layer 10.1016/S0141-0296(02)00147-5.
turbulence. Q J R Meteorolog Soc 1972;98:563–89. https://doi.org/10.1002/qj. [49] Zhai W, Wang K, Cai C. Fundamentals of vehicle-track coupled dynamics. Veh Syst
49709841707. Dyn 2009;47(11):1349–76. https://doi.org/10.1080/00423110802621561.
[36] Lumley JL, Panofsky HA. The structure of atmospheric turbulence. New York, USA: [50] ERRI D 214/RP 5. Rail bridges for speeds > 200 km/h: Numerical investigation of
Interscience Publishers; 1964. the effect of track irregularities at bridge resonance, European Rail Research
[37] Cai CS, Chen SR. Framework of vehicle–bridge–wind dynamic analysis. J Wind Eng Institute, Utrecht, 1999.
Ind Aerodyn 2004;92(7–8):579–607. https://doi.org/10.1016/j.jweia.2004.03.

16

You might also like