You are on page 1of 18

Journal of Sound and Vibration 383 (2016) 277–294

Contents lists available at ScienceDirect

Journal of Sound and Vibration


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

Effect of vertical ground motion on earthquake-induced


derailment of railway vehicles over simply-supported bridges
Zhibin Jin a,c,n, Shiling Pei a,b, Xiaozhen Li a,c, Hongyan Liu b, Shizhong Qiang a
a
National Engineering Laboratory for Technology of Geological Disaster Prevention in Land Transportation, Southwest Jiaotong University,
Chengdu, PR China
b
Department of Civil and Environmental Eng., Colorado School of Mines, Golden, USA
c
MOE Key Laboratory of High-Speed Railway Eng., Southwest Jiaotong University, Chengdu, PR China

a r t i c l e i n f o abstract

Article history: The running safety of railway vehicles on bridges can be negatively affected by earthquake
Received 16 April 2016 events. This phenomenon has traditionally been investigated with only the lateral ground
Received in revised form excitation component considered. This paper presented results from a numerical inves-
14 June 2016
tigation on the contribution of vertical ground motion component to the derailment of
Accepted 30 June 2016
vehicles on simply-supported bridges. A full nonlinear wheel–rail contact model was used
Handling Editor: L.G. Tham
Available online 4 August 2016 in the investigation together with the Hertzian contact theory and nonlinear creepage
theory, which allows the wheel to jump vertically and separate from the rail. The wheel–
Keywords: rail relative displacement was used as the criterion for derailment events. A total of 18
Railway vehicle
ground motion records were used in the analysis to account for the uncertainty of ground
Bridge–train coupled vibration
motions. The results showed that inclusion of vertical ground motion will likely increase
Derailment
Vertical ground motion the chance of derailment. It is recommended to include vertical ground motion compo-
nent in earthquake induced derailment analysis to ensure conservative estimations. The
derailment event on bridges was found to be more closely related to the deck acceleration
rather than the ground acceleration.
& 2016 Elsevier Ltd. All rights reserved.

1. Introduction

During an earthquake, the train running safety on bridges may be governed by derailment due to excessive vibration
before the structural failure of the bridge. For example, a Shinkansen train derailed on the bridge during the Niigata
earthquake of 2004 in Japan without major bridge structural damage [1]. In modern high-speed railway networks, since the
percentage of total railway length on bridges often exceeds 50%, there is a high probability that the trains will be running
bridge structures when an earthquake hits. This makes earthquake-induced derailment on bridges a practical concern for
the seismic resiliency of high-speed railway networks. For example, 80.5% the total length (1318 km) of high-speed line from
Peking to Shanghai in China are covered by bridges [2], most of them are simply-supported structures. For modern high-
speed railway networks, it is of great importance to develop a fundamental understanding of the earthquake-induced
vehicle derailment on simply-supported bridge. This provides the motivation of this study.

n
Corresponding author at: Department of Bridge Engineering, Southwest Jiaotong University, Chengdu 610031, PR China. Tel.: þ 8602886466917; fax:
þ 86 02887603323.
E-mail address: jinzhibin@home.swjtu.edu.cn (Z. Jin).

http://dx.doi.org/10.1016/j.jsv.2016.06.048
0022-460X/& 2016 Elsevier Ltd. All rights reserved.
278 Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294

The derailment of vehicles on earthquake-excited tracks has been investigated both numerically and experimentally in
past studies. Matsuura [3] proposed a 3D multi-body vehicle model to simulate the derailment process. In this model, the
rails are simplified as oscillators moving along with the wheels. Using this model, the derailment limit in term of ground
motion (simplified as sine waves) for high speed train on oscillating tracks was obtained. Nishimura et al. [4] used a similar
vehicle model to simulate the derailment of vehicles subjected to ground motion. In this study, the stiffness and damping of
the rail pads were considered, while the mass of the rail itself was omitted. Miyamoto et al. [5] carried out full-scale
experiments of a Shinkansen vehicle's derailment caused by earthquakes, and the experimental results confirmed the
effectiveness of the numerical simulations used in the design of railway structures in Japan. Xiao et al. [6] investigated the
derailment mechanism of a railway vehicle on slab tracks, where the rails were modeled as Timoshenko beams and the
slabs were modeled by solid finite elements. In all these studies, displacement based derail criteria were adopted, where the
uplift of wheel relative to the rail was used to identify derailment condition. Sogabe et al. [7] studied the riding quality of
vehicles on long span bridges with high piers, where the rail was modeled as beam elements connected to the bridge deck
by springs representing the rail fasteners. Numerical investigation found that increasing the damping ratio or pier rigidity
can help to increase the running safety of vehicles during earthquakes. Ju [8] investigates the derailment of high-speed
trains moving on multi-span simply-supported bridges. It was found that both higher train speed and larger gaps between
adjacent girders led to larger derailment coefficients. Kawanishi et al. [9] considered the traveling wave effects of earth-
quakes due to ground irregularity to the derailment of the vehicles. Yang et al. [10] developed a vehicle–rail–bridge
interaction (VRBI) system to simulate the responses of the vehicle under earthquake excitation. In their study, earthquake
motions were normalized to a moderate intensity level, and the derailment was evaluated based on the derailment coef-
ficients. Xia et al. [11] studied the vehicle running safety on bridge considering seismic wave propagation, and identified the
safety boundary for different running speeds and ground motion intensities. Yau and Frýba [12] investigated the vertical
responses of suspension bridges loaded simultaneously by moving vehicles and vertical ground motion.
Currently, most studies on vehicle–bridge vibration during earthquakes only considered the lateral [7–11] or the vertical
ground motion component [12] separately. The combined effects of vertical and lateral ground motion on the derailment
were not investigated. However, experimental study of derailment on oscillating ground has shown that the vertical
component can give rise to the uplift of the wheel especially in higher frequency range [5]. Thus there is a current
knowledge gap in understanding the train–bridge dynamic system behavior under simultaneous lateral and vertical ground
motion excitation. This study seeks to better understand the combined effect of the vertical and lateral ground motion on
the derailment of railway cars on simply-supported bridges. The railway vehicle in this study was modeled as a multi-body
system with 31 DOFs. The moving modal method of rail model [13] was adopted to simulate the vibration of the rail with
much less DOFs than using the finite element method (FEM). Similar to existing studies, a displacement based criterion was
used to assess the derailment. The responses of the vehicle, rail and bridge were simulated with and without vertical ground

Fig. 1. Vehicle–rail–bridge dynamic interaction system model. (a) Front view. (b) Side view.
Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294 279

motion respectively through the Zhai's explicit integration method. For each ground record, multiple simulations were
carried out where the PGA for each simulation was gradually increased following the procedure of incremental dynamic
analysis (IDA) [14].

2. Vehicle–rail–bridge dynamic interaction model for derailment simulation

The study of derailment of railway cars on bridges during earthquake involves the dynamic responses of vehicle, rail and
bridge systems as well as their interactions. The modeling diagram in Fig. 1a briefly describes of the interactive dynamic
model used in this study. This model was briefly introduced by the authors [15], and was expanded in this study to include
the vertical ground motion excitation.
The governing equation of vehicle–rail–bridge system is time-varying, since the location of the vehicle changes con-
tinuously when the vehicle moves forward. However, the governing equation of the whole system needs not to be actually
formulated owning to the adoption of the explicit integration algorithm in this study. Instead, the equations for each sub-
system of the vehicle, rail and bridge were formulated individually. For explicit algorithm, the dynamic equation can be
solved in elemental or nodal level. The time-varying nature of the whole system was represented by changing contact points
between the rail and bridge with time in calculation of interactive forces.

2.1. Vehicle model

The vehicle is idealized as seven inter-connected rigid bodies, i.e., one car-body, two trucks and four wheel-sets. The DOF
along the train traveling direction was omitted by assuming the vehicle travels at constant speed. The remaining five DOFs
of the car-body is Y c ¼ ðyc ; zc ; ϕc ; θc ; ψ c ÞT , those of the i-th truck (i¼ 1,2) is Y ti ¼ ðyti ; zti ; ϕti ; θti ; ψ ti ÞT . Symbols y; z; ϕ; θ; ψ
denote the lateral, vertical, rolling, pitching and yawing motion respectively; subscripts ‘c’ and ‘t’ stand for car-body and
truck respectively. The wheel-sets were assumed to rotate about the y axis at constant a speed V/r0, where V is the running
speed of the vehicle, and r0 is the contact radius of the wheel-set. So, the pitching motions of the four wheel-sets about the y
axis are omitted here. Thus, the remaining four DOFs of the j-th (j¼1–4) wheel-set is Y wj ¼ ðywj ; zwj ; ϕwj ; ψ wj ÞT . Since the
vertical and rolling motion of each wheel-set are both independent DOFs, the wheels can separate from the rail in order to
simulate the derailment conditions. The motion of a railway car are described by 31 DOFs as
Y V ¼ ðY c ; Y t1 ; Y t2 ; Y w1 ; Y w2 ; Y w3 ; Y w4 ÞT .
The car-body is connected to two trucks via the secondary suspension. The secondary suspension is composed of four
sets of springs and dampers. Each set of secondary suspension spring has three components, i.e, k2x ; k2y and k2z , in long-
itudinal, lateral and vertical directions respectively. Similarly, the dampers of the secondary suspension arec2x ; c2y and c2z .
The primary suspension connecting the trucks and wheel-sets is composed of eight sets of springs and dampers. For each
set of spring and damper, the stiffness and damping are denoted as k1x ; k1y ,k1z , and c1x ; c1y ,c1z , respectively.
In addition to the suspensions, an anti-roll bar with stiffness kR is installed between the car-body and each truck to
control the rolling motion. A stopper in the lateral direction with stiffness kS between the trucks and the car-body and is
included with an initial gap of 0.02 m.
Given the 31 DOFs of the vehicle and all springs and dampers, the motion equation of the vehicle can be written as

M V Y€ V þC V Y_ V þ K V Y V ¼ F gV þF rV (1)

where MV , C V and K V are the mass, damping and stiffness matrix of the vehicle, Y V is the displacement vector of the vehicle
with respect to the ground in motion. The overhead dot denotes the derivative operation with respect to time, F rV is the force
vector acting on the vehicle from the rails, and F gV is the seismic force applied on the vehicle. The mass, damping and
stiffness matrix can be derived automatically using the virtual work method proposed by the authors in an earlier study.
Interested readers can refer to Z. Jin et al. [15] for details.
The seismic force vector of the vehicle is the production of the ground acceleration and the vehicle mass components.
The seismic force vector of the vehicle F gV can be written as
 T
F gV ¼ F gc T; F gt1 T; F gt2 T; F gt1 T; F gw1 T; F gw2 T; F gw3 T; F gw4 T (2)

where F gc , F gti (i¼ 1,2), F gwj (j¼1–4) are the seismic force vectors applied on the car-body, the i-th truck and the j-th wheel-set
respectively with their explicit form as follows,
 T
F gc ¼  mc ay ðt Þ; mc az ðt Þ; 0; 0; 0 ;
 T
F gti ¼  mt ay ðt Þ; mt az ðt Þ; 0; 0; 0 ;
F w ig ¼  ½mw ay ðtÞ; mw az ðtÞ; 0; 0T : (3)

In Eq. (3), ay ðt Þ and az ðt Þ are the lateral and the vertical acceleration of the ground motion at time “t”;mc , mt and mw are
the mass of the car-body, truck and wheel-set respectively.
280 Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294

Table 1
Vehicle parameters.

Variable Meaning Value Unit

Mw Mass of wheel-set 1900 kg


Iwx, Iwz Roll and Yaw mass moment of wheel-set 1120 kg m2
Mt Mass of truck 3175 kg
Itx Roll mass moment of truck 2040 kg m2
Ity Pitch mass moment of truck 2710 kg m2
Itz Yaw mass moment of truck 4650 kg m2
Mc Mass of car-body 30060 kg
Icx Roll mass moment of car-body 8.0e4 kg m2
Icy, Itz Pitch and yaw mass moment of car-body 2.0e6 kg m2
K1x Primary longitudinal stiffness 1.05e7 N/m
K1y Primary lateral stiffness 1.20e7 N/m
K1z Primary vertical stiffness 1.26e7 N/m
C1z Primary vertical damping 2.00e4 N s/m
KΦ Anti-roll stiffness 1.60e6 N m/rad
C2x Secondary longitudinal damping 1.22e3 N s/m
C2y Secondary lateral damping 1.22e3 N s/m
C2z Secondary vertical damping 1.65e4 N s/m
Lt A half of axle distance 1.25 m
Lc A half of truck distance 9.00 m
b1, b2 A half of lateral distance of suspension springs 0.978 m
h1, h2 ,h3 Vertical distances 1.415, 0.10, 0.14 m
b Half of the distance between the two rolling circles 0.7465 m
r Radius of rolling circle of the wheel 0.450 m

Fig. 2. Wheel–rail forces.

The rail force vector applied on the vehicle F rV reads

F rV ¼ ½0151 ; F rw1 ; F rw2 ; F rw3 ; F rw4 T (4)

The zero vector 0151 was used in Eq. (4) to reflect that the rail forces were not applied on the DOFs of the car-body and
trucks. F rwj is the rail force vector applied to the DOFs of the j-th wheel-set, and it can be calculated from the wheel–rail
forces by
h      i
F rwj ¼ H Lwr;j þ H Rwr;j ; V Lwr;j þ V Rwr;j ; r H Lwr;j þH Rwr;j  b V Lwr;j V Rwr;j ; b LLwr;j  LRwr;j (5)

where H wr;j , V wr;j and Lwr;j denote the horizontal, vertical and longitudinal wheel–rail forces at their contact patch; the
superscript “L” and “R” denote the left and right rail respectively. The dimension “r” denotes radius of rolling circle of the
wheel, and “b” is half of the distance between the two rolling circles of the same wheel-set, and their values are listed in
Table 1. The wheel–rail forces and dimensions in Eq. (5) are shown in Fig. 2.
In this study, the dimensions and dynamic parameters of a typical railway vehicle in high-speed rail system of China are
listed in Table 1. Since the mass of train is small compared to that of the bridge, and the length of a car (26 m) is comparable
to the typical span of the simply-supported railway bridge considered in this study (32 m), the effect of multiple vehicles on
a bridge is not considered.

2.2. Rail model

The rail structure is flexible and can serve as a dynamic impact buffer between the vehicle and bridge. Previous studied
have shown that the rigid track model which omits the deformation of the track will result in unrealistic and excessive
wheel–rail forces. Hence, the effect of track dynamic deformation was considered in this study.
Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294 281

Fig. 3. Moving rail method. (a) Shape functions of the rail. (b) Moving rail model.

There are a number of track modeling methods used in the past, including generalized finite element approach [9] and
modal superposition method [8]. These two methods are very computationally intensive as they involve large number of
DOFs in the track model. In this study, an alternative moving track modeling method [13] that significantly decreases the
DOFs of the track was used (see Fig. 3). This method used the steady deflection of the rail induced by a moving concentrated
force as the shape function of the rail [16]. This shape function is attached to each wheel and is assumed to move forward
with the train.
In this work, the mode shape was further approximated by the static deformation of the rail due to a stationary force. The
difference between the steady-state and static deformation of the UIC60 rail is almost indistinguishable for ordinary
operation speed (below 300 km/h). Numerical results show that the error in modal properties of the rail (modal mass,
stiffness and frequency) induced by the approximation of the mode shape by the static deformation is less than 3%.
The moving mode is the deformation of the rail due to a static concentrated load,
 
wðxÞ ¼ e  βjxj ½ cos βjxjÞ þ sin βjxjÞ (6)
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where β ¼ krv ⁄ ð4Er I rv Þ, and krv is the vertical stiffness of the supporting pad under the rail per unit length, Er is the
elasticity modulus of rail material (steel) and I rv the bending moment of inertia about the lateral axis of the rail section.
The displacement of the rail with respect to the ground produced by the wheel–rail force can be expressed as
zr ðx; t Þ ¼ wðxÞqrv ðtÞ (7)
where qrv ðtÞ is the generalized displacement of the rail. Using the virtual work method, the motion equation of the rail with
respect to qrv ðtÞ can be derived as
Mrv q€ rv ðt Þ þ C rv q_ rv ðt Þ þ K rv qrv ðt Þ ¼ F r (8)
where Mrv , C rv and K rv are the modal mass, damping and stiffness; F r is the modal force from the wheel and the bridge deck.
The modal mass, damping and stiffness of the rail as well as the external force can be written as
Z 1 Z 1 Z 1  2 2
dw ðxÞ
M rv ¼ mr w2 ðxÞdx; K rv ¼ krv w2 ðxÞdx þ Er I rv dx;
1 1 1 dx2
Z 1
2
C rv ¼ crv w ðxÞdx; F r ¼ wð0ÞðV wr þV rb Þ ¼ ðV wr þV rb Þ: (9)
1

In Eq. (9), mr is the mass of the rail per unit length, V wr is vertical component of the wheel–rail force. The improper
integrals in Eq. (9) all converge and their analytical solution can be obtained.
The vertical force acting on the wheel from the bridge V rb can be approximated as

V rb ¼ K rv Z d þ C rv Z_ d (10)
where Z d is the vertical deck displacement beneath the moving rail. In all, the vertical vibration equations of the moving
rails beneath the eight wheels of the vehicle are
L;R
M rv q€ rv;i þ C rv q_ L;R L;R L;R L;R
rv;i þ K rv qrv;i ¼ V wr;i þ V rb;i  M rv az ði ¼ 1  4Þ (11)

In Eq. (11), the subscript ‘i’ denotes the i-th wheel-set, the superscript ‘L’ and ‘R’ represent the left and the right rail
respectively.
Since the shape function w(x) at x¼0 equals 1, the physical displacement zr ð0; t Þ equals the modal displacement qvr ðtÞ
from Eq. (7). In the following part of this paper, zr ð0; t Þ and qvr ðtÞ will be used interchangeably.
The lateral motion equations of the moving rail can be derived in a parallel way with the vertical rail equations as
L;R
Mrh q€ rh;i þ C rh q_ L;R L;R L;R L;R
rh;i þ K rh qrh;i ¼ H wr;i þH rb;i  M rh ay ði ¼ 1  4Þ (12)
282 Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294

Fig. 4. Dimensions of bridge deck.

Table 2
Rail parameters.

Variable Meaning Value Unit

mr Mass of the rail per unit length 60.64 kg/m


Er Elasticity modulus of rail material 2.0  1011 Pa
I rv Bending moment of inertia about the lateral axis of the rail section 3.217  10  5 m4
I rh Bending moment of inertia about the vertical axis of the rail section 5.24  10  6 m4
krv Vertical stiffness of rail pad per unit length 8.33  107 N/m2
c rv Vertical damping of rail pad per unit length 1.25  105 N s/m2
krh Lateral stiffness of rail pad per unit length 5.00  107 N/m2
c rh Lateral damping of rail pad per unit length 1.00  105 N s/m2

In Eq. (12), the subscript ‘h’ denotes the lateral direction, H denotes the lateral forces, and the meaning of other symbols
is the same to those in Eq. (11). In derivation of Eq. (12), the lateral mode shape of the rail is also assumed as the ffi function in
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Eq. (6). However, the rail constant β was calculated using the lateral properties of the rail, or β ¼ krh ⁄ ð4Er I rh Þ, where krh is
the lateral stiffness of the supporting pad under the rail per unit length, I rv is the bending moment of rail about the vertical
axis of the rail section.
The lateral forces applied on the rail from the bridge Hrb can be written as

H rb ¼ K rh Y d þ C rh Y_ d (13)

where Y d is the lateral deck displacement beneath the moving rail. The deck displacement beneath the moving rail can be
calculated from the motion of bridge deck by

Y L;R
d;i
¼ Y b ðxi Þ  θb ðxi Þhr ; Z Ld;i ¼ Z b ðxi Þ  θb ðxi Þbr ; Z Rd;i ¼ Z b ðxi Þ þ br (14)

The geometric relations in Eq. (14) are illustrated by Fig. 4, where the dimensions Y b ðxi Þ, Z b ðxi Þ and θb ðxi Þ are the lateral,
vertical and torsional displacement at the centroid of the bridge section, respectively, and xi is the longitudinal location of
the i-th wheel-set on the bridge deck which varies with time as the vehicle moves along the bridge. The dimensions hr and
br are the distance between the bottom of the rail and the centroid of the bridge section in vertical and lateral directions
respectively.
In this study, the parameters of the rail used in Chinese high-speed railways [17] are adopted and listed in Table 2.

2.3. Bridge model

The bridge was modeled using the finite element method (FEM). The bridge structure was discretized into 3D beam
elements with each node having 6 DOFs. The equation of the motion of the bridge represented by FEM models reads,
Mb u€ b þC b u_ b þ K b ub ¼  Mb Ry ay ðt Þ M b Rz az ðt Þ þ F br (15)

where M b , C b and K b are the mass, damping and stiffness matrix of the bridge, and ub is the displacement vector (with
respect to the ground) of the bridge. Vector Ry and Rz were used to apply the seismic force on the DOFs of the bridge in the
lateral and vertical direction, respectively. The i-th element Ry;z ðiÞ of the vector Ry;z can be written as
(
1; if the i  th DOF represents a translation in lateralðverticalÞ direction
Ry;z ðiÞ ¼ (16)
0; else

In Eq. (15), F br denotes the load vector applied on the bridge DOFs from the rail. When the i-th rail is located at a node of
the bridge deck at time t, the node force vector F N br is

FN
br ¼ ½0; H br;i ; V br;i ; T br;i ; 0; 0; 0
T
(17)

where H br;i ,V br;i and T br;i are the lateral, vertical and torsional loads acting on this node from the rails. These loads were
Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294 283

6700
1250 2100

500
200

340
600
R=500

250
×344
1033 Sym.

3000
Sym.

1800
500×500

500
300
2990 3400

Fig. 5. Cross section of the bridge, unit in mm. (a) Beam. (b) Pier.

Table 3
Sectional properties of the girder and the pier.

Element Area (m2) Torsional moment (m4) Bending moment in plane (m4) Bending moment out of plane (m4)

Girder 8.63 15.21 6.58 85.86


Pier 8.80 30.84 11.18 45.42

calculated by the following expressions,


   
H br;i ¼ K rh qLrh;i þqRrh;i Y Ld;i Y Rd;i þ C rh q_ Lrh;i þ q_ Rrh;i  Y_ d;i  Y_ d;i
L R
(18a)

   
V br;i ¼ K rv qLrv;i þ qRrv;i Z Ld;i  Z Rd;i þ C rv q_ Lrv;i þ q_ Rrv;i  Z_ d;i  Z_ d;i
L R
(18b)

   
T br;i ¼ H br;i hr  K rv br qLrv;i  Z Ld;i  qRrv;i þ Z Rd;i  C rv q_ Lrv;i  Z_ d;i  q_ Rrv;i þ Z_ d;i
L R
(18c)

When the wheel–rail contact point is located between two nodes, the displacements at the contact point can be determined
based on the interpolation of the shape functions.
As mentioned earlier, a large portion of bridges on the railway lines is simply supported. In this study, we investigate the
running safety of the train car on a series span of simply supported bridges. The beam span is 32 m (commonly seen in
Chinese high-speed rail network) with a box section plotted in Fig. 5(a). The hollow section of the piers was plotted in Fig. 5
(b), their sectional properties are listed in Table 3. The concrete modulus of elasticity for the girder is 35.5 GPa, and 30.0 GPa
for that the pier. The height of the piers is set to be 20 m. The beams and piers were modeled as 3D beam elements using the
finite element method (see Fig. 6). The connection between beam ends and pier tops were represented by the master–slave
constraints. The bridge is composed of 30 successive spans to create a scenario that the vehicle runs on the bridge during the
full duration of the earthquake.
The fundamental natural modes in the lateral direction and vertical direction are shown in Figs. 7 and 8. The natural
frequencies of the fundamental lateral and vertical modes are 3.465 Hz and 5.205 Hz respectively.

2.4. Wheel–rail interactions and derailment criterion

Because of the intensive interactions between the wheel and rail during earthquakes, a fully nonlinear wheel–rail
interaction model must be used to represent the nonlinear wheel–rail interaction. The model should be able to capture the
separation of the wheel from the rail and the impact of the uplifted wheel against the rail (i.e. wheel jump and impact). In
this study, the nonlinear interaction model uses the Hertz contact theory [18] to represent the wheel–rail contact in the
normal direction, and the Shen–Hedrick formula [19] for friction in the tangent direction. The Hertzian contact forces are
calculated from the wheel–rail virtual penetration. The contact geometries in calculating of wheel–rail contact forces are
determined by a contact point searching subroutine using wheel and rail profiles. In the end, the wheel–rail forces were
expressed as functions of the velocities and displacements of the wheel-set and rail,
 
_ L;R _ L;R _ L;R
V L;R L;R
wr;i ¼ f V ywi ; y wi ;qrh;i ; qrv;i ; q rh;i ; q rv;i ; (19.a)

 
_ L;R _ L;R _ L;R
H L;R L;R
wr;i ¼ f H ywi ; y wi ;qrh;i ; qrv;i ; q rh;i ; q rv;i ; (19.b)

 
_ L;R _ L;R _ L;R
LL;R L;R
wr;i ¼ f L ywi ; y wi ;qrh;i ; qrv;i ; q rh;i ; q rv;i ; (19.c)
284 Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294

Fig. 6. FEM modal of the bridge.

Fig. 7. Lateral mode of bridge (top view, frequency ¼3.465 Hz).

Fig. 8. Vertical mode of bridge (front view, frequency ¼ 5.205 Hz).

Fig. 9. Wheel displacement based derailment criterion (Unit in mm).

Table 4
Earthquake records inputted to the train–bridge system.

ID Number Name Magnitude Recording Station

1 Northridge 6.7 Beverly Hills -14145 Mulhol


2 Northridge 6.7 Canyon Country-W Lost Cany
3 Duzce, Turkey 7.1 Bolu
4 Hector Mine 7.1 Hector
5 Imperial Valley 6.5 Delta
6 Kobe, Japan 6.9 Nishi-Akashi
7 Kobe, Japan 6.9 Shin-Osaka
8 Kocaeli, Turkey 7.5 Duzce
9 Kocaeli, Turkey 7.5 Arcelik
10 Landers 7.3 Yermo Fire Station
11 Loma Prieta 6.9 Capitola
12 Loma Prieta 6.9 Gilroy Array #3
13 Manjil, Iran 7.4 Abbar
14 Superstition Hills 6.5 El Centro Imp. Co. Cent
15 Cape Mendocino 7.0 Rio Dell Overpass – FF
16 Chi-Chi, Taiwan 7.6 CHY101
17 Chi-Chi, Taiwan 7.6 TCU045
18 Friuli, Italy 6.5 Tolmezzo

This full nonlinear wheel–rail relation allows two widely adopted displacement based derailment criteria to be used,
namely, the vertical derailment and lateral derailment criteria [4]. The vertical derailment occurs when the lowest point of
the wheel reaches the highest point of the rail (wheel lifts by hmax as shown in Fig. 9). In this situation, a small perturbation
in the lateral direction may make the wheel fall out of the range of the rail. The lateral derailment occurs as the wheel-set
Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294 285

0.3 No.1

0.0
Lateral
-0.3 Vertical
No.3
0.3
0.0
-0.3

0.3 No.6

0.0
-0.3

0 10 20 30 40

Fig. 10. Acceleration time history of ground motion.

slides laterally a distance umax causing the wheel profile fall out the range of rail profile laterally (see Fig. 9). This lateral
derailment situation can be seen as the “real” derailment, while the vertical derailment criterion is only a necessary but not
sufficient condition for true derailment, since the uplifted wheel may fall back to the rail if lateral movement is not
excessive.
In this study, the geometric parameters for Chinese LM wheel tread and CHN60 rail profile [20] were used. The nominal
roll radius of the wheel r is 450 mm, the inside gauge between the flanges is 1353 mm and the rail gauge is 1435 mm. The
rails have a cant of 1/40. For such a wheel–rail pair, the uplift value hmax is 28 mm for uplift derailment. The lateral motion of
the wheel–rail umax is 94 mm for lateral derailment. The calculation of the derailment displacement hmax and umax is illu-
strated in Fig. 9 with the standard wheel and rail setup shown together.

2.5. Ground motion input to the vehicle–rail–bridge system

A total of 18 recorded 3D ground motion records from various sites were used as input to the vehicle–rail–bridge system.
The use of a ground motion suite helps to draw a more general conclusion, i.e., not specific for one given time history of
ground motion. The ground motion records used here are from various earthquake events around the world [21], with their
magnitudes, names and recording stations listed in Table 4.
Fig. 10 shows the lateral and vertical acceleration of the Nos. 1, 3, and 6 ground motions. In this figure, ground motions
were scaled to the peak acceleration of 0.5 g in the lateral direction. In this study, this ground motion suite will be scaled and
applied to the train–rail–bridge system following an incremental dynamic analysis (IDA, [14]) format. The ground motion
intensity at which derailment occurs will be captured. This will be done for both cases with and without vertical ground
motion component. By comparing the change in derailment intensity with and without vertical ground motion, the impact
of including vertical component on derailment evaluation will be quantified.

3. Solution method to the vehicle–rail–bridge dynamic system

The dynamic interaction of the vehicle–rail–bridge system was solved by an explicit integration algorithm (W M Zhai,
1996) [22] which was originally designed for the simulation of railway vehicle dynamics. Explicit algorithm requires no
iterations between the wheel and rail. For a general structure dynamic system,
MX€ þC X_ þKX ¼ F ext ðX; X_ Þ (20)
where X is the displacement of the system, M, C and K are the mass, damping and stiffness matrix; F ext is the external force.
The Zhai's method firstly predict the current state of the system from the previous states by
X n ¼ X n  1 þ X_ n  1 Δt þ ðX€ n  1 1=2X€ n  2 ÞΔt 2 (21.a)

X_ n ¼ X_ n  1 þ ð3=2X_ n  1 1=2X_ n  2 ÞΔt (21.b)


where the subscript ‘n’ denotes the n-th time step, Δt is the time step size.
Then using the predicted displacement X n and velocity X_ n to calculate the sum of external and internal forces in Eq. (20)
as
 
F n ¼ F ext X n ; X_ n  C X_ n  KX n (22)

The acceleration at time step n þ1 was then calculated by


X€ n ¼ M  1 F n (23)
286 Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294

This explicit integral algorithm was proved to have the same stability limit as that of the central differential method. In
this study a time step size of 0.0001 s was used which was recommended for the dynamic analysis of vehicle and track
system based on numerical experiments [22].
Using this integration method, the solution procedure for the vehicle–rail–bridge system under earthquake was sum-
marized in the following solving steps.

Step 1: Assign initial zero conditions to the vehicle, rail and bridge at time step  2Δt and  Δt.
Step 2: Predict the displacement and velocity at time step n using Eq. (21).
n oEq: 17n o
Y€ V ðn 2Þ; Y_ V ðn 2Þ; Y V ðn  2Þ; Y€ V ðn  1Þ; Y_ V ðn  1Þ; Y V ðn  1Þ ) Y_ V ðnÞ; Y V ðnÞ ;
n oEq: 17n o
L;R
q€ r;i ðn  2Þ; q_ L;R L;R € L;R _ L;R L;R
r;i ðn  2Þ; qr;i ðn 2Þ; qr;i ðn 1Þ; qr;i ðn  1Þ; qr;i ðn  1Þ ) q_ L;R L;R
r;i ðnÞ; qr;i ðnÞ ;


Eq: 17

u€ b ðn 2Þ; u_ b ðn  2Þ; ub ðn  2Þ; u€ b ðn 1Þ; u_ b ðn  1Þ; ub ðn 1Þ ) u_ b ðnÞ; ub ðnÞ :

Step 3: Calculate rail–wheel forces from Eq. (19), using the predicted displacement and velocity of the rail and vehicle.
Then calculate the rail–wheel forces and the rail load on the bridge.
n oEq: 19n oEq: 5 Eq: 4
Y_ V ðnÞ; Y V ðnÞ;q_ L;R L;R
r;i ðnÞ; qr;i ðnÞ ) V L;R L;R L;R r r
wr;i ; H wr;i ; Lwr;i ) F wj ) F V :

Step 4: Renew the locations of all wheel-sets on the bridge xi at time step nþ1; then extract the displacement on the deck
beneath all rail elements. Calculate the forces applied on the rail from the bridge.

xi
Eq: 14n L;R oEq:10 and 13n o
u_ b ðnÞ; ub ðnÞ ) Y b ðxi ; nÞ; Z b ðxi ; nÞ; θb ðxi ; nÞ ) Y_ d;i ðnÞ; Y L;R ðnÞ; Z_ d;i ðnÞ; Z L;R
L;R
d;i d;i
ðnÞ ) V L;R
rb;i
; H L;R
rb;i

Step 5: Calculate the load applied on the bridge by the rail according Eq. (18). Then add this load on the load vector of the
bridge in Eq. (17).
n oEq: 18
Eq: 17
Y_ d;i ðnÞ; Y L;R
L;R
ðnÞ; Z_ d;i ðnÞ; Z L;R
L;R
ðnÞ;q_ L;R L;R
) H br;i ; V br;i ; T br;i ) F N P NF :
d;i d;i r;i ðnÞ; qr;i ðnÞ br ) F
br
br

Step 6: Calculate the seismic forces applied on the vehicle, rail and bridge.

Eq: 2;3
Eq: 11;12
Eq: 15;16
ay ; az ) F gV , ay ; az )  Mrv az ;  M rh ay , ay ; az )  M b Ry ay  M b Rz az :

Step 7: Calculate the acceleration of the system at time step n using Eq. (23).
h i
Y€ V ðnÞ ¼ M V 1 C V Y_ V ðnÞ  K V Y V ðnÞ þ F gV þ F rV ;
h i
L;R
q€ rv;i ðnÞ ¼  C rv q_ L;R L;R L;R L;R
rv;i ðnÞ K rv qrv;i ðnÞ þ V wr;i þ V rb;i =M rv  az ;

u€ b ðnÞ ¼  Ry ay ðt Þ  Rz az ðt Þ þM b 1 ½F br  C b u_ b ðnÞ  K b ub ðnÞ:

Step 8: Go to step 2, repeat the cycles 2–7 for the next time step.

Table 5
Ground motion input in Ref. [23].

Ground motion case ID Amplitude-A (mm) Frequency-f (Hz)

1 320 0.5
2 105 0.8
3 100 1.5
Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294 287

2 00

Forces (kN)
1 00

2 00

Forces (kN)
1 00

0 2 4 6 8 10 12

Fig. 11. Wheel–rail forces for ground input Case 1 (A ¼320 mm, f¼ 0.5 Hz).

2 00
Forces (kN)

1 00

200
Forces (kN)

100

2 4 6 8

Fig. 12. Wheel–rail forces for ground input Case 2 (A ¼ 105 mm, f¼ 0.8 Hz).

4. Response characteristics of the vehicle–rail–bridge due to ground motion

Simulated results from the vehicle–rail–bridge system are demonstrated here in order to illustrate the response
characteristics of the dynamic system under ground motion excitation. These results can also be used to show that the
theory and simulation routines used in this study can produce reasonable response results by comparing these to existing
studies.

4.1. Wheel–rail interaction due to earthquakes

Since the bridge and vehicle systems are linear, the most complicated part in the model lies in the highly nonlinear
wheel–rail contact. In order to check the wheel–rail interaction subjected to earthquake, results from our model are
compared with those from Kazuhiko Nishimura et al. [23]. The lateral ground motion in their simulation is a five-cycle sine
wave. All three cases of ground motions simulated in [23] were repeated here. The amplitude and frequency of ground
motion input were listed in Table 5.
The comparison between wheel–rail forces from our simulations with those in [23] was plotted in Figs. 11–13.
It can be seen from Figs. 11–13 that both the waveform and the amplitude of wheel–rail forces from present simulations
resemble those in [23].

4.2. Responses of the bridge, rails and the vehicle

The lateral acceleration at the middle of the 16th span simulated using our program with a vehicle running on the bridge
(totally 30 spans in the model) excited by lateral ground motion No.1 in Table 4 was plotted in Fig. 14(a). The parameters of
the vehicle, rail and bridge were taken from Tables 1–3. Fig. 14(b) shows the bridge acceleration at the same location
obtained from a commercial software MIDAS [24] without vehicle–bridge interaction. The bridge acceleration using our
simulation agrees well with that from MIDAS, although the vehicle vibration in the MIDAS simulation was not accounted for.
Since the mass of the vehicle is very small compared with that of the bridge, the response of bridge was governed by the
inertial forces of the bridge itself.
In Fig. 15, the lateral displacements of the rail and the wheel-set were also shown. It can be seen that the lateral dis-
placements of the rail are very close to those of the bridge deck. However, wheel-set moves further laterally than the rail
due to the lateral gap between the wheel and rail.
288 Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294

Forces (kN)
2 00

1 00

200

Forces (kN)
100

0 1 2 3 4

Fig. 13. Wheel–rail forces for ground input Case 3 (A ¼ 100 mm, f¼ 1.5 Hz).

Fig. 14. Bridge acceleration due to earthquake, (a) From current simulation including vehicle–bridge interaction, (b) From commercial software without
vehicle–bridge interaction.

Fig. 16 shows the vertical deflection of the bridge deck. The solid line is the deck deflection beneath the moving 1st
wheel-set of the vehicle. Violent vibration of the bridge in the vertical direction occurs near the time of the peak lateral
ground motion. This violent vertical vibration of the bridge was due to the dynamic interaction between the vehicle and
bridge shaken by earthquakes. When the intensity of the ground motion attenuated, the deck deflection was mainly con-
trolled by the moving weight of the vehicle. Thus, outside of the violent interaction region, the bridge deflects approxi-
mately as half sine waves. This can be more clearly demonstrated by the deflection at the midpoint of the 16th span (the
dashed line in Fig. 16).
The wheel and rail vertical displacement on the bridge were plotted in Fig. 17. From this figure, it can be seen that the
vertical location of wheel and rail follows that of the bridge surface outside the region of violent wheel–rail interaction
where the wheel jump occurs frequently.
Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294 289

0.05
Deck under the 1st wheel-set
0.04 Lateral displacemen of the rail
The 1st wheel-set of the vehicle
0.03

Lateral displacement (m)


0.02

0.01

0.00

-0.01

-0.02

-0.03

-0.04

-0.05
5 10 15
Tme (s)

Fig. 15. Lateral displacement of the bridge, rail and wheel-set.

0.0020
Deck deflection under the moving wheel-set
0.0015 Deck deflection at 16th mid-span
Vertical displacement of bridge(m)

0.0010

0.0005

0.0000

-0.0005

-0.0010
Approximative
-0.0015 half sine waves
Violent vehicle bridge interaction
-0.0020
0 5 10 15 20 25
Time (s)

Fig. 16. Vertical displacement of the bridge.

Wheel
Vertical displacement of wheel and rail (m)

0.015 Rail Wheel jump (detail)

0.010

0.005
Wheel displacement due
to bridge deflection
0.000

-0.005
0 5 10 15 20 25
Time (s)

Fig. 17. Vertical displacement of the bridge, rail and wheel-set.

5. Earthquake-induced Derailment simulation of vehicles on bridges

In order to quantify the effect of ground motion intensity on derailment, incremental dynamic analysis procedure was
adopted to find the PGA of the ground motion at which derailment happens. This PGA value is termed herein as the
Derailment Intensity at the Ground (DIG). Similarly, the peak deck acceleration at which the car derails is defined as the
Derailment Intensity at the Deck (DID). The two derailment intensities (DIG and DID) were introduced here to find a better
intensity index for derailment events on bridges. The Derailment Intensity for different ground motion will be different. For
this study, all components of the ground motion were scaled proportionally in the IDA process.
290 Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294

120 120
14 1 2 L 2 4 14 L
4 5 5 6
12 15 7 10
9
Wheel uplift / mm

Wheel uplift / mm
13 17 18
90 6 7 9 90 11 12
13 15 17
11
18 1
8
3 8
60 10 60
3
16
1.2g 0.49g
0.20g 30 Derailment boundary 16
30 [ ] [ ]
Derailment boundary
0.91g

0 0
0.0 0.3 0.6 0.9 1.2 1.5 0.0 0.3 0.6 0.9 1.2 1.5
PGA / g PDA / g

Fig. 18. Wheel uplift v.s. PGA and PDA (with lateral ground motion only) (The numbers in the figures denote the ground motion ID number).

Fig. 19. Time history of the lateral acceleration (a) at the ground; (b) at the bridge deck.

0.010 0.10
3.5 Hz
GM No.3
0.008 GM No.11 0.08
Amplitude (g)
Amplitude (g)

GM No.3
0.006 0.06 GM No.11
3.5 Hz

0.004 3.71E-3 0.04

0.002 0.02
3.56E-4
0.000 0.00
0 2 4 6 8 10 2 4 6 8 10
Frequency (Hz) Frequency (Hz)

Fig. 20. Frequency contents of the lateral acceleration (a) at the ground; (b) at the bridge deck.

5.1. Vehicle derailment analysis without vertical ground motion component

Fig. 18 showed the results from IDA using only lateral ground motion components. The amount of wheel uplift increases
as the peak ground acceleration (PGA) or peak deck acceleration (PDA) increases. It is quite clear from Fig. 18 that derailment
of the vehicle has a better correlation with PDA (the motion of the deck) than with PGA (the ground motion). This is
expected because the vehicle is directly excited by the bridge deck instead of the ground.
During an earthquake, the bridge structure transmits the ground motion as a filter. The degree of ground motion
amplification (or suppression) depends on both the dynamic properties of the bridge and the frequency contents of the
ground motion. Thus the amplification effect of a bridge for different ground motion records may diverge greatly. Fig. 19
(a) shows the lateral acceleration at the ground (for ground motion records Nos. 3 and 10) and Fig. 19(b) shows the lateral
acceleration at the bridge deck. For the same PGA of 0.5 g for ground motion Nos. 3 and 10, the PDAs are 0.57 and 1.97 g
respectively. In other words, the ground motion Nos. 3 and 10 were amplified by the bridge about 1.4 and 3.9 times
respectively. This large discrepancy of the amplification factor may lies in the frequency contents of the ground motion.
Fig. 20(a) and (b) show the frequency contents for ground motion (record Nos. 3 and 10) and that of the deck acceleration
Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294 291

120 120
L+V 14 4 5 L+V
1 2 14 7
4 5 6 10 13
12 15 11 12
13 9
90 6 7 9 17 18 90 15 17
Wheel uplift / mm

Wheel uplift / mm
20
8
8 1
60 11
10 60 2
3
3
16
0.48g 16
0.17g
30 [ Derailment boundary
] 30 Derailment boundary
[ ]
1.20g 0.89g

0 0
0.0 0.3 0.6 0.9 1.2 1.5 0.0 0.3 0.6 0.9 1.2 1.5
PGA / g PDA / g
Fig. 21. Wheel uplift v.s. PGA and PDA (with both lateral and vertical ground motions).

Table 6
Derailment Intensity with and without vertical ground motion.

Ground motion ID 1 2 3 4 5 6 7

Ratio of vertical to lateral PDA 0.933 0.738 0.769 0.493 0.491 1.521 0.435
Ratio of vertical to lateral PGA 0.629 0.668 0.246 0.450 0.388 0.743 0.278
DID/g (L) 0.659 0.83 0.825 0.688 0.727 0.805 0.629
DID/g (L þV) 0.641 0.823 0.767 0.672 0.733 0.701 0.633
DIG/g (L) 0.272 0.369 0.718 0.250 0.363 0.359 0.318
DIG/g (L þV) 0.265 0.366 0.667 0.245 0.366 0.313 0.320

Ground motion ID 8 9 10 11 12 13 14
Ratio of vertical to lateral PDA 0.518 0.718 0.892 2.806 0.449 1.243 0.676
Ratio of vertical to lateral PGA 0.644 0.572 0.897 1.234 0.761 1.084 0.489
DID/g (L) 0.661 0.911 0.797 0.815 0.8 0.87 0.487
DID/g (L þV) 0.665 0.866 0.671 0.707 0.801 0.821 0.475
DIG/g (L) 0.212 0.352 0.203 0.397 0.264 0.291 0.252
DIG/g (L þV) 0.213 0.334 0.171 0.344 0.264 0.275 0.246

Ground motion ID 15 16 17 18 Mean Standard deviation


Ratio of vertical to lateral PDA 0.37 0.526 0.467 0.984 0.835 0.579
Ratio of vertical to lateral PGA 0.356 0.375 0.764 0.846 0.635 0.273
DID/g (L) 0.802 0.903 0.806 0.651 0.759 0.110
DID/g (L þV) 0.806 0.895 0.811 0.617 0.728 0.105
DIG/g (L) 0.255 1.211 0.354 0.268 0.373 0.238
DIG/g (L þV) 0.257 1.200 0.356 0.254 0.359 0.235

Fig. 22. Ir v.s. ratio of vertical to lateral deck acceleration.


292 Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294

1.0

Cumulative distribution function


0.8

0.6

0.4

Fitted curve
0.2
Simulated data

0.0
0 2 4 6 8 10 12 14 16 18
Ir (%)

Fig. 23. CDF of Ir for the 18 ground motion records.

70

60
Wheel uplifts (mm)

2 4 7
50
9 10 11
13 15 17
40

30

20

10

0
0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Ratio of vertical Acc. to lateral Acc. at deck level
Fig. 24. Wheel uplifts v.s. the ratio of vertical Acc. to lateral Acc. at deck level (The numbers in the figures denote the ground motion ID number).

respectively. The frequency components corresponding to the first natural frequency of the bridge in lateral direction
(3.5 Hz) were singled out in Fig. 20. The frequency component at 3.5 Hz of ground motion No. 3 is significantly larger than
that of the ground motion No. 10. Thus the bridge amplifies ground motion No. 3 much greater than ground motion No. 10.

5.2. Effect of vertical ground motion on the derailment of vehicles on bridges

In order to quantify the effect of vertical ground motion on the derailment of vehicles on bridges, a IDA analysis similar to
the one conducted in Section 5.1 was performed to identify the derailment intensity with the vertical ground motion
component included. The recorded ground motion in vertical direction was applied on the system simultaneously with that
in lateral direction. As a simplified treatment, the same scale factor was applied to both the vertical (V) and lateral (L) ground
motion components (i.e. the ratio V/L was kept constant during the IDA). The dependence of wheel uplifts on the PGA (peak
ground acceleration) and PDA (peak deck acceleration) in presence of vertical ground motion is shown in Fig. 21. Similar to
the IDA without vertical component, the wheel uplift amount showed an increasing trend with PGA and PDA. However, the
Derailment Intensities (DIG and DID) leading to derailment for these two cases are different. Table 6 summarizes the
Derailment Intensities in cases without vertical ground motions (denoted by “L”) and with vertical ground motions (denoted
by “LþV”). The mean and standard deviation in the last two columns of Table 6 are statistics of the results from the 18
ground motion records. On average, the presence of the vertical ground motion reduced the derailment intensity in terms of
PGA and PDA. The standard deviation of the DID (derailment intensity at deck level) is smaller than that of the DIG
(derailment intensity at ground level). This shows again that the PDA is a more effective indicator for derailment than PGA.
Based on the simulation results, we define the intensity reduction factor I r by the derailment intensity with vertical
ground motion and that without vertical ground motion as
DIG with both lateral and vertical ground motion
Ir ¼ 1  (24)
DIG with lateral ground motion only
Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294 293

I r quantifies the percentage reduction of the derailment intensity when vertical ground motion is considered (i.e. how
much easier will derailment happen when including vertical component). The value of factor I r can also be negative if
vertical motion “helps” to stabilize the train. Fig. 22 plots the dependence of I r on the ratio of the vertical to lateral PDA for
the 18 ground motion records. The intensity reduction factor shows to increase as the vertical to lateral PDA increases. The
intensity reduction factor is more closely correlated to the ratio of the vertical to lateral PDA (with correlation coefficient
0.75) than to the ratio of the vertical to lateral PGA (with correlation coefficient 0.556).
As indicated in Fig. 22, in eight cases among the 18 ground motion records, the vertical ground motion has nearly no
effect on the derailment threshold (jI r j o 1%), and in ten cases the vertical ground motion makes the vehicle more prone to
derail. In some cases, vertical component actually helps with train running safety (negative I r ). This is possible as the
derailment process is highly nonlinear and dynamic.
Considering the uncertainty of earthquake ground motions, the effect of the vertical ground motion on derailment was
evaluated statistically. Using the 18 ground motions adopted in this study, the cumulative distribution function (CDF) of I r
was obtained (shown in Fig. 23). The 95% confidence level value of I r is calculated to be 0.17, in other words, if the vertical
ground motion was to be ignored, the derailment intensity overestimation will be very likely below 20% (1/0.83–1).
For a given ground motion record, if only the vertical ground motion was scaled while the lateral ground motion remains
constant, the wheel uplifts roughly increase with the intensity of vertical ground motions (see Fig. 24). One can notice the
trend is not always increasing in some cases due to the dynamic nature of the problem. Generally, observations from Fig. 24
indicated that vertical ground motion has some negative impact on train running safety, but the effect is not very significant.

6. Conclusion

The safety related to derailment of railway vehicles running on bridges during earthquakes was investigated through
numerical simulations, with a focus on identifying the impact of vertical ground motion components which are typically
ignored in many existing studies. The motion of the vehicle was represented by multiple body dynamic equations, and the
bridge was modeled using the finite element method. The rail structure was modeled using the moving modes method in
order to decrease the DOFs in the model.
In the numerical simulation, a fully nonlinear wheel–rail contact model was adopted in order to simulate the “jumping”
of the wheel relative to rail during dynamic excitation. Existing relative displacement based derailment criteria was used to
identify derailment event during the simulation. In order to account for earthquake uncertainty, a suite of 18 3D ground
motion records was used in an incremental dynamic analysis for derailment limit state with and without vertical ground
motion component. The simulated results indicate that derailment events were more closely related to the lateral deck
accelerations than to the accelerations of the ground. The peak acceleration at deck level was suggested as a good intensity
indicator for evaluating vehicle derailments on bridges. Based on the simulated threshold deck acceleration for derailment,
it is discovered that vertical ground motion component has a notable negative impact on train running safety on bridges.
The wheel uplifts or the derailment risk generally increases with the ratio of the lateral deck acceleration to the vertical deck
acceleration. The simulation without considering vertical ground motion component can overestimate the threshold
intensity measured in lateral deck acceleration by 20% with 95% confidence level. If this non-conservativeness becomes
critical to train running safety, it is recommended that vertical ground motion component be considered in train running
safety evaluation on bridges during seismic events.
The scope of this study is limited in that only numerical simulation was used. It will be ideal if experimental data can be
generated to further validate the numerical results. Also, this study only considered one type of bridge and vehicle con-
figuration. It will also be beneficial for future studies to include other structure-train combinations commonly encountered
in high-speed train networks in the regions of interest.

Acknowledgment

This study is supported by the National Natural Science Foundation of China (Grant nos. 51408506 and 51308470); Open
Fund of the National Engineering Laboratory for Technology of Geological Disaster Prevention in Land Transportation (Grant
no. SWJTU-GGS-2014004); Key Program of National Natural Science Foundation of China (Grant no. U1434205); and the
National Basic Research Program of China (2013CB036301).

References

[1] Masahiko Ogura, The Niigata Chuetsu earthquake–railway response and reconstruction, Japan Railway Transport Review 43/44 (2006) 46–63.
[2] H. Xia, N. Zhang, W.W. Guo, Application of train–bridge-interaction analysis to bridge design of high-speed railways in China, Proceedings of the 1st
International Workshop on High-speed and Intercity Railways, Springer-verlag, Berlin Heidelberg, 2012, pp. 355–371.
[3] A. Matsuura, Simulation for analyzing direct derailment limit of running vehicle in oscillating tracks, Structural Engineering and Earthquake Engineering
15 (1998) 63–72.
294 Z. Jin et al. / Journal of Sound and Vibration 383 (2016) 277–294

[4] K. Nishimura, Y. Terumichi, T. Morimura, et al., Analytical study on the safety of high speed railway vehicle on excited tracks, Journal of System Design
and Dynamics 4 (2010) 211–225.
[5] M. Miyamoto, N. Matsumoto, M. Sogabe, et al., Railway vehicle dynamic behavior against large-amplitude track vibration-a full-scale experiment and
numerical simulation, Quarterly Report of RTRI 45 (2004) 111–115.
[6] X. Xiao, L. Ling, X. Jin, A study of the derailment mechanism of a high speed train due to an earthquake, Vehicle System Dynamics: International Journal
of Vehicle Mechanics and Mobility 50 (2012) 449–470.
[7] M. Sogabe, M. Ikeda, Y. Yanagisawa, Train-running quality during earthquakes and its improvement for railway long span bridge, Quarterly Report of
RTRI 48 (2007) 183–189.
[8] S. Ju, Nonlinear analysis of high-speed trains moving on bridges during earthquakes, Journal of Nonlinear Dynamics 69 (2012) 173–183.
[9] T. Kawanishi, Y. Murono, T. Miyamoto, et al., The influence of seismic motion with ground irregularity on the dynamic behavior of railway vehicles,
Quarterly Report of RTRI 50 (2009) 89–94.
[10] Y.B. Yang, Y.S. Wu, Dynamic stability of trains moving over bridges shaken by earthquakes, Journal of Sound and Vibration 258 (2002) 65–94.
[11] H. Xia, Y. Han, N.0 Zhang, et al., Dynamic analysis of train–bridge system subjected to non-uniform seismic excitations, Earthquake Engineering
Structural Dynamics 35 (2006) 1563–1579.
[12] J.D. Yau, L. Frýba, Response of suspended beams due to moving loads and vertical seismic ground excitations, Engineering Structures 29 (2007)
3255–3262.
[13] C. Rosario, L.E. José, M.R. Antonio, Stability analysis of multi-body systems with long flexible bodies using the moving modes method and its
application to railroad dynamics, Journal of Computer and Nonlinear Dynamics 9 (2013). 011005-1-10.
[14] Dimitrios Vamvatsikos, C. Allin Cornell, Incremental dynamic analysis, Earthquake Engineering Structural Dynamics 31 (2002) 491–514, http://dx.doi.
org/10.1002/eqe.141.
[15] Jin Zhibin, Pei Shiling, Qiang Shizhong, Study on derailment of railway vehicles on bridges during earthquakes based on IDA analysis, China Civil
Engineering Journal 47 (2014) 234–239. in Chinese.
[16] Y.H. Chen, Y.H. Huang, C.T. Shi, Response of an infinite Timoshenko beam on a viscoelastic foundation to a harmonic moving load, Journal of Sound and
Vibration 241 (2001) 809–824.
[17] W.M. Zhai, Vehicle-Track Coupling Dynamics (Third Edition), Science press, Beijing, 2007, 379–380 (in Chinese).
[18] W.M. Zhai, H. Xia, C.B. Cai, et al., High-speed train–track–bridge dynamic interactions, Part I: theoretical model and numerical simulation, International
Journal of Rail Transportation 1 (2013) 3–24.
[19] Z.Y. Shen, J.K. Hedrick, J.A. Elkins, A comparison of alternative creep force models for rail vehicle dynamic analysis, Vehicle System Dynamics: Inter-
national Journal of Vehicle Mechanics and Mobility 12 (1983) 79–83.
[20] J.J. Ding, F. Li, Y.H.0 Huang, et al., Application of the semi-Hertzian method to the prediction of wheel wear in heavy haul freight car, Wear 314 (2014)
104–110.
[21] C.A. Kircher, C.B. Haselton, G.G. Deierlein, Overview of Ground Motions, ATC-63 Project, COSMOS Annual Meeting, Berkeley, CA, 2006.
[22] W.M. Zhai, Two simple fast integration methods for large-scale dynamic problems in engineering, International Journal for Numerical Methods in
Engineering 39 (1996) 4199–4214.
[23] Kazuhiko Nishimura, Yoshiaki Terumichi, Tsutomu Morimura, et al., Development of vehicle dynamics simulation for safety analyses of rail vehicles on
excited tracks, Journal of Computational and Nonlinear Dynamics 4 (1) . 011001-0111.
[24] MIDAS CIVIL [Computer software]. Midas Information Technology, Seongnam-si, South Korea.

You might also like