dynamicanalysisofcoupledvehiclebridgesystem 2013

Upload: leo2902

Post on 08-Jul-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/19/2019 DynamicAnalysisOfCoupledVehicleBridgeSystem 2013

    1/9

    Dynamic analysis of coupled vehicle–bridge system based on inter-system

    iteration method

    Nan Zhang ⇑, He Xia

    School of Civil Engineering, Beijing Jiaotong University, Beijing 100044, China

    a r t i c l e i n f o

     Article history:

    Received 4 December 2011

    Accepted 10 October 2012

    Available online 7 November 2012

    Keywords:

    Vehicle–bridge interaction system

    Railway bridges

    Numerical history integral

    Iteration method

    a b s t r a c t

    An inter-system iteration method is proposed for dynamic analysis of coupled vehicle–bridge system. In

    this method, the dynamic responses of vehicle subsystem and bridge subsystem are solved separately,

    the iteration within time-step is avoided, the computation memory is saved, the programming difficulty

    is reduced, and it is easy to adopt the commercial structural analysis software for bridge subsystem. The

    calculation efficiency of the method is discussed by case study and an updated iteration strategy is sug-

    gested to improve the convergence characteristics for the proposed method.

     2012 Elsevier Ltd. All rights reserved.

    1. Introduction

    The dynamic effect of the vehicle is an important problem in

    railway bridge design, especially for high-speed railway and hea-

    vy-haul railway bridges. In recent years, the dynamic analysis of 

    vehicle–bridge interaction system has been carried out for lots of 

    cases to ensure the safety of bridge structure and running train

    vehicles and the riding comfort of passengers. For example, the lat-

    eral amplitude of steel plate girders with 20–40 m spans was found

    too large after the raise of train speed during 2000–2003 in China.

    To enhance the lateral stiffness of the girders, Xia et al.  [1] per-

    formed numerical analysis on vehicle–bridge system to over 100

    reinforcement measures and decided the final ones. Through in site

    experiments, the reinforcement measures were validated that they

    can effectively reduce the lateral amplitude as predicted.

    In most of the researches, the vehicle is modeled by the multi-

    body dynamics, while the bridge is modeled by the FEM (finite ele-

    ment method) discretized with the direct stiffness method or the

    modal superposition method. In these analyses, the wheel–railinteraction assumptions are quite different, which they can be di-

    vided into three categories:

    (1)  Moving loads. By neglecting the local vibration and the mass

    effect, the vehicle can be simplified into a series of moving

    loads. The method is widely used in analytical studies and

    the cases with low bridge stiffness. Only the bridge model

    is adopted in the method and the system can be analyzed

    by a time history integral method.

    (2)   Compatible motion relationship. The vehicle and the bridge

    are linked with the wheel–rail relative motion relationship.

    In vertical direction, the wheel-set is commonly assumed

    to have the same motion with the track at the wheel–rail

    contact point. In lateral direction, Xia et al.  [1]   and Xu

    et al. [2]   used the hunting movement to define the wheel–

    rail relative motion, while Guo et al. [3]  took the measured

    bogie hunting movement as the lateral system exciter.

    (3)  Force–motion relationship. The wheel–rail interaction force is

    defined as the function of wheel–rail relative motion. Zhai

    et al.   [4] adopted the Kalker’s linear theory and the Hertz

    contact theory to define the wheel–rail interaction force, in

    which the lateral/tangent wheel–rail force is the product of 

    the creep coefficient and the wheel–rail relative velocity,

    the vertical/normal wheel–rail force has a non-linear rela-

    tionship to wheel–rail relative compression deformation.

    Zhang et al. [5] simplified the Zhai’s definition to meet the

    linear wheel–rail relation both in lateral and vertical direc-

    tions. Torstensson et al.  [6]   and Fayos et al.   [7]   modeledthe rotating wheel-set and derived the wheel–rail interac-

    tion force by kinematics methods.

    Some researches focused on the effect of the parameters in the

    vehicle–bridge interaction system, including the effects of the ratio

    of train/bridge natural frequency, the ratio of train/bridge mass,

    the ratio of train/bridge length [8], the track irregularity, the bridge

    skewness [9], the bridge stiffness and the bridge damping [10].

    The numerical method in solving the vehicle–bridge interaction

    equations is dependent on the wheel–rail interaction assumption.

    Gao and Pan [11], Li et al. [12] and Jo et al. [13] modeled the vehicle

    and the bridge subsystem separately, and solved them with time

    0045-7949/$ - see front matter   2012 Elsevier Ltd. All rights reserved.http://dx.doi.org/10.1016/j.compstruc.2012.10.007

    ⇑ Corresponding author. Tel.: +86 1051683786; fax: +86 1051684393.

    E-mail address:   [email protected] (N. Zhang).

    Computers and Structures 114–115 (2013) 26–34

    Contents lists available at  SciVerse ScienceDirect

    Computers and Structures

    j o u r n a l h o m e p a g e :   w w w . e l s e v i e r . c o m / l o c a t e / c o m p s t r u c

    http://dx.doi.org/10.1016/j.compstruc.2012.10.007mailto:[email protected]://dx.doi.org/10.1016/j.compstruc.2012.10.007http://www.sciencedirect.com/science/journal/00457949http://www.elsevier.com/locate/compstruchttp://www.elsevier.com/locate/compstruchttp://www.sciencedirect.com/science/journal/00457949http://dx.doi.org/10.1016/j.compstruc.2012.10.007mailto:[email protected]://dx.doi.org/10.1016/j.compstruc.2012.10.007

  • 8/19/2019 DynamicAnalysisOfCoupledVehicleBridgeSystem 2013

    2/9

    history integral method TSI (time-step iteration), where the two

    subsystems meet the equivalent equations within each time-step

    by iteration. Xia et al.   [1], Antolin et al.   [14]   and Yang and Yau

    [15] coupled the two subsystems into global equations with vary-

    ing coefficients by adopting the wheel–rail interaction into the

    non-diagonal sub-matrices. Feriani et al.   [16]  and Shi et al.   [17]

    used a complete time history iteration method in which the two

    subsystems was analyzed separately and linked by an interfaceprogram, but their studies only concerned the vertical interaction

    force for highway bridges and trucks.

    The lateral and the torsional interaction forces are not necessary

    for analysis of highway bridges but are very important for railway

    bridges. In this paper, an iteration method for solving the railway

    vehicle–bridge interaction system is proposed, considering the ver-

    tical, lateral and torsional interaction between the bridge and the

    railway vehicle, and adopting the track irregularity and the

    wheel–rail force–motion relationship (inter-system iteration, ISI).

    In the ISI method, firstly, the bridge subsystem is assumed rigid,

    while the vehicle motion and wheel–rail force histories are solved

    by the independent vehicle subsystem for the complete simulation

    time; next the bridge motion can be obtained by applying the pre-

    viously obtained wheel–rail force histories to the independent

    bridge subsystem. Following, the updated bridge deck motion his-

    tories are combined with the track irregularities to form the new

    excitation to the vehicle subsystem for the next iteration process,

    until the given error threshold is satisfied.

    2. The ISI analysis method for vehicle–bridge interaction system

    The vehicle–bridge interaction system is composed by the vehi-

    cle subsystem and the bridge subsystem; the two subsystems are

    linked by the wheel–rail interaction; the given track irregularity

    is taken as an additional system exciter.

    The same coordinate systems are adopted for the both subsys-

    temsand the track irregularity: X denotesthe train runningdirection,

     Z upward, and Y is defined by the right-hand rule. U , V and W denotethe rotational directions about the axes X , Y  and Z , respectively.

    Thecoordinate systems of both vehicle andbridge subsystemare

    absolute, and they have the same coordinate direction and length

    unit. Each rigid body in the vehicle has its independent coordinate

    system, with theorigin in Y and Z directions at the static equilibrium

    position of each rigid body. According to the assumptions in Sec-

    tion 2.1, there is no X-DOF considered in the vehicle subsystem, so

    it is no need to define the origin of coordinates in  X  direction.

     2.1. Vehicle model

    The following assumptions are adopted for the vehicle model

    and the wheel–rail interaction:

    (A1) The train runs over the bridge at a constant speed.

    (A2) The train can be modeled by several independent vehicles by

    neglecting the interaction among them.

    (A3) Each vehicle is composed of one car-body, two bogies, four

    or six wheel-sets and the spring-damper suspensions

    between the components.

    (A4) By the Kalker’s Linear theory, the lateral (Y ) displacement of 

    the wheel-set is the product of the creep coefficient and the

    wheel–rail relative velocity.

    (A5) By the wheel–rail corresponding assumption, the wheel-set

    and the rail have the same vertical ( Z ) and rotational (U ) dis-

    placements at the wheel–rail contact point.

    (A6) Each car-body or bogie has five independent DOFs in direc-

    tions   Y ,   Z ,   U ,   V   and   W ; each wheel-set has 1 independentDOF in direction Y and 2 dependent DOFsin directions Z and U .

    Some measured results indicated that the wheel-set yaw angle

    in high-speed trains is much smaller than that in the traditional

    trains, partly due to the special structure of yaw dampers mounted

    on the high-speed trains, thus the wheel-sets’ DOF in W  direction

    (yaw angle) is not considered in the vehicle model.

    From assumption (A2), the vehicle subsystem can be considered

    as several vehicles separately. Thus the dynamic equations for an

    individual vehicle are:

    MV €X V þ CV  _X V þ KVX V  ¼  PV   ð1Þ

    where MV, CV  and  KV  are the mass, damping and stiffness matrices

    of the vehicle, which are constant matrices [5]; PV   is the force vec-

    tor; X V is the displacement vector, containing the independent DOFs

    of the car-body, the bogies and the wheel-sets. There are 19 inde-

    pendent DOFs and 8 dependent DOFs for a 4-axle vehicle; 21 inde-

    pendent DOFs and 12 dependent DOFs for a 6-axle vehicle. For

    example, the displacement vector  X V  of a 4-axle vehicle is:

    X V ¼ ½ yC; z C;uC; v C;wC; yT1; z T1;uT1; v T1;wT1; yT2; z T2;uT2; v T2;

    wT2; yW1; yW2; yW3; yW4T

    where the subscript C stands for the car-body, T1 and T2 for the

    front and rear bogie, W1 and W2 for the wheel-set linked to thefront bogie, W3 and W4 for the wheel-set linked to the rear bogie,

    respectively.

     2.2. Bridge model

    The bridge model can be established by the FEM. The dynamic

    equations for the bridge subsystem can be written as:

    MB €X B þ CB  _X B þ KBX B  ¼  FB   ð2Þ

    where   MB,   CB   and   KB  are the global mass, damping and stiffness

    matrices,  FB  and  X B  are the force and displacement vectors of the

    bridge subsystem, respectively.

    It is very important to note that the lumped mass method can-

    not be adopted for the mass matrix. Because if the diagonal ele-ments related to the torsional (U ) DOFs in   MB   is zero, the

    torsional moment of the vehicle may cause unreasonable angular

    acceleration for the bridge deck.

    In some cases, the modal superposition method may be used in

    modeling the bridge subsystem to reduce the number of DOFs. The

    equations of the bridge subsystem are expressed as:

    €X B þ 2nBxB _X B þ x

    2BX B  ¼ U

    TBFB   ð3Þ

    where nB and  xB are the damping ratio and circular frequency diag-onal matrices, respectively; UB   is the modal matrix.

    For the same reason, if lumped mass method is adopted, there is

    no torsional mode in UB  and the torsional moment and angle can-

    not be included in calculation. Therefore, the consistent mass ma-

    trix for the bridge subsystem is used to reflect the torsionaldynamic characteristics of the bridge.

     2.3. Track irregularity

    The track irregularity is the distance of the actual position and

    the theoretical position of the rail. According to the definition in

    rail engineering, the track irregularities are defined as:

     yI  ¼ yR þ yL 

    2

     z I  ¼ z R þ z L 

    2

    uI  ¼ z R  z L  g 0

    8><>: ð4Þ

    where yL  and yR  are the lateral irregularities for the left and the right

    rail; z L  and  z R  are the vertical irregularities for the left and the rightrail; g 0  is the rail gauge;  yI,  z I  and  uI  are the align (lateral), vertical

    N. Zhang, H. Xia / Computers and Structures 114–115 (2013) 26–34   27

  • 8/19/2019 DynamicAnalysisOfCoupledVehicleBridgeSystem 2013

    3/9

  • 8/19/2019 DynamicAnalysisOfCoupledVehicleBridgeSystem 2013

    4/9

     2.5. Interaction equations and inter-system iteration

    The dynamic equilibrium equations for the vehicle–bridge

    interaction system can be formed by the equations of the vehicle

    subsystem and the bridge subsystem. When the direct stiffness

    method is adopted for the bridge, the interaction equations are:

    MV1 €

    X V1 þ ðCV1 þ CC1Þ _

    X V1 þ KV1X V1 ¼  FV1MV2

     €X V2 þ ðCV2 þ CC2Þ _X V2 þ KV2X V2 ¼  FV2

    .

    .

    .

    MVn €X Vn þ ðCVn þ CCnÞ  _X Vn þ KVnX Vn ¼  FVn

    MB €X B þ CB  _X B þ KBX B  ¼  FB

    8>>>>>>><>>>>>>>:

    ð14Þ

    where n  is the vehicles number of the train. The first  n  equations in

    Eq. (14) are for the vehicle subsystem, the last equation is for the

    bridge subsystem.

    The mass, damping, stiffness and additional damping matrices

    in the left-hand side of Eq.   (14)   are constants. The force vector

    FVi is the function of  yIm, z Im, uIm, yBm, z Bm  and  uBm in Eq. (12), with

    the subscript Im indicating the track irregularity and Bm the bridge

    motion, respectively. The force vector   FB   is the function of the

    above exciters and the vehicle motion in Eq.   (13), with the sub-

    script Wm  indicating the wheel set motion and Jk  the bogie mo-

    tion, respectively. Thus Eq. (14) are coupled and can be solved by

    an iteration procedure.

    For the iteration strategies TSI and ISI, the iteration procedures

    are compared in Fig. 2.

    The wheel–rail interaction force histories are adopted for the

    convergence check in ISI, because they reflect the dynamic status

    of both the vehicle and the bridge. The operation of ISI consists

    of the following procedures:

    Step 1: Solve the vehicle subsystem by assuming the bridge sub-

    system rigid, setting the bridge motion to zero, and using

    the track irregularities as the excitation, to obtain the time

    histories of wheel–rail forces/moments for all wheel-sets;

    Step 2: Solve the bridge subsystem by applying the wheel–rail

    interaction force histories obtained in the previous itera-

    tion loop (or Step 1) on bridge deck, to obtain the updated

    time histories of bridge deck movement at all joints;Step 3: Solve the vehicle subsystem by combining the updated

    bridge deck movements obtained in Step 2 with the track

    irregularities as the updated system excitation, to obtain

    the updated time histories of wheel–rail forces/moments

    for all wheel-sets;

    Step 4: Calculate the errors between the updated wheel–rail inter-

    action force histories of all the wheel-sets obtained in Step

    3 and those in the previous iteration loop (or Step 1) for

    the convergence check;

    If the maximum instantaneous absolute differences for all

    wheel-sets in the whole integral time satisfy the given threshold,

    the convergence check is OK, meaning the calculation is com-

    pleted; otherwise, return to Step 2 to start a next iteration loop.

    This iteration procedure is completely different to that of TSI. In

    TSI, the vehicle subsystem and the bridge subsystem are solved

    simultaneously through the iteration process in each time-step,

    and the convergence check is upon the dynamic responses at the

    end of each time-step. While in ISI, the two subsystems are solved

    separately over the complete simulation time in each iteration

    loop, and the convergence check is performed afterwards using

    the continuously updated histories of wheel–rail forces/moments

    until the error threshold is satisfied.

    Based on the wheel–rail interaction assumption, the wheel–rail

    interaction force is the function of the wheel–rail relative motion.

    Fig. 2.   Iteration procedures of TSI (left) and ISI (right).

    N. Zhang, H. Xia / Computers and Structures 114–115 (2013) 26–34   29

  • 8/19/2019 DynamicAnalysisOfCoupledVehicleBridgeSystem 2013

    5/9

    The masses, damping and stiffness of both vehicle and bridge arequite large, while the energy inputted to the vehicle–bridge inter-

    action system is limited, which cannot excite intense vibration in

    high frequency, so the high frequency components in the wheel–

    rail force are small. Without importing the numerical dissipation,

    the Newmark-b method is adopted in solving the vehicle and the

    bridge subsystems, with c  = 1/2 and  b  = 1/4.It can be seen in Fig. 2 that the ISI method is simpler in iteration

    procedure. The convergent results can be obtained in each iteration

    step when an unconditionally convergent iteration method is used.

    But ISI is not an unconditionally convergent procedure. The diver-

    gent results may be found even when an unconditional convergent

    iteration method is used in solving the vehicle and the bridge sub-

    systems, which will be found in Section 3. Alsoin Ref. [18], in which

    it is concluded that a convergent result cannot be obtained for thecoupled vehicle–bridge system even by using a smaller time-step

    when the wheel–rail interaction is defined by the wheel–rail rela-tive motion, such as the Assumption (A5) in this paper.

    One of the main advantages of adopting ISI is that the commer-

    cial structural analysis software can be used for the bridge subsys-

    tem, it is equivalent to solve Eqs.  (2) or  (3), making the analysis

    easier and more accurate. While for TSI, it is difficult to invoke

    external programs within the time-step, thus the matrices of 

    bridge must be calculated explicitly.

    The vehicles are coupled through the bridge and must be solved

    simultaneously, which may lead larger memory consuming and

    programming difficulty in TSI. While for ISI, the vehicles run on a

    constant system exciter and can be analyzed separately, with the

    equilibrium equations of Eq. (10). The method of ISI is illustrated

    in Fig. 3.

    The mass, damping and stiffness of each DOF in the vehicle andthe bridge subsystems are quite large, so the high frequency

    Fig. 3.   Illustration of ISI.

     Table 1

    Vehicle parameters.

    Item Value (m) Item Value

    Distance of wheel-sets 2.50 Car-body, X-inertia 100t  m2

    Distance of bogies 17.50 Car-body, Y-inertia 1500t  m2

    Transverse spana of primary suspension 2.00 Car-body Z-inertia 2500t  m2

    Transverse span of secondary suspension 2.00 Primary suspension X-damp/side 0

    Car-body to secondary suspension 0.80 Primary suspension Y-damp/side 0

    Secondary suspension to bogie   0.20 Primary suspension Z-damp/side 20 kN s/m

    Bogie to wheel-set 0.10 Secondary suspension X-damp/side 60 kN s/m

    Wheel radius 0.43 Secondary suspension Y-damp/side 60 kN s/m

    Wheel-set mass 2t    Secondary suspension Z-damp/side 30 kN-s/m

    Wheel-set X-inertia 2t  m2 Primary suspension X-spring/side 5000 kN/m

    Bogie mass 3t    Primary suspension Y-spring/side 5000 kN/m

    Bogie X-inertia 3t  m2 Primary suspension Z-spring/side 800 kN/m

    Bogie Y-inertia 8t  m2 Secondary suspension X-spring/side 200 kN/m

    Bogie Z-inertia 8t  m2 Secondary suspension Y-spring/side 200 kN/m

    Car-body mass 40t    Secondary suspension Z-spring/side 200 kN/ma Transverse span: the transverse distance between the spring/damper in suspension system,  b1  is shown in  Fig. 1.

     Table 2

    Bridge parameters.

    Beam type   f H/Hz   f V/Hz   G1/kN m1 G2/kN m

    1 I X/m4

    High-speed railway beam (A) 15 7 170 80 25

    Speed-raised railway beam (B) 6 5 130 50 0.06

    Common railway beam (C) 3 4 110 40 0.05

    Common railway low-height beam (D) 2.5 3 80 40 0.03

    30   N. Zhang, H. Xia / Computers and Structures 114–115 (2013) 26–34

  • 8/19/2019 DynamicAnalysisOfCoupledVehicleBridgeSystem 2013

    6/9

    component is small in the vehicle–bridge interaction system. The

    convergent result can be obtained in several iteration steps. While

    for some other problems with multi coupling subsystem (other

    than the problem of vehicle–bridge system), it may become diffi-

    cult to get the convergent result by the ISI method owing to the

    high frequency components in vibration. The problem can be

    partly solved by using a smaller time-step or larger threshold in

    convergence check, or adopting the numerical dissipation to re-

    duce the high frequency vibration artificially.

    3. Case study and discussion

     3.1. General information of cases

    For simplicity, an individual vehicle and a bridge with single

    span beam are analyzed in this section. The parameters of the vehi-

    cle and the bridge are listed in Tables 1 and 2, respectively.

    The vehicle parameters are not from a certain type of train. Four

    types of beams are considered for the bridge, which are all pre-

    stressed and single bound, 32 m in span. Beam A is box-sectional,

    while Beam B, C and D are T-sectional. In  Table 2,   f H   and   f V   are

    the lateral and vertical fundamental frequency,  G1   is the beam

    weight per unit length, G2   is the secondary weight (including therail structure and the additional devices) per unit length, and   I  X 

    is the inertia moment of beam section about X -axis, respectively.

    All the four types of beams are straight ones with single-bound

    track laid along the centerline of the bridge. The beam is divided

    into 32 spatial beam elements of 1 m in length, which is restrained

    in  X ,  Y ,  Z  and  U  directions at the fixed support and in  Y ,  Z  and  U 

    directions at the movable support. The motion equations of the

    bridge subsystem are expressed by the direct stiffness method, as

    in Eq. (14). Thus the total DOF number of the bridge subsystem is

    33  ⁄  6 7 = 191. By adopting the Poisson’s ratio 0.2, the torsional

    Fig. 4.  Lateral, vertical and torsional track irregularity samples.

    Fig. 5.  Initial and final positions of the vehicle traveling through the bridge.

     Table 3

    Responses of vehicle–bridge subsystem of ISI and TSI.

    Item Beam A Beam B Beam C Beam D

    Mid span lateral disp./mm 0.010 0.029 0.079 0.125

    Mid span vertical disp./mm 0.432 1.204 2.345 5.104

    Mid span torsional disp./mrad 0.001 0.027 0.072 0.137

    Mid span lateral acc./m s2 0.082 0.056 0.106 0.111

    Mid span vertical acc./m s2 0.050 0.100 0.193 0.357

    Mid span torsional acc./m s2 0.012 0.174 0.487 0.767

    Lateral w/r  force/kN 10.43 10.40 10.27 10.38

    Vertical w/r  force/kN 144.5 144.6 144.4 144.9

    Torsional w/r  moment/kN m 15.61 15.51 15.50 15.20

    Car-body lateral acc./m s2 0.175 0.176 0.172 0.171

    Car-body vertical acc./m s2 0.121 0.119 0.117 0.114

    N. Zhang, H. Xia / Computers and Structures 114–115 (2013) 26–34   31

  • 8/19/2019 DynamicAnalysisOfCoupledVehicleBridgeSystem 2013

    7/9

    fundamental frequencies of Beam A, B, C and D are 20.78 Hz,

    2.54 Hz, 3.97 Hz and 4.40 Hz, respectively.

    The track irregularity data generated from the German Low Dis-

    turb Spectrum is adopted; the power spectrum density is ex-

    pressed in Eq. (15).

    S aðXÞ ¼  AaX

    2c

    ðX2þX2r ÞðX2þX2c Þ

    S vðXÞ ¼  AvX

    2c

    ðX2þX2r ÞðX2þX2c Þ

    S cðXÞ ¼  Avb

    2X2cX

    2

    ðX2þX2r ÞðX2þX2c ÞðX

    2þX2s Þ

    8>>>><>>>>:

    ð15Þ

    where   S a(X),   S v(X) and   S c(X) are align, vertical and cross-lever

    irregularities, respectively, with   S a(X) and   S v(X) i n m2/(rad/m)

    and   S c(X) in 1/(rad/m). The parameters are taken as  Xc = 0.8246

    rad/m,   Xr = 0.0206 rad/m,   Xs = 0.4380 rad/m,   Aa = 2.119 107

    cm2 rad/m,   Av = 4.032 cm2 rad/m,   b = g 0/2 = 0.7465 m.   X   is the

    spatial angular frequency calculated by  X = 2p/Lt,  where   Lt   is the

    wavelength of the track irregularity, ranging from 1 m to 80 m.

    The maximum value for the lateral (Y ), vertical ( Z ) and torsional

    (U ) irregularities adopted in the case study are 7.57 mm, 7.15 mm

    and 4.79 mrad, respectively. The samples of irregularity are shown

    in Fig. 4.

    The complete histories of the train traveling through the bridge

    are analyzed, with the train speed of 120 km/h, the damping ratioof the bridge 0.02, and the time-step 0.005 s. The initial and final

    positions of the train are shown in Fig. 5.

    It must be pointed that the ISI method, since it adopts the direct

    time integration for both the two subsystems, has its inherent fil-

    tering characteristics. Thus the system with high frequency vibra-

    tion may be underestimated when the time-step length or the

    element size is not small enough in the calculation. In the case

    study, the bridge in simplified into 191 DOFs, the maximum

    (191st) frequencies of Beam A, B, C and D are 78.1 kHz, 31.3 kHz,

    20.8 kHz and 15.6 kHz, respectively. The time-step is 0.005 s or

    the calculated sampling frequency is 200 Hz. They are enough to

    meet the accuracy requirement for a railway engineering problem.

    Of course, it is assumed in (A5) that there’s no relative motion be-

    tween the wheel-set and the bridge, which may lead to the local

    Fig. 6.  Lateral displacement history of bridge mid span using ISI (left) and TSI (right).

    Fig. 7.  Vertical displacement history of bridge mid span using ISI (left) and TSI (right).

    Fig. 8.  Torsional displacement histories of bridge mid span using ISI (left) and TSI (right).

    32   N. Zhang, H. Xia / Computers and Structures 114–115 (2013) 26–34

  • 8/19/2019 DynamicAnalysisOfCoupledVehicleBridgeSystem 2013

    8/9

    vertical vibration underestimated. Therefore, if only the macro mo-

    tion status of the vehicle and the bridge are concerned, the pro-

    posed model is acceptable, but if the local motion is also

    concerned, the more accurate wheel–rail interaction assumption

    must be used.

     3.2. Iteration process and result analysis

    The maximum instantaneous absolute difference thresholds are

    10 N for the lateral and vertical wheel–rail force and 10 N m for the

    torsional wheel–rail moment for each wheel-set and at each time-

    step.

    Only in the time period when the wheel-sets or the car-body arecoupled with the bridge, the wheel–rail force and acceleration are

    considered in maximum value statistics. For the methods of TSI

    and ISI, the maximum responses of the bridge mid span and vehicle

    car-body are listed in  Table 3, in which the ISI and TSI have the

    same results. The classic displacement histories when the vehicle

    transverses Beam D are shown in Figs. 6–8.

    It is found from Table 3 that the dynamic responses of bridge

    decrease with the bridge stiffness increasing. The wheel–rail inter-

    action force and the car-body acceleration vary quite little for dif-

    ferent beams, because the bridge motion contributes very little to

    the wheel-set motion, compared to the track irregularity does, as

    shown in Figs. 9–13, which indicate the relative proportion of the

    bridge motion (solid line) and the track irregularity (dotted line)

    at the 1st wheel-set position when the train traverses the Beam D.From the above figures, it is found that the irregularities are

    much larger than the bridge motion in the lateral and torsional dis-

    placements, while the bridge motion has relative larger proportion

    in the vertical displacement, but it is in quite low frequency and

    has small effect on the wheel–rail interaction force or the vehicle

    response. In vertical velocity and acceleration history, the irregu-

    larities are still much larger the bridge motion.

     3.3. Influence of iteration step number 

    The iteration numbers of ISI and TSI are shown in Table 4, where

    only the step numbers when the vehicle and the bridge are coupled

    (from Step 181 to 493) are taken into account. The column ‘‘Itera-

    tion N.’’ refers to the total number of iterations between steps 181and 493 for both the methods. It is obvious that the ISI and TSI have

    similar iteration steps in the four cases. In other words, they have

    similar calculation efficiency.

    Wu [18] proved the wheel–rail displacement compatibility con-

    dition is the main reason of divergence, and the additional mass in

    both sides of the system equations is helpful to get the convergent

    result for the vehicle–bridge interaction system. It implies that the

    bridge mass affects the number of iteration steps to meet the con-

    vergence check. The difference of bridge stiffness causes input

    exciters difference between time-steps for the vehicle subsystem

    and may also lead to different convergent conditions. Thus, further

    analysis is performed for the cases with different bridge distrib-

    uted mass and bridge stiffness. The number of iteration steps for

    different masses and stiffnesses are shown in  Fig. 14, where 10–100% of stiffness and 30–100% of distributed mass of Beam D are

    concerned.

    It is found that the number of iteration steps increases with de-

    crease of distributed mass, from 8 with 100% mass to 23 with 30%

    mass. The relationship between the iteration number and the

    stiffness is not monotonic. When the bridge mass is over 40%,

    the number of iteration steps changes little with the bridge

    stiffness, while when the bridge mass is 30%, the number of itera-

    tion steps increases obviously with the stiffness, from 17 with 10%

    stiffness to 23 with 100% stiffness.

     3.4. Convergence strategy

    For the cases with 10–20% of distributed mass for Beam D, theiteration procedure is divergent, no matter ISI or TSI is used. It is

    Fig. 9.  Lateral displacement histories of bridge motion and track irregularity.

    Fig. 10.  Vertical displacement histories of bridge motion and track irregularity.

    Fig. 11.  Vertical velocity histories of bridge motion and track irregularity.

    Fig. 12.  Vertical acceleration histories of bridge motion and track irregularity.

    Fig. 13.  Torsional displacement histories of bridge motion and track irregularity.

     Table 4

    Number of iteration steps.

    Beam ISI ISI TSI TSI TSI

    Steps Iteration N. Max steps Min steps Iteration N.

    Beam A 4 1252 5 2 1203

    Beam B 4 1252 5 3 1279

    Beam C 5 1565 6 3 1449

    Beam D 8 2504 10 6 2712

    N. Zhang, H. Xia / Computers and Structures 114–115 (2013) 26–34   33

  • 8/19/2019 DynamicAnalysisOfCoupledVehicleBridgeSystem 2013

    9/9

    common to illustrate that the convergent characteristics is decided

    by the ‘‘convergent radius’’: if the evaluated system response is

    within the convergent region, or the error is small enough, the con-

    vergent result must be obtained by iteration, otherwise the itera-

    tion is divergent.When the difference of the evaluated response between two

    time-steps is too large, the convergent region may be missed. In or-

    der to reduce the step length to meet the convergent region; or to

    avoid skipping it, the wheel–rail force acted on the bridge subsys-

    tem in step   N   can be regarded as a linear combination of the

    wheel–rail force calculated from the vehicle subsystem in step N 

    and step N  1:

    F BN  ¼  kF VN þ ð1  kÞF 

    VN1   ð16Þ

    where  F BN  is the wheel–rail force acted on the bridge subsystem in

    step  N , which stands for the interaction force of any wheel-set in

    any direction. F V N  and  F V N 1 are the wheel–rail forces calculated from

    the vehicle subsystem in step   N   and step   N  1, respectively.

    0 <  k 6 1 is the combination factor, and  k  = 1 is adopted in the cal-

    culations. The number of iteration steps for the beams with 100%

    stiffness and 5–30% distributed mass of Beam D are listed in Table 5.

    The relationship between k  and the number of iteration steps is

    quite complex, but it can be seen that the smaller combination fac-

    tor   k   helps to get a convergent result. However, smaller   k   also

    causes smaller updating of the evaluated response value, which

    may need more iteration steps to meet the convergence check.

    4. Conclusions

    In this paper, an inter-system iteration method (ISI) is proposed

    for dynamic analysis of coupled vehicle–bridge system, whose re-

    sult is very close to the widely-used time-step iteration method(TSI). The characteristics of ISI are as follows:

    (1) Comparing to traditional methods, the iteration within time-

    step is avoided in ISI, so it is convenient to use the commer-

    cial structural analysis software for the bridge subsystem

    instead of calculating the bridge matrices directly, each vehi-

    cle can be analyzed separately, the computation memory is

    saved, and the programming difficulty is reduced.

    (2) An updated iteration strategy is proposed for ISI to improve

    the convergent characteristics in solving the vehicle–bridgeinteraction system, in which the wheel–rail force acting on

    the bridge subsystem is regarded as a linear combination

    of the wheel–rail force calculated from the vehicle subsys-

    tem in current and previous iteration steps.

    (3) In the ISI method, more iteration step number is needed to

    meet the convergence check when the bridge has smaller

    distributed mass.

     Acknowledgements

    The research is sponsored by the Major State Basic Research

    Development Program of China (‘‘973’’ Program: 2013CB036203),

    the 111 project (Grant No. B13002), the National Science Founda-

    tion of China (Grant Nos. 51178025 and 50838006) and the Funda-

    mental Research Funds for the Central Universities (Grant No.

    2009JBZ016-4).

    References

    [1] Xia H, De Roeck G, Zhang HR, Zhang N. Dynamic of train–bridge system and itsapplication in steel girder reinforcement. Comput Struct 2001;79(21–22):1851–60.

    [2] Xu YL, Zhang N, Xia H. Vibration of coupled train and cable-stayed bridgesystems in cross winds. Eng Struct 2004;26(10):1389–406.

    [3] Guo XR, Deng ZM, Luo H. Dynamic responses of time-dependent system of Tianxingzhou bridge and train under wind load. In: Proceedings of environmental vibrations: prediction, monitoring, mitigation and evaluation,Beijing, China; 2009. p. 1208–13.

    [4] Zhai WM, Cai CB, Guo SZ. Coupling model of vertical and lateral vehicle–trackinteractions. Veh Syst Dyn 1996;26(1):61–79.

    [5] Zhang N, Xia H, Guo WW, De Roeck G. A vehicle–bridge linear interacted modeland its validation. Int J Struct Stab Dyn 2010;10(2):335–61.

    [6] Torstensson PT, Nielson JCO, Baeza L. Dynamic train–track interaction at highvehicle speeds – modeling of wheelset dynamics and wheel rotation. J SoundVib 2011;330(22):5309–12.

    [7] Fayos J, Baeza L, Denia FD, Tarancon JE. An Eulerian coordinate-based methodfor analyzing the structural vibrations for a solid of revolution rotating aboutits main axis. J Sound Vib 2007;306(3–5):618–35.

    [8] Majka M, Hartnett M. Effects of speed, load and damping on the dynamicresponse of railway bridges and vehicles. Comput Struct 2008;86(6):556–72.

    [9] Majka M, Hartnett M. Dynamic response of bridges to moving trains: a studyon effects of random track irregularities and bridge skewness. Comput Struct2009;87(19–20):1233–52.

    [10] Liu K, De Roeck G, Lombaert G. The effect of dynamic train-bridge interactionon the bridge response during a train passage. J Sound Vib 2009;325(8):240–51.

    [11] Gao MM, Pan JY. Coupling vibration analysis for train–track–bridge system. In:Proceedings of EURODYN 2005, Paris, France; 2005. p. 1069–75.

    [12] Li XZ, Ma WB, Qiang SZ. Coupling vibration analysis of vehicle–bridge systemby iterative solution method. J Vib Shock 2002;21(3):21–5 [in Chinese].

    [13] Jo JS, Jung HJ, Kim H. Finite element analysis of vehicle–bridge interaction byan iterative method. Struct Eng Mech 2008;30(2):165–76.

    [14] Antolin P, Goicolea JM, Astiz MA, Alonso A. A methodology for analyzing lateralcoupled behavior of highspeed railway vehicles andstructures.In: Proceedingsof 9th world congress on computational mechanics and 4th Asian Pacificcongress on computational mechanics, Sydney, Australia; 2010. p. 1–10.

    [15] Yang YB, Yau JD. An iterative interacting method for dynamic analysis of theMaglev train-guideway/foundation–soil system. Eng Struct 2011;33(3):1013–24.

    [16] Feriani A, Mulas MG, Aliprendi C. Time domain iterative procedures forvehicle–bridge dynamic interaction. In: Proceedings of ISMA 2006, Leuven,Belgium; 2006. p. 1179–93.

    [17] Shi Y, Song YF, Wang RB. Analysis on coupled vibration of vehicle and curvedcontinuous beam bridge based on FEM. J Highw Transp Res Dev2010;27(4):95–100 [in Chinese].

    [18] Wu DJ, Li Q, Chen AR. Numerical stability of iteration scheme for solution of vehicle–bridge coupling vibration. Chin Q Mech 2007;28(3):405–11 [in

    Chinese].

    Fig. 14.  Iteration step number with different bridge coefficients.

     Table 5

    Iteration steps versus distributed mass and combination factor.

    Distributed mass   k = 1   k = 0.5   k = 0.2   k = 0.1

    30% 23 14 32 64

    25% 192 15 36 71

    20% Divergence 18 42 82

    15% Divergence 21 48 95

    10% Divergence 33 58 112

    5% Divergence Divergence 118 218

    34   N. Zhang, H. Xia / Computers and Structures 114–115 (2013) 26–34