Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

Journal of Vibration and Control

http://jvc.sagepub.com

Analytical Modeling of Bridge-Road-Vehicle Dynamic Interaction System


Hani H Nassif and Ming Liu
Journal of Vibration and Control 2004; 10; 215
DOI: 10.1177/1077546304033950

The online version of this article can be found at:


http://jvc.sagepub.com/cgi/content/abstract/10/2/215

Published by:

http://www.sagepublications.com

Additional services and information for Journal of Vibration and Control can be found at:

Email Alerts: http://jvc.sagepub.com/cgi/alerts

Subscriptions: http://jvc.sagepub.com/subscriptions

Reprints: http://www.sagepub.com/journalsReprints.nav

Permissions: http://www.sagepub.co.uk/journalsPermissions.nav

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


Analytical Modeling of BridgeRoadVehicle Dynamic
Interaction System

H. H. NASSIF
Department of Civil and Environmental Engineering, Rutgers, The State University of New
Jersey, 131 A-Wing, 98 Brett Road, Piscataway, NJ 08854-8014, USA

M. LIU
Black & Veatch Corp., 11900 East Cornell Avenue, Suite 300, Aurora, CO 80014, USA
(Received 28 October 2002; accepted 5 February 2003)

Abstract: We present a three-dimensional (3D) dynamic model for the bridgeïroadïvehicle interaction sys-
tem. A slab-on-girder bridge is modeled as a grillage system subjected to multiple moving truck loads. Multi-
axle semi-tractorïtrailer is idealized as a 3D vehicle model with a nonlinear tireïsuspension system, having
eleven independent degrees of freedom. Road roughness profiles are generated from the random Gaussian
process as well as limited measurements of actual road profiles. Truck wheel loads are applied at any point
and then transferred to nodes as equivalent nodal forces. The Newmark-q integration method is applied as a
numerical algorithm for solving the bridgeïroadïvehicle dynamic interaction equations. The major parame-
ters affecting the bridge dynamic response (or the dynamic load factor) include road roughness, truck weight,
speed and mechanical properties of the tire-suspension system and bridge stiffness and boundary conditions.
Results from other dynamic models as well as field tests are compared with those from the current 3D model.
The results show that the dynamic load factor is highly dependent on road roughness, vehicle suspension, and
bridge geometry.

Key Words: Bridges, grillage method, dynamic load, road roughness, suspension, truck

NOMENCLATURE
d distance between the mass center and front axle;
D cross-sectional area
idj nodal acceleration vector
dJ arbitrary proportionality factors, 4 ? l ? .4
GG D
dG D @ where GG D is the distance between the front tireïaxle set and the mass
GG N
center of the tractor
GG Q
dG Q @ where GG Q is the distance between the front tireïaxle set and pivot point
GG N

Journal of Vibration and Control,  215ï241, 2004 DOI: 10.1177/1077546304033950


f 2004
? Sage Publications

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


216 H. H. NASSIF and M. LIU

Dpc
apc = where Dpc is the distance between the pivot point and the mass center of
Dpr
the semi-trailer
αf , αm , αr twisting rotations of the front, middle and rear tire–axle sets
αT , αR twisting rotations for the mass centers of the tractor and semi-trailer
α constant roughness coefficient
αi · β i weighting factors
b distance between the mass center and rear axle
β input parameter used for describing the rate at which the suspension force
within a hysteresis loop approaches the outer boundary of the envelope
B width of truck body
c distance between the truck body edges and axles
D spacing between the relevant longitudinal grillage members
Dfm distance between the front and middle tire–axle sets
Dpr distance between the pivot point and rear tire set
δi suspension deflection at the current time step
δ i−1 suspension deflection at the last time step
E modulus of elasticity
ξ longitudinal distance
G modulus of rigidity
γ spatial frequency or wavenumber
I0 polar mass moment of inertia
I flexural bending moment of inertia
IT θ , IRθ mass moments of inertia subjected to the pitching rotations of the tractor
and semi-trailer
ITα, IRα mass moments of inertia subjected to the twisting rotations of the tractor
and semi-trailer
If , Im , Ir mass moments of inertia subjected to the twisting rotations of the front,
middle and rear tire–axle sets
Ip polar moment of inertia of the cross-section
J torsional constant
Jg torsional constant of the girder
Jslαb torsional constant of the equivalent transverse beam
[K] global stiffness matrix
L bridge span length
LB length of truck body
λ wavelength
mT , mR masses of the tractor and semi-trailer
[M] global mass matrix
mf , mm , mr masses for the front, middle and rear tire–axle sets
m̄ mass per unit length of beam element
m mean of random process X (t)
ωn natural frequency associated with any specified mode
{P} equivalent nodal force vector
Pj connection force at the pivot point
P applied wheel load

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 217

! O > *O independent random phases distributed uniformly between 0 to 5


s> t assumed parameters
^T` square matrix containing different powers of natural frequencies
UYY autocorrelation function of the independent random process [ +w,
UOO autocorrelation function of the independent random process Q +w,
UZZ autocorrelation function of the correlated random process \ +w,
VIJ suspension force at the current time step
VIJ4 suspension force at the last time step
VI&/7J suspension force corresponding to the upper and lower boundaries of the
envelope of the measured spring suspension characteristics at the deflection  J
VYY +$, power spectral density function of [ +w,
VOO +$, power spectral density function of Q +w,
VZZ +$, power spectral density function of \ +w,
V spacing between the relevant equivalent transverse beams
VG distance between the right and left suspensions
W period
WNJK elastic spring constant of tire lm
tg tangent of an angle
4 rotation between the local and global coordinate systems
5 > 3 pitching rotations of the mass centers of the tractor and semi-trailer
 bridge skew angle
WIJK > VIJK tire and suspension forces, in which l @ i, p and u denote front, middle
and rear tireïaxle sets, respectively
3 time delay
ZG distance between the right and left tires
Y velocity
iyj nodal velocity vector
{JK road roughness (elevation) at tire lm
}5 > }3 vertical displacements of the mass centers of the tractor and semi-trailer
}eJK bridge deflection at tire lm
}JK vertical displacement at tire lm
}G > }N > }S vertical displacements of the front, middle and rear tireïaxle sets

1. INTRODUCTION

The dynamic load on a bridge structure, as defined in Figure 1, could be a significant


component of the imposed loads; it should be assessed accurately in the design and evaluation
process. Most of the existing bridges built in the United States are designed according to
the Standard American Association of State Highway Transportation Officials (AASHTO,
1994) bridge design specification, where the design load was based on the HS20-44 truck
model (a three-axle, 20-ton truck, adopted in 1944). Due to the increase in intensity of
heavy truck traffic and to changes in truck configurations since the adoption of the AASHTO
truck model, it has become essential to determine the actual dynamic component of live
loads on existing bridges. Moreover, the identification of dynamic load spectra for girder-
type bridges is critical for evaluation, rehabilitation and maintenance of under-capacity and

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


218 H. H. NASSIF and M. LIU

)LJXUH  6WDWLF DQG G\QDPLF UHVSRQVHV RI D EULGJH XQGHU WKH HIIHFW RI D YHKLFOH

deficient structures. Field tests and computer simulations are used to determine the dynamic
behavior of a bridge under moving vehicles. Advancements in computer technology make
computer simulations a viable, effective, and economical method of evaluating bridgeï
vehicle interactions.
In this paper, we present a three-dimensional (3D) dynamic computer model that
describes the behavior of girder bridges under the effect of truck loads. The model considers
the dynamic interaction between the truck, bridge, and road roughness. The interaction
between each component is investigated through a detailed 3D model. The bridge model
is based on the efficient grillage approach. The five-axle semi-tractorïtrailer vehicle model
is a 3D representation of the most common trucks on the highways as observed by Nassif
(1993). The five-axle truck is assumed to be equipped with multileaf suspensions and linear
elastic tires, moving at a constant speed, with a rigid body motion. The bridge response
or simply dynamic load factors (DLFs) are computed from the 3D computer program using
the Newmark- method. Nassif et al. (2002) presented the experimental validation of this
3D model and compared computed and experimental results to the specified value of the
DLF in the new AASHTO-LRFD (1998) specification. The experimental DLFs are based
on field tests carried out by Nassif and Nowak (1995). The results of this paper show that
the variation of the DLF is highly dependent on the road roughness, truck suspension, and
bridge stiffness. Moreover, due to the interactive behavior in the bridge, vehicle, and road
parameters, it is essential to use 3D dynamic models that will accurately capture the actual

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 219

dynamic behavior of the bridgeïroadïvehicle system. In this paper we present a model


that uses the fast converging grillage approach with enhanced equivalent nodal loads that
represents the actual behavior of bridges under normal traffic.

2. THEORETICAL STUDIES

A review of numerous early theoretical studies has revealed that a bridgeïvehicleïroad


surface roughness system had been simplified to be a simple beam or plate under a moving
load or unsprung/sprung mass. A model for a three-axle tractorïtrailer-type truck was
developed at the University of Illinois during a systematic and comprehensive investigation
from 1950 to 1970 (Veletsos and Huang, 1970). The truck was represented by two planar rigid
bodies interconnected at the pivot point (the so-called fifth wheel), taking into account the
effects of interleaf friction in the tire-suspension systems. Honda and Kobori (1980) treated
the bridge as a continuous system of several simple beams. The discrete system can be in
the form of a simple beam, a simple beam with torsion (Gupta and Trail-Nash, 1980), or an
orthotropic plate (Mulcahy, 1983). Recently, the 3D grillage model (Wang et al., 1992) and 3D
finite element model of bridge (Fafard et al., 1997) have been used together with 3D vehicle
models owing to the improvements in computer technology. The road roughness profiles are
usually represented by artificial bumps on the bridge entrance and simple a sinusoidal model
(Gupta and Trail-Nash, 1980), and random processes together with Fourier series (Honda
and Kobori, 1980; Hwang and Nowak, 1991). The vehicle models consist of one degree of
freedom (DOF; Chatterjee et al., 1994), two DOFs (Genin and Chung, 1979), or multiple
DOFs (Veletsos and Huang, 1970; Wang et al., 1992). The types of bridges studied include
single span bridges (Gupta and Trail-Nash, 1980), multispan continuous bridges (Hawk and
Ghali, 1981; Green and Cebon, 1994; Chatterjee et al., 1994), cantilever bridges (Veletsos
and Huang, 1970), and suspension bridges (Chatterjee et al., 1994). Table 1 summarizes the
theoretical work performed by various researchers since 1970.
Recently, more sophisticated bridge and vehicle models using the finite element method
(FEM) have been developed. Humar and Kashif (1993) presented control parameters and
design recommendations for bridgeïvehicle interactions, which were based on an analytical
investigation of a simple beam model traversed by a moving mass system. Kawatani and
Kim (1999) used the FEM and a vehicle model with eight DOFs. Simultaneous differential
equations for a coupling vibration bridge and moving vehicle including roadway roughness
were derived by means of modal analysis. Cheung et al. (1999) presented a method in
which the bridge and the vehicle are considered separately with their interaction dealt with
subsequently. Based on the structural impedance method, the equations of vibration of a slab
bridge subjected to a convoy of moving vehicles were obtained.
All models reviewed in the literature did not acquire actual data for road roughness, truck
suspension characteristics, and were not validated experimentally through field dynamic
testing of bridges. In this paper we present a dynamic model that incorporates all the
components of the vibration phenomena in the bridgeïroadïvehicle system. The bridge
model uses the fast converging grillage model (i.e. an assembly of longitudinal and transverse
beam elements), is flexible and can be used for simple as well as continuous (multispan)
bridges. Actual road roughness from measurement as well as simulated profiles is used to
simulate road conditions. The vehicle model is a 3D representation of the type of trucks
observed on US highways.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


220 H. H. NASSIF and M. LIU

7DEOH  6XPPDU\ RI WKHRUHWLFDO PRGHOV RI EULGJH±URDG±YHKLFOH LQWHUDFWLRQV 7KH


DEEUHYLDWLRQV XVHG DUH DV IROORZV 6 WKUHHVSDQ 6% VLPSOH EHDP 77 WUDFWRU±WUDLOHU
60 VPRRWK 66 VLQJOH VSDQ 6%7 VLPSOH EHDP ZLWK WRUVLRQ 67 VLQJOH WUXFN %0 EXPS
&/ FDQWLOHYHU 23 RUWKRWURSLF SODWH / OLQHDU 6,1 VLQXVRLGDO & FRQWLQXRXV )623 ILQLWH
VWULS RUWKRWURSLF SODWH 5 ULJLG * JULG 36' UDQGRP SURILOH 63 VXVSHQVLRQ 1 QRQOLQHDU
63& VSHFLILHG )(0 ILQLWH HOHPHQW PHWKRG 9' YDOLGDWLRQ
Reference Bridge Skew Bridge Vehicle Axle Vehicle Suspension Tire Road VD
model model
Veletsos and Huang (1970) 3-S CL No SB TT 3 2D Bilinear L SM No
Gupta and Trail-Nash (1980) SS No SBT; OP ST 2 2D Bilinear L BM No
Hawk and Ghali (1981) 3-S C No G TT 3 3D sprung Nonlinear R SIN No
Mulcahy (1983) SS No FSOP TT 3 2D Bilinear L BM No
Hwang and Nowak (1991) SS No SB TT 3 2D Nonlinear L PSD No
Wang et al. (1992) SS No G TT 3 3D Nonlinear N PSD No
Humar and Kashif (1993) SS No SB ST 2 1D, 2D Nonlinear R BM No
Savard (1993) SS No FEM TT 3 3D Nonlinear N PSD No
Green and Cebon (1994) 3-S C No SB; FEM TT 4 1D, 2D Nonlinear N SPC No
Chatterjee et al. (1994) 3-S SP No SBT TT 3 1D, 2D, 3D Bilinear L PSD No
Marchesiello et al. (1999) SS No OP ST 2 3D Nonlinear L PSD No
Kawatani and Kim (1999) SS No FEM ST 3 3D Nonlinear L PSD No
Nassif et al. (2002) Multi- Yes G ST, TT Multi- 3D Nonlinear L Actual Yes
span axle /PSD

3. BRIDGE MODEL

The bridge model is based on the grillage method and represents single as well as multiple
continuous span bridges. The grillage model regards the bridge as an assemblage of one-
dimensional (1D) beams, which are subjected to loads acting in the direction perpendicular
to the plane of the assembly. The bridge consists of a set of steel I-girders and a composite
reinforced-concrete deck slab as shown in Figure 2. Unlike a plane frame, this assemblage
of beams incorporates the beam torsional stiffness. For a slab-on-girder bridge, the girders
span longitudinally between abutments, with a composite deck spanning transversely across
the top of the girders, as shown in Figure 2. Since the slab has only a fraction of the flexural
bending stiffness relative to the girders, the slab flexes in the transverse direction with a much
greater curvature than in the longitudinal direction, that is, the slab behaves much like a large
number of planks spanning transversely. Normally, the longitudinal grillage members are
placed coincident with the centerlines of bridge girders, while the bridge slab is divided into
equivalent transverse beams. The number of these beams should be as large as possible with
spacing not more than 1.5 times that of the longitudinal grillage members (West, 1973; Jaeger
and Bakht, 1982).
The section properties of the longitudinal grillage members are calculated in a manner
similar that of composite T-beams. For re-entrant shapes, such as I and L type, the beam
torsional stiffness is calculated by separating the cross-section into several rectangular
elements, adding the torsional constants of these elements to obtain the total torsional
constant, MH .

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 221

)LJXUH  *ULOODJH EULGJH PRGHO

3.1. Matrix Generation for Grillage System

We derive the local stiffness matrix, ^N- `, for a typical beam element with three DOFs at each
end, considering the combination of flexural and torsional effects. Subsequently, the global
stiffness matrix for the grillage system, ^N`, is obtained from an assemblage of local stiffness
matrices. Similarly, the global mass matrix, ^P`, for the grillage system can be obtained from
an assemblage of local consistent mass matrices, ^P- `. Since the local coordinate system
is not always coincident with the global coordinate system, transformation of coordinates is
essential to assemble local matrices into global matrices.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


222 H. H. NASSIF and M. LIU

)LJXUH  :KHHO ORDG WUDQVIRUPDWLRQ

The damping matrix for a grillage system, ^F`, is considered to be proportional to the
global mass and stiffness matrices (Clough and Penzien, 1993). Damping matrix, ^F`, may
generally be of the form:
[  J
4
^F` @ ^P` J ^P` ^N` = (1)
J

In order to determine the arbitrary proportionality factors, J , the modal damping ratio,
 O for any specified mode, q, must be assigned in advance such that
4
idj @ 5 ^T` i j (2)

The modal damping ratio, i j, can be estimated closely from previous experience, such
as laboratory experiments and field testing of existing structures. Although the experimental
MDR scatter over a wide range, it is often conservative to assign a value corresponding to the
first natural frequencies of the structures, namely 0.01 to 0.02 for steel structures and 0.03 to
0.05 for reinforced-concrete structures. The damping ratio for higher modes is assumed to
increase in proportion to higher natural frequencies. The procedure to determine the damping
matrix for a grillage system from the modal damping ratio i j is presented in Liu (1996).

3.2. Equivalent Nodal Loads

The grillage model deals with nodal forces and displacements. Therefore, wheel loads need
to be transferred into equivalent nodal forces when the positions of the wheel loads do not
coincide with the grillage nodes. Moreover, stresses and displacements at any point on
the bridge are also obtained from the nodal displacements through the same wheel load
transformation. Figures 3(a) and (b) show a representation of a point load (wheel load)

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 223

applied in between the four adjacent nodes and the corresponding equivalent nodal forces,
respectively. Jaeger and Bakht (1982) recommended that wheel loads could be distributed
linearly in the longitudinal direction ignoring any moments, and nonlinearly in the transverse
direction to account for bending moments. Also, in an attempt to oversimplify the loading
transformation, Chen (1994) simply assumed the wheel load distribution to be proportional to
the tributary area without taking into account any moments in both longitudinal and transverse
directions.
Clearly, the wheel load transformation is dependent on the structural behavior of the
bridge deck. The wheel load transmission path is from slab (or equivalent transverse
beams) through girders to the abutments. Therefore, wheel loads are first transferred from
the equivalent transverse beams to the longitudinal grillage members, assuming that each
equivalent transverse beam has two fixed ends acting on the longitudinal grillage members.
Thereafter, the concentrated loads are transferred into the equivalent nodal forces (Liu, 1996).
Based on this assumption the following loads will act on the longitudinal members between
the nodes under consideration:

Sg 5 +5f . V, Sfg 5 Sfg 5


SS4 @ > WS4 @  > PS4 @  tg (3)
V6 V5 V5
Sf5 +5g . V, Sf5 g Sf5 g
SS5 @ > WS5 @ > PS5 @ tg= (4)
V6 V5 V5

Afterwards, the concentrated loads SSJ , WSJ and PSJ (where l @ 4 and 2) are transferred
into the equivalent nodal forces as follows

Sg 5 +5f . V,  6 5
 9Sfg 5  5 
S4 @ 5  6 . 4 .    tg> (5a)
V6 GV 5
Sg 5 +5f . V, G  6 5
 Sfg 5  5 
P4 @    5 .   6  7 . 4 tg> (5b)
V6 V5
Sfg 5
W4 @  +4   , > (5c)
V5
Sg 5 +5f . V,  6 5
 9Sfg 5  5 
S5 @ 5 . 6     tg> (6a)
V6 GV 5
Sg 5 +5f . V, G  6 5
 Sfg 5  5 
P5 @      6  5 tg> (6b)
V6 V5
Sfg 5
W5 @  > (6c)
V5
Sf5 +5g . V,  6  9Sf5 g  5 
S6 @ 5  6 5 . 4 .    tg> (7a)
V 6 GV 5

Sf +5g . V, G  6
5
5
 Sf5 g  5 
P6 @    5 .   5 6  7 . 4 tg> (7b)
V 6 V

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


224 H. H. NASSIF and M. LIU

Sf5 g
W6 @  +4   , > (7c)
V5
Sf5 +5g . V,  6 5
 9Sf5 g  5 
S7 @ 5 . 6     tg 4 > (8a)
V 6 GV 5

5
Sf +5g . V, G 6  5
 Sf5 g  5 
P7 @      6  5 tg 4 > (8b)
V6 V5
Sf5 g
W7 @  > (8c)
V5
where  @ d@G=

3.3. Dynamic Equations of Motion for the Bridge Model

The dynamic equations of a bridge subjected to wheel loads can be easily established from
the global mass ^P`, stiffness ^N` matrices and the damping matrix, ^F`, that is

^P` i 33 j . ^F` i 3 j . ^N` ij @ iSj = (9)

Once the nodal displacements are obtained from equation (9), static as well as dynamic
displacements at any other point on the bridge are derived from the beam bendingï
displacement functions. Moreover, the strains at any point on the bridge are obtained from
the derivatives of the displacement function.

4. VEHICLE MODEL

The vehicle model considered in this study is a five-axle semi-tractor-trailer that is assumed
to be composed of three components: (1) tire, (2) suspension and (3) truck body as shown in
Figures 4(a) and (b). Five rigid masses represent the tractor, semi-trailer, and three tireïaxle
sets (i.e. front, middle and rear axles). The tractor and semi-trailer are each assigned three
DOFs, corresponding to the vertical displacement, pitching rotation about the transverse axis
and twisting rotation about the longitudinal axis. The tire-axle sets are each provided with
two DOFs pertaining to the vertical and twisting rotational movements. The tractor and semi-
trailer are interconnected at the pivot point. Therefore, although the total number of DOFs of
this truck is twelve, the number of independent DOFs is reduced to eleven. Because of the
complexity of tire-suspension mechanical properties, a few assumptions are made to simplify
the vehicle model as follows:
(1) truck bodies and tireïaxle sets are rigid;
(2) mass centers of the tractor and semi-trailer are assumed to be at the same level of the
pivot point;
(3) vehicle components move with the same constant speed in the longitudinal direction;
(4) each tire contacts the bridge at a single point;
(5) only vertical interaction forces between the bridge and vehicle are considered, i.e., the
horizontal friction forces are ignored.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 225

)LJXUH  )UHH ERG\ GLDJUDP RI YHKLFOH PRGHO

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


226 H. H. NASSIF and M. LIU

The truck model is also expanded to address trucks with multiple axles such as three and
eleven axles. However, in this paper we present only the truck model with five axles. The
mechanical properties of tires are assumed to be linear, ignoring the effect of tire damping.
The truck bodies are modeled as masses subjected to rigid body motions. The vertical
displacements, pitching and twisting rotations are accounted for masses and mass moments
of inertia of the truck bodies.

4.1. Dynamic Properties of Vehicle Components

The five-axle semi-tractorïtrailer is assumed to be equipped with multileaf spring


suspensions. In essence, multileaf spring suspensions are nonlinear devices that dissipate
energy during each cycle of oscillation. The suspension frequency of oscillation usually
occurs in a range between 0 and 15 Hz. Fancher (1980) measured suspension characteristics,
and showed that forceïdeflection relationships of suspensions are independent of its
frequencies of oscillation, but are dependent on the amplitudes of suspension motions and
nominal applied loads. Figure 5 shows a typical forceïdeflection diagram for the middle
axle suspension. The equation for the forceïdeflection relationship representing a leaf spring
suspension suggested by Fancher (1980) is as follows:

VIJ @ VI&/7J . +VIJ4  VI&/7J , hmml ml4 ml= (10)

In this equation, VI&/7J and  have different values for the front, middle, and rear
tireïaxle sets, corresponding to their forceïdeflection diagrams. Equation (10) is easy to
implement in time-related integration methods, however, it requires iterations to find the
forces and deflections at the specified time. The mechanical properties of tires, ignoring
the effect of tire damping, are assumed to be linear.
The truck bodies are modeled as masses subjected to rigid body motions. The masses and
mass moments of inertia of the truck bodies account for the vertical displacements, pitching
and twisting rotations of the load model. As shown in Figure 6, the mass moments of inertia
of the truck bodies are obtained by assuming a trapezoidal and uniform weight distributions
in the xx and yy directions, respectively. The truck mass moments of inertia in the xx
direction, LYY , and in the yy direction, LZZ , are expressed in terms of the constant mass
density,  , physical dimensions, and total weight of the truck as follows:

E6 O# + s . t,
LYY @ (11)
57
EO6# s5 . 7st . t5
LZZ @ = (12)
69 s.t

From the geometric properties of the trapezoid shape shown in Figure 6, and calculating
the volume, Y, the relationship between s and t, and the total weight Z, will lead to the
following equations:

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 227

)LJXUH  7\SLFDO IRUFH±GHIOHFWLRQ UHODWLRQVKLS RI PLGGOH VXVSHQVLRQ

ZE5
LYY @ (13)
45
Z  5    
LZZ @ 7 u . u . 4 de . 5u 5 . 5u  4 d5 . e5 = (14)
9

4.2. Dynamic Equations of Motion for Vehicle Model

Considering the equilibrium state of the tractorïtrailer at the pivot point, SK , as shown in
Figures 4(a) and (b), the equations of motion in three different directions (rotation, pitch, and
roll) were derived for the tractor and trailer (Liu, 1996). The equilibrium equations of motion
for the vehicle were then assembled and expressed as a :  : matrix as follows

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


228 H. H. NASSIF and M. LIU

)LJXUH  7UXFN PDVV GLVWULEXWLRQV

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 229

^D`  ^‚}` @ ^E` (15)

where

5 6
‚}G TS
9 ‚}NTS :
9 :
  9 ‚}STS :
‚ 9 :
] @ 9 ‚}G TM :
9 :
9 ‚}NTM :
7 ‚}STM 8
SK
5 6
+4  dG D, GG N +VIG M . VINM ,  dG D GG N +VIG S . VIG S,
9 +4  dG D , GQS +VISS . VISM , :
9 VG :
9 :
9 +VIG S . VINS  VIG M  VINM , :
9 5 :
and ^E` @ 9
9
VG :
:
9 +VISS  VISM , :
9 5 :
9  +VIG S . VINS . VIG M . VINM , :
7  +VISS . VISM , 8
3
5 L5 r L5 r
 3
9 G
 GN  G GN
9 L3r 4 L3r L3r
9
9  dG Q dG Q 
9 GQS 5 GQS 5GQS
9 L5 k
9  3 3
9
9 VG
^D` @ 9
9 3 3 
L3k
9   VG
9 4
9 p5  dG D p5 dG D 3
9
9  5 
9 4  dQD dQD
9 p  dG Q +4  dQD , p3 dG Q +4  dQD , p3
7 3
5 5
4 4 3
6
3 3 3 +dG D  dG Q , GG N
L3r L3r :
3  dQD GQS :
5GQS 5GQS :
L5 k :
3 3 3 :
:
VG :
L3k :
3 3 3 :=
VG :
p5 :
3 3 4 :
 5  :
4  dQD dQD :
p3 3 p3 4 :
5 5 8
4 4 3 3

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


230 H. H. NASSIF and M. LIU

The matrix ^D` depends on truck mass distribution, its body moment of inertia, and
its geometry; matrix ^E` represents
 the suspension forces and the geometry of the truck
configuration; and matrix ‚ ] represents the degrees of freedom and unknown forces VIJK
in the suspension system and wheels. The suspension forces can be calculated from equation
(10) and the tire forces, WIJK , are calculated as follows:

WIJK @ WNJK +}JK . {JK  }eJK , = (16)

5. ROAD ROUGHNESS PROFILES

Road roughness usually refers to an uneven, impaired or bumpy pavement on a bridge. The
American Society of Testing and Materials (ASTM) defines road roughness as ññthe deviations
of a pavement surface from a true planar surface with characteristic dimensions greater than
15.24 mm (0.6 in.)òò in order to distinguish road roughness from pavement texture. The
power spectral density (PSD) will be used to generate road roughness profiles if the actual
road roughness (elevation) profiles are not available. Hwang and Nowak (1991) concluded
that the generation of road roughness profiles might be considered as stationary Gaussian
random processes. The stationary Gaussian random process [ +w, can be considered as a
simple periodic cosine function of time with amplitude , circular frequency $, and phase
angle  :

[ +w, @  frv +$w  , = (17)

It is assumed that the phase angle  O is an independent random variable distributed


uniformly in the range from 0 to 2. Further development of [ +w, would result in the
following expression

[
/
s
[ + , @ 7VYY + O ,  O frv +5 O    O , (18)
O@4

where  @ Yw and
 O.4   O4
 O @ = (19)
5
The number of components Q in equation (18) should be large enough so that exact
values of the spatial frequency (or wavenumber),  , within a frequency window  are
not significant. The spatial frequency (or wavenumber),  , usually covers a range from
<=;7  436 to 3.28 cycle m4 (0.0033 to 1 cycle ft4 ), or wavelengths from 0.31 to 91.46
m cycle4 (1 to 300 ft cycle4 ), with a  interval of <=;7  436 cycle m4 (0.0033 cycle
ft4 ) for low frequencies and 0.33 cycle m4 (0.01 cycle ft4 ) for higher frequencies (Sayers,
1988). A typical single road roughness profile generated from the PSD is shown in Figure 7.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 231

)LJXUH  5DQGRPO\ JHQHUDWHG VLQJOH URDG URXJKQHVV SURILOH

5.1. Generation of Two Correlated Road Roughness Profiles

For the 3D computer model, two correlated road roughness profiles ought to be developed
instead of a single ññbicycle modelòò road profile. In this case, the procedure developed by
Hwang and Nowak (1991) is modified to determine the three spectral densities: PSD for each
profile and cross-spectral density. Furthermore, the road surface is assumed to be isotopic (i.e.
road roughness profiles have the same statistical properties in any directions). Therefore, the
correlated random process \ +w, corresponding to the independent random processes [ +w, and
Q +w, may be generated as follows:

\ +w, @ J [ +w   3 , .  J Q +w, (20)


t
\ +w, @  YZ +$, [ +w, . 4   5YZ +$,Q +w, = (21)

For two correlated random processes [ +w, and \ +w, with the spacing distance e on an
isotopic surface as shown in Figure 8, the coherence function  YZ +$O , is expressed as
U4
mVYZ +$O ,m U + , frv $O  g
 YZ +$O , @ @ U34 (22)
V +$O , 3
U + , frv $O  g
s
where  @  5 . e5 =

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


232 H. H. NASSIF and M. LIU

)LJXUH  5RDG SDUDPHWHUV RQ DQ LVRWRSLF VXUIDFH

The procedure to calculate  YZ +$O , is presented in Liu (1996). Clearly,  YZ +$O , equals
to unity when e @ 3, i.e. [ +w, and \ +w, are the same random processes.  YZ +$O , equals
zero when e @ 4, i.e. [ +w, and \ +w, are totally independent.
Equation (21) can be written in discrete form as follows:

[
/
s
\ +w, @  YZ +$O , 7VYY +$O , $O frv +$O w  ! O ,
O@4

/ t
[ s
. 4   5YZ +$O , 7VYY +$O , $O frv +$O w  *O , = (23)
O@4

6. NEWMARK-q METHOD

The Newmark- method developed by Newmark (1959) is one of the best direct numerical
integration procedures. Based on the linear acceleration method, the forward integration of
velocities and displacements are
 
 3U.U @  3U . +4  ,  33U .   33U.U w (24)

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 233

  
4
 U.U @  U .  3U w .    U .  U.U w 5 =
33 33
(25)
5
It was pointed out (Bathe and Wilson, 1973) that a comprehensive analysis of stability and
numerical integration produces the optimum values of  and  for the Newmark- method,
4 4
namely,  @ and  @ . For convergence, determination of the time increment w is
5 7
4
very important. Normally, w  , where i is the highest natural frequency of interest.
53i
In slab-on-girder bridges, i ranges from 5 to 20 Hz and it was found by Liu (1996) that an
acceptable simulation time increment, w, can be taken as 0.001 s.

7. COMPUTER ALGORITHM

The parameters used in the computer model include bridge parameters, road profiles, and
truck configurations. The computer model includes matrix generation and decomposition;
eigenvalue and eigenvector extraction; and a step-by-step transient response. The
Newmark- method is used to integrate dynamic equations of the bridge under nonlinear
and non-periodic wheel loads. Figure 9 shows a flow chart of the computer algorithm steps
for the solution of the bridgeïvehicleïroad interaction system of equations. First, the data
are entered into the 3D computer program, including the bridge stiffness, mass and damping
matrices, vehicle configuration and parameters, and road roughness profiles. Simulation
starts when the front axle of the truck enters the bridge and it stops after the rear axle exits the
bridge structure. In order to capture the dynamic behavior of the bridge within a reasonable
accuracy, the interval between successive time increments is taken as 0.001 s. Thus, a reliable
dynamic bridge behavior up to 30 Hz is obtained. At each time interval, the wheel position
is located using truck speed and time increments. Wheel loads of the current time increment
are dependent on the conditions of the previous time increment. Furthermore, the initial
conditions of the current time increment are the same as the computed results of the previous
time step. The assumed wheel loads are transferred into the nodal loads in the grillage mesh,
taking into account the applied forces, bending and torsional moments. The grillage nodal
accelerations are obtained by solving the bridge dynamic equations of motion in equation
(9). Using these nodal accelerations, the grillage nodal velocities, as well as displacements
are integrated by the Newmark- method. The bridge displacements at each wheel position
are calculated from the four adjacent nodal displacements, using displacement shape function.
Thereafter, the vehicle dynamic equations of motion in equation (15) are also solved by the
Newmark- method. The new wheel loads are computed from the resulting girderïdeck
displacements at the wheel positions and the road roughness profile. If the difference between
the new and assumed wheel loads is less than the designated tolerance, the program will
continue to the next time increment. Otherwise, the new wheel loads are considered as the
next assumed wheel loads, and the program will start from the beginning of the current step.
Once the iteration converges, the girder deflection at any point can be calculated using the
displacement function. Similarly, the bridge strains and stresses at any point in any girder are
also obtained.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


234 H. H. NASSIF and M. LIU

)LJXUH  )ORZ FKDUW RI FRPSXWHU VLPXODWLRQ SURJUDP

8. PARAMETRIC STUDY

A parametric study of slab-on-girder steel bridges was performed to identify the effects of
truck, road, and bridge parameters on DLF values. Some of these factors include multiple
truck configuration on the bridge, truck weight, speed, tire properties, axle spacing, road
roughness, cross-section properties, length, frequency and boundary conditions. Bridge I
shown in Figure 2 was used as a typical bridge section. The truck properties are selected
by obtaining the mean values of 7,000 five-axle truck data collected by Nassif and Nowak

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 235

)LJXUH  (IIHFW RI URDG SURILOH RQ EULGJH G\QDPLF VWUHVVHV

(1995). Five multiple truck configurations (i.e., two trucks side-by-side, two trucks staggered,
two trucks following each other with front truck being heavy, two trucks following each other
with back heavy, and two trucks in opposite directions) and one single truck configuration
were considered. The DLF value of each configuration was evaluated by changing one
parameter at a time and keeping all other parameters constant.

8.1. Effect of Road Roughness Profile

Road roughness conditions are described by the PSD of the actual road roughness profiles.
Since an infinite set of road roughness profiles can be generated from the same PSD, two
randomly generated road roughness profiles as well as the actual road roughness profiles
have been investigated. Figure 10 illustrates that the peak dynamic stress increased from
6.28 to 9.36 MPa (49% increase) when changing the road roughness profile from ññsmoothòò
to ññpooròò. Moreover, there is an increase in the stress range that gives rise to more severe
cyclic fatigue behavior in the steel girder.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


236 H. H. NASSIF and M. LIU

)LJXUH  (IIHFW RI ERXQGDU\ FRQGLWLRQV RQ EULGJH G\QDPLF VWUHVVHV

8.2. Effect of Boundary Conditions

Many existing bridges have as-built end conditions that are drastically different from the
original plans, due to freezing of bearings, continuity of concrete decks over supports, end
diaphragms, etc. The boundary conditions of the bridge are an important factor that will
affect the dynamic response. Figure 11 shows bridge response due to various end conditions.
Moreover, it was also observed that damping has an effect on dynamic stresses, however, it
is not pronounced as in the case of the road roughness parameter. In general, it is difficult
to estimate damping in bridges due to connections, railing, parapets, and other mechanical
attachments or even the inherent damping properties of steel bridges.

8.3. Effect of Truck Suspension Type

As shown in Figure 12, it is observed that the mechanical properties of suspension of the
vehicle affect the bridge dynamic response significantly. Moreover, it was found that the
tractor weight and pivot position as well as tire stiffness have little effect on the dynamic
stresses.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 237

)LJXUH  (IIHFW RI VXVSHQVLRQ SURSHUWLHV RQ EULGJH G\QDPLF VWUHVVHV

8.4. Effect of Truck Multiple Presence

For various truck configurations it was assumed that heavy trucks travel in the right lane
while light trucks travel in the left lane. Several truck combinations were formed and the
corresponding DLF values were determined. When the design stresses (i.e. DLF x maximum
static stress) were computed, it was observed that the case of two trucks side-by-side governs.
Figure 13 shows the DLF versus the maximum static stress in the most loaded girder of
the bridge. It was observed that as the static stress increases the DLF decreases. Highest
DLF values were obtained in the case when two trucks are following each other. Although
not shown for brevity, the lowest DLF values were obtained for the case when two trucks
are traveling in the opposite directions. For the two trucks following each other, the critical
formation is when the heavy truck is following the light one.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


238 H. H. NASSIF and M. LIU

)LJXUH  6WDWLF VWUHVV YHUVXV '/) IRU PXOWLSOH WUXFNV FRPSDUHG ZLWK D VLQJOH WUXFN HYHQW

8.5. Effect of Truck Speed

As shown in Figure 14, for all the cases (single and multiple presence) at very low speeds
as well as high speeds, DLF values are low. However, between speed values of 60 and 120
kmph, DLF values were significantly. It is suggested that at low speeds (e.g. crawling speed)
the response is closer to the static response and the interaction between vehicle suspension,
road roughness, and bridge dynamics is not yet pronounced. A similar observation is made
regarding very high speeds.

9. SUMMARY

A new 3D model is developed based on the dynamic interaction between bridge, vehicle, and
road roughness. The dynamic models of the truck include body, suspensions, and tires. Road
roughness profiles are obtained by discrete Fourier transform of PSD function and from actual
measurement of road roughness. The bridge model uses the grillage approach in modeling
longitudinal and transverse beam and slab elements. Equations of motion are solved by the

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 239

)LJXUH  6SHHG YHUVXV '/) IRU PXOWLSOH WUXFNV FRPSDUHG ZLWK D VLQJOH WUXFN HYHQW

Newmark- method to find the maximum dynamic stresses and deflections. A parametric
study is performed to find the effects of various parameters in the bridge, vehicle, and road
model. Vehicle parameters include weight, speed, axle distance, and tire stiffness. Bridge
parameters include damping coefficient, and boundary conditions. Road parameters include
roughness profiles and bumps at bridge entrance.
The results of this study lead to the following conclusions.
1. The 3D model represents the actual dynamic behavior of a girder bridge more closely
than a two-dimensional (2D) model. In particular, the effect of the presence of multiple
trucks (i.e. side by side, following, etc.) can be addressed using the 3D model only.
2. The road roughness profile is an important factor in finding the bridge dynamic re-
sponse or DLF. Since infinite road roughness profiles can be generated from the same
PSD, knowledge of the actual road conditions (e.g., poor, good, very good, excellent)
can be incorporated into the current model. Therefore, in the evaluation of existing
bridge actual road conditions should be incorporated into the rating and evaluation and
calculation of DLF procedures.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


240 H. H. NASSIF and M. LIU

3. Truck suspension properties have a significant effect on the dynamic behavior of the
bridge. Moreover, the effect of truck speed on the dynamic response of the bridge can
be significant at 90 kmph. Field measurements collected in this study but not shown
in paper as well as other research in the literature seem to support this conclusion.
However, statistical methods such as Monte Carlo simulation are needed to address the
inherent variabilities in truck type and speed.

Acknowledgments. This study was funded in part by the New Jersey Department of Transportation (NJDOT).
Also, the help of Leo DeFrain, of the Michigan DOT, in collecting the road roughness data from bridge
sites is gratefully acknowledged. The findings expressed in this paper are those of the authors and do
not necessarily reflect the view of NJDOT or MDOT. Their financial support and technical assistance are
gratefully acknowledged.

REFERENCES
AASHTO, 1994, ‘‘Standard specifications for highway bridges,’’ American Association State Highway and Transporta-
tion Officials, Washington, DC.
AASHTO-LRFD, 1998, ‘‘LRFD specifications for highway bridges,’’ American Association State Highway and Trans-
portation Officials, Washington, DC.
Bathe, K. J. and Wilson E. L., 1973, ‘‘Stability and accuracy analysis of direct integration methods,’’ Journal of Earth-
quake Engineering and Structural Dynamics 1, 283–291.
Chatterjee, P. K., Datta, T. K., and Surana, C. S., 1994, ‘‘Vibration of continuous bridges under moving vehicles,’’ Journal
of Sound and Vibration 169, 619–632.
Chen, Y., 1994, ‘‘Modeling q bridges subjected to moving loads,’’ Mathematical Modelling & Scientific Computing, 4,
567–572.
Cheung, Y. K., Au, F. T. K., Zheng D. Y. and Cheng, Y., 1999, ‘‘Vibration of multi-span non-uniform bridges under moving
vehicles and trains by using modified beam vibration functions,’’ Journal of Sound and Vibration 228(3): 611–
628.
Clough, R. W., and Penzien, J., 1993, ‘‘Dynamics of Structures,’’ 2nd edition, McGraw Hill.
Fafard, M. Bennur, M. Savard., 1997, ‘‘General multi-axle vehicle model to study the bridge–vehicle interaction,’’ En-
gineering computations 14(5), 491–508.
Fancher, H., 1980, ‘‘Force–deflection of multileaf vehicle suspensions,’’ University of Michigan, Truck Transportation
Research Institute UMTRI, Ann Arbor, MI.
Genin, J. and Chung, Y. I. 1979, ‘‘Response of a continuous guideway on equally spaced supports traversed by a moving
vehicle,’’ Journal of Sound and Vibration 67(2), 245–251.
Green, M. F. and Cebon, D., 1994.‘‘Dynamic response of highway bridges to heavy vehicle loads: Theory and experi-
mental validation,’’ Journal of Sound and Vibration 170, 51–78.
Gupta, R. K. and Trail-Nash, R.W., 1980. ‘‘Bridge dynamic loading due to road surface irregularities and braking of
vehicle,’’ Earthquake Engineering and Structural Dynamics 8, 83–96.
Hawk, H. and Ghali, A., 1981, ‘‘Dynamic response of bridges to multiple truck loading,’’ Canadian Journal of Civil
Engineering 8, 392–401.
Honda, H. and Kobori, T., 1986, ‘‘Numerical data base on roadway roughness of highway bridges and roughness char-
acteristics,’’ in Proceedings of the Japan Society of Civil Engineers, 398 I-10, pp. 385–394.
Humar, J. L. and Kashif, A. M., 1993, ‘‘Dynamic response of bridges under traveling loads,’’ Canadian Journal of Civil
Engineering 20, 287–298.
Hwang, E. S. and Nowak, A. S., 1991, ‘‘Simulation of dynamic load for bridges,’’ ASCE Journal of Structural Engi-
neering 117(5), 1413–1434.
Jaeger, L. G. and Bakht, B., 1982, ‘‘The grillage analogy in bridge analysis,’’ Canadian Journal of Civil Engineering 9,
224–235.
Kawatani, M. and Kim, C. W., 1999, ‘‘3D simulation for dynamic wheel loads of heavy vehicles,’’ in Proceedings of the
5th Korea–Japan Joint Seminar on Steel Bridges, October 14–15, Pusan, Korea, pp. 593–606.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010


ANALYTICAL MODELING OF DYNAMIC INTERACTION SYSTEM 241

Liu, M., 1996, ‘‘A three-dimensional dynamic model for bridge–road–vehicle system,’’ Master’s Thesis, Bradley Uni-
versity, Peoria, IL.
Marchesiello, S., Fasana, A., Garibaldi, L., and Piombo, B. A. D., 1999, ‘‘Dynamics of multispan continuous straight
bridges subject to multi-degrees-of-freedom moving vehicle excitation,’’ Journal of Sound and Vibration 224(3),
541–561.
Mulcahy, N. L., 1983, ‘‘Bridge response with tractor–trailer vehicle loading,’’ Earthquake Engineering and Structural
Dynamics 11, 649–665.
Nassif, H. H., 1993, ‘‘Live load spectra for girder bridges,’’ Ph.D. Dissertation, The University of Michigan, Department
of Civil and Environmental Engineering, p. 250.
Nassif, H. H. and Nowak, A. S., 1995, ‘‘Dynamic load spectra for girder bridges,’’ Journal of Transportation Research
Board, 1476, TRB, National Research Council, Washington, DC, pp. 69–83.
Nassif, H., Liu, M., and Ertekin, O., 2003, ‘‘Model validation for bridge–road–vehicle dynamic interaction system,’’
American Society of Civil Engineering (ASCE) Journal of Bridge Engineering, March/April, 281–294.
Newmark, N. M., 1959, ‘‘A method of computation for structural dynamics,’’ ASCE Journal of Engineering Mechanics
85, 67–94.
Sayers M. W., 1988, ‘‘Dynamic terrain inputs to predict structural integrity of ground vehicles,’’ University of Michigan
Transportation Research Institute, Report No. UMTRI-88–16.
Veletsos, A. S. and Huang, T., 1970, ‘‘Analysis of dynamic response of highway bridges,’’ ASCE Journal of Engineering
Mechanics 96, 593–620.
Wang, T. L., Huang. D., and Shahawy, M., 1992, ‘‘Dynamic response of multigirder bridges,’’ Journal of Structural
Engineering, ASCE, 118(8), 2222–2239.
West, R., 1973, ‘‘Recommendations on the use of grillage analysis for slabs and pseudo-slab bridge decks,’’ Report
46.017, Cement and Concrete and CIRIA, London.

Downloaded from http://jvc.sagepub.com at RUTGERS UNIV on July 5, 2010

You might also like