Вы находитесь на странице: 1из 17

170 MODELING AND ANALYSIS

long as the boundary flexibilities are known. Also, rotational member end
stiffnesses are readily available in any textbook on structural analysis, in cases
where rotational DOFs are considered in the analysis.
For squat bridge piers, where the clear column height is no longer
significantly larger then the column depth D, shear deformations can become

significant in comparison with the flexural deformations. The shear deforma-

tion for a unit load, or the shear flexibility, in Fig. 4.7(c) can be expressed as

fv H e (4.21)
AG

where A ve represents the effective shear area (to be discussed later) and G the
shear modulus of the pier cross section. As a general rule of thumb, shear
deformations can become significant when the shear span MN of the pier is
less than three times the pier depth D, or

3D
V (4.22)

where M and V represent the maximum moment M and corresponding shear


force V in the bridge pier, respectively. The shear flexibility f, can be combined
with the flexural or bending flexibility

= 1 = H e
fb
kb aEIe (4.23)

to form the combined stiffness

1 1
k
3
fb H elaEIe + HelAveG
(4.24)
Since bridge columns are, as outlined in Chapter 1, expected to respond
under the design earthquake inelastically, effective member properties He, le,
and A ye , which reflect the extent of concrete cracking and reinforcement
yielding, should be used in the modeling and analytical member characteriza -
tion to obtain realistic seismic response quantification.
From Eq. (4.17) it can be seen that rather than gross-section-based moments of
inertia I g and shear areas A, effective properties I e and Ave are specified, and for
the height H in the stiffness formulation an effective height H e is used.
Constitutive parameters E and G are assumed to be constant for most types of
bridge analyses and can be determined for concrete piers from the nominal
concrete compression strength f: based on standard ACI or CEB procedures.
The modulus of elasticity E can be determined as outlined in
FUNDAMENTALS OF SEISMIC BRIDGE BEHAVIOR 171

Section 5.1, Eqs. (5.1) and (5.2), and the shear modulus G with a Poisson ratio

v for concrete between 0.15 and 0.2 for homogeneous material assumptions as
G E
2(1 + v) (4.25)

While, strictly speaking, even E and G vary depending on loading or unloading


and on the orthogonal strain states, these variations are not accounted for in
most bridge analyses except in very detailed nonlinear finite-element analyses,
discussed in Section 4.5.
The effective column height He is different from the clear column height
He for longitudinal or double bending bridge response, or from the cantilever
height to the mass centroid for transverse response, since based on the design
philosophy developed in Chapter 1, plastic column hinges are expected to
form at one or both ends of the column member. With increasing levels
of ductility in these members and hinges, yield penetration of the column
reinforcement into the adjacent footing or cap beam occurs, which provides
additional flexibility to these yield penetration regions. This added flexibility
can conveniently be expressed by an increase of effective column height to
He for SDOF formulations or with additional nodes and members for MDOF
models. For plastic hinges directly adjacent to a concrete cap or footing, the
amount of yield penetration into the joint can be estimated as outlined in
Chapter 5 from Eq. (5.39) as
_
L 0.15M b r (f y in ksi) ( (4.26)
P' 0.022fydbi fy in MPa)

This yield penetration length 41 can be added to the clear column height
as outlined in Fig. 4.8 for longitudinal and transverse response, respectively,
to form the effective column height 1-1,. In addition, footing springs, as shown
in Fig. 4.8, modeling the effects of soil deformations (discussed in Section
4.4.2) should be employed to obtain the correct stiffness characteristics for
bridge bent models.
To reflect the cracked state of a concrete bridge column in the seismic
response analysis, an effective or cracked-section moment of inertia 4 should
be employed. The effective stiffness El, does not reflect only the effect of
cracking but also the state of the bridge column determined at first theoretical
yield of the reinforcement and can be determined from sectional moment
curvature analyses as
M
El, = (4.27)
cloyi

where Myi and (13 yi represent the ideal yield moment and curvature for a

bilinear momentcurvature approximation as discussed further in Section


172 MODELING AND ANALYSIS

a) Prototype b) Model
Mass Center
FIG. 4.8 Effective column height.
1
1_1
2
t
5.3.2. The effective stiffness I e depends on the axial load ratio Paxiail(A gfc) and
3
the longitudinal reinforcement ratio A st/Ag, where A, and A represent the
gross concrete area and the total longitudinal reinforcement areas, respec -
tively. These effective stiffnesses are represented in graphical form in Fig. 4.9
for typical circular and square column cross sections and show that for typical
column reinforcement ratios between 1 and 3% and axial load ratios between
10 and 30%, a reduction in effective section moment of inertia to between 35
ofnetration
Yield Pe T
and 60% the gross section moment of inertia HIg is not L uncommon.
He A similar
He
reduction in effective stiffness applies to other concrete
e bridge members, such
as cap beams and superstructure girders, and appropriate values for I e can
also be determined from Eq. (4.27) or Fig. 4.9 for corresponding axial load
levels and reinforcement ratios.
Finally, an effective shear stiffness GA rather than the shear stiffness
based on the shear area A, should
4 be employed to reflect the increased shear
deformations in flexurally cracked concrete
/ / / / /members.
/ // Again, a dependency
similar to the effective flexural stiffness EI, on the axial load and the
reinforcement ratio can be expected prior to significant shear distress, which
0.20 ______

FUNDAMENTALS
OF SEISMIC BRIDGE BEHAVIOR 173

11
4;
0.70
cS.
17-1, Ast/Ag = .04
0.60=
rn
C13
0.50
Ast/Ag .03
0.40 i
ra
C.) = .02
0.30-
Ast/Ag = .01

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35


AXIAL LOAD RATIO 1 3 /f A
g
a) Circular Sections

40,
0.80-
cp
0. 70
PA 0.50
A 0.60 AstiAg = .03

Ast/Ag = .02
Ast /A 5 = .
01

Fa 0.40
L)
g0.30

0.20.......................I 1 1 1111 l .
1 " 1 ____________________________________ll________I

0 . 0 0 0 . 0 5 0 . 1 0 0 .....15 0.20 0.25 0.30 0.35


AXIAL LOAD RATIO P/fe'Ag
b) Rectangular Sections

FIG. 4.9 Effective stiffness of cracked reinforced concrete sections [N4].

should not occur in capacity design-protected bridge members. Due


to lack of specific research data, it can be assumed that the effective
stiffness
reduction in shear can be considered proportional to the effective stiffness
reduction in flexure:
El
G A = G A , o r Av e = A (4.28)
E I g

until more experimental and analytical research data become available.

(c) Damping. The third term in the general equation of motion is a viscous
or velocity proportional damping force fd(t) = cas(t), which is used primarily
174 MODELING AND ANALYSIS

for mathematical or numerical convenience and stability rather than for the
phenomenological modeling of actual bridge damping characteristics. Viscous
damping is physically correct only for an oil-filled dashpot and is difficult to
rationalize for other forms of damping actually encountered in bridge or other
structural systems. More common damping types in bridges are (1) Coulomb
damping, (2) radiation damping, and (3) hysteretic damping. Coulomb or
friction damping occurs primarily in bridge superstructure bearings and move-
ment joints and is independent of velocity or displacement. To a lesser degree,
friction damping can occur in cracks of reinforced concrete structures. Radia-
tion damping in bridges occurs due to soil structure interaction (SSI) and
energy dissipated by waves radiating out into the half-space of soil surrounding
the bridge footings. The most common and physically most obvious form of
damping or energy dissipation in bridge structures is in the form of hysteresis
of the forcedeformation response.
To conform to the simple mathematical form of the equation of motion,
other forms of damping, in particular, hysteretic damping, encountered in
bridge systems are conveniently expressed in the form of an equivalent viscous
damping coefficient ceq. The equivalent viscous damping coefficient ceq is
commonly expressed by the equivalent damping ratio t eq and the critical
damping coefficient c,, which is the smallest amount of damping for which
no oscillation occurs in free dynamic response [CO]:

C e q = teqCcr (4.29)

The hysteretic damping or energy loss per cycle, represented by the area
Ah in Fig. 4.10 for one complete idealized loaddisplacement hysteresis loop,
can then be converted for the same displacement amplitude to an equivalent
viscous damping ratio:

A h A h
eq = (4.30)
27TV,A, 477 -A,

where V, and Am represent the average peak force and displacement values
[C10]. The area A, represents the elastic strain energy stored in an equivalent
linear elastic system under static conditions with effective stiffness

V
keff = --'n (4.31)
Am

The equivalent viscous damping coefficient can then be obtained from Eq.
(4.29). To assess the magnitude of equivalent viscous damping from hysteretic
damping, typical forcedisplacement hysteresis loop shapes for various bridge
members are depicted in Fig. 4.11.
FUNDAMENTALS OF SEISMIC BRIDGE BEHAVIOR 175

FORCE
k eff

vnia.

Oliarj
Vm=1/2( Ivi,1+

1
A = Elastic Strain
am=1/2( Irma m i n e

A llEMN M E n e rg y

' a l min
A A 1114:112441. DISPLACEMENT
6
` max
= E n e rg y D i s s i p a t i o n
Per Cycle (Shaded Area)

min

FIG. 4.10 Hysteretic energy dissipation and effective


stiffness for cyclic response.

It is obvious from Fig. 4.10 that the maximum


equivalent viscous damping ratio which can be obtained with Eq. (4.30) is EN
= 2/r = 0.64 for a system that cycles with rigidperfectly plastic force
deformation characteristics. This rigidperfectly plastic loop shape is not very
realistic for typical local inelastic mechanisms in bridges that are depicted in
Fig. 4.11 for one representative cycle. Even frequently used elasticplastic
response idealizations such as those shown in Fig. 4.11(a) apply in very few
cases. Only friction slider bearings as shown in Fig. 4.11(c) can approach this
value. Beam hinges with no or low axial load levels can also exhibit
significant hysteretic energy absorption, as shown schematically by the large
loops in Fig. 4.11(b), and can result in equivalent viscous damping ratios of
30% or higher. High axial loads on a bridge member such as columns or
prestressed cap beams result in pinched hysteretic loop shapes, as shown in Fig.
4.11(d resulting in reduced equivalent viscous damping between 10 and 25%.
),

The rocking response of a bridge pier as depicted in Fig. 4.11(e) is essentially


nonlinear elastic without noticeable hysteresis and thus very little equivalent
hysteretic damping. However, in the case of foundation rocking, additional
energy is dissipated in the form of radiation damping in the surrounding soil,
as outlined in Section 6.4.
For bridge members with nonsymmetric response characteristics, as shown
in Fig. 4.11(f), the displacement-dependent equivalent viscous damping coef-
ficient approach no longer applies strictly since displacements in the two
directions are of different magnitude, but an averaged procedure, as suggested
176 MODELING AND ANALYSIS

(a) Idealized Elastic/ (b) Beam Hinge (c) Friction Slider


Plastic Response

(d) Column Hinge with (e) Rocking of Pier (f) Knee Joint
High Axial Load
FIG. 4.11 Typical hysteretic response in bridge components.

in Fig. 4.10, can still be employed. The same difficulty exists with any dynamic
analysis where equivalent viscous damping is employed since the damping
values are derived for the maximum amplitudes and are subsequently applied
in the analytical model also to all smaller-amplitude cycles.
From the discussions above it is obvious that determination of a correct
equivalent viscous damping coefficient for analytical bridge models is difficult
at best. Thus empirical damping values are frequently employed to reflect the
sum of all possible damping contributions, as well as the fact that most of the
cyclic dynamic response in an earthquake is expected to occur at smaller -
than-expected maximum displacement levels.
Thus, for steel structures, damping values between 2 and 5% of critical
damping are commonly assumed, while for concrete structures a range from
2 to 7% is used to reflect the most representative dynamic response range.
In light of these uncertainties, the commonly assumed 5% viscous damping
coefficient in structural dynamic analysis can hardly be argued with. Only in
FUNDAMENTALS OF SEISMIC BRIDGE BEHAVIOR 177

cases where (1) soilstructure interaction plays an important role, (2) special
energy absorption devices are employed, and (3) high hysteretic energy dissi -
pation is relied upon should higher damping coefficients be employed. It
should also be noted that most analytical models will be based on initial elastic
stiffness, as discussed in the preceding section, and the damping adopted
should represent the elastic phase of response. Where inelastic time-history
analysis is used, the hysteretic energy dissipation will be directly modelled by
the force-displacement hysteresis rules adopted in the analysis. Only with the
substitute structure analysis procedure, discussed further in Section 4.5.2(b),
where the effective stiffness represents that at maximum displacement, rather
than at yield, should the effective damping be increased to include the effects
of hysteretic damping.

4.3.2 Bridge Dynamic Response Characteristics


Independent of the specific dynamic input, each bridge system is represented
within the elastic range by dynamic response modes typically referred to as
the natural modes of vibration, characterized by independent mode shapes (D i
with corresponding periods of vibration T,. The number of characteristic mode
shapes and vibration periods of a bridge model depend on the selected number
of dynamic degrees of freedom defined during the analytical model discreti -
zation.
While the prototype bridge features an infinite number of vibration modes,
bridge analysis models feature a selected finite number of DOFs and associated
modes of vibration. However, the governing dynamic response of a bridge
can typically be captured by the contribution of a limited number of vibration
modes. The fundamental or lowest mode of vibration can often provide a
good indication of the dynamic response of a bridge, making single-degree of-
freedom models, which approximate the fundamental dynamic response of
the prototype bridge, invaluable design tools.

(a) Single-Degree-of-Freedom Characteristics. The fundamental or first


mode of vibration characteristics can be found for simple systems from a
single-degree-of-freedom (SDOF) model like the one shown in Fig. 4.3 once
the lumped mass and stiffness characteristics are known. As long as damping
is significantly less than critical damping, which is the case for essentially
elastic response, damping has very little influence on the dynamic response
characteristics and is typically ignored.
For a single-degree-of-freedom bridge model with lumped mass m and
effective stiffness k, the undamped free vibration can be expressed from Eq.
(4.3) as

mu(t) + ku(t) = 0 (4.32)


178 MODELING AND ANALYSIS

Assuming that the displacement u(t) with time follows a harmonic motion,
as shown in Fig. 4.12, of the form

u(t) = A sin(cot a) (4.33)

where o.) is the circular natural frequency, a a phase shift for the sinusoidal
response, and A a scaling factor that determines the amplitude of the harmonic
motion, Eq. (4.33) and its second time derivative can be substituted into Eq.
(4.32), resulting in the characteristic equation

ku co2mu = (k co2m)u = 0 (4.34)

For arbitrary displacements u, Eq. (4.34) can be satisfied when

lk w2m1 = 0 (4.35)

which occurs only for a specific circular frequency w or the eigenvalue of Eq.
(4.35). The eigenvalue solution of this scalar equation represents the circular
frequency w at which Eq. (4.35) is satisfied as

CO = (4.36)

and from the undamped natural circular frequency co in Eq. (4.36) the cyclic
natural frequency f and natural period of vibration T for a SDOF bridge
model can be found as

(4.37)
u(t)

(4.38)

FIG. 4.12 Free undamped harmonic response of SDOF system.


FUNDAMENTALS OF SEISMIC BRIDGE BEHAVIOR 179

Combining Eqs. (4.36) and (4.38), the fundamental or natural undamped


period of vibration can be expressed as

T = 27r (4.39)

The corresponding natural mode shape consists of a displacement at the


designated SDOF (i.e., the transverse motion of the lumped mass in Fig. 4.3),
of arbitrary sign and magnitude.
In bridge systems where high damping expressed in the form of an equiva-
lent viscous damping ratio eq is present, the damped circular frequency w d can
be expressed as

wd = (4.40)

and the corresponding damped natural period of vibration of a SDOF sys-


tem as
Td = 27r m (4.41)
k(1 eq)

From Eqs. (4.40) and (4.41) the small influence of the damping ratio on the
dynamic response characteristics is evident. For example, an equivalent viscous
damping ratio of 10% only increases the natural period of vibration by 0.5%,
and even a 50% damping ratio only results in a 15% natural period elongation.
Since typically maximum expected equivalent viscous damping ratios in bridges
are less than 15%, the influence of damping on the dynamic response
characteristics can be neglected.
As discussed above, m and k in Eq. (4.39) represent the effective seismic
mass WsIg and the effective stiffness keff with reference to the single displacement
degree of freedom at the mass centroid. This single-degree-of-freedom concept
to estimate the fundamental dynamic bridge response characteristics can also
be applied to structures with distributed parameters in the form of a
generalized single-degree-of-freedom system as outlined by Eq. (4.4).
Expressions for a generalized single-degree-of-freedom response of a dis-
tributed parameter systems can be found in [C10,N1], and only the special
case of lumped stiffness and mass characteristics along the length x of a bridge
structure are discussed below since discrete springs (modeling the stiffness of
individual bents) and discrete masses (modeling lumped superstructure inertia)
can readily be identified for typical bridge systems. In the case of discrete
translational and rotational masses m1 and j1, the principle of virtual work can
be applied to a generalized single-degree-of-freedom system in the form of
Eq. (4.4) such that a new generalized mass m* can be derived as
180 MODELING AND ANALYSIS

m* = it14xi) i;45(1) (4.42)

and a new generalized stiffness k* as

k* = k Ag a) (4.43)

The analytical relationships for the dynamic response characterization of a


SDOF model are illustrated in the following with the example of the four-
span bridge bent in Fig. 4.4.
In this example, realistic bridge conditions, such as abutment stiffness and
flexibility, as well as movement joint constraints with the adjacent frame, are
replaced by idealized free or fixed boundary conditions to demonstrate and
emphasize the analytical procedures. More realistic prototype characteristics
are discussed in Section 4.4.2.

Example 4.1. For the example of the multicolumn bent in Fig. 4.4 with an
in-plane rigid bridge deck, the generalized mass and stiffness parameters can
be found as follows, based on the idealized boundary conditions.

Case 1: Abutment 1 Free to Move. With idealized free boundary conditions


of the three-column bent as shown in Fig. 4.4(c) and the assumption of the
same effective stiffness ke in all three bents, the generalized parameters for the
transverse translation of the mass centroid at x l = L/2 can be expressed with

tp(x) = 1 = const. (4.44)

as

m* = K2
= 0 = TriL (4.45)
and

k* k, 1 = 3ke

with the effective stiffness k e for each bent derived from Eq. (4.18). The (4.46)
fundamental period of vibration can then be found from Eq. (4.39) as

T=277 k m* * =27r (4.47)


mL
Case 2: Abutment 1 Pinned. For the case of an idealized pinned in-plane

abutment boundary condition as shown in Fig. 4.4(d ), an infinite transverse


FUNDAMENTALS OF SEISMIC BRIDGE BEHAVIOR 181

stiffness at the abutment is added to the system and the generalized displaced
shape can be expressed as

(x) = 1 x (4.48)

where x* = L/2 represents the distance from the center of stiffness to the
generalized coordinate or displacement DOF u*(t). Lumped-mass properties m
and j can be found at x = L/2 from Eqs. (4.7) and (4.8), respectively. The
generalized mass can now be found as
*_ -
L (L/ 2 ) 2 WIL 3 ( 1 2

M _m
"IL4-
X* 12 x* = 12 (LL/23)2 = nlL (4.49)

and the generalized stiffness as


)2 7
k * = E k = E x . = ke
x* 2 (4.50)

resulting in a natural period of vibration of


\/* 8 inL
-m
T = 277 27r 71-
21 ke
(b) Multidegree-of-Freedom Characteristics. For multidegree-of-freedom (4.51)
(MDOF) bridge models, Eq. (4.32) still describes the undamped free vibration
response, but now k and m are no longer scalars but matrices containing
stiffness and mass coefficients, which correspond to the vector of chosen
displacement degrees of freedom u and their interaction. Denoting vector and
matrix quantities by bold type, the general undamped free vibration of a
MDOF system can be expressed as

mu(t) + ku(t) = 0 (4.52)

The stiffness matrix k can be obtained from standard static displacement-


based analysis models [G31 and features coupling between DOFs in the form
of off-diagonal terms, whereas the mass matrix m, due to the negligible effect
of mass coupling, can best be expressed in the form of tributary lumped masses
to the corresponding displacement degree of freedom, resulting in a diagonal or
uncoupled mass matrix.
The characteristic dynamic equation for the n-MDOF bridge model can
now be expressed as

(k co2m)u = 0 (4.53)
182 MODELING AND ANALYSIS

which for arbitrary displacements requires that

1k (02m1 = 0 (4.54)

and n roots or solutions o can be found that satisfy Eq. (4.54), representing
the n natural circular frequencies of motion of the bridge model.
Once the vector a) with characteristic or modal frequencies is obtained, the
corresponding mode shapes or natural modes of vibration can be obtained
by substituting the individual modal frequencies into Eq. (4.53), and arbitrarily
prescribing the magnitude of one of the DOFs for the harmonic motion, since
with time, any magnitude of displacements between zero and the maximum
values is possible.

Example 4.2. The basic concepts of the dynamic response characterization


of a MDOF bridge model is explained by the example of Fig. 4.13, which
represents an idealized bridge bent for a double-deck viaduct and the corres-
ponding simplified two-degree-of-freedom analysis model. In this model, for
illustration purposes, axial and shear deformation of members are neglected
and the bridge superstructures are assumed to be rigid flexural elements. More
realistic bridge modeling assumptions for multicolumn bents are provided in
Section 4.4.2.
From Eq. (4.19) the flexural stiffness for each upper column can be deter-
mined as 12EIIH3 and from Eq. (4.18) the individual lower column transla-
tional stiffness is 3E11113, resulting in a stiffness matrix k of

24E/ 24E1

H3 H3 = 6E1 [ 4 4
k= (4.55)
H3 4 5
24E/
H3 H3
30E/

DOF
'ioEcoEr' 1 1.0 1.0

H Eleff

1


1
rigid

H EIS Eleff

B B MODE 1 MODE 2

(a) Structure (b) Idealization (c) Displacement Modes


FIG. 4.13 Double-deck viaduct bridge bent.
FUNDAMENTALS OF SEISMIC BRIDGE BEHAVIOR 183

The mass matrix can be expressed with the cap beam length B and the
assumed distributed constant mass Tri = m/B as
[mB 0 = m B [1 0 _ 0
(4.56)
m
1
= =m
0 mB 01 01

and the characteristic Eq. (4.54) requires that the determinant of


(6113E/ [ 44 541 _ co2m 11 011
=0
LO 1j

Substituting A = co2mH3/6E/ results in


6E1 [4 A 4 _0
H3 4 5 A

and A2 9A + 4 = 0 with roots Al = 0.469 and A2 = 8.531.


The circular natural frequencies can now be obtained as

wl
and the mode shapes 4:1) i can be determined from Eq. (4.53) assuming that u1
= 1 as
ct, = 1.0 ci)2 = { 1.0

and
0.883 1.133

These mode shapes are depicted schematically in Fig. 4.13(c) and show
that in mode 1 the two lumped masses move in phase, whereas the sign change
in mode 2 constitutes out-of-phase motion of the two mass points.

4.3.3 Elastic Seismic Response of Bridges: Maximum Response Values


The general forced vibration of a structure in the case of an earthquake was
expressed by Eq. (4.3) with the mass times the acceleration as the forcing
function. For a discrete displacement-based analytical bridge model, Eq. (4.3)
can be written in matrix form as

mii(t) + 6(0 + ku(t) = p eff(t) (4.57)

with the forcing function peff(t) expressed as

Peff(t) = mrag(t) (4.58)


184 MODELING AND ANALYSIS

The vector r represents the displacement of each of the DOFs for a


unit ground displacement ug and determines which mass DOFs are directly
excited by the earthquake ground motion. For only translational DOFs in
the direction of earthquake ground motion, the influence coefficient vector
r takes on the form of a unit vector {1}. Equations (4.57) and (4.58) apply
equally to lumped SDOF, generalized SDOF, or MDOF systems with the
number of simultaneous equilibrium conditions or the problem size to be
solved represented by the number of selected displacement DOFs in the
analytical bridge model.
The dynamic bridge response analysis can be carried out in various forms:
through (1) a rarely used frequency-domain analysis, (2) a direct integration
of the coupled equilibrium equations in the time domain, or (3) a transforma -
tion to normal or modal coordinates and solution and superposition of uncou-
pled or orthogonal modal response. All three techniques lead to a complete
time-history displacement response of the bridge model, and details of the
various solution strategies can be found in the technical literature on structural
dynamics [C10,N1]. In general, since both the frequency-domain and modal
time-domain analyses rely on the principle of superposition, these techniques
are only applicable to linear or linearized bridge systems, whereas the direct
integration can also be employed for nonlinear time-history analyses. How -
ever, linearized iterative analysis techniques based on load-dependent Ritz
vectors [W8] rather than natural made shapes can be used to solve for non -
linear time-history response. Since earthquake ground motions tend to excite
the lowest modes of vibration more than the higher modes, good approxima-
tions of the earthquake response of a bridge can be obtained from only a few
modes, making the modal superposition analysis a powerful tool for bridge
systems, particularly when large numbers of DOFs are involved.
The seismic assessment or design of bridges is generally based on extreme
or maximum dynamic response quantities and does not necessarily require a
complete time-history response. Although these maximum response quantities
can be obtained by scanning time-history. response records, it is typically
sufficient and more convenient to determine maximum modal response quanti-
ties by means of response spectra, as discussed in Section 2.2, once the dynamic
response characteristics in the form of natural periods of vibration and mode
shapes have been determined.
In the combination of maximum response values for individual modes of
vibration, two issues need to be considered: (1) that each of the modes of
vibration has a different earthquake participation factor, and (2) that the
maximum response in each mode does not occur at the same time and in the
same direction during the earthquake duration. For a bridge model with mass
matrix m, normalized mode shapes (Z and ground motion influence coefficient
,

vector r, the participation of each mode can be obtained as the modal participa-
tion coefficient

4imr
Pi= T (4.59)

Вам также может понравиться