Sie sind auf Seite 1von 9

Engineering Structures 32 (2010) 29312939

Contents lists available at ScienceDirect

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

Best-fit models for nonlinear seismic response of reinforced concrete frames


Andres Lepage a, , Michael W. Hopper a , Sebastian A. Delgado b , Jeff J. Dragovich c
a

The Pennsylvania State University, 104 Engineering Unit A, University Park, PA, 16802, USA

Departamento de Estructuras, Facultad de Ingeniera, Universidad del Zulia, Av. Goajira, Ala D, Maracaibo, Zulia, Venezuela

National Institute of Standards and Technology, Building and Fire Research Laboratory (MS 8604), 100 Bureau Drive, Gaithersburg, MD 20899-8604, USA

article

info

Article history:
Received 19 February 2010
Received in revised form
14 April 2010
Accepted 24 May 2010
Available online 23 June 2010
Keywords:
Frequency domain error
Hysteresis
Nonlinear dynamic analysis
Timehistory analysis

abstract
This paper identifies the optimal combination of hysteresis-modeling and damping parameters for use in
practical nonlinear dynamic analysis to obtain satisfactory correlations in both amplitude and waveform
between the calculated and measured seismic response of reinforced concrete frames. In this study, frame
members are characterized by five modeling parameters: initial stiffness, bondslip rotations, post-yield
stiffness, unloading stiffness, and viscous damping. The calculated response is compared with measured
data from three small-scale shake-table multistory test structures and from a seven-story instrumented
building. The three test specimens (structures MF1, MF2, and FNW) are each analyzed for two different
base acceleration tests whereas the seven-story building (Holiday Inn at Van Nuys, CA) is analyzed using
a single recorded seismic event (1994 Northridge) in each of the two principal directions of the building
(structures HNS and HEW). Analyses for all five structures are carried out using three different computer
programs. The goodness-of-fit of the computed response to the recorded experimental data is measured
by the Frequency Domain Error (FDE) index. Simplified rules are presented to derive the best modeling
characterizations that give consistent low values of FDE for the various structures and structural analysis
programs considered.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
Practicing engineers are increasingly using nonlinear static or
dynamic analysis to evaluate the response of a structure subjected
to seismic events. When designing new buildings, structural engineers refer to the governing building code for loads and analysis procedures. When designers choose a non-compliant system to
resist lateral loads, sufficient analytical and test data must be submitted to assure that the non-compliant system has adequate seismic response comparable to that of a conventional code-compliant
system. Deviating from the building code is likely to trigger at
least two actions: first, a traditional code-level design is performed
where engineers use a linear response analysis and relate it to
traditional lateral force-resisting systems; second, the building is
evaluated using nonlinear dynamic analysis for its response to a
severe seismic event. These actions are commonplace in the seismic evaluation of existing buildings [1].
Although a nonlinear dynamic analysis is an excellent way to
evaluate the performance of structures subjected to strong ground
motions, the modeling complexities involved in characterizing the

number and type of material nonlinearities often lead to erroneous


models and inadequate solutions. Simple rules are needed to identify and characterize assorted modeling parameters and to provide
guidance for their implementation using standard structural analysis software. This paper stems from the search for practical computer models capable of representing a realistic nonlinear seismic
response of reinforced concrete frames with a satisfactory correlation between calculated and measured response, not only in amplitude but also in waveform.
This study is limited to mid-rise reinforced concrete frames that
are subjected to strong ground motions. The response-history analyses include geometric and material nonlinearities with hysteretic
response defined at a macroelement level. The primary aim is to
identify the optimal values of the hysteresis modeling parameters
after comparing calculated data with data recorded from the University of Illinois earthquake simulator [2] on three small-scale test
specimens and from the instrumented Holiday Inn building in Van
Nuys, California [3]. For this purpose, nonlinear dynamic analyses
were performed using five basic modeling parameters: initial stiffness, bondslip deformations, post-yield stiffness, unloading stiffness, and type of damping.
2. Nonlinear dynamic analysis

Corresponding address: The Pennsylvania State University, Department of


Architectural Engineering, 104 Engineering Unit A, University Park, PA, 16802, USA.
Tel.: +1 814 865 3013; fax: +1 814 863 4789.
E-mail address: lepage@psu.edu (A. Lepage).
0141-0296/$ see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.engstruct.2010.05.012

The seismic analysis in this study uses a numerical model which


directly incorporates the nonlinear loaddeformation characteristics of individual frame members. The model is subjected to

2932

A. Lepage et al. / Engineering Structures 32 (2010) 29312939

earthquake shaking represented by ground-motion acceleration


histories. The computed response includes loaddeformation histories for individual frame members as well as frame global response parameters: base shear, base overturning moment, and
story drifts. Analyses are carried out with three computer programs: LARZ [4], PERFORM 3D [5] and SAP 2000 [6].
The following general assumptions and simplifications were involved in the development of the numerical model for the frames
considered: (1) the structure, loads, and response are defined in
one vertical plane; (2) the ground motion is considered in the
horizontal direction; (3) the foundation of the structure is rigid;
(4) masses are lumped at the floor levels and constrained by rigid
diaphragms; (5) each frame element is a line element with a linearelastic middle portion bounded by nonlinear springs connected
to rigid beamcolumn joints; (6) linear-elastic response accounts
for shear and flexural deformations; (7) axial deformations are assumed to be negligible; (8) material nonlinearity occurs only due
to flexure; (9) slip of reinforcement is considered at beamcolumn
connections; (10) P effects are incorporated into the basic formulation of the structural stiffness matrix using a geometric stiffness correction that accounts for the effects of axial load on the
chord rotation of the columns. All of the three computer programs
considered allow the implementation of the above assumptions.
The modeling assumptions target ductile flexural response at
the member ends. Localized brittle failure modes such as shear failure and beamcolumn joint failure are not directly accounted for
in the analysis; the sudden strength loss typically associated with
the nonlinear modeling of a brittle failure may be computationally
difficult and prone to solution convergence problems. The probability that these types of failure may occur can be assessed by
post-processing the analysis results to check the demands at the
locations in question.
In both LARZ and PERFORM 3D, the solution technique for integration of the equation of motion uses the Newmark method [7]
with = 1/4. In SAP 2000, the solution is based on the Fast Nonlinear Analysis (FNA) method by Wilson [8] with participation of
all translational modes of vibration (one per story). Stiffness characteristics are assumed constant over the user-defined time step
and the programs divide the time step into smaller substeps if a
change of stiffness is detected. The computed response is calculated and reported at time steps no larger than the smaller of 1/20
of the fundamental period of the uncracked structure or 1/4 of the
time step used to define the ground acceleration. These limits are
recommended to facilitate convergence of the solution technique
and to capture the high-frequency content of the response.
The damping matrix, [C ], is constructed by a linear combination
of the mass [M ] and stiffness [K ] matrix, both defined at global
structural degrees of freedom:
[C ] = [M] + [K ]

(1)

where and are defined using

i =

1
2i

+ i2

3. Modeling parameters
Practicing engineers rarely test the assumptions built-in their
computer models with measured data. With the intention of
helping engineers create meaningful models and computer output,
this section describes the key modeling parameters considered and
gives two relatively extreme values to each. Each value typifies
a modeling assumption that is to be tested in combination with
the values assigned to the other parameters. The test is based on
determining the goodness-of-fit between the calculated nonlinear
seismic response and the measured data for selected structures as
discussed later in Sections 5 and 6.
It is assumed that the loaddeformation characteristics of the
members are dominated by flexure. The effect of axial load was
taken into account when defining the momentrotation relationship for each column. Five basic modeling parameters are considered, four of which directly relate to the primary backbone
momentrotation relationship of frame members: initial stiffness,
bondslip deformations, post-yield stiffness, and unloading stiffness. The fifth parameter, viscous damping, also has a significant
impact on the dynamic response. Table 1 shows a summary of the
parameters considered.
3.1. Initial stiffness
Momentrotation primary curves are derived from moment
curvature relationships with the assumption that inflection points
are at midspans. The initial slope of the momentcurvature diagrams is determined using uncracked (U ) or cracked (C ) section
properties. For U cases, the initial stiffness is based on gross section
properties and the cracking
p point is calculated
p
based on a modulus of rupture, fr , of 1/2 fc0 (MPa) 6 fc0 (psi) . For C cases, the
initial stiffness is defined using the secant to yield. Both U and C
cases share the same yield point. These cases explore if satisfactory
correlations may be obtained when the cracking point is ignored.
The cracking and the yield point are derived from sectional analyses based on the concretes stressstrain relationship by Hognestad [10] and the steel is taken as perfectly elastoplastic. Although
the cracking and yield moments are determined for the initial axial
load due to gravity, both moments are assumed to be independent
of the change in axial load induced by the ground motion.
3.2. Bondslip rotation
Bondslip rotation accounts for the softening effect due to the
elongation of the longitudinal reinforcement beyond the beam
column interface into the joint. Bondslip rotation, 0 , is defined
herein as a function of the development length expressed as a multiple of the bar diameter, db :

0 =

1
2

y db
f

(2)

and i is the fraction of critical damping for mode i of frequency i .


It is convenient to define damping as a function of the first-mode
frequency, 1 , based on uncracked section properties. For the case
of mass-proportional damping, = 21 1 and = 0, where
affects the unchanging [M ] matrix. For stiffness-proportional
damping, = 0 and = 21 /1 , where affects the continuously
updated [K ] matrix.
Additional details are presented in Ref. [9] including the input
data required for the definition of the nonlinear spring properties
for each frame member in all of the structural models considered
herein.

M
My

2
(3)

where = 4uy , fy is the yield stress of reinforcement, u is the average bond stress, y is the curvature at yielding, M is the applied
moment, and My is the yield moment.
Eq. (3) was derived based on four simplifying assumptions,
similar to those found in Refs. [4,11]: (i) pullout does not occur
when the reinforcement develops fy ; (ii) the steel stress varies
linearly from a maximum at the beamcolumn interface to zero
inside the joint; (iii) the rotation due to bond slip is measured
with respect to the neutral axis; and (iv) the tensile stress in the
reinforcement is proportional to the moment.
Two cases of bondslip rotation are considered: a case of tight
(T ) bond where is taken as 20 bar diameter lengths, and a case
of loose (L) bond where is taken as 40 bar diameter lengths. The
values of closely correspond to practical values of fy and u. For
instance, a case where fy = 410 MPa (60 ksi) and u = 5.2 MPa
(0.75 ksi), gives = 20, and a case where fy = 410 MPa (60 ksi)

A. Lepage et al. / Engineering Structures 32 (2010) 29312939

2933

Table 1
Parameter identification.
Parameter

Characteristic value

Symbol

Initial stiffness:

Uncracked
Cracked

fr = 1/2
fr 0

Bondslip:

Tight
Loose

= 20
= 40

T
L

Post-yield stiffness:

Hard
Soft

Kp = 0.10 Ke
Kp = 0.02 Ke

H
S

Unloading stiffness:

Nonreducing
Reducing

=0
= 0.6

N
R

Viscous damping:

2% mass proportional
5% mass proportional
2% stiffness proportional
5% stiffness proportional

2
5
2
5

and u = 2.6 MPa (0.375 ksi), gives = 40. The large values of 0
that correspond to the L cases may also be interpreted as including
some joint flexibility.
3.3. Post-yield stiffness
The momentrotation post-yield stiffness, Kp , is expressed as a
fraction of the secant-to-yield stiffness, Ke = My /e where e is the
effective yield rotation which includes bondslip effects. Two cases
are considered: a soft case (S ) where Kp = 0.02Ke and a hard case
(H ) where Kp = 0.10Ke . The values of Kp are directly defined in
programs PERFORM 3D and SAP 2000 via M curves while LARZ
requires a special definition of the bondslip beyond the yield point
to control Kp .
The moment and rotation at ultimate depend on the assumed
material properties, the curvature distribution along the member
length, and the plastic-hinge length. Instead of the precise calculation of the ultimate point, it is convenient to define the post-yield
stiffness as a function of the secant to yield. A finite positive slope
is assigned to the stiffness after yielding to incorporate the strain
hardening characteristics of the longitudinal reinforcement. This
approach also gives the analyst the option of defining the primary
momentrotation curve mainly as a function of the yield point.
Because the structures considered in this study predominantly
responded with rotation demands below the ultimate point, the
suggested approach proved satisfactory without need for consideration of strength degradation.
3.4. Unloading stiffness
Measured loaddeformation hysteresis data for reinforced
concrete members have shown that the unloading stiffness
decreases as the maximum deformation increases [12]. The
unloading stiffness has been defined in relation to a stiffnessreducing exponent, , as implemented by the Takeda hysteresis
model [13]:


Kr = Ke

e
m


(4)

where Kr is the unloading stiffness, Ke is the secant stiffness to the


yield point, e is the effective yield rotation including bondslip
effects, and m is the maximum rotation attained. The values of Kr

= 0.02
= 0.05
= 0.02
= 0.05

f0
c

(MPa)

U
C

when unloading from a positive direction depend on the values of


Ke , e , and m for the positive direction.
In relation to the representative momentrotation curves
shown in Table 1, the yield point is identified with an open circle of
coordinates (e , My ) where My = Ke e and Ke is the slope from the
origin to the yield point; Kr is the slope of the unloading segments
R or N; and m is defined by the largest rotation along segments S
or H.
Two values of are considered: zero to represent a non-reducing (N ) unloading stiffness and 0.6 to represent a reducing (R)
case. The case of = 0 is investigated because engineers often
choose oversimplified hysteresis models where the unloading stiffness is assumed constant throughout the response history analysis.
The value of = 0.6 was selected over the original 0.4 recommended by Takeda [13] because pilot studies [9] showed an improved correlation with the measured data considered.
A different hysteresis model is used by each of the programs
considered: the Takeda model [13] in program LARZ; the energyratio model [5] in PERFORM 3D; and the pivot model [14] in SAP
2000. All of these models are stiffness-reducing models. In LARZ,
the value of is directly assigned as a Takeda hysteresis model
parameter but in PERFORM 3D and SAP 2000 the target value of
has to be emulated through other hysteresis parameters.
In PERFORM 3D, the N case is represented by an energydissipation ratio of 0.5 and the R case by a ratio of 0.2. In SAP
2000, the N case is represented with the pivot model using pivot =
10, pivot = 0.5, and pivot = 0; for the R case, the pivot model is
defined using pivot = 1.0, pivot = 0.3, and pivot = 10. These
variables are described in detail in Refs. [5,6,14].
3.5. Viscous damping
When inelastic behavior is explicitly considered, damping is
meant to represent the energy dissipation associated with preyield stages of response and therefore is usually taken into account by means of a damping factor expressed as a fraction of
critical damping. Damping forces are assumed to be proportional
to the relative velocities of the points where the translational degrees of freedom (at floor levels) are defined. The damping matrix
is assumed to be a linear combination of the mass and stiffness
matrices, as expressed in Eq. (1). This means that if damping is

2934

A. Lepage et al. / Engineering Structures 32 (2010) 29312939

b
Fig. 1. Test structures, (a) MF1 and MF2 and (b) FNW.

defined as stiffness proportional then the damping matrix is updated every time an element cracks or yields, whereas if damping is
defined as mass proportional then it remains unchanged throughout the analysis. In addition, stiffness-proportional damping implies increased damping values on modes of vibration other than
the fundamental mode whereas mass-proportional damping implies reduced damping values on the higher modes.
Four cases for the amount and type of viscous damping are considered. Two cases use mass-proportional damping at 2% and 5% of
critical damping (2 , 5 ) and two cases use stiffness-proportional
damping at 2% and 5% of critical damping (2 , 5 ). The constants
and (Eq. (1)) are determined using Eq. (2) with the fundamental
frequency of vibration always based on uncracked section properties.
Table 1 presents a summary of the five modeling parameters
considered. Four parameters adopt two values and a fifth parameter adopts four values, giving at most a total of 64 combinations
per structure for a single ground motion and for each program of
analysis.
4. Structures and ground motions
Nonlinear dynamic analyses are performed on three small-scale
test structures responding to two separate shake-table base motions and a real building subjected to the ground motions recorded
at its base during the 1994 Northridge earthquake. The analyses
incorporate the combinations of parameter values presented in
Section 3.
4.1. Test structures: MF1, MF2, and FNW
The tests were performed using the University of Illinois earthquake simulator [2]. Geometries of the specimens are shown in
Fig. 1. Structures MF1 [15], MF2 [16] and FNW [17] were composed
of two planar frames working in parallel. The ten-story structures,
MF1 and MF2, had identical geometries except for the discontinued beam at the first story in MF2. The nine-story structure, FNW,
had a tall first story. Nominal cross sectional dimensions of beams
and columns were identical for all three frames, with all columns
stiffer and stronger than the beams. The heavy base girders and
the specimen-to-simulator connections were designed to simulate
a fixed-base condition. The story weights were nominally 4.45 kN
(1 kip) with the exception of the first story in MF2 where nearly
1/3 of the weight was removed. The weights were transferred directly to the column centerlines such that each column carried an
equal fraction of the story weight.

Generally, test runs of a given specimen included repetitions


of free vibration tests, to determine low-amplitude natural frequencies, followed by earthquake simulation. The sequence was
repeated with the intensity of the earthquake simulation being increased in successive runs. Only the first two runs are considered.
These runs targeted a roof displacement of approximately 1% the
frame height for run 1 and 2% for run 2. The base motions were
patterned after El Centro 1940 NS with the time scale compressed
by a factor of 2.5 to obtain realistic ratios between input frequency
contents and natural frequencies of the specimens. Recorded data
include displacements relative to the ground and absolute accelerations, all measured at beam lines.
Specimens were cast using small-aggregate concrete with a
nominal compressive strength of 35 MPa (5 ksi) and a modulus of
elasticity of about 21,000 MPa (3000 ksi). All members were reinforced by steel wire gage No. 13 and 16, longitudinally and transversely, respectively. Longitudinal reinforcement with a nominal
yield stress of 350 MPa (50 ksi) was continuous through all joints
and extended into exterior joint stubs for anchorage. Transverse reinforcement provided by rectangular spirals prevented shear failure before flexural failure. Beamcolumn joints were reinforced
with helical spirals to prevent joint failures. More details about the
fabrication of the specimens are found in Refs. [1517]. Tabulated
data in Ref. [9] describe all member properties in relation to the
modeling parameters presented in Section 3.
4.2. Instrumented building: HEW and HNS
The seven-story Holiday Inn building in Van Nuys (California)
was built in 1966 on the northeast side of the Los Angeles basin
at 6 km (4 miles) from the epicenter of the 1994 Northridge
earthquake. The structure is essentially symmetrical in elevation
and in plan about both axes (Fig. 2), except for light framing
members supporting the stairway and elevator openings on the
southwest end. The nonlinear dynamic analysis of the building
is performed using two-dimensional models for each direction
of analysis: the EastWest frames, or structure HEW, and the
NorthSouth frames, or structure HNS. Given the symmetrical
layout and the presence of perimeter moment frames, the merits
of a 2D model have been explored and proven satisfactory [3].
The floor system is a reinforced concrete flat plate that is
220 mm (8.5 in.) thick at typical floors, 250 mm (10 in.) at the first
floor, and 200 mm (8 in.) at the roof. Spandrel beams, 410 mm
(16 in.) wide in the EW direction and 360 mm (14 in.) in the
NS direction, are typically 570 mm (22.5 in.) deep, except at the

2935

Imaginary

A. Lepage et al. / Engineering Structures 32 (2010) 29312939

to top of slab

(Roof)

(Rc, Ic)
(Base)

Err
or

Elevation NS
(Section BB)
Calculate

Elevation EW
(Section AA)

(Rm, Im)

d
sure

Mea

NORTH
Plan

Real

Fig. 2. Holiday Inn building in Van Nuys, California.

Fig. 3. FDE representation.

first floor and roof with depths of 760 mm (30 in.) and 560 mm
(22 in.), respectively. Seismic design of the building was based on
the assumption that the code-specified lateral forces were resisted
by the combined action of the interior slabcolumn frames and
exterior beamcolumn frames. The interior column cross sections
are 510 mm (20 in.) square in the first story and 460 mm (18 in.)
square above. Exterior columns are 360 510 mm (14 20 in.) for
the full building height with the 510 mm (20 in.) dimension along
the NS direction. Although the column longitudinal reinforcement
ratios range between one and three percent of the column gross
area, the transverse reinforcement, spaced at 300 mm (12 in.), is
noncompliant with modern special moment frame requirements.
All concrete is normal-weight with the first-story columns
having a specified compressive strength of 35 MPa (5 ksi). The
first-level floor slab and columns between levels 1 and 2 were
built with 28 MPa (4 ksi) concrete. All other concrete above grade
had a specified concrete strength of 21 MPa (3 ksi). Columns are
reinforced with steel having a specified yield of 410 MPa (60 ksi),
while beams and slabs used 280 MPa (40 ksi) steel. The foundation
system consists of groups of cast-in-place reinforced concrete
friction piles with pile caps connected by grade beams.
The structure damage due to the 1994 Northridge earthquake
was serious. The perimeter frames had extensive cracking of
concrete related to shear and bond stresses, especially at columns.
A total of 16 accelerometers located at the roof, fifth, second, first,
and ground levels recorded motions in the EW, NS, and vertical
directions. The recorded horizontal peak ground accelerations
were 0.45 g (EW) and 0.42 g (NS). The building experienced roof
displacements of about 1.3% the building height.
Tabulated data are included in Ref. [9] describing all member
properties in relation to the modeling parameters presented in
Section 3.

domain. For each frequency, the real and imaginary components


of the complex number can be thought of as xy coordinates
(Argand diagram) defining the vectors shown in Fig. 3. The FDE index is based on the error vector (Fig. 3) normalized by the sum of
the vector magnitudes of the measured and calculated signals:

5. FDE index
The FDE index [18] measures the correlation between two
waveforms. The index is used herein to help identify the combinations of modeling parameters that lead to the best correlations
between measured and calculated roof displacement response for
the structures considered. Roof displacement is chosen as the key
response parameter because it has proven satisfactory for characterizing the global and local response of frames having individual
elements without abrupt changes in stiffness and with columns
proportioned to develop limited yielding [4].
The FDE index quantifies amplitude and phase deviations between two signals and gives a value between 0 and 1, where 0 is
for a perfect correlation and 1 is for 180 out of phase. The index
uses the Fourier transform of both the measured and calculated
signals to represent them as complex numbers in the frequency

f2
P

FDE =

RMi RCi

2

+ IMi IC i

2

i=f1
f2 q
P
i=f1

2
R2Mi + IM
i

(5)

q

R2Ci + IC2i

where RMi and IMi are the real and imaginary components of the
measured signal at frequency i; RCi and ICi are the real and imaginary components of the calculated signal at frequency i; f1 and
f2 are the starting and ending frequencies for summation. In this
study, frequencies f1 and f2 are defined as a function of the fundamental period, T1 , of the uncracked structure, using 1/(4T1 )
and 1/(0.1T1 ), respectively. This range is expected to capture the
most relevant frequency content of the signals representing inelastic structural response. The duration of the response history over
which the index is calculated is set to approximately 30T1 .
6. Analysis cases and results
To identify the combinations of modeling parameters that
lead to a realistic response, the FDE index is determined for the
roof displacement history obtained from each of the nonlinear
dynamic analysis cases considered. First, trends are identified in
the data corresponding to the test structures; and second, the
Holiday Inn building is analyzed to verify if the combinations that
give satisfactory calculated-to-measured correlations for the test
structures also work for the full-scale building.
Program LARZ was used in all three test structures (MF1, MF2,
and FNW) and in the Holiday Inn building (HEW and HNS) for all
64 possible parameter value combinations (see Table 2). Programs
PERFORM 3D and SAP 2000 were also used to model the same
structures but with a reduced number of cases due to program limitations. When using discrete nonlinear springs at member ends,
both PERFORM 3D and SAP 2000 cannot appropriately represent
the U models using a breakpoint before the yield point.
For each analysis program, the FDE index data were sorted
for each type of structure and damping and presented in the
form of FDE clocks (Figs. 49). An FDE clock [18] is a graphical
representation resembling a dart board, where the center of the
chart (FDE = 0) indicates a perfect correlation between the
measured and calculated response histories. A point near the
periphery (FDE = 0.75) represents a very poor correlation. The
circle is divided into 16 sectors, where each sector represents one
of the 16 models after the combination of U /C , H /S, N /R, and T /L.

2936

A. Lepage et al. / Engineering Structures 32 (2010) 29312939

Table 2
Analysis cases.
Program

Structure

Base motion

Parameters

Cases

LARZ

MF1/MF2/FNW
HEW/HNS

run 1/run 2
1994 Northridge

U /C , H /S , N /R, T /L, 2 /5 /2 /5
U /C , H /S , N /R, T /L, 2 /5 /2 /5

3 2 64 = 384
2 1 64 = 128

PERFORM 3D

MF1/MF2/FNW
HEW/HNS

run 1/run 2
1994 Northridge

C , H /S , N /R, T /L, 2 /5 /2 /5
C , H /S , N /R, T /L, 2

3 2 32 = 192
2 1 8 = 16

SAP 2000

MF1/MF2/FNW
HEW/HNS

run 1/run 2
1994 Northridge

C , H /S , N /R, T /L, 2 /5 /2 /5
C , H /S , N /R, T /L, 2

3 2 32 = 192
2 1 8 = 16

Fig. 4. FDE clocks, structure MF1, program LARZ.

Fig. 5. FDE clocks, structure MF2, program LARZ.

Fig. 6. FDE clocks, structure FNW, program LARZ.

calculated roof displacement is 80% or less than the maximum


measured roof displacement. Note that the 5 models are more
likely to underestimate the roof displacement. For run 1, the lowest FDE index averages corresponding to the data in Figs. 46
include: CHRT-2 , CHRL-5 , UHRT-2 , and UHRL-5 , with FDE index averages of 20%, 26%, 19% and 19%, respectively. For run 2, the
only model with FDE indexes below 25% for all test structures was
CHRL-2 .
FDE clocks corresponding to the calculated response using
PERFORM 3D and SAP 2000 for 2 models are shown in Figs. 79.
The trends identified with LARZ were corroborated by data from
PERFORM 3D and SAP 2000: for run 1, model CHRT-2 was the top
performer, whereas for run 2, model CHRL-2 averaged the lowest
FDE index.
Roof displacement histories (Figs. 10 and 11), base shear histories (Fig. 12), and story drift ratios (Fig. 13) for a few of the models
representative of the lowest FDE index averages, indicate a satisfactory correlation between the calculated and measured response.
There is a tendency for the analytical models to underestimate the
maximum story drift ratios in some stories, especially for the run 2
cases, as shown in Fig. 13 for the first and top stories. This suggests
that the analytical models deviate from the idealized fixed-base
condition and do not properly account for the high-mode effects
in cases where the structures had previously yielded.

6.1. Test structures: MF1, MF2, and FNW

6.2. Instrumented building: HEW and HNS

The goodness-of-fit data for the calculated roof displacement


response for all 384 LARZ cases (see Table 2) are summarized in
the FDE clocks shown in Figs. 46. The plotted data include a cross
marker to identify cases where the amplitude of the maximum

The goodness-of-fit data for the calculated roof displacement


response for all 128 LARZ cases (see Table 2) are summarized in
the FDE clocks shown in Fig. 14. The FDE clock data suggest that
the U models in LARZ have a better correlation with the measured

A. Lepage et al. / Engineering Structures 32 (2010) 29312939

2937

Fig. 7. FDE clocks, structure MF1, (a) PERFORM 3D and (b) SAP 2000.

Fig. 10. Roof displacement histories, MF2 run 1, program LARZ.

Fig. 8. FDE clocks, structure MF2, (a) PERFORM 3D and (b) SAP 2000.

Fig. 9. FDE clocks, structure FNW, (a) PERFORM 3D and (b) SAP 2000.

response when compared to the C models (except for the 5


models). The FDE clocks obtained for 2 damping (Fig. 15) with
programs PERFORM 3D and SAP 2000 also support the relatively
poor correlation of the C models. This may be an indication that the
nonstructural components of the building affected the response by
providing an initial stiffening effect and/or additional damping.
The Holiday Inn building had suffered minor damage during the
1971 San Fernando earthquake but was subsequently repaired to
a nearly uncracked state. Therefore, a comparison with the test
structures subjected to run 1 is appropriate. When comparing the
data from the analysis of the test structures (run 1) with that of the
Holiday Inn building, only two models give an FDE index below 25%
for all structures: models UHRT-2 and UHRL-5 . Roof displacement response of these models when applied to HEW and HNS are
shown in Fig. 16, indicating a satisfactory correlation between the
measured and calculated data, with only a slight overestimation of
the maximum displacement.

Fig. 11. Roof displacement histories, test structures run 2, program LARZ.

7. Conclusions
To identify the modeling assumptions that lead to the best correlation between calculated and measured seismic responses, a
series of nonlinear dynamic analyses was performed on three
small-scale shake-table test structures (MF1, MF2, and FNW) and
on the orthogonal structural systems (HNS and HEW) of an instrumented seven-story building located in Van Nuys, California. The
structures represent multibay multistory reinforced concrete moment frames. The nonlinear dynamic analyses were implemented
using three programs (LARZ [4], PERFORM 3D [5] and SAP 2000 [6]).
Each program was used to represent the influence of five modeling parameters: initial stiffness, bondslip rotations, post-yield
stiffness, unloading stiffness, and viscous damping. Each of the
modeling parameters is assigned two extreme but plausible values to help identify the direction of the parameter value that
most favorably represents a realistic response. In all, for a single

2938

A. Lepage et al. / Engineering Structures 32 (2010) 29312939

Fig. 15. FDE clocks, structures HEW and HNS, (a) PERFORM 3D and (b) SAP 2000.

Fig. 12. Base shear histories, test structures run 2, program LARZ.

Fig. 13. Displacement response envelopes, MF2 run 2, model CHRL-2 .

Fig. 16. Roof displacement histories, structures HEW and HNS, program LARZ.

Fig. 14. FDE clocks, structures HEW and HNS, program LARZ.

structure and ground motion, up to 64 analysis cases were considered per program. The Frequency Domain Error (FDE) index [18]
helped identify the analytical models having the best and most

consistent correlation with the measured data. The calculated response wasprocessed using the FDE clocks, which is an effective
tool for visualizing the influence of multiple parameter values on
the correlation between the calculated response and the measured
response histories.
From the limited number of structures, modeling assumptions,
modeling parameter range of values, and computer programs considered, the following are concluded (refer to Table 1 for parameter
identification):
(1) For the test structures (MF1, MF2, and FNW) subjected to
run 1, the best models for each type of damping were: CHRT-2 ,
CHRL-5 , UHRT-2 , and UHRL-5 .
These models are derived by the following three rules: (i) use
10% post-yield stiffness ratio and an unloading stiffness parameter
of 0.6 (i.e. use models with H and R); (ii) for models using uncracked initial stiffness select stiffness-proportional damping and
for models using cracked initial stiffness select mass-proportional
damping (i.e. use U models with or C models with ); and
(iii) for models where bondslip deformations are based on development lengths of 20 bar diameters use 2% damping and for development lengths of 40 bar diameters use 5% damping (i.e. use T
models with 2% damping or L models with 5% damping). Note that
constant damping (mass proportional) works well with the softer

A. Lepage et al. / Engineering Structures 32 (2010) 29312939

C models and that higher values of damping work well with the
softer L models.
(2) For the test structures (MF1, MF2, and FNW) subjected to
run 2, model CHRL-2 outperformed all others. This model may be
viewed as a logical outcome from the rules given for run 1 because
before run 2 the structures had already yielded and experienced
some damage. The change from T to L is justified because the
structure is effectively softer.
(3) For the full-scale Holiday Inn building (structures HNS and
HEW) two of the best models were consistent with the rules given
above in conclusion (1): UHRT-2 and UHRL-5 .
In general, the models derived with the first two rules of
conclusion (1) had satisfactory accuracy when representing the
global measured response of the test structures for both run 1 and
run 2. Calculated roof displacement, base shear, and overturning
moment histories successfully tracked the measured response.
Although the models for the test structures in run 1 were able to
satisfactorily represent the local measured response, indicated by
story drift ratios, the models for run 2 were not as accurate. This
is possibly due to limitations in the analytical models to properly
account for high-mode effects.
The findings from this study suggest that a valuable contribution to practicing engineers would be to have developers of structural software incorporate moment frame model templates with
pre-assigned nonlinear springs. The default spring properties may
be based on conclusion (1) for an undamaged structure and on
conclusion (2) for a previously damage structure. These properties would set the stage for a more reliable evaluation of the relative merits of several preliminary designs before the final design
is reached.
Research is needed to develop more definite recommendations
of parameter values following statistical analysis of a larger number of structures and a more refined grid of parameter values.
Acknowledgements
The writers are grateful to Prof. Mete Sozen (Purdue University)
for sharing the data from the University of Illinois earthquake
simulator.

2939

References
[1] American Society of Civil Engineers/Structural Engineering Institute. Seismic
evaluation of existing buildings. ASCE/SEI 31-03. Reston (VA); 2003.
[2] Sozen MA, Otani S, Gulkan P, Nielsen NN. The University of Illinois earthquake
simulator. In: Proc. of the fourth world conference on earthquake engineering.
vol. 3. 1969. p. 13650.
[3] Browning JP, Lepage A. Discussion of nonlinear analyses of an instrumented
structure damaged in the 1994 Northridge earthquake (by Li, Y, Jirsa, JO).
Earthquake Spectra 1999;15(1):1759.
[4] Saiidi M, Sozen MA. Simple and complex models for nonlinear seismic
response of reinforced concrete structures. Struct. research series, no. 465.
Univ. of Illinois at Urbana-Champaign; 1979.
[5] Computers and Structures, Inc. PERFORM 3D: nonlinear analysis and performance assessment for 3D structures. Version 4. Berkeley (CA); 2006.
[6] Computers and Structures, Inc. SAP 2000: static and dynamic finite element
analysis of structures. Nonlinear Version 14.0.0. Berkeley (CA); 2009.
[7] Newmark NM. A method of computation for structural dynamics. J Eng Mech
Div, ASCE 1959;85(3):6986.
[8] Wilson EL. A new method of dynamic analysis for linear and nonlinear systems.
Finite Elem Anal Des 1985;1(1):213.
[9] Hopper MW. Analytical models for the nonlinear seismic response of
reinforced concrete frames. M.S. thesis. University Park (PA): The Pennsylvania
State Univ.; 2009.
[10] Park R, Paulay T. Reinforced concrete structures. John Wiley and Sons, Inc.;
1975. p. 115.
[11] Sozen MA. Seismic behavior of reinforced concrete buildings. In: Bozorgnia Y,
Bertero VV, editors. Earthquake engineering: from engineering seismology to
performance-based engineering. Boca Raton (FL): CRC Press; 2004 [chapter
13].
[12] Otani S. Hysteresis models of reinforced concrete for earthquake response
analysis. J Fac Eng Univ Tokyo 1981;36(2):12559.
[13] Takeda T, Sozen MA, Nielsen NN. Reinforced concrete response to simulated
earthquakes. J Struct Div, ASCE 1970;96(12):255773.
[14] Dowell RK, Seible F, Wilson EL. Pivot hysteresis model for reinforced concrete
members. ACI Struct J 1998;95(5):60717.
[15] Healey TJ, Sozen MA. Experimental study of the dynamic response of a tenstory reinforced concrete frame with a tall first story. Struct. research series,
no. 450. Univ. of Illinois at Urbana-Champaign; 1978.
[16] Moehle JP, Sozen MA. Earthquake-simulation tests of a ten-story reinforced
concrete frame with a discontinued first-level beam. Struct. research series,
no. 451. Univ. of Illinois at Urbana-Champaign; 1978.
[17] Moehle JP, Sozen MA. Experiments to study earthquake response of R/C
structures with stiffness interruptions. Struct. research series, no. 482. Univ.
of Illinois at Urbana-Champaign; 1980.
[18] Dragovich JJ, Lepage A. FDE index for goodness-of-fit between measured and
calculated response signals. Earthq Eng Struct Dyn 2009;38(15):17518.

Das könnte Ihnen auch gefallen