Sei sulla pagina 1di 10

ACI STRUCTURAL JOURNAL

TECHNICAL PAPER

Title no. 109-S35

Dynamic Collapse Analysis of a Concrete Frame


Sustaining Column Axial Failures
by Wassim M. Ghannoum and Jack P. Moehle
A one-third-scale, three-bay, three-story reinforced concrete (RC)
planar frame with lightly reinforced columns was subjected to base
motions that simulated earthquake shaking. Dynamic responses
ranged from essentially linear-elastic response through column flexural yielding to column shear and axial failure. A structural analysis
model was implemented in computer software to demonstrate the
modeling and analysis requirements necessary to achieve accurate response simulation. A new bond-slip model is developed and
implemented to represent effects of longitudinal bar slip from joint
and footing anchorage regions. Existing models for column shear
strength and failure are assessed based on the measured and calculated dynamic responses. Recommendations for dynamic response
simulation and column shear-failure assessment are given.
Keywords: axial; collapse; column; earthquake; failure; frame; laboratory
testing; reinforced concrete; shear.

INTRODUCTION
Shear failure of inadequately proportioned and detailed
columns is a leading cause of the damage and collapse of
older concrete buildings subjected to earthquakes. Assessment of the potential for shear failure requires an estimate
of expected earthquake ground shaking, expected dynamic
response of the building frame, and expected column capacity.
This study focused on the estimation of the building frame
dynamic response and column capacity. Analytical models
are developed, calibrated, and assessed by comparison with
the observed performance of a one-third-scale reinforced
concrete (RC) frame tested on an earthquake simulator. The
companion paper (Ghannoum and Moehle 2012) presents the
details of the experimental program and experimental results.
This study builds on important developments in analytical modeling and dynamic response simulation of RC in
recent years. The modeling and simulation are implemented
using OpenSees (McKenna et al. 2000), a software framework for developing applications to simulate the performance of structural and geotechnical systems subjected to
earthquakes. The software includes various types of finite
elements for modeling the nonlinear behavior of RC frame
elements subjected to earthquake loading, including the
force formulation fiber-section elements originally proposed
by Spacone et al. (1996). This element model takes advantage of nonlinear concrete and steel material properties also
incorporated in OpenSees.
More recently, Elwood (2004) introduced the limit state
material model for shear and axial failures of RC columns
with light transverse reinforcement. This model introduces
shear and axial springs attached in series with the fibersection elements, with limiting drifts associated with shear
and axial failure defining when these springs are activated.
It was of interest in this study to examine the efficacy of the
limit state model for a multi-story, multi-bay frame.
ACI Structural Journal/May-June 2012

The slip of longitudinal reinforcement from beam-column


joints and foundation elements results in additional flexibility that should be modeled (Sezen and Setzler 2008).
Several practical methods to estimate increased flexibility due to bar slip have been proposed. Elwood and
Eberhard (2009) proposed a stiffness reduction for the
framing member that is a function of axial load. Several
researchers (Otani 1974; Ibarra et al. 2005) have proposed
adding rotational springs at the frame member ends; the
absence of slip-axial-load interaction is a common shortcoming of this approach. Monti and Spacone (2000) used
fiber elements to represent the flexibility of frame elements
as a function of axial load, increasing the flexibility of the
end fibers to indirectly account for added flexibility due to
bar slip. Berry (2006) and Zhao and Sritharan (2007) introduced a zero-length fiber-section element placed in series
with the framing element. This paper introduces a new zerolength fiber-section element that is intended as an improvement of the aforementioned models.
RESEARCH SIGNIFICANCE
Engineers routinely use structural analysis models implemented in computer software to conduct performance assessments of existing buildings subjected to earthquake shaking.
The opportunity to develop and calibrate such models using
measured test data from physical structures tested to collapse
is rare. The observations and conclusions from this study
can be used to improve the accuracy of analytical simulation of RC structures subjected to high-intensity earthquake
shaking, from which improved estimates of structural deformations and related damage can be obtained. More accurate
performance assessments of existing buildings ensue.
TEST STRUCTURE
The test structure is a three-story, three-bay, planar frame
dimensioned to partially represent typical 1960s office
building construction in California (Ghannoum and Moehle
2012). Two of the columns (at Axes A and B; Fig. 2 in the
companion paper [Ghannoum and Moehle 2012]) had a
longitudinal reinforcement ratio rl of 0.0245 and widely
spaced ties closed with 90-degree hooks, which is typical
of older construction. These columns had a transverse
reinforcement ratio rt of 0.0015 (rt = Av/bs, where Av is the
total area of transverse reinforcement parallel to the framing
direction with spacing s; and b is the column width). The
ACI Structural Journal, V. 109, No. 3, May-June 2012.
MS No. S-2010-209.R2 received June 2, 2011, and reviewed under Institute
publication policies. Copyright 2012, American Concrete Institute. All rights
reserved, including the making of copies unless permission is obtained from the
copyright proprietors. Pertinent discussion including authors closure, if any, will be
published in the March-April 2013 ACI Structural Journal if the discussion is received
by November 1, 2012.

403

ACI member Wassim M. Ghannoum is an Assistant Professor at the University of


Texas at Austin, Austin, TX. He is a member of ACI Committee 369, Seismic Repair and
Rehabilitation, and Joint ACI-ASCE Committee 441, Reinforced Concrete Columns.
His research interests include shear in reinforced concrete, large deformation behavior
of reinforced concrete members, and collapse of structures.
Jack P. Moehle, FACI, is a Professor of Civil and Environmental Engineering at the
University of California, Berkeley, Berkeley, CA. He is Chair of ACI Subcommittee
318-H, Seismic Provisions, a member of the ACI Board of Direction, and a member of
ACI Committees 318, Structural Concrete Building Code; 369, Seismic Repair and
Rehabilitation; and Joint ACI-ASCE Committee 352, Joints and Connections in Monolithic
Concrete Structures. He is also a past member of the Technical Activities Committee.

other two columns (at Axes C and D; Fig. 2 in the companion


paper [Ghannoum and Moehle 2012]) had a longitudinal
reinforcement ratio rl of 0.0109 and hoops that were welldetailed and closely spaced, as required for columns of
special moment frames (ACI Committee 318 2008). These
columns had a maximum transverse reinforcement ratio rt
of 0.011 at their ends. All longitudinal reinforcement was
continuous. Columns with wide hoop spacing were intended
to yield in flexure initially and then sustain shear and axial
failures. In contrast, columns with close hoop spacing were
intended to respond primarily in flexure and thereby slow
the progression of collapse once it initiated in the other
columns. Beams and beam-column joints were dimensioned
and detailed to produce a strong-beam, strong-joint, weakcolumn system. The concrete cylinder compressive strength
f c at the time of testing was 3.56 ksi (24.6 MPa) and the
Youngs modulus Ec was 2800 ksi (19,000 MPa). The longitudinal reinforcing steel had a measured yield stress fy of
64.5 ksi (445 MPa). Additional information about the test
frame can be found in the companion paper (Ghannoum and
Moehle 2012) and in Ghannoum (2007).
Columns, beams, and joints are identified by the
column line, level, and bay in which they are located
(Fig. 1). For example, Column A1 refers to the column along
Column Line A in Story 1, Beam AB1 refers to the beam
bounded by Column Lines A and B and situated at the first
level, and Joint A1 refers to the joint along Column Line A
at Level 1.

Fig. 1Flexure-shear-critical column analytical representation.


404

SHAKE-TABLE TESTS
The test structure was bolted atop load cells fixed to the
earthquake simulator at the Richmond Field Station of the
University of California, Berkeley, Berkeley, CA. The test
structure was braced out of plane, and then lead weights
were attached to the beams to increase inertial effects and
column axial stresses. The resulting gravity axial load was
approximately 0.16Ag f c on the first-story interior columns
and 0.08Ag f c on the first-story exterior columns (Ag is the
column gross section area). Horizontal base motions representative of earthquake shaking were then imparted horizontally within the plane of the test structure. One shaking test
(designated the half-yield test) produced approximately
half the yield stress in the longitudinal steel of the first-story
columns. This test was followed by three high-intensity tests
that caused the partial collapse of the test frame. Only the
half-yield test and the first of the three high-intensity tests
(Test 1) are described in this paper. Additional details about
frame behavior during the strong shaking can be found in the
companion paper (Ghannoum and Moehle 2012).
ANALYTICAL MODEL
Beam and column elements
Force formulation fiber-section elements (Spacone et al.
1996) were used to model columns and beams with secondorder P-D effects included for columns. The fiber-section
force formulation is relatively simple to implement, as it
only requires section geometry and material properties as
input variables. These elements can account for axial load
variations while simulating the full range of cyclic flexural
behavior, including concrete cracking, steel yielding, and
concrete crushing.
Three sub-elements were used per column (Fig. 1). The
sub-elements at column ends had length h, where h is the
column section depth in the direction of loading. These
elements consisted of two fiber sectionsone at each
endwhich correspond to the two integration points of the
elements. The middle column elements were discretized
into five fiber sections. This configuration results in most,
if not all, of the inelastic rotations occurring within the
sub-elements at the column ends and produces an effective
analytical plastic hinge length of approximately h/2. Because
the underlying goal is to develop an analytical model that
requires the least amount of user calibration possible while
providing a high degree of accuracy, a more elaborate evaluation of the plastic hinge length was not deemed necessary,
and the simpler choice of h/2 was selected.
The beams were discretized into six force formulation
sub-elements containing two fiber sections each. These
sub-elements spanned between discrete loading points at
lead-weight attachment locations and out-of-plane bracing
connections. This approach produced sub-elements at the
ends of beams 12 in. (305 mm) long. The effects of beam
end element lengths on the model response were found to
be minimal.
Elastic shear deformations of frame elements were
introduced into the analytical model using zero-length
elastic spring elements at both ends of the columns and
beams (Fig. 1). The stiffness of these springs was given by
(5/6GAg)/(L/2), where G is the shear modulus, Ag is the gross
section area, and L is the element clear length. Shear deformations were minimal compared with flexural ones.
Zero-length springs (Elwood 2004) were introduced
at the base of the columns with wide hoop spacing
ACI Structural Journal/May-June 2012

Fig. 2Bar slip illustration.


(Column Lines A and B) to model possible shear and axial
failure in those columns (Fig. 1); the springs were not used
for columns with close hoop spacing (Column Lines C and
D). A rigid rotational spring was added in parallel with the
shear and axial elements to lock rotational motion. The accuracy
of the shear failure model predictions will be discussed in
the following.
Bar slip
The fiber-section elements described previously account for
the elongation of longitudinal reinforcement along the beam
and column lengths, but not elongation associated with the
slip of longitudinal reinforcement from beam-column joints
and foundation elements (Fig. 2). Bar slip generates rigidbody rotations at the ends of frame elements that can substantially increase member flexibility (Sezen and Moehle 2006).
The following text introduces a new model for this effect.
The proposed bar-slip model uses a zero-length fiber
section (Fig. 1). Fiber models have the advantage of
adjusting the neutral axis location as a function of axial load
and loading direction, thereby fostering a more accurate
representation of bar slip as affected by these loading parameters. The proposed model is an extension of earlier work by
Berry (2006) and Zhao and Sritharan (2007). Those models
produce a discontinuity in steel stresses and neutral axis location between the bar slip and adjacent column fiber sections,
as they use material properties in the bar-slip fiber sections
that are not directly scaled from those of the adjacent column
fiber sections (Ghannoum 2007). The proposed model introduces modifications to avoid those discontinuities.
The bar-slip fiber section has the same geometry as the
fiber sections of the element it is attached to but different
uniaxial material properties for its steel and concrete fibers.
Figure 3 illustrates the two adjacent fiber sections with equal
shear force V, axial load P, and moment M applied. For
a given loading, the bar slip rotation is Obs = Ss/c, where
Ss is the bar slip of longitudinal bars and c is the distance
from the neutral axis to the longitudinal bars in tension. The
frame-element, fiber-section curvature is Kfe = es/c, where
es is the longitudinal bar strain. Thus, Obs = Kfe(Ss/es) for
any loading condition. This implies that concrete strains in
any fiber of the bar-slip fiber section ecbs must be related to
concrete strains in frame element fibers by the ratio r = (Ss/es)
to maintain compatibility of material stresses and neutral
axis location between the bar-slip and frame-element fiber
sections. The same applies to steel fibers. For fiber sections
with bilinear material models, the scaling ratio can be taken
as ry = (Sy/ey), where Sy is the bar slip at yield and ey is the bar
yield strain. Hence, compatibility of material stresses and
ACI Structural Journal/May-June 2012

Fig. 3Bar-slip fiber-section equilibrium, strain profiles,


and materials.

Fig. 4Longitudinal bar anchorage stresses and strains.


neutral axial location can be achieved by selecting identical
steel and concrete material models in bar-slip and frameelement fiber sections and scaling bar-slip material strains
from frame-element material strains by the ratio ry.
The relation between bar slip and stress is evaluated by
integrating the strain profile within the anchorage region
(Fig. 4) under the assumption of bi-uniform bond stress
(based on Lehman and Moehle [2000]). For a bar with sufficient anchorage length, this approach gives
Ss =

s f s db
, s y
8ue

(1)
405

Ss =

y f y db
8ue

+ y

)( f

f y db

8u p

, s > y

(2)

where fy is the bar yield stress; fs is the bar stress at the interface; ey is the bar yield strain; es is the bar strain at the interface; db is the bar diameter; ue is the elastic bond stress; and
up is the plastic bond stress.
For a bilinear steel stress-strain relation, the bar stressversus-slip relation derived from Eq. (1) and (2) is parabolic
in both the elastic and plastic regions (Fig. 5). This relation
can be linearized in both regions without much loss in accuracy (Fig. 5). The bilinear stress-slip relation is then defined
by the yield stress fy, yield slip Sy, and hardening ratio (ratio
of plastic modulus to elastic modulus). From Eq. (1), the
longitudinal bar slip at yield is
Sy =

y f y db

(3)

8ue

This study defines the elastic bond stress as ue = 12

fc

psi (1 fc MPa) based on the recommendation of Lehman


and Moehle (2000). Lowes et al. (2003) suggest a much
higher elastic bond stress value of 21 fc psi (1.74 fc MPa)
at joint interfaces; however, due to the short development
lengths in the test frame joints and the high levels of damage
and deformations observed in the joints, the lower elastic
bond stress value is adopted. The hardening ratio is taken
equal to the longitudinal steel hardening ratiothat is, 0.01.
Given the proposed uniform elastic bond stress and the hardening ratio, Eq. (2) imposes a plastic bond stress value of up
= 3.24 fc psi (0.27 fc MPa) (with linear approximation
matching the parabolic post-yield stress-slip relation at the
ultimate bar stress fu = 1.25 fy) (Fig. 5).

Fig. 5Longitudinal bar stress-versus-slip relation for


No. 3 (10 mm) bar.

Joints
Joints were modeled as rigid elements. This approach
produced reasonably accurate analytical results, as shown
in the following, which suggests that the bar-slip model
described previously accounted for much of the deformation
associated with the joints.
Material properties
The concrete material model was a Kent-Scott-Park
concrete material with degraded linear unloading/reloading
stiffness according to Karsan-Jirsa (Scott et al. 1982). The
envelope relation was calibrated to the stress-strain relation
obtained from the cylinder tests. Confinement effects on core
concrete were estimated from Mander et al. (1988). Other
input concrete properties were: 1) concrete tensile strength
of 7.5 fc psi (0.62 fc MPa); and 2) concrete tensile softening slope of E0/5, where E0 is the concrete material tangent
modulus at zero strain = (2f c/ec) (where ec is the strain at f c).
The longitudinal reinforcing steel material model was a
Giuffre-Menegotto-Pinto (Menegotto and Pinto 1973) steel
material with isotropic strain hardening. Measured material
properties were used for the uniaxial behavior, except that
the yield strength was increased to account for strain-rate
effects. Maximum longitudinal reinforcement strain rates
recorded from strain gauges were found to be in the range
of 0.05 to 0.2 (1/s) in Test 1. Based on these rates and data
by Malvar (1998), the yield stress for longitudinal steel was
increased to 1.25 times the measured static yield stress.
The strain-rate-adjusted steel stresses produced calculated
sectional responses that more closely matched experimental
results than did calculations using the static properties. The
modulus of elasticity Es was not modified for strain-rate
effects (Malvar 1998). The strain-hardening ratio was 0.01.
Damping
Damping was introduced as stiffness and mass proportional
Rayleigh damping and was based on the first and second
modes of vibration of the structure. The lowest modes were
used to minimize the generation larger than defined mass
proportional damping. Modal frequencies used for damping
calculation were obtained from eigenvalue analysis updated
at each time increment. The tangent stiffness evaluated at
each time increment was also used in evaluating the damping
matrix. The damping ratio was taken as 2% of critical based
on laboratory test observations (Ghannoum 2007). Updated
tangent stiffness and modal properties were used to minimize the introduction of erroneous damping when the structure sustained large inelastic deformations (Charney 2008;
Petrini et al. 2008; Priestley and Grant 2005).

Table 1Analytical versus experimental frame dynamic properties


Experimental

Analytical

Initial uncracked

Post-half-yield test cracked

Initial uncracked

Post-half-yield test cracked

First period, seconds

0.31

0.35

0.32

0.34

Second period, seconds

0.101

0.115

0.107

0.114

Third period, seconds

0.069

0.08

0.068

0.073

406

ACI Structural Journal/May-June 2012

COMPARISON BETWEEN ANALYSIS AND


EXPERIMENTAL RESULTS
Dynamic properties
Table 1 lists the first three modal periods of the test frame
obtained from low-amplitude snapback tests and from eigenvalue analysis of the analytical model. The test data were

obtained before and after the half-yield dynamic testthat


is, for the essentially uncracked and partially cracked frame.
Analytical values were obtained before and after the analytical model was subjected to the measured base motion from
the half-yield test. The test and analytical data show a slight
lengthening of the period due to the half-yield test, with
excellent agreement between the test and analytical results.

Fig. 6Global response: half-yield test.

Low-to-moderate deformation levelshalf-yield test


The test structure was subjected to horizontal base motion
during the half-yield test and the analytical model was
subjected to the same base motion recorded on the shaketable platform. Figure 6 shows characteristic global response
results. The analytical model faithfully reproduces the main
characteristics of the top- and first-level drift and base shear
histories up to 22 seconds into the motion, after which some
discrepancy in phase and amplitude occurs. The analytical
model adequately represents first-story initial stiffness
and the softening that occurs at higher amplitude motions
(Fig. 7). The maximum first-story drift ratio was 0.46%
experimentally and 0.56% analytically (a 19% difference).
At the maximum analytical drift ratio of 0.46%, the base
shear was 12.3 kips (54.7 kN) experimentally and 13.0 kips
(57.8 kN) analytically (a 5.7% difference). Figure 8 compares
the measured and calculated axial load ratios at the base of
the first-story columns, showing generally excellent correlation. The axial load ratio is defined as the axial load divided
by the column gross section area and concrete compressive
strength. Small discrepancies between the experiment and
analysis in member stiffnessparticularly when flexural
cracking occurs (Fig. 8 and 9), combined with the long duration of the imposed motioncontributed to larger differences observed between the analysis and experiment beyond
the 22-second mark (Fig. 6).
At the column element level, good agreement between the
analysis and experiment is also achieved by the proposed
analytical model. Figure 9 shows that both initial tangent
and postcracking column stiffnesses are reasonably wellmodeled for first-story columns. Analytical and experimental column shears shown in Fig. 10 have less than a 5%
difference at the maximum analytical drift ratio and approximately a 25% difference at the minimum analytical drift
ratio. This plot suggests that the assumed elastic bond stress

Fig. 7First-story horizontal drift ratio versus shear: halfyield test.

of 12

fc psi (1

fc MPa), concrete ultimate tension stress

of 7.5

fc psi (0.62

fc MPa), and concrete linear tension-

Fig. 8First-story horizontal drift ratios versus axial loads: half-yield test.
ACI Structural Journal/May-June 2012

407

Fig. 9First-story critical column horizontal drift ratios versus shear: halfyield test.

Fig. 10Global response Test 1: vertical line indicates shear


failure initiation in column.
softening slope of E0/5 are appropriate parameters for use in
the analytical model.
Figure 11 compares measured and calculated column end
rotations. The rotations were measured over a height h in both
the test and analytical models and include rotations due to
longitudinal bar slip from adjacent anchorage regions (joints
or footings). The moments in Fig. 11 are below the calculated yield moments for all frame columns. Close correlation
between the measured and calculated results is noted.
Unique values of the effective flexural stiffness EI could
not be obtained from the tests because of varying end fixity
and varying axial loads during the dynamic tests. Instead,
the analytical model, which generally produced excellent
agreement with the measured column behavior, was used
to estimate the effective EI. For this purpose, a cantilever
column having length equal to half the column clear length
was subjected to the at-rest gravity loads and the maximum
interstory drift ratio measured during the test. The calculated effective EI was 0.30EIgross, 0.37EIgross, 0.24EIgross, and
0.21EIgross for Columns A1, B1, C1, and D1, respectively
(E = 2800 ksi [19,000 MPa] and Igross is the gross section
moment of inertia).
Large deformation levelsTest 1
To account for initial damage, the analytical model was
subjected to the recorded base motion for the half-yield test
408

and then the recorded base motion for Test 1. Figure 10 shows
characteristic global response results. The analytical model
reproduces with reasonable accuracy the main characteristics of the top- and first-level drift and base shear histories up
to the initiation of shear degradation in Column B1. Mainly
due to discrepancies between when shear failure occurred
in Column B1 and when it was triggered by the analytical
model, larger errors were observed in the analytical results
beyond that shear failure point. The analytical model also
adequately represents the observed first-story elastic stiffness and plastic shear levels (Fig. 12).
At the column element level, good agreement between the
analysis and experiment is achieved by the proposed analytical model up to the initiation of shear failure in Column B1
(Fig. 13). Both elastic stiffness and plastic shear levels are
reasonably estimated by the model, which supports the 25%
increase in the longitudinal steel yield strength for strainrate effects. Figure 14 compares the measured and calculated column end rotations. A close correlation between
the measured and calculated results is noted again up to the
initiation of shear failure in Column B1.
During Dynamic Test 1, Column B1 sustained severe shear
damage and strength degradation, along with gradual but
partial axial failure. Column A1 did not experience apparent
shear or axial degradations, although it did sustain minor
shear cracking. The analytical model, however, showed
shear failures of both Columns A1 and B1, with axial failure
initiation in Column B1. The analytical shear failures of
Columns A1 and B1 occurred a half cycle prior to the experimentally recorded shear failure initiation of Column B1
(Fig. 13). Columns A1 and B1 sustained shear failures in the
analytical model at drift ratios of 2.4% and 2.2%, respectively. These drifts are lower than the experimental drift ratio
at shear failure of Column B1 of 3.2%.
The shear spring attached to Column B1 appears to model
the shear degrading slope of that column fairly well, even
though the failure in the analysis occurred half a cycle prior
to the actual failure. After significant shear degradation, the
measured shear behavior of Column B1 appears to oscillate
randomly about a low residual value owing to the complex
geometry of the failed section and the continually varying
applied axial load. The shear failure model simulates that
behavior by limiting shear strength to a constant residual
shear value.
The axial failure model assumes that column axial failure
initiates when residual shear strength is reached. The
ACI Structural Journal/May-June 2012

Fig. 11Column end rotations versus moments: half-yield test.


axial failure of Column B1 is thus initiated analytically in
Test 1 at the point where it reaches residual shear strength.
In the shake-table test, the initiation of axial failure of
Column B1 corresponded roughly to reaching the residual
shear strength (Ghannoum 2007; Ghannoum and Moehle
2012). Column A1 did not reach the residual shear strength
in the test and did not sustain axial failure.
COLUMN SHEAR CAPACITY MODELS
Strength-based shear capacity models
The shear strengths of the columns with widely spaced
hoops (Column Lines A and B) were evaluated using
equations from ACI 318-08 (ACI Committee 318 2008)
and ASCE/SEI 41-06 (ASCE/SEI Committee 41 2007).
ACI 318-08 shear strength equations that account for the
shear-moment ratio and the effects of axial load were used
(ACI 318-08, Eq. (11-2), (11-5), (11-6), and (11-15)).
The following ASCE/SEI equation was used
Vn = Vs + Vc = k

Av f yt d
s

6[0.5] f

Nu
c
0.8 Ag
+ k
+ 1+
M
6[3.45e-3 ] fcAg
Vd

(4)

(psi [MPa] units)


where k = 1.0 for displacement ductility less than or equal to 2;
k = 0.7 for displacement ductility greater than or equal to 6; k
varies linearly for displacement ductility between 2 and 6; Av
is the area of shear reinforcement in the direction of loading
placed at spacing s; fyt is the measured yield stress of transverse steel; d is the distance from the extreme compression
fiber to the centroid of longitudinal tension reinforcement;
f c is the measured concrete compressive strength; M/Vd is
the largest ratio of moment to shear times effective depth;
ACI Structural Journal/May-June 2012

Fig. 12First-story horizontal drift ratio versus shear:


Test 1 (plotted from time t = 14.5 to 23.5 seconds).
Nu is the axial force normal to the cross section (positive in
compression); and Ag is the gross section area.
Both the ACI and ASCE/SEI equations require the input
of applied shear, moment, and axial loadall of which
varied substantially during Test 1. Thus, the nominal shear
strength for Columns A1 and B1 needed to be evaluated
at each time increment and checked against the applied
shears. The times when Vu first exceeds Vn are indicated in
Fig. 15. The ACI equations produce Vn = 7.92 and 7.76 kips
(35.2 and 34.5 kN) for Columns B1 and A1, respectively,
whereas the ASCE/SEI equation produces Vn = 8.87 and
8.73 kips (39.4 and 38.8 kN). These values compare with
the maximum measured shear forces of 9.87 and 9.48 kips
(43.8 and 42.1 kN) for Columns B1 and A1, respectively.
Thus, both methods produced conservative estimates of
shear strength.
Deformation-based shear capacity models
Deformation-based models aim to estimate the drift at
which shear failure occurs. Drift estimates derived from
409

Fig. 13Test 1: Horizontal drift ratios versus shear of Columns A1


and B1.

Fig. 14Test 1: Column end rotations versus moments. (X


indicates instrumentation failure).

Fig. 15 Test 1: Shear capacity models shear failure estimates.


models presented by Elwood and Moehle (2005), Kato and
Ohnishi (2002), Pujol et al. (1999), and Sasani (2007) are
compared with measured drifts in the following paragraphs.
Elwood and Moehle (2005) proposed an empirical, regression-based, shear-drift model based on results from 50 lightly
confined columns sustaining shear failure following flexural yielding. This same model was implemented in the
OpenSees simulations presented previously in this paper.
410

This model is based on observations that the drift ratio at


shear failure (Ds/L) is positively correlated with transverse
reinforcement ratio rt and negatively correlated with the
shear-stress ratio (u/ fc ) and axial load ratio (P/Ag f c).
Shear failure is defined when the shear resistance capacity
drops to 80% of the maximum shear. The model, which is
intended to produce a mean response estimate, is given by
ACI Structural Journal/May-June 2012

s
3
1

1 P
1
=
+ 4t

L 100
500 [ 42 ] fc 40 Ag fc 100

Sasani (2007) used a Bayesian parameter estimation technique to develop a probabilistic drift capacity model for RC
columns failing in shear. Sasani (2007) observed that the
drift ratio DR at shear failure increased with an increasing
volumetric transverse reinforcement ratio rs and shear spansection height ratio a/h and decreased with an increasing
axial load ratio (ho = P/Ag f c). Sasani (2007) also differentiated between the case of single and double curvature column
deformations through the factor q, which equals 1.0 for
columns deformed in single curvature and 0.85 for columns
deformed in double curvature. Shear failure is defined when
the shear resistance capacity drops to 80% of the maximum
shear. The model, which is intended to produce a mean
response estimate, is given by

(5)

(psi [MPa] units)

Kato and Ohnishi (2002) proposed that the plastic drift


capacity can be estimated based on the maximum edge strain
in the core concrete, axial load ratio, and cross-sectional
dimensions. The total drift ratio is defined by summing the
drift ratio at yield of longitudinal reinforcement Dy/L and the
calculated plastic drift ratio Dp/L. Shear failure is defined
when the shear resistance capacity drops to 80% of the
maximum shear. The model, which is intended to produce a
mean response estimate, is given by
s y p
=
+
L
L
L

0.18
DR = 0.74
s o

(6)

(0 < e

<1

( 13 e

<2

(7)

where D is the depth of the gross cross section; je is the


depth of the core; ecp is the strain at the maximum stress for
the core; m is the ratio of the concrete strain at the edge of
the core concrete to ecp at shear or axial failure (determined
empirically for shear [m = 2.3] and axial [m = 3.6] failures);
and en is an equivalent axial load ratio as defined in Kato and
Ohnishi (2002).
Pujol et al. (1999) proposed an empirical model based on
results from 92 columns. The model is based on observations
that the maximum drift ratio at shear failure (DMAX/L) tended
to increase with an increasing aspect ratio (a/d, where a is a
column half-span) and increasing transverse reinforcement
index (rt fyt /vMAX, where vMAX is the column shear stress at
shear failure = shear force at failure/bwd). Shear failure is
defined when the shear resistance capacity drops to 80%
of the maximum shear. The model, which is intended to
produce a lower-bound response estimate, is given by
t f yt a a
MAX
100 =
,4
L
vMAX d d

(9)

Figure 15 overlays the various model shear failure initiation points on the test results of Columns A1 and B1 for
Test 1. Table 2 lists the measured and calculated drift ratios
at shear failure and the ratios of calculated-to-measured
drift ratios. The measured drift ratio at shear failure is
defined as the absolute value of the drift ratio at which shear
resistance drops to 80% of the maximum shear. Note that
Column A1 sustained shear cracking for the largest positive
displacement excursion, although reversal of deformations at
the peak apparently protected the column from shear failure.
In contrast, nominal shear failure for Column B1 occurred
during the subsequent negative drift excursion, resulting in
clearly visible shear distress in the column (refer to Ghannoum and Moehle [2012]).
All four models produced conservative estimates of drifts
at shear-failure initiation. The models of Kato and Ohnishi
(2002) and Sasani (2007) produced the most accurate estimates of the nominal shear-failure drifts. The Kato and
Ohnishi (2002) model requires the input of drift ratio at yield
which, for this study, was taken directly from data measured
during the test, whereas the results for the Sasani (2007)
model were based entirely on data external to the shaketable test.

where

m cp 2
D

je 3en
p
=
L m cp
4

D j 2 3 5en 3
e

a
h

CONCLUSIONS
A detailed analytical model was developed to simulate the
dynamic behavior of a three-story, three-bay, RC frame up
to the point of severe shear-strength degradation and axial
failure of critical columns. A new zero-length fiber-section
representation of longitudinal bar slip from anchorage
regions was introduced. Good correlation between measured
and calculated load-deformation behaviors was observed at

(8)

Table 2Calculated versus measured drift ratios at nominal shear failure*


Column A1

Column B1

Measured or calculated

Calculated/measured

Measured or calculated

Calculated/measured

Measured

3.65

4.66

Elwood and Moehle (2005)

2.39

0.65

2.18

0.47

Kato and Ohnishi (2002)

3.39

0.93

3.90

0.84

Pujol et al. (1999)

1.90

0.52

1.82

0.39

Sasani (2007)

3.89

1.07

4.00

0.86

Drift ratio at nominal shear failure is absolute value of drift ratio at which shear resistance capacity drops to 80% of maximum shear.

From Test 2.

ACI Structural Journal/May-June 2012

411

the frame system and element levels. The model was capable
of simulating the important aspects of the measured dynamic
response from low deformation levels to well beyond flexural yielding. The following observations and recommendations are derived from this analytical exercise:
1. Force formulation fiber-section elements and zero-length
fiber-section bar-slip models can reproduce RC frame element
flexural behavior with high fidelity, as they are able to adapt
to variations in frame element boundary conditions (that is,
end restraints and axial loads). Such analytical models are
relatively simple to implement, as they only require section
geometry and material properties as input variables.
2. A bilinear longitudinal bar stress-versus-slip relation that
is based on a bi-uniform anchorage bond stress provided a
good basis for bar-slip-induced rotation modeling. An elastic
anchorage bond stress of 12

fc psi (1

fc MPa) and asso-

ciated plastic bond stress of 3.24 fc psi (0.27 fc MPa)


produced adequate estimates of bar-slip-induced rigid body
rotations in frame elements.
3. A concrete material ultimate tension stress of 7.5 fc psi
(0.62 fc MPa) and linear tension-softening slope of E0/5
(where E0 is the concrete material tangent modulus at zero
strain) produced adequate estimates of column flexural
cracking moments and softening.
4. The longitudinal steel yield stress should be increased
appropriately in cases where high strain rates are anticipated.
5. Currently available column strength and deformationbased shear-failure models produced shear-failure estimates
with large scatter. The models studied herein produced
conservative estimates.
6. The test frame response could not be modeled with
high fidelity past shear-failure initiation of critical columns,
although the failure sequence of critical columns was
correctly modeled. Further refinement in available shearfailure models is required.
ACKNOWLEDGMENTS
This work was supported in part by the Pacific Earthquake Engineering
Research Center (PEER) through the Earthquake Engineering Research
Centers Program of the National Science Foundation under Award No.
EEC-9701568 and by NSF Award No. 0618804. The opinions, findings,
and recommendations expressed in this paper are those of the authors
and do not necessarily reflect those of NSF. The laboratory tests were
conducted in the research laboratories of PEER at the University of
California, Berkeley, Berkeley, CA.

REFERENCES
ACI Committee 318, 2008, Building Code Requirements for Structural
Concrete (ACI 318-08) and Commentary, American Concrete Institute,
Farmington Hills, MI, 473 pp.
ASCE/SEI Committee 41, 2007, Seismic Rehabilitation of Existing
Structures (ASCE/SEI 41/06), American Society of Civil Engineers,
Reston, VA, 428 pp.
Berry, M. P., 2006, Performance Modeling Strategies for Modern
Reinforced Concrete Bridge Columns, PhD dissertation, University of
Washington, Seattle, WA.
Charney, F. A., 2008, Unintended Consequences of Modeling Damping
in Structures, Journal of Structural Engineering, ASCE, V. 134, No. 4,
pp. 581-592.
Elwood, K. J., 2004, Modeling Failures in Existing Reinforced Concrete
Columns, Canadian Journal of Civil Engineering, V. 31, pp. 846-859.
Elwood, K. J., and Eberhard, M. O., 2009, Effective Stiffness of
Reinforced Concrete Columns, ACI Structural Journal, V. 106, No. 4,
July-Aug., pp. 476-484.

412

Elwood, K. J., and Moehle, J. P., 2005, Drift Capacity of Reinforced


Concrete Columns with Light Transverse Reinforcement, Earthquake
Spectra, V. 21, No. 1, pp. 71-89.
Ghannoum, W. M., 2007, Experimental and Analytical Dynamic
Collapse Study of a Reinforced Concrete Frame with Light Transverse
Reinforcements, PhD dissertation, University of California, Berkeley,
Berkeley, CA, 362 pp.
Ghannoum, W. M., and Moehle, J. P., 2012, Shake-Table Tests of
a Concrete Frame Sustaining Column Axial Failures, ACI Structural
Journal, V. 109, No. 3, May-June, pp. 393-402.
Ibarra, L. F.; Medina, R. A.; and Krawinkler, H., 2005, Hysteretic
Models that Incorporate Strength and Stiffness Deterioration, Earthquake
Engineering & Structural Dynamics, V. 34, No. 12, pp. 1489-1511.
Kato, D., and Ohnishi, K., 2002, Axial Load Carrying Capacity of R/C
Columns under Lateral Load Reversals, The Third U.S.-Japan Workshop on
Performance-Based Earthquake Engineering Methodology for Reinforced
Concrete Building Structures, Seattle, WA, pp. 247-255.
Lehman, D., and Moehle, J. P., 2000, Seismic Performance of WellConfined Concrete Bridge Columns, 1998/01, UCB/PEER, University of
California, Berkeley, Berkeley, CA.
Lowes, L. N.; Mitra, N.; and Altoontash, A., 2003, A Beam-Column
Joint Model for Simulating the Earthquake Response of Reinforced
Concrete Frames, 2003/10, UCB/PEER, University of California,
Berkeley, Berkeley, CA, 59 pp.
Malvar, L. J., 1998, Review of Static and Dynamic Properties of
Steel Reinforcing Bars, ACI Materials Journal, V. 95, No. 5, Sept.-Oct.,
pp. 609-616.
Mander, J. B.; Priestley, M. J. N.; and Park, R., 1988, Theoretical StressStrain Model for Confined Concrete, Journal of Structural Engineering,
ASCE, V. 114, No. 8, pp. 1804-1826.
McKenna, F.; Fenves, G. L.; Scott, M. H.; and Jeremie, B., 2000, Open
System for Earthquake Engineering Simulation, OpenSees, Berkeley, CA.
Menegotto, M., and Pinto, P. E., 1973, Method of Analysis for Cyclically Loaded Reinforced Concrete Plane Frames Including Changes in
Geometry and Nonelastic Behavior of Elements under Combined Normal
Force and Bending, Proceedings of IABSE Symposium on Resistance and
Ultimate Deformability of Structures Acted on by Well-Defined Repeated
Loads, Lisbon, Portugal.
Monti, G., and Spacone, E., 2000, Reinforced Concrete Fiber Beam
Element with Bond-Slip, Journal of Structural Engineering, ASCE,
V. 126, No. 6, pp. 654-661.
Otani, S., 1974, Inelastic Analysis of R/C Frame Structures, Journal of
the Structural Division, ASCE, V. 100, pp. 1433-1449.
Petrini, L.; Maggi, C.; Priestley, M. J. N.; and Calvi, G. M., 2008, Experimental Verification of Viscous Damping Modeling for Inelastic Time
History Analyzes, Journal of Earthquake Engineering, V. 12, pp. 125-145.
Priestley, M. J. N., and Grant, D. N., 2005, Viscous Damping in Seismic
Design and Analysis, Journal of Earthquake Engineering, V. 9, No. 2,
pp. 229-255.
Pujol, S.; Ramirez, J. A.; and Sozen, M. A., 1999, Drift Capacity of
Reinforced Concrete Columns Subjected to Cyclic Shear Reversal, Seismic
Response of Concrete Bridges, SP-187, American Concrete Institute, Farmington Hills, MI, pp. 255-274.
Sasani, M., 2007, Life-Safety and Near-Collapse Capacity Models for
Seismic Shear Behavior of Reinforced Concrete Columns, ACI Structural
Journal, V. 104, No. 1, Jan.-Feb., pp. 30-38.
Scott, B. D.; Park, R.; and Priestley, M. J. N., 1982, Stress-Strain
Behavior of Concrete Confined by Overlapping Hoops at Low and High
Strain Rates, ACI JOURNAL, Proceedings V. 79, No. 1, Jan.-Feb., pp. 13-27.
Sezen, H., and Moehle, J. P., 2006, Seismic Tests of Concrete Columns
with Light Transverse Reinforcement, ACI Structural Journal, V. 103,
No. 6, Nov.-Dec., pp. 842-849.
Sezen, H., and Setzler, E. J., 2008, Reinforcement Slip in Reinforced
Concrete Columns, ACI Structural Journal, V. 105, No. 3, May-June,
pp. 280-289.
Spacone, E.; Filippou, F. C.; and Taucer, F. F., 1996, Fibre Beam-Column
Model for Non-Linear Analysis of R/C Frames: Part I. Formulation, Earthquake Engineering & Structural Dynamics, V. 25, No. 7, pp. 711-725.
Zhao, J., and Sritharan, S., 2007, Modeling of Strain Penetration Effects
in Fiber-Based Analysis of Reinforced Concrete Structures, ACI Structural
Journal, V. 104, No. 2, Mar.-Apr., pp. 133-141.

ACI Structural Journal/May-June 2012

Potrebbero piacerti anche