ACI SP 297 Nonlinear Modeling Parameters and Acceptance Criteria

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 210

SP-297—1

Nonlinear modeling parameters and acceptance criteria for concrete columns


Ghannoum, W. M., Matamoros, A. B.

Abstract
A database of 490 pseudo-static tests of reinforced concrete columns subjected to load reversals
was used to evaluate nonlinear modeling parameters that define the lateral force versus lateral
deformation envelope relation of columns under seismic excitations. Based on the modeling
parameters, criteria that identify acceptable deformation levels at various performance objectives
are proposed. The effects of bi-directional loading and number-of-cycles of the displacement
history on the drift ratio at axial failure are discussed, and recommendations are given to account
for such effects. Modeling parameters and acceptance criteria are provided in a format that is
consistent with provisions of the ASCE 41-06 Standard entitled “Seismic Rehabilitation of
Existing Structures”.

Keywords: concrete, columns, nonlinear, modeling parameters, acceptance criteria, bi-


directional, cyclic, seismic

1. Introduction
When estimating the performance of existing structures subjected to seismic events, it is
often necessary to conduct nonlinear dynamic analyses that require the definition of the lateral
force versus lateral deformation relation of frame members. Based on a database comprised of
490 reinforced concrete column tests [1, 2], relations are proposed to define key limiting
deformations in the lateral force-deformation relation of reinforced concrete columns subjected
to seismic demands. More specifically, relations for evaluating plastic rotations at incipient
lateral-strength degradation and incipient axial degradation are proposed. Furthermore,
recommendations are provided for selecting acceptable deformations or Acceptance Criteria
(AC) below which the performance of reinforced concrete columns is deemed acceptable for
selected target performance objectives.
The proposed relations and recommendations are given in a format that is compatible with
the standard for seismic rehabilitation ASCE 41-06 Supplement 1 [3, 4]; hereafter referred to as
ASCE 41. The standard defines the nonlinear force-deformation backbone relations of concrete
columns and other members as illustrated in Fig.1a. In the Figure, the plastic rotation at
incipient lateral-strength degradation is given through the Modeling Parameter (MP) a. The
plastic rotation in a concrete column at incipient axial degradation is given through the MP b.
The residual lateral strength of a column is given by the MP c. MP a and b are given in the
current standard as conservative lower-bound estimates of experimental values (Fig. 1b) [5]. To
avoid skewing the results of nonlinear simulations, the proposed MP a and b will target median
experimental values. AC in the standard were based on fixed percentages of the MP for various
performance objectives. Because the error on the estimates of MP exhibit different dispersions
for various parameters and members, selecting a fixed fraction of those MP values for AC
results in varying probabilities of exceedance for the AC (e.g., 75% of a MP does not always
provide the same probabilities of exceedance because that probability depends on the standard

1.1
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

deviation on the MP estimate) . Thus, proposed AC are defined through fixed probabilities of
exceedance for various performance objectives and the corresponding factions of MP values
that achieve those probabilities are given.

Lower bound “b”


M
Lower bound “a”

Rotation
a) b)
Figure 1: a) ASCE 41-06 backbone response for nonlinear modeling of RC components (from [3]);
b) illustration of MP deviation from median values and resulting backbone for modeling RC elements.

2. Column Database
The database used to support the development of the proposed MP and AC contains 319
rectangular column tests and 171 circular column tests for a total of 490 tests [1, 2]. Much of
the data was derived from the PEER column database [6]. All tests in the database were
conducted quasi-statically. The new database is webcast and accessible to the public ([1, 2])
and additional information about the database can be found in Sivaramakrishnan [7]. The
distributions of key parameters of the columns in the data sets are illustrated in the bar charts of
Fig. 2. Out of the rectangular columns in the database, 37 can be considered to satisfy the
requirements of ACI 318-11 [8] for Special Moment Resisting Frames (SMRF). Out of the
circular columns, 24 can be considered to satisfy these requirements. A limited number of
rectangular columns (25 out of 171) have reported ties with 90o hooks. The tie hook-angle is
unknown for 26 rectangular columns, while 269 rectangular columns have ties with 135o hooks
or welded ends. A limited number of circular columns had ties with lapped ends (13 out of
171). Tie details were unknown for one circular column, while the remainder of the circular
columns had spirals, ties with welded ends, or hooks anchored into the core. Limited
information on hook extensions was available for column tests of the database. Splices and
anchorage deficiencies are not within the scope of the work presented, and as such no tests on
columns containing anchorage deficiencies were used.

1.2
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

90 140
Rectangular Rectangular
80 Circular Circular
120
70
100
Number of Columns

Number of Columns
60

50 80

40 60
30
40
20
20
10

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.25 0.5 0.75 1 1.25 1.5 1.75
Axial Load Ratio s/d

90
Rectangular
Circular
80
90
Rectangular
70 80 Circular

60 70
Number of Columns

Number of Columns

60
50

50
40

40
30
30
20
20

10 10

0 0
0 0.0025 0.005 0.0075 0.01 0.0125 0.015 0.0175 0.02 0.0225 0.025 0.0275 0.03 0.0325 0.035 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Transverse Reinforcement Ratio Vy/Vo

Figure 2: Distribution of key parameters in column database – see Notation section for term definitions
3. Data Extraction
The values of a and b were extracted for all column tests. The a values were taken as: a =
(Δ0.8-Δy)/L with Δ0.8 = lateral drift at which the lateral strength of an element degrades by 20%
from peak, Δy = lateral drift at onset of significant inelastic deformations, and L = column clear
length. The drift at onset of inelastic deformation, Δy, was obtained as recommended by Sezen
and Moehle [9]. A secant line was extended on the lateral force-deformation plot of a column
test from the origin to the point on the backbone curve at 70% of the maximum shear (0.7 Vmax).
Δy was then taken as the drift at the intersection of the secant line with the horizontal line drawn
at Vmax. Δy was also evaluated using a secant line passing through 0.6 Vmax. Little difference was
observed between results from the two intercepts. Results based on a 0.7 Vmax intercept were
used in parameter extractions.
The drifts at yield obtained as described above differ from estimates that would be obtained
using column stiffness relations provided in the ASCE 41 standard. ASCE 41 stiffness relations
were not used for extracting paraerters a and b to avoid skewing the extracted plastic rotations by

1.3
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

errors in stiffness estimates inherent in the ASCE 41 stiffness relations [10] (especially for
columns with low deformation capacities). The objective of the study was to extract the best
estimate of plastic rotations a and b such that the proposed relations defining them would
provide median estimates of the plastic rotations and the trends described in those relations
would not be artificially altered. The outcome for users of the standard that opt to use ASCE 41
stiffness relations will be a slightly skewed total deformation capacity for elements, but the
estimated behavior of the elements would not be skewed (i.e., the degree of inelastic
deformations prior to loss of strength estimated by the relations would be the median estimate).
Moreover, the proposed relations are intended to be used not only with the ASCE 41 elastic
stiffness relations but also with other methods for evaluating stiffness (such as fiber-section
models).
Due to the scarcity of column tests conducted to collapse, two sets of b values were produced.
The first set originates from column tests that were conducted to axial collapse. If a test was
conducted to axial collapse, the plastic rotation at axial failure was taken as b1 = (Δaxial-Δy)/L;
with Δaxial = drift at onset of axial collapse. Only 36 rectangular and 9 circular columns in the
database were pushed to axial failure. The webcast database was further bolstered by 12 recent
rectangular-column collapse tests [11, 12]. The second set of b values, b2, includes the first set
but also introduces b values for all other columns in the database based on the following:
- If a test reached a drift at which the lateral strength degraded to 25% of the peak but no
axial failure was reported, the plastic rotation at that drift was taken as b2. This limit on
deformations at axial failure was introduced to mitigate possible errors or omissions in
the reporting of axial failure in database tests. Columns that lost 75% of their lateral
strength were deemed unstable and close to axial collapse.
- If a columns was not tested to collapse or to a deformation causing a reduction in lateral
strength to below 25% of the peak lateral strength, the plastic rotation at the largest drift
a column was pushed was taken as b2.
This methodology provides a lower bound estimate on the parameter b but utilizes all the
available column tests. Additional details about the process by which the plastic rotations were
extracted from the database can be found in [7].
Values of Vy were extracted from experimental data through analytical means. In all
calculations, measured material properties were used. Vy is defined as the shear demand
corresponding to the development of moment strength in a column. Vy = My/La with My = column
moment capacity, and La = column shear span. My was calculated using fiber-section analyses
that utilized the parabolic stress-strain relation for concrete proposed by Hognestad [13] for
concrete in compression. The limiting strain in compression for the concrete was adopted as
0.003. Confinement effects on concrete material properties, as well as the tensile strength of
concrete were neglected. An elastic perfectly plastic material model was used for steel fibers and
capped at the measured yield stress of longitudinal bars.
ACI 318-11 specifies the use of the ultimate stress of longitudinal bars (approximated as 1.25
the specified yield stress) when evaluating the flexural strength of plastic hinges in flexural
members. The increment of 1.25 on yield stress is an arbitrary limit to ensure that an upper
bound estimate of the shear demand is used for proportioning the transverse reinforcement. In
the study that was performed, measured values of yield stress were used in calculating Vy for
following reasons:

1.4
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

1) In reinforced concrete columns, longitudinal bars will only reach ultimate stress in well
confined columns that are pushed to large inelastic deformations. For columns that
sustain shear failure prior to flexural yielding or at relatively low inelastic deformations,
evaluating Vy using the yield stress of bars is more appropriate.
2) Deriving the modeling parameter and acceptance criteria relations using lower Vy values
obtained using the yield stress of bars would provide lower modeling parameters and
acceptance criteria should users opt to evaluate Vy using 1.25 the yield stress of
longitudinal bars.
The beneficial effects of confinement on flexural strength were not included in calculations of
Vy for similar reasons.

Vo was evaluated based on the shear-strength equation of ASCE 41-06 for reinforced concrete
columns. The shear strength equation used was:

Φ ⁄
1 0.8 Eq. 1

where α = 6 in psi units and 0.5 in MPa units; Nu = axial compression force (= 0 for tension
force); M/Vd is the ratio of moment to shear times the effective depth and was bounded by the
values of 4 and 2; d is the effective depth; and Ag is the gross cross-sectional area of the column.
For circular columns d = 0.8D, where D is the diameter of the column. To account for the
inefficacy of transverse reinforcement in resisting shear when spaced beyond 75% of the
effective depth of a section, the steel contribution to shear strength (first term in Eq. 1) was
modified by a factor of Φ that was taken as 1.0 for s/d ≤ 0.75, zero for s/d ≥ 1.0, and linearly
interpolated between the two values of s/d.

4. Regression-Based Modeling Parameters (MP)


4.1 Relations
Linear regressions using the most influential parameters were conducted to calculate a median
estimate for parameters a and b. In a dataset that has a reasonable spread across parameters, a
linear regression plane should intersect the data at around the median (i.e., the regression plane
should divide the data into two approximately equal groups, with half the data points above and
half the data points below the plane). This was found to be the case for this database. Only minor
adjustments on the intercept of the regression equations were necessary to achieve a median fit.
Regression equations rather than a tabulated format were selected to avoid the problem of
stepping functions at parameter boundaries.
Most rectangular columns in the database contained ties with 135o hooks or welded ties with
the welded cases accounting for a small minority of tests. Most circular columns in the database
were spirally reinforced. A handful of circular columns had welded or core-anchored ties that are
considered to perform similar to spirals. In the following, the “spirally” reinforced column group
was considered to contain circular columns reinforced with welded circular ties and those with
circular ties adequately anchored in the core.

1.5
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

Regression analyses indicated that spirally reinforced circular columns responded differently
to variations in influential parameters than rectangular columns reinforced with rectangular ties.
To account for such differences, separate regression equations were developed for spirally
reinforced circular columns than for all other columns. The most influential parameters for
rectangular and circular columns were: the axial load ratio, the transverse reinforcement ratio
(ρt), and the ratio Vy/Vo. All three parameters are currently used in the standard ASCE 41 to
estimate plastic rotation limits for concrete columns. The maximum shear stress applied to a
column section is also used in the standard as a parameter to estimate MP. That parameter was
not found to have a determining role in column responses analyzed in this work.
The proposed relations between plastic rotations at incipient loss of lateral strength and axial
failure are given below. The subscripts R and C in the equations indicate plastic rotations for
rectangular and circular columns, respectively. Only relations for b2 were derived using linear
regression because there are too few b1 values to provide meaningful regression-based estimates.
Plastic rotation capacities for columns other than spirally reinforced circular columns
P Vy
a R  0.042  0.043  0.63 t  0.023  0.0 (rad) Eq.2
Ag f c' Vo

P Vy
b2 R  0.051  0.051 '
 1.3 t  0.023  a R (rad) Eq.3
Ag f c Vo

Plastic rotation capacities for columns for spirally reinforced circular columns
P Vy
aC  0.06  0.058 '
 1.3 t  0.037  0.0 (rad) Eq.4
Ag f c Vo

P Vy
b2C  0.064  0.07  2.85  t  0.03  a C (rad) Eq.5
Ag f c' Vo

When evaluating the above parameters, ρt should not be taken greater than 0.0175 and Vy/Vo
should not be taken smaller than 0.2. The equation is not applicable for ρt ≤0.0005. An upper
bound on the transverse reinforcement ratio of 0.0175 is prescribed because few columns in the
database contained a ratio exceeding that limit. Equations for modeling parameters cannot be
used for columns with a transverse reinforcement ratio below 0.0005 because the equations are
not intended for unreinforced columns. A lower limit on Vy/Vo of 0.2 is prescribed because few
columns in the database had lower values of Vy/Vo.
Values of a and b2 for rectangular and spirally reinforced column are presented in Table 1 at
practical parameter boundaries. In Table 1, the notation is consistent with that used in Eq. 2 to 5
in which the subscript r is used in reference to rectangular columns while the subscript c is used
in reference to circular columns.. As can be observed in the table, for a given set of parameters,
spirally reinforced columns exhibited significantly larger deformation capacities than rectangular
columns.

1.6
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

Table 1: Values of a and b at parameter boundaries


P/(Agf’c) ρt Vy/Vo aR (rad) aC (rad) b2R (rad) b2C (rad)
0 0.0005 0.2 0.038 0.058 0.047 0.060
0 0.0005 2.0 0.0 0.0 0.006 0.005
0 0.0175 0.2 0.048* 0.071* 0.069* 0.108*
0 0.0175 2.0 0.007 0.0 0.028 0.054
0.7 0.0005 0.2 0.008 0.016 0.011 0.016
0.7 0.0005 2.0 0.0 0.0 0.0 0.0
0.7 0.0175 0.2 0.018 0.029 0.033 0.059
0.7 0.0175 2.0 0.0 0.0 0.0 0.005
* Maximum permissible values
4.2 Analyses of Fit
4.2.1 Estimates of a
The cumulative distribution of the error difference between experimental and regression
estimates of a are plotted for rectangular and circular columns in Fig. 3. Errors are plotted for
estimates evaluated using the proposed regression equations as well as for those estimated using
the ASCE 41 standard. As can be observed in the figures, the proposed regression equations
shifted estimates from conservative ones based on the standard to median estimates; i.e., error =
0 at 0.5 probability of exceedance. For aR, both methods produced similar spread on the error as
evidenced by the similar slopes of the cumulative distribution curves and standard deviations
presented in Fig. 3. For aC however, the proposed relation produced a large shift in estimates
from those derived using the standard. This is not surprising given that : 1) the table in the
standard defining the a values for reinforced concrete columns was based only on data from
rectangular column tests [5], and 2) spirally reinforced circular columns in the data set showed
increased deformation capacity over rectangular columns with similar parameters.
1 1

0.9 0.9
Equation estimates
0.8 0.8
ASCE 41-06 estimates
Equation estimates
0.7 0.7
Cumulative Distribution

Cumulative Distribution

ASCE 41-06 estimates


0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3
Number of Tests = 319 Number of Tests = 171
0.2 0.2
Standard Deviation Equation = 0.015 Standard Deviation Equation = 0.019
0.1 0.1
Standard Deviation ASCE41-06 = 0.015 Standard Deviation ASCE41-06 = 0.022
0 0
-0.04 -0.02 0 0.02 0.04 0.06 0.08 -0.04 -0.02 0 0.02 0.04 0.06 0.08
Error in aR = Experiment - Estimate (rad) Error in aC = Experiment - Estimate (rad)

Figure 3: Cumulative distribution of the error between experimental values and estimates for aR and aC

1.7
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

The cumulative distribution of the error difference between experimental and regression
estimates of aR are plotted for various bins of rectangular columns in Fig. 4. Test data are divided
into bins to illustrate the fit of the proposed relation in various parameter quadrants. As can be
observed in the figure, the proposed relation produced estimates that were very close to the
median in all bins. Standard deviations on the error of the proposed equation vary from 0.005 to
0.018; which indicates variable accuracy of the proposed equation across bins. Figure 5 presents
similar data as Fig. 4 but for circular columns. Similar conclusions can be drawn from Fig. 5 for
spirally reinforced circular columns, although the estimates from the proposed equation deviate
conservatively from the median estimate for columns with low transverse reinforcement ratio
and intermediate values of Vy/Vo (ranging between 0.6 and 1.0). The standard deviations were
slightly higher for circular columns than rectangular ones and ranged from 0.005 to 0.029 across
bins.

Axial Load Ratio  0.2, Transv. Reinf. Ratio  0.006, Vy/VO  0.6 Axial Load Ratio>0.2, Transv. Reinf. Ratio  0.006, Vy/VO  0.6
1 1

0.5 0.5
Number of Tests = 21 Number of Tests = 21
0 Standard Deviation = 0.015 0 Standard Deviation = 0.018
-0.04 -0.02 0 0.02 0.04 0.06 -0.04 -0.02 0 0.02 0.04 0.06
Axial Load Ratio  0.2, Transv. Reinf. Ratio  0.006, 0.6<Vy/VO  1.0 Axial Load Ratio>0.2, Transv. Reinf. Ratio  0.006, 0.6<Vy/VO  1.0
1 1

0.5 0.5
Number of Tests = 39 Number of Tests = 34
0 Standard Deviation = 0.016 0 Standard Deviation = 0.01
-0.04 -0.02 0 0.02 0.04 0.06 -0.04 -0.02 0 0.02 0.04 0.06
Axial Load Ratio  0.2, Transv. Reinf. Ratio  0.006, Vy/VO>1.0 Axial Load Ratio>0.2, Transv. Reinf. Ratio  0.006, Vy/VO>1.0
1 1

0.5 0.5
Number of Tests = 26 Number of Tests = 18
Standard Deviation = 0.0084 Standard Deviation = 0.005
Cumulative Distribution

0 0
-0.04 -0.02 0 0.02 0.04 0.06 -0.04 -0.02 0 0.02 0.04 0.06
Axial Load Ratio  0.2, Transv. Reinf. Ratio>0.006, Vy/VO  0.6 Axial Load Ratio>0.2, Transv. Reinf. Ratio>0.006, Vy/VO  0.6
1 1

0.5 0.5
Number of Tests = 47 Number of Tests = 95
0 Standard Deviation = 0.016 0 Standard Deviation = 0.016
-0.04 -0.02 0 0.02 0.04 0.06 -0.04 -0.02 0 0.02 0.04 0.06
Axial Load Ratio  0.2, Transv. Reinf. Ratio>0.006, 0.6<Vy/VO  1.0 Axial Load Ratio>0.2, Transv. Reinf. Ratio>0.006, 0.6<Vy/VO  1.0
1 1

0.5 0.5
Number of Tests = 11 Number of Tests = 6
0 Standard Deviation = 0.018 0 Standard Deviation = 0.016
-0.04 -0.02 0 0.02 0.04 0.06 -0.04 -0.02 0 0.02 0.04 0.06
Axial Load Ratio  0.2, Transv. Reinf. Ratio>0.006, Vy/VO>1.0 Axial Load Ratio>0.2, Transv. Reinf. Ratio>0.006, Vy/VO>1.0
1 1
Proposed Equation
0.5 ASCE 41-06 Sup 1 Table 0.5
Number of Tests = 1 Number of Tests = 0
0 Standard Deviation = 0 0
-0.04 -0.02 0 0.02 0.04 0.06 -0.04 -0.02 0 0.02 0.04 0.06
Error in aR = Experiment - Estimate (rad) Error in aR = Experiment - Estimate (rad)

Figure 4: Cumulative distributions of the error between experimental values and estimates for aR; data
split into bins covering various ranges of parameters and standard deviations given for equation estimates
only.

1.8
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

Axial Load Ratio  0.2, Transv. Reinf. Ratio  0.006, Vy/VO  0.6 Axial Load Ratio>0.2, Transv. Reinf. Ratio  0.006, Vy/VO  0.6
1 1

0.5 0.5
Number of Tests = 40 Number of Tests = 8
0 Standard Deviation = 0.027 0 Standard Deviation = 0.018
-0.04 -0.02 0 0.02 0.04 0.06 0.08 -0.04 -0.02 0 0.02 0.04 0.06 0.08
Axial Load Ratio  0.2, Transv. Reinf. Ratio  0.006, 0.6<Vy/VO  1.0 Axial Load Ratio>0.2, Transv. Reinf. Ratio  0.006, 0.6<Vy/VO  1.0
1 1

0.5 0.5
Number of Tests = 27 Number of Tests = 5
0 Standard Deviation = 0.015 0 Standard Deviation = 0.0046
-0.04 -0.02 0 0.02 0.04 0.06 0.08 -0.04 -0.02 0 0.02 0.04 0.06 0.08
Axial Load Ratio  0.2, Transv. Reinf. Ratio  0.006, Vy/VO>1.0 Axial Load Ratio>0.2, Transv. Reinf. Ratio  0.006, Vy/VO>1.0
1 1

0.5 0.5
Number of Tests = 46 Number of Tests = 8
0 Standard Deviation = 0.0099 0 Standard Deviation = 0.0094
Cumulative Distribution

-0.04 -0.02 0 0.02 0.04 0.06 0.08 -0.04 -0.02 0 0.02 0.04 0.06 0.08
Axial Load Ratio  0.2, Transv. Reinf. Ratio>0.006, Vy/VO  0.6 Axial Load Ratio>0.2, Transv. Reinf. Ratio>0.006, Vy/VO  0.6
1 1

0.5 0.5
Number of Tests = 13 Number of Tests = 20
0 Standard Deviation = 0.029 0 Standard Deviation = 0.018
-0.04 -0.02 0 0.02 0.04 0.06 0.08 -0.04 -0.02 0 0.02 0.04 0.06 0.08
Axial Load Ratio  0.2, Transv. Reinf. Ratio>0.006, 0.6<Vy/VO  1.0 Axial Load Ratio>0.2, Transv. Reinf. Ratio>0.006, 0.6<Vy/VO  1.0
1 1
Proposed Equation
0.5 0.5 ASCE 41-06 Sup 1 Table
Number of Tests = 3 Number of Tests = 1
0 Standard Deviation = 0.024 0 Standard Deviation = 0
-0.04 -0.02 0 0.02 0.04 0.06 0.08 -0.04 -0.02 0 0.02 0.04 0.06 0.08
Axial Load Ratio  0.2, Transv. Reinf. Ratio>0.006, Vy/VO>1.0 Axial Load Ratio>0.2, Transv. Reinf. Ratio>0.006, Vy/VO>1.0
1 1

0.5 0.5
Number of Tests = 0 Number of Tests = 0
0 0
-0.04 -0.02 0 0.02 0.04 0.06 0.08 -0.04 -0.02 0 0.02 0.04 0.06 0.08
Error in aC = Experiment - Estimate (rad) Error in aC = Experiment - Estimate (rad)

Figure 5: Cumulative distributions of the error between experimental values and estimates for aC; data
split into bins covering various ranges of parameters and standard deviations given for equation estimates
only.

The effects of hook details of transverse reinforcement are explored in Fig. 6. In this figure,
the error between experimental values and equation-based estimates of a are plotted versus s/d.
Rectangular columns with ties having 135o hooks and 90o hooks are highlighted. Columns with
unknown hook details are also indicated. Spirally reinforced circular columns as well as those
with lapped ties are highlighted. As can be seen in the figure, there appears to be little overall
bias in equation estimates with respect to hook details. In fact, for rectangular columns with 90o
hooks the median of the error on the estimate of aR is 0.005 and the associated standard deviation
is 0.011, while for 135o hooks the median of the error on the estimate of aR is zero and the
associated standard deviation is 0.015. Similarly, for circular columns with lapped ties the mean
error of the estimate of aC is -0.002 and its associated standard deviation is 0.011, while for
spirally reinforced columns the mean error of the estimate of aC is 0.006 and its associated
standard deviation is 0.019. For large s/d ratios (above 0.75) the proposed equation for
rectangular columns provided conservative estimates even for columns with ties having 90o
hooks; 90% of all data points in that range have conservative estimates of aR. One should note
that columns in the database with lapped ties or ties having 90o hooks had relatively low
transverse reinforcement ratios, ρt (as is the case typically in older existing buildings). The
median ρt for rectangular columns with 90o ties is 0.0017, while that for circular columns with
lapped ties is 0.0008. At such low reinforcement ratios, the contribution of ties to the overall

1.9
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

deformation capacity of columns is relatively low, which may explain the observed insensitivity
of estimate errors to hook details of ties.
0.08 0.08
135o hooks spirals
Error in aR = Experiment - Estimate (rad)

Error in aC = Experiment - Estimate (rad)


o lapped
90 hooks
0.06 0.06 Unknown
Unknown

0.04 0.04

0.02 0.02

0 0

-0.02 -0.02

-0.04 -0.04
0 0.2 0.4 0.6 0.8 1 1.2 1.4 0 0.2 0.4 0.6 0.8 1 1.2 1.4
Rectangular column s/d Circular column s/d

Figure 6: Error between experimental values and equation-based estimates for a plotted versus s/d
with transverse reinforcement type highlighted
4.2.2 Estimates of b
The cumulative distribution of the error difference between experimental values and estimates
of b2 are plotted for rectangular and circular columns in Fig. 7. Errors are plotted for estimates
evaluated using the proposed regression equations as well as for those estimated using the ASCE
41 standard. The trends observed in Fig. 7 for values of b2 were similar to those observed for a
values. It is apparent from Fig. 7 that the proposed regression equations shifted estimates from
conservative ones based on the standard to median estimates; i.e., error = 0 at 0.5 probability of
exceedance. For b2R, both methods produced similar spread on estimate error as evidenced by the
similar slopes of the cumulative distribution curves and the standard deviations. For b2C
however, the proposed relation produced a large shift in estimates from those derived using the
standard. For brevity, figures similar to Fig. 4 and Fig. 5 are not presented for b2 values. Similar
trends as those for a values were observed in the various bins for b2 values.
1 1

0.9 0.9

0.8 0.8

0.7 0.7
Cumulative Distribution

Cumulative Distribution

0.6 0.6

0.5 0.5
Equation estimates Equation estimates
0.4 0.4
ASCE 41-06 estimates ASCE 41-06 estimates
0.3 0.3
Number of Tests = 319 Number of Tests = 171
0.2 0.2
Standard Deviation Equation = 0.019 Standard Deviation Equation = 0.018
0.1 0.1
Standard Deviation ASCE41-06 = 0.022 Standard Deviation ASCE41-06 = 0.02
0 0
-0.04 -0.02 0 0.02 0.04 0.06 0.08 -0.04 -0.02 0 0.02 0.04 0.06 0.08
Error in b2R = Experiment - Estimate (rad) Error in b2C = Experiment - Estimate (rad)

Figure 7: Cumulative distribution of the error between experimental values and estimates for b2R and b2C

1.10
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

In Fig. 8, equation-based estimates of b2 are plotted versus experimental values. Highlighted


in the figures are tests that were conducted to axial collapse. As can be seen in Fig. 8, the
equation for b2R provided a conservative estimate of the deformation at axial failure for the
columns that were tested to collapse. The equation for b2c on the other hand shows an
approximate median estimate for tests conducted to axial collapse.
0.14 0.14
Tests not to axial collapse Tests not to axial collapse
Tests to axial collapse Tests to axial collapse
0.12 0.12

0.1 0.1
Estimate of b 2R (rad)

Estimate of b 2C (rad)
0.08 0.08

0.06 0.06

0.04 0.04

0.02 0.02

0 0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Experimental b2R (rad) Experimental b2C (rad)

Figure 8: Comparison between experimental and model estimates for plastic rotations at axial collapse b2

5. Model-Based Modeling Parameter at Onset of Axial Failure b


The limited amount of data makes it difficult to use statistical methods to calibrate an
expression to estimate the drift ratio at axial failure of reinforced concrete columns. A behavioral
model that estimates the plastic rotation at axial collapse of rectangular columns is proposed
instead. The model is mechanics-based and as such is better suited than statistics-based relations
for use outside the range of available test data. The proposed behavioral model simplifies the
Elwood and Moehle model [14] and is verified using recently produced collapse-test data [12,
15-18] and those available in the database.
The axial failure model by Elwood and Moehle [14] states that the lateral drift ratio at axial
failure of a reinforced concrete column is given by:

(terms defined in the notations section) Eq. 6

In Eq. 6, θ is the angle of the critical shear crack from horizontal. Elwood and Moehle
recommend adopting a value of θ of 65̊. If this value is used, the equation is reduced to:
∆ .
Eq. 7
.

The previous equation is further simplified to:


∆ .
Eq. 8

When originally developed, the provisions in the ASCE 41 standard were intended to
provide a conservative estimate of the drift ratio at axial failure. The conservative nature of the
standard is reflected in the Fig. 9, in which the drift ratio at axial failure calculated following the

1.11
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

provisions in the standard and the measured drift ratio at axial failure for several collapse tests
[12, 15-18] are compared. The data set used for the comparison shown in Fig. 9 includes
rectangular columns that are both slender and short, columns that are shear-critical, and columns
that are flexure-shear critical (i.e., yield in flexure prior to sustaining shear failure).
Figure 9 corroborates that drift ratios at axial failure calculated according to the provisions
in ASCE 41 are very conservative for rectangular slender and short columns subjected to
uniaxial loading protocols with three cycles per target lateral drift. Contrary to that, estimates of
the drift ratio at axial failure for specimens subjected to biaxial loading with three cycles per
target drift in each direction or uniaxial loading with six cycles per target drift correspond
approximately to the mean, which is consistent with the observation that columns with these
types of loading protocols experience axial failure at lower drift ratios than columns subjected to
fewer loading cycles [12, 19].

Figure 9: Comparison between measured and calculated drift ratio at axial failure; calculations made
using the ASCE 41 provisions. Test data from [12, 15-18].
A similar comparison between experimental drift ratio at axial failure and that calculated
using the proposed simplified model for rectangular columns is presented in Fig. 10. A different
trend was observed in Fig. 10 than in Fig. 9, whereby calculated drift ratios at axial failure for
slender and short columns subjected to uniaxial loading protocols (3 cycles per drift level)
corresponded approximately to the mean, while estimates of drift ratio for specimens subjected
to biaxial loading and higher number of cycles were unconservative.

1.12
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

Figure 10: Comparison between measured and calculated drift ratio at axial failure; calculations
made using the proposed model. Test data from [12, 15-18].

On the basis of these observations, a reduction factor of 0.5 was applied to drift estimates
derived using the proposed model to calculate parameter b for columns subjected to biaxial
loading or columns subjected to loading protocols with large numbers of cycles per drift ratio.
The results of applying the reduction factor are shown in Fig. 11. As can be seen in the figure,
the reduction factor improved the fit of the model drastically. The statistical results for the data
set of rectangular columns evaluated are presented in Table 2.

Figure 11: Comparison between measured and calculated drift ratio at axial failure; calculations
made using the proposed model with reduction factor for loading protocol. Test data from [12, 15-18].
Table 2: Statistical comparison of measured and calculated drift ratio at axial failure
Model Average Measured/ Standard Coefficient of
Calculated drift ratio Deviation Variation
at axial failure
ASCE 41-06 Suppl. 1 1.97 0.93 0.47
Elwood-Moehle 0.97 0.44 0.46
Simplified Elwood – Moehle
1.13 0.38 0.33
with loading factor

1.13
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

Parameter b2 values for all rectangular and circular columns in the database are shown in
Fig. 12. Estimates of parameter b were obtained using the proposed simplified equation for drift
ratio at axial failure. Parameter b was calculated by subtracting the experimentally observed drift
ratio at yield Δy/L from the drift ratio at axial failure estimated by the model. Tests conducted to
axial collapse are highlighted in the figure.
0.14 0.14
Tests not to axial collapse Tests not to axial collapse
Tests to axial collapse Tests to axial collapse
0.12 0.12
Proposed behavioral model b (rad)

Proposed behavioral model b (rad)


0.1 0.1

0.08 0.08

0.06 0.06

0.04 0.04

0.02 0.02

0 0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
Experimental b2R (rad) Experimental b2C (rad)

Figure 12: Comparison between experimental and model estimates for plastic rotations at axial
collapse; model estimates not adjusted by a factor of 1.3 for circular columns
For rectangular columns, the mean error ratio for b1 = 1.07 (mean experimental b1 / model
estimate), and the corresponding coefficient of variation was 0.71. For circular columns the mean
error ratio for b1 was 1.64 and the corresponding coefficient of variation was 0.57. For a majority
of column tests not conducted to collapse (201 rectangular and 117 circular), the model produced
higher estimates of the b2 values (tests to the left of the diagonal lines in Fig. 12). However,
given that the experimental b2 values constitute a lower-bound of the actual plastic rotations at
axial failure, it is not possible to conclude whether the model produced conservative or
unconservative estimates for those tests. The model produced lower estimates of the b2 values for
many tests that were not conducted to collapse (82 rectangular and 45 circular tests to the right of
the diagonal lines in Fig. 12), which indicates conservative estimates of the plastic rotations at
axial failure for those tests. Such numbers suggest that the model provided a reasonable estimate
of the mean plastic rotations at axial failure for tests conducted to collapse and may not be
unconservative over the entire database. For circular columns however, the analysis indicates
that the model provided a conservative estimate of the parameter b, particularly considering that
the mean error ratio of b1 values was 1.64. In light of these results, and given the paucity of
experimental results to axial failure of circular column, a factor of 1.3 is proposed to increase
model estimates for spirally reinforced circular columns. This factor was chosen by comparing
the terms of statistically-derived equations for the parameter b2 of circular and rectangular
columns. While the 1.3 factor does generate a mean error ratio still significantly larger than 1.0
(=1.26) for b1 values, it is consistent with the approximate 30% difference between the
regression based b2R and b2C estimates. Increasing model estimates by more than 30% was not
deemed appropriate given the limited test data (only nine tests on circular columns were
conducted to collapse).

1.14
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

6. Proposed Modeling Parameters for Reinforced Concrete Columns


It is proposed that the modeling parameter a for determining the plastic rotation at onset of
lateral-strength degradation for reinforced concrete columns be given by equations derived
through regression methods. Given the large experimental dataset for parameter a values,
statistical methods provide a useful tool for developing expressions for median estimates of that
parameter. The alternative to deriving expressions through regressions on influential parameters
is to develop a fundamental expression based on mechanics and calibrate it using test data.
Several mechanisms however can trigger lateral-strength degradation in reinforced concrete
columns such as yielding of the transverse reinforcement crossing a shear crack, or the buckling
or fracture of longitudinal bars. Because multiple modes of failure must be considered, a
potential mechanics-based model is more complex to develop and calibrate than the simple
equations proposed. It is therefore recommended that Eq. 2 and Eq. 4 be used for the evaluation
of the plastic rotations at onset of lateral-strength degradation (a) for reinforced concrete
columns.
For the plastic rotations at onset of axial collapse (b), the paucity of test data makes difficult
the use of statistical methods to estimate that parameter. The accuracy of a mechanics-based
model [14] is explored through the available dataset. A simplified version of the model was
shown to provide a reasonable estimate of the b parameter for rectangular columns tested to
collapse and a conservative estimate for spirally-reinforced circular columns tested to collapse. It
is therefore recommended to use the mechanics-based model with the proposed simplifications to
estimate the b parameter for all columns with the exception of spirally reinforced circular
columns. For the latter type of columns, it is recommended that the b parameter estimated using
the mechanics-based model be increased by 30% to account for the observed improved
performance over other types of columns. In no case should the a value for a column be taken
larger than the b value.
Test results from columns subjected to more than 3 cycles per drift level in one direction or
3 cycles per drift level in two orthogonal directions have been shown to have a drift ratio at axial
failure up to 50% lower than observed in similar columns subjected to 3 cycles per drift level in
one direction [11, 18]. The loading protocols used in those tests imposed a severe displacement
demand on the test columns that is difficult to correlate with displacement histories induced by
seismic excitations from various types of earthquake sources. The loading protocols of most tests
in the database consisted of three cycles at incrementally increasing lateral drifts and constitute
loading protocols that are commonly considered to be severe enough to encompass a large
majority of seismic demands [20]. For these reasons, no further adjustments to the proposed
relations are introduced in this paper, although engineers are cautioned to reduce modeling
parameter b when evaluating the performance of buildings near sources with the potential to
generate unusually long duration earthquakes or seismic demands with similar intensities in the
fault normal and parallel components.
7. Comparison Between Proposed and ASCE 41 Estimates for Modeling Parameters
In Fig. 13, the difference between a values estimated using the ASCE 41 standard and the
proposed equations are illustrated. The figure shows that for the majority of rectangular columns,
the proposed Eq. 2 alters estimates from the standard by +/- 0.01 rad. For circular columns, the
change in estimate is greater and biased towards increasing deformation capacities; which is

1.15
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

expected given the observed higher deformation capacities of circular columns that are not
accounted for in the standard.
0.05 0.07

0.045
0.06
0.04

0.035 0.05
ASCE 41-06 aR (rad)

ASCE 41-06 aC (rad)


0.03
0.04
0.025
0.03
0.02

0.015 0.02

0.01
0.01
0.005

0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Proposed equation aR (rad) Proposed equation aC (rad)

Figure 13: Difference in estimates of a between ASCE41 and the proposed equations
A comparison between b values estimated using of the proposed mechanics-based model
and those of the ASCE 41 standard are shown in Fig. 14. For both rectangular and spirally
reinforced circular columns, the proposed model estimates increased deformation capacities
significantly over standard estimates. This was to be expected because the standard targets
conservative deformation capacities while the model targets median estimates.
0.12 0.14
Tests not to axial collapse Tests not to axial collapse
Tests to axial collapse Tests to axial collapse
0.12
Proposed behavioral model b (rad)
Proposed behavioral model b (rad)

0.1

0.1
0.08

0.08
0.06
0.06

0.04
0.04

0.02
0.02

0 0
0 0.02 0.04 0.06 0.08 0.1 0.12 0 0.02 0.04 0.06 0.08 0.1 0.12
ASCE 41-06 b(rad) ASCE 41-06 b (rad)

Figure 14: Difference in estimates of b between ASCE 41 and the proposed model (including the
30% adjustment for circular columns)
8. Acceptance Criteria
Acceptance criteria (AC) should reflect the amount of damage that is acceptable for the
Immediate Occupancy (IO) performance objective (defined in [3]). For more severe performance
objectives related to safety and structural stability, considerations for selecting AC should not
only reflect acceptable damage levels but also the probability of exceeding a threshold
behavioral state such as a column initiating lateral-strength degradation and exhibiting a negative
lateral stiffness. . The ASCE 41 standard provides procedures for defining AC based on
experimental data. For all members, the standard defines AC for the Immediate Occupancy (IO)
performance objective as the critical deformation at which permanent visible damage occurs in a

1.16
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

member. The standard defines AC differently for primary and secondary members at the Life-
Safety (LS) and Collapse-Prevention (CP) performance objectives. For primary members, AC
are defined as 75% of the MP a for LS and equal to a but not greater than 75% of b for CP. For
secondary members, AC are defined as 75% of MP b for LS and equal to b for CP. It is useful to
recall here that current MP in the standard were selected conservatively and thus by limiting AC
to the MP a or b, the standard in effect aims to limit deformations to below values at which loss
of lateral strength or member failure occur.
It is proposed here that AC for IO be based on a percentage of the a value since columns
with better detailing and higher deformation capacities tend to exhibit visible damage at larger
deformations than less well detailed columns. A conservative value of 10% of the a parameter is
selected as the limiting deformation at which a reinforced column is deemed to need repair and
no longer satisfy the IO performance objective.
For more severe performance objectives (i.e., Life-Safety (LS) and Collapse-Prevention
(CP)), structural stability and safety are of paramount concern. For these performance objectives,
ensuring a fixed probability of exceeding the deformation at onset of lateral-strength degradation
(a) or the onset of member failure (b) is recommended. Thus for primary structural members, AC
should be defined such that the deformation at onset of lateral-strength degradation (MP a) has a
low probability of being exceeded. For secondary structural members, AC should be defined so
the deformation corresponding to member loss of strength (MP b) has a low probability of being
exceeded. Because the errors on the estimates of MP exhibit different dispersions for various
parameters, selecting a fixed fraction of those MP values for AC results in different probabilities
of exceeding MP deformations for members with deformations not exceeding the AC.
Consequently, it is recommended that AC be defined through specific percentiles rather than
percentages of MP. Percentiles for selecting AC were selected to remain consistent with the
intent of the current ASCE 41standard definitions for AC (i.e., percentiles were selected to
provide percentage values of MP equivalent to those currently defined in the standard). The
following summarizes the selected percentiles recommended for the various performance
objectives:

a. For LS of primary members, it is proposed that plastic rotations should not exceed
the 20th percentile of a. For a member critical to the stability of a structure,
satisfying the AC for LS would indicate a high level of confidence (80%) that the
member under consideration has not initiated lateral-strength degradation.
b. For CP of primary members, it is proposed that plastic rotations should not exceed
the 35th percentile of a. The ASCE 41 standard specifies that this AC be equal to
the MP a. Given that MP provided in the ASCE 41 standard are conservative
estimates of the behavioral milestones rather than median estimates, the AC is
given here as lower than the proposed median estimate of that parameter.
c. For LS of secondary members, it is proposed that plastic rotations should not
exceed the 10th percentile of b nor be less than AC for LS of primary members.
Due to the more critical nature of the behavioral milestone identified by b, a lower
percentile was selected for this AC than for primary members.
d. For CP of secondary members, plastic rotations should not exceed the 25th
percentile of b nor be less than AC for CP of primary members.
e. In no case should the AC for primary members be larger than those for secondary
members.

1.17
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

The percentages of MP a and b that satisfy the selected percentiles for reinforced concrete
columns are given in Table 3. As can be seen from the table, the resulting AC are on the order of
30 to 50% of a for LS of primary members, 65 to 75% of a for CP of primary members, 60% of
b for LS of secondary members, and 75% of b for CP of secondary members. The range of
percentages needed to achieve the same percentiles of MP illustrates the need to move to fixed
percentiles and variable percentages to remain consistent with the performance-based philosophy
of the ASCE 41 standard.

Table 3: Recommended acceptance criteria


IO LS Primary CP Primary LS Secondary CP Secondary
Rectangular 0.1 aR 0.50 aR 0.75 aR 0.60 bR 0.75 bR
Circular 0.1 aC 0.30 aC 0.65 aC 0.60 bC 0.75 bC

9. Summary and Conclusions


A database of 490 pseudo-static tests of reinforced concrete columns was used to develop
relations for nonlinear modeling parameters that define the lateral force-deformation envelope of
such columns under seismic excitation. Expressions were developed to estimate the plastic
rotation in a concrete column at incipient lateral-strength degradation (Modeling Parameter MP
a) and the plastic rotation at incipient axial degradation (MP b). The proposed relations are given
in a format that is compatible with the standard for seismic rehabilitation ASCE 41-06
Supplement 1 [3, 4]. Contrary to provisions in the standard, that produce conservative estimates
of MP, the proposed equations provide median estimates of MP with the intent to avoid skewing
analytical results. Separate relations were derived for spirally reinforced circular columns and all
other columns due to observed larger deformation capacities of circular columns.
Regression-based expressions were derived for the MP a and b based test data. Influential
parameters for rectangular and circular columns were: the axial load ratio, the transverse
reinforcement ratio, and the ratio Vy/Vo. Regression-based relations for a produced median
estimates of the parameter and standard deviations on the same order as those of estimates
produced using the ASCE 41 standard. The proposed circular-column expression for a produced
values significantly higher than those calculated using the ASCE-41 standard because the
provisions in the standard were derived based on rectangular column data and do not account for
the improved performance of spirally reinforced circular columns.
Due to the scarcity of data on columns tested to axial collapse, the regression-based
expressions for the parameter b could only produce a lower-bound estimate of that parameter. An
alternative to estimate deformations at axial failure was proposed through a behavioral model.
The model is mechanics based and as such is better suited than statistics-based relations for use
outside the range of available test data. The model constitutes a simplification of the model by
Elwood and Moehle [14]. Model estimates were increased by 30% to account for the higher
deformation capacity of spirally reinforced circular columns. The proposed model was shown to
produce approximately median estimates of the b parameters for column tests conducted to
collapse. Increased numbers of lateral cycles or significant bi-directional excitation have been

1.18
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

shown to reduce the drift at axial collapse by up to 50%. Because the loading protocols that
generated such reductions are unusually severe and it is difficult to correlate them to earthquake
duration, no further adjustments to the proposed relations were introduced. Engineers are advised
to reduce model estimates of drift ratio at axial failure should an unusually demanding
displacement history be expected to occur.
Acceptance criteria (AC) were proposed that identify acceptable deformation levels at various
performance objectives. Limiting deformations for Life-Safety and Collapse-Prevention
performance objectives were defined in terms of the a and b MP such that reaching the AC
deformations would correspond to a low probability of exceeding deformations at onset of lateral
or axial degradation.

Acknowledgements
The authors would like to thank Dr. Kenneth Elwood for his insightful advice and support during
the development of the proposed relations. The authors would also like to acknowledge the help
of Insung Kim, Garrett Hagen, and Degenkolb Engineers for providing information regarding the
compliance of columns in the database with ACI318-11 seismic provisions.

1.19
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

Notations
a = modeling parameter representing the plastic rotation at incipient lateral-strength degradation;
experientially extracted as a = (Δ0.8-Δy)/L
aC = modeling parameter a for circular columns
aR = modeling parameter a for rectangular columns
Ag = concrete gross section area
Av = area of transverse reinforcement in the direction of loading spaced at s
b = modeling parameter representing the plastic rotation at incipient axial degradation
b1 = modeling parameter b for column tests conducted to axial collapse = (Δaxial-Δy)/L
b2 = modeling parameter b for all column tests including those conducted to axial collapse
b2C = modeling parameter b2 for circular columns
b2R = modeling parameter b for rectangular columns
bw = section web width
c = modeling parameter representing the residual lateral strength of a column
d = section effective depth form edge of compression region to centroid of tension steel. d is
taken as 0.8D (D=section diameter) for circular columns
dc = depth of core (centerline to centerline of ties)
f’c = concrete compressive strength
fyt = transverse steel yield strength
L = column clear length
La = column shear span
M/Vd = ratio of applied moment to applied shear times the effective depth and is bounded by the
values of 4 and 2
My = column moment strength
Nu = applied axial load used for shear strength evaluation (= 0 for tension loads)
P = axial load (positive in compression)
P/ Ag f’c = axial load ratio
s = transverse reinforcement spacing
Vmax = maximum applied shear
Vo = nominal shear strength of columns as defined by ASCE 41-06 but with the following
modification: the shear strength provided by steel is fully accounted for up to a s/d ratio of
0.75, is taken equal to zero for a s/d ratio above 1.0, and linearly interpolated between full
strength and zero strength between s/d = 0.75 and s/d = 1.0
Vy = shear corresponding to the development of flexural yield moments = Mp / La

1.20
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

α = unit adjustment factor = 6 in psi units and 0.5 in MPa units


Δ = drift at axial failure estimated using the proposed mechanics-based model
Δ0.8= lateral drift at which the lateral strength of an element degrades by 20% from peak
Δaxial = drift at onset of axial collapse
Δy = lateral drift at yield of longitudinal bars or at onset of significant inelastic deformations
Φ = factor for adjusting steel shear contribution and is taken as 1.0 for s/d ≤ 0.75, zero for s/d ≥
1.0, and linearly interpolated between the two values of s/d
ρt = transverse reinforcement ratio = Av/(bws)
θ = angle of critical shear crack from horizontal

1.21
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

References

1. Ghannoum, W.M. and B. Sivaramakrishnan, ACI 369 Rectangular Column Database,


2012: Network for Earthquake Engineering Simulation (database).
2. Ghannoum, W.M. and B. Sivaramakrishnan, ACI 369 Circular Column Database, 2012:
Network for Earthquake Engineering Simulation (database).
3. American Society of Civil Engineers/Structural Engineering Institute (ASCE/SEI)
Committee 41, Seismic Rehabilitation of Existing Structures, in ASCE Standard2007:
Reston, VI. p. 428.
4. American Society of Civil Engineers/Structural Engineering Institute (ASCE/SEI)
Committee 41, Supplement 1 to ASCE 41, in ASCE Standard2007.
5. Elwood, K.J., et al., Update to ASCE/SEI 41 Concrete Provisions. Earthquake Spectra,
2007. 23(3): p. 493-523.
6. Berry, M.P., M. Parrish, and M.O. Eberhard, PEER Structural Performance Database
User’s Manual, 2004, Pacific Earthquake Engineering Research Institute,
http://nisee.berkeley.edu/spd/.
7. Sivaramakrishnan, B., Non-Linear Modeling Parameters for Reinforced Concrete
Columns Subjected to Seismic Loads, in Department of Civil, Architectural, and
Environmental Engineering2010, University of Texas at Austin: Austin. p. 75.
8. American Concrete Institute (ACI) Committee 318, Building Code Requirements for
Reinforced Concrete (318-11), 2011, American Concrete Institute: Farmington Hills, MI.
p. 503.
9. Sezen, H. and J.P. Moehle, Shear Strength Model for Lightly Reinforced Concrete
Columns. Journal of Structural Engineering, 2004. 130(11): p. 1692-1703.
10. Elwood, K.J. and M.O. Eberhard, Effective stiffness of reinforced concrete columns. ACI
Structural Journal, 2009. 106(4): p. 476-484.
11. Woods, C., Mitigation of Collapse Risk in Vulnerable Concrete Buildings, 2009,
University of Kansas: Lawrence, Kansas.
12. Henkhaus, K., Axial Failure of Vulnerable Reinforced Concrete Columns Damaged by
Shear Reversals, 2010, Purdue University: West Lafayette, IN.
13. Hognestad, E., A study of combined bending and axial load in reinforced concrete
members, in Bulletin 3991951, University of Illinois. p. p. 128.
14. Elwood, K.J. and J.P. Moehle, Axial Capacity Model for Shear-Damaged Columns. ACI
Structural Journal, 2005. 102(4): p. 578-587.
15. Lynn, A.C., Seismic Evaluation of Existing Reinforced Concrete Building Columns, in
Department of Civil and Environmental Engineering2001, University of California,
Berkeley. p. iv, 359 leaves.
16. Matamoros, A.B., L. Matchulat, and C. Woods, Axial Load Failure of Shear Critical
Columns Subjected to High Levels of Axial Load, in 14th World Conference on
Earthquake Engineering2008: Beijing, China.
17. Sezen, H. and J.P. Moehle, Seismic Tests of Concrete Columns with Light Transverse
Reinforcement. ACI Structural Journal, 2006. 103(6): p. 842-849.
18. Woods, C. and A.B. Matamoros, Effect of Longitudinal Reinforcement Ratio on The
Failure Mechanism of R/C Columns Most Vulnerable to Collapse, in 9th Us National and
10th Canadian Conference on Earthquake Engineering2010: Toronto, Canada.

1.22
@Seismicisolation
@Seismicisolation
Nonlinear modeling parameters and acceptance criteria for concrete columns

19. Simpson, B. and A.B. Matamoros, Criteria for Evaluating the Effect of Displacement
History and Span-to-Depth Ratio on the Risk of Collapse of R/C Columns, in 15th World
Conference on Earthquake Engineering2012: Lisbon, Portugal.
20. American Concrete Institute (ACI) Committee 374, Acceptance Criteria for Moment
Frames Based on Structural Testing and Commentary (374.1-05), 2005, American
Concrete Institute: Farmington Hills, MI. p. 9.

1.23
@Seismicisolation
@Seismicisolation
W.M. Ghannoum and A.B. Matamoros

1.24
@Seismicisolation
@Seismicisolation
SP-297—2

Assessment of ASCE/SEI 41 Concrete Column Provisions using


Shaking Table Tests
Y. Li1, K.J. Elwood2, and S.-J. Hwang3
ABSTRACT

A database comprised of 59 reinforced concrete columns subjected to strong ground shaking using earthquake
simulators (or shaking tables) is compiled. This paper will focus on insights provided by the database related to the
concrete column provisions in ASCE/SEI 41. In particular, the Shaking Table Test Column Database is used to
evaluate the accuracy of column effective stiffness models, column classification criteria, and the level of
conservatism provided by the plastic rotation capacities specified in ASCE/SEI 41. It is found that the Standard
generally overestimates the column effective stiffness, while providing a mean value estimate of the column shear
strength regardless of tie spacing. The modeling parameters specified in the standard provide conservative estimate of
the column drift capacities and are consistent with the targeted probability of failure. Refinements of the shear
strength model and the criteria for column classifications are suggested. This study also compares the measured
response of columns subjected to quasi-static cyclic loads and shaking table tests.

Keywords: Shaking table tests; ASCE/SEI 41; reinforced concrete columns; shear failure, axial-load failure

1
Design Engineer, Read Jones Christoffersen, Vancouver, Canada
2
Professor, Department of Civil Engineering, University of British Columbia, Canada
3
Professor, Department of Civil Engineering, National Taiwan University, Taiwan

2.1
@Seismicisolation
@Seismicisolation
Y. Li et al.

INTRODUCTION

ASCE/SEI 41 (2013) is commonly used throughout the United States and internationally for the seismic
assessment and retrofit of existing concrete buildings. In 2007, the concrete provisions of ASCE/SEI 41 were
updated to better reflect the observed performance of concrete components from quasi-static cyclic laboratory tests
(Elwood et al. 2007). These updates have been incorporated in the latest release of the seismic rehabilitation
standard, ASCE/SEI 41 (2013).

Figure 1 ASCE/SEI 41 Backbone Model

ASCE/SEI 41 recommends generalized load-deformation relation (or backbone) for displacement-controlled


components including concrete columns, as shown in Figure 1. According to the ASCE/SEI 41 backbone model, a
column is assumed to remain elastic for shear demands below the effective yield strength of the column, Vy, which is
taken as the minimum of the plastic shear capacity associated with plastic hinging at both ends of the column, Vp,
and the nominal shear strength at low ductility demands, V0.
After the rotation demand exceeds the yield rotation θy, the column experiences plastic deformations and
ASCE/SEI 41 provides the length of the plastic plateau, a, depending on the expected failure mode of the column.
For primary components, a also defines the maximum allowable plastic rotation (acceptance criteria) for the
Collapse Prevention (CP) performance level. (Note that ASCE/SEI 41-13 no longer provides the nonlinear
acceptance criteria for primary components since all components can be considered as secondary when using
nonlinear procedures in which any strength degradation is captured in the nonlinear model. The length of the yield
plateau, a, is still the basis for the m-factor for primary components used in linear procedures.)
When plastic rotation demands exceed a, the column shear resistance drops to a residual strength of cVy, and
remains at this level of resistance until the ultimate plastic rotation, b, is reached. Quantities b and c are also
specified in ASCE/SEI 41 based on the expected failure mode of the column. The ultimate plastic rotation is
intended to represent the plastic rotation after which the axial-load carrying capacity of the column should not be
relied upon. For secondary components, b also defines the maximum allowable plastic rotation (acceptance criteria)
for the Collapse Prevention (CP) performance level.

Examples of other backbone models or acceptance criteria for existing concrete columns include
Panagiotakos and Fardis (2001), Elwood and Moehle (2006) and Haselton et al. (2008). Panagiotakos and Fardis
(2001), used as the basis for acceptance criteria in Eurocode 8 Part 3 (EN 1998-3 2005), provided empirical
equations for estimating the chord rotation at yield and 20% loss of strength. Elwood and Moehle (2006) provided a
backbone model for columns experiencing shear failure after flexural yielding, including estimates of the column
effective stiffness and drift ratio at shear and axial load failure. Haselton et al. (2008) provided a backbone model
for flexure-controlled columns based on a database of 255 quasi-static cyclic column tests.

Due to the prevalence of application in engineering practice, this study will focus on the backbone model and
provisions for concrete columns in ASCE/SEI 41. In particular, a database comprised of 59 reinforced concrete
columns subjected to strong ground shaking using earthquake simulators (or shaking tables) has been compiled to
evaluate the accuracy of column effective stiffness models, the performance of column classification criteria, and the
level of conservatism provided by the plastic rotation capacities specified in ASCE/SEI 41. A comparison of the
measured response of columns subjected to quasi-static cyclic loads and shaking table tests is also included.

2.2
@Seismicisolation
@Seismicisolation
Assessment of ASCE/SEI 41 Concrete Column Provisions using Shaking Table Tests

EXPERIMENTAL DATABASE

To assess the ASCE/SEI 41 backbone model, a Shaking Table Test Column Database is compiled (for
brevity, this database will be referred to as “Dynamic Database” herein). The Dynamic Database includes results of
shaking table tests that have been conducted on simple two-dimensional frames with relatively large-scale
reinforced concrete columns (length scale factor no less than 1:3.5), for which critical test data (column axial load,
column shear, horizontal drift ratio and column shortening/lengthening) are available. Based on this criteria, seven
shaking table test programs were identified, three of which were conducted at UC Berkeley (Elwood 2002,
Ghannoum 2007, and Shin 2007) and the remaining four were carried out at National Center for Research on
Earthquake Engineering (NCREE) in Taiwan (Su 2007, Kuo 2008, Wu et al. 2006, Yavari 2011). Most of the
selected shaking table tests were conducted up to collapse of the frame specimens. Typical test configurations are
shown in Figure 2. All specimens were single story frames with the exception of specimens tested by Yavari (2011)
and Ghannoum (2007) which were two and three stories, respectively. Details of the shaking table tests are
summarised in Li (2012).

Fifty-nine (59) first-story columns out of a total of 97 columns are included in the Dynamic Database (upper
story column shear forces and axial loads were not recorded directly and cannot be determined with sufficient
reliability to be included in the database). Figure 3 provides the range of column geometry, material properties and
reinforcement details for columns in the database; data for each column are provided in Appendix A.

Columns in the Dynamic Database were classified into three types based on the observed failure modes:
flexure critical, flexure-shear critical and shear critical columns. The classification criteria based on column
properties and recorded response during testing are summarized in Table 1.

Fifteen of the flexure-shear-critical columns were part of a larger multi-story frame and the top of the
columns were connected to flexible beams similar to the configuration shown in Figure 2(a). This is relevant to the
assessment below since the flexible beam results in additional deformations from beams and joints, contributing to
the measured interstory drift. For specimens with rigid beams (e.g. setups similar to Figure 2(b)), the additional
deformation from beams and joints can be neglected. All shear-critical columns were connected to rigid beams.

(a) (b)

Figure 2 Typical configuration of tests in Dynamic Database (a) columns connected with flexible beams
(source: Yavari 2011); (b) columns connected with rigid beam (source: Shin 2007)

2.3
@Seismicisolation
@Seismicisolation
Y. Li et al.

S FS F S FS F S FS F
40 40 40

35 35 35

Number of Columns
Number of Columns
30 30
Number of Columns

30
25 25 25

20 20 20

15 15 15

10 10 10

5 5 5

0 0 0
20~24 24~28 28~32 32~36 36~38 <300 <350 <400 <450 <500 <400 <500 <600 <700
Concrete Strength (MPa) Long. Rein. Yield Strength (MPa) Trans. Rein. Yield Strength (MPa)

S FS F S FS F S FS F
40 40 40
35 35 35
Number of Columns

Number of Columns

Number of Columns
30 30 30
25 25 25
20 20 20
15 15 15
10 10 10
5 5 5
0 0 0
<2.0 <2.5 <3.0 <3.5 <4.0 <4.5 <1.5% <2.0% <2.5% <3.0% <3.5% <0.20% <0.60% <1.00% <1.40%
Aspect Ratio a/d Long. Rein. Ratio Trans. Rein. Ratio

S FS F S FS F
40 40
35 35
Number of Columns

Number of Columns
30 30
25 25
20 20
15 15
10 10
5 5
0 0
<0.1 <0.2 <0.3 <0.4 <0.1 <0.2 <0.3 <0.4 <0.5
Initial Axial Load Ratio Pini/Agfc' Max. Axial Load Ratio PMax/Agfc'

Note: F – flexure-critical columns; FS – flexure-shear-critical columns; S – shear-critical columns

Figure 3 Column Properties

Figure 4 shows the normalized envelopes of measured lateral response for columns experiencing shear or
flexure-shear failure modes in the Dynamic Database. Column lateral force is corrected for P-Delta effects to give
the effective lateral force instead of recorded shear force of each column (unless specified otherwise, the lateral
force used throughout this paper is the column shear force corrected for P-Delta effects). The x-axis is the ductility
demand, based on the measured yield displacement determined from the maximum force observed during the test
and an effective stiffness based on a straight line from the origin through the measured response of the column at 75%
of the maximum lateral force. Detailed descriptions of the tests and the development of the envelopes can be found
in Li (2012). Despite the fact that all columns in Figure 4 experienced shear failures (before or after flexural
yielding), it is noted that all columns achieved ductilities greater than unity and the majority of flexure-shear-critical
columns were able to sustain lateral loads beyond a ductility demand of 2.0.

2.4
@Seismicisolation
@Seismicisolation
Assessment of ASCE/SEI 41 Concrete Column Provisions using Shaking Table Tests

1.2 1.2 1.2

0.8 0.8 0.8


Normalized Force (V/Vmax)

Normalized Force (V/Vmax)
0.4 0.4 0.4

Normalized Force (V/Vmax)
0.0 0.0 0.0

‐0.4 ‐0.4 ‐0.4

‐0.8 ‐0.8 ‐0.8

‐1.2 ‐1.2 ‐1.2


‐6 ‐4 ‐2 0 2 4 6 ‐6 ‐4 ‐2 0 2 4 6 ‐6 ‐4 ‐2 0 2 4 6
Ductility (μ=Δ/Δyield) Ductility (μ=Δ/Δyield) Ductility (μ=Δ/Δyield)
(a) (b) (c)
Figure 4 Experimental envelopes for (a) flexure-shear-critical columns connected with rigid beam;
(b) flexure-shear-critical columns connected with flexible beam; (c) shear-critical columns.

Table 1 Column classification

Failure mode
Flexure-critical Flexure-shear critical Shear-critical
• Inadequate transverse • Non-ductile columns with
reinforcement for ductile inadequate transverse
• Adequate transverse response (ρt≤0.002; 90° hooks) reinforcement
Column
reinforcement • Considered to be (transverse reinforcement ratio
Properties
(ρt>0.002; 135° hooks) representative of columns found ρt≤0.002; 90° hooks)
in older reinforced concrete • Usually have low aspect ratio
structures. (i.e. "short column")

• Demonstrate extended • Vulnerable to damage • Vulnerable to damage


deformation after yielding associated with shear failure associated with shear failure
• Experience failure associated (e.g., diagonal crack, fracture of (e.g., diagonal crack, fracture of
with flexural damage (e.g., transverse reinforcement) transverse reinforcement)
Column concrete spalling, buckling of •Develop limited inelastic • Experience sudden shear
Response longitudinal reinforcement) deformation beyond flexural failure before flexural yielding
• Shear failure is unlikely to occur yielding • Lose gravity load support
• Have a plastic plateau after • As shear-failure planes shortly after shear failure
initial elastic response in lateral develop, axial load support is • Develop very limited plastic
force-deformation relations generally lost drift before column failure
• A total of 33 columns are
included, of which 27 columns • A total of 5 columns with
experienced shear failure and aspect ratio ≤2.0 all from
• A total of 21 columns are axial-load failure. The NCREE 2007 tests (Su 2007).
included in Dynamic Database remaining 6 columns did not • All five columns experienced
• No column failed during the fail in tests. sudden shear failure before
Observation
test (i.e. no more than 20% of • Severe lateral force yielding of longitudinal
lateral force degradation after P- degradation (loss of more than reinforcement
∆ correction) 20% of its peak resistance) • Axial-load failure occurred
• longitudinal reinforcement shortly after column shear
buckled at the top or bottom of failure
some columns
Examples (Refer • NCREE 2009-MUF-A1 • NCREE 2009-MCFS-A1 • NCREE 2007-S1-C2
to Appendix A) • Ghannoum 2006-C1 • Elwood 2002-S1-CenC • NCREE 2007-S3-C1

2.5
@Seismicisolation
@Seismicisolation
Y. Li et al.

2
Shear‐critical
Flexure‐Shear‐critical
Flexure‐Shear‐critical (Beam)
1.5
Flexure‐critical
EIMeas /EIASCE

Flexure‐critical (Beam)
Mean
1

0.5

0
0 0.1 0.2 0.3 0.4
Initial axial load ratio Pini /Agfc'
Figure 5 Evaluation of column effective stiffness estimated by ASCE/SEI 41

2
Shear‐critical
Flexure‐Shear‐critical
Flexure‐Shear‐critical (Beam)
1.5
Flexure‐critical
EIMeas /EIEE

Flexure‐critical (Beam)
Mean
1

0.5

0
0 0.1 0.2 0.3 0.4
Initial axial load ratio Pini /Agfc'
Figure 6 Evaluation of column effective stiffness estimated by Elwood and Eberhard (2009)

EFFECTIVE STIFFNESS

Most columns included in the Dynamic Database could be considered as fixed against rotation at both ends,
thus the relationship between measured column lateral stiffness, kMeas, and effective flexural rigidity, EIeff, can be
expressed by Equation 1:

kMeas Vy /y 12EIeff / L2 Equation 1

where Vy is the effective yield strength of the column; θy is the yield rotation; L is the column clear height.

The effective flexural rigidity (EIeff) is often expressed as a fraction of the gross-section flexural rigidity EIg. In the
interest of using common terminology from engineering practice, the ratio (EIeff /EIg) will herein be referred to as
effective stiffness ratio.

In this research, the measured effective stiffness ratio, EIMeas/EIg, and yield drift for columns in the Dynamic
Database are based on a secant stiffness to 75% of the maximum lateral force and corresponding drift on the
measured lateral force-drift envelope. According to ASCE/SEI 41, the effective stiffness ratio (EIeff /EIg) for
columns can be estimated as 0.3 for axial load ratios (P/Agfc’ ) less than 0.1, and 0.7 for axial load ratios higher than
0.5. Linear interpolation of the effective stiffness ratio is used between these limits.

Results of measured to calculated effective stiffness ratio, EIMeas/EIASCE, are presented in Figure 5 and
grouped by column failure mode. For specimens subjected to multiple ground motions, only the first test is
considered here to avoid consideration of stiffness reductions associated with previous testing. Specimens connected
to a flexible beam are indicated by a different symbol in Figure 5 (i.e. “Flexure-critical (Beam)” and “Flexure-shear-
critical (Beam)”), since such cases are expected to result in a lower bound estimate of the column stiffness due to the
additional flexibility of the beam and joint. Statistics are provided in Table 2 for all specimens and just those
specimens with rigid beams to remove the additional variability introduced by the beam flexibility.

2.6
@Seismicisolation
@Seismicisolation
Assessment of ASCE/SEI 41 Concrete Column Provisions using Shaking Table Tests

Based on the test results of 150 columns subjected to quasi-static cyclic loads, Elwood and Eberhard (2009)
proposed an effective stiffness model taking into account the deformation due to flexure, shear and bar slip, as
shown in Equation 2:

EI EE 0.45  2.5 P Ag f c
'

0.2    1.0
EI g  db   D  Equation 2
1  110    
 D  a 
where P is the column axial load; Ag is the gross cross-section area of the column; f’c is the concrete compressive
strength; db is the nominal diameter of longitudinal bars; D is the diameter of circular column or column depth in
direction of loading and a is the shear span of the column.

Ratio of measured to calculated effective stiffness EIMeas/EIEE by Elwood and Eberhard (2002) for the Dynamic
Database is presented in Figure 6 and Table 2.

ASCE/SEI 41 tends to overestimate the stiffness of nearly all columns regardless of failure mode. Elwood
and Eberhard (2009) gives an improved estimation (closer to mean value of the test results) but still generally
overestimates the measured stiffness, except for shear critical columns. Since the Elwood and Eberhard model was
developed to provide a close estimate of the mean response from quasi-static cyclic tests, these results suggest that
columns subjected to real ground motions tend to exhibit a slightly lower effective stiffness compared to quasi-static
cyclic tests. This result may be attributed to the large number of below-yield cycles typically imposed on columns
during real ground motions; however, further testing would be required to generalize this conclusion.

Table 2 Statistical results of effective stiffness ratio

Standard Coefficient of
   Min Max Mean Median Deviation Variation(COV)
Flexure-critical columns
EIMeas/EIASCE 0.27 0.98 0.58 0.57 0.14 25%
EIMeas/EIASCE (Rigid beam only) 0.27 0.76 0.58 0.60 0.12 21%
EIMeas/EIEE 0.41 1.24 0.75 0.73 0.20 26%
EIMeas/EIEE (Rigid beam only) 0.41 1.07 0.75 0.74 0.16 21%
Flexure -shear-critical columns
EIMeas/EIASCE 0.41 1.21 0.70 0.67 0.17 24%
EIMeas/EIASCE (Rigid beam only) 0.50 1.21 0.71 0.65 0.17 24%
EIMeas/EIEE 0.52 1.33 0.89 0.87 0.18 20%
EIMeas/EIEE (Rigid beam only) 0.63 1.31 0.88 0.86 0.16 18%
Shear-critical columns
EIMeas/EIASCE (Rigid beam only) 0.65 1.27 0.79 0.75 0.18 23%
EIMeas/EIEE (Rigid beam only) 0.97 1.90 1.19 1.13 0.27 22%

COLUMN CLASSIFICATION

To establish the deformation capacity of a concrete column, ASCE/SEI 41 classifies columns into three
“conditions”, or failure modes, based on the ratio of the shear strength at low ductility demand, V0, to the plastic
shear demand, Vp. The plastic shear demand, Vp, is the shear force corresponding to flexural yielding of plastic
hinges. As shown in Figure 1, for columns with both ends fixed (as assumed for columns in the Dynamic Database),
Vp is calculated by dividing the maximum moment strength obtained from section analysis by shear span, a=L/2.
V p  2M p / L Equation 3
ASCE/SEI 41 adopts the shear strength model developed by Sezen and Moehle (2004) and recommends
calculating column shear strength, Vn, by summing the contribution from concrete, Vc, and transverse reinforcement,

2.7
@Seismicisolation
@Seismicisolation
Y. Li et al.

Vs, as shown in Equation 4. The term V0= Vn /k is the shear strength excluding the reduction factor related to
ductility demand.

Vn  kV0  k Vc  Vt 
 0.5 f ' Nu 
Vc    c
1  0.8 Ag
 M / Vd 0.5 f '
A 
 c g 

 Av f y d 
   s / d   0.5  Equation 4
 s 
 Av f y d 
Vs  0.5  0.5   s / d   1.0
 s 
0   s / d   1.0 
 
 
As indicated in Equation 4, ASCE/SEI 41 specifies that if the spacing of column transverse reinforcement is
larger than half of the effective depth of column section but less than effective depth (d/2< s ≤ d) in the direction of
applied shear force, the transverse reinforcement should be taken as no more than 50% effective when resisting
shear or torsion, and thus Vs is set to Vs = 0.5Avfyd / s. If the spacing is larger than the effective depth of the column
section (s > d), the transverse reinforcement shall be considered ineffective and Vs =0. This modification of Vs is
not included in the original model by Sezen and Moehle (2004).

Su (2007) observed that the shear strengths for shear-critical columns in the NCREE 2007shake table tests
were higher than for columns subjected to quasi-static test with similar configurations. This difference may be
attributed to the higher strain rate and fewer pre-failure cycles that shear-critical columns experienced during the
dynamic tests. This observation suggests that the Sezen and Moehle model may be conservative for the columns in
the Dynamic Database as this original model was developed based on a database of quasi-static cyclic tests.

Based on the calculated shear strength, V0, and plastic shear demand, Vp, ASCE/SEI 41 classifies columns
with 135-degree hooks on transverse reinforcement into three conditions: condition i, ii and iii. These conditions are
intended to correspond to the following failure modes:

 Condition i: flexure failure;


 Condition ii: flexure-shear failure, and;
 Condition iii: shear failure

Table 3 ASCE/SEI 41 column classification

Transverse Reinforcement Details


ACI Conforming Closed hoops with 90° Other (including lap
details with 135° hooks hooks spliced trans. rein.)
Vp/V0 ≤ 0.6 i ii ii
0.6 < Vp/V0 ≤ 1.0 ii ii iii
Vp/V0 >1.0 iii iii iii

Poor transverse reinforcement detailing reduces the column lateral deformation capacity and therefore,
columns with transverse reinforcement including 90-degree hooks or having lap spliced transverse reinforcement are
adjusted to a poorer condition (e.g., from condition i to condition ii) as shown in Table 3.

The accuracy of the ASCE/SEI 41 classification method is evaluated using the matrix shown in Table 4. Only
columns that were observed to experience column failure (i.e. 20% drop in lateral resistance) during the tests are
included. The data along the diagonal in the blue cells indicates the number of columns that are classified properly
by ASCE/SEI 41. The column data that falls in the green cells provides a “conservative” estimation of the column
failure type in terms of column drift capacity (for example “23” indicates that there were 23 columns that are
classified as shear-critical columns, yet were observed to have experienced flexure-shear failures during the tests).

2.8
@Seismicisolation
@Seismicisolation
Assessment of ASCE/SEI 41 Concrete Column Provisions using Shaking Table Tests

While “green zone” classification may result in unnecessary retrofit costs, it is unsafe and hence unacceptable to
have column tests falling in the red cells in Table 4. Results in Table 4 show that ASCE/SEI 41 is over-conservative
since the 23 of the 27 flexure-shear-critical columns are categorized in Condition iii.

Table 4 Evaluation of ASCE/SEI 41 column classification method

ASCE/SEI 41 Classification

  Flexure Flexure-Shear Shear


Total
(Condition i) (Condition ii) (Condition iii)
Flexure 0 2 0 2 
Failure Type
Observed

Flexure-Shear 0 4 23 27 
Shear 0 0 5 5 
Total 0 6 28 34 

Table 5 Evaluation of modified ASCE/SEI 41 classification method

Modified ASCE/SEI 41 Classification


Flexure Flexure-Shear Shear
Total
(Condition i) (Condition ii) (Condition iii)
Flexure 0 2 0 2
Failure Type
Observed

Flexure-Shear 0 27 0 27
Shear 0 0 5 5
Total 0 29 5 34

To provide a better match with the observed failure modes from the Dynamic Database, two modifications to
the classification methodology in ASCE/SEI 41 are proposed. First, the shear strength, V0, is not reduced for widely
spaced ties (d > s > d/2). This is consistent with the development of the shear strength model in Sezen and Moehle
(2004) where columns with widely spaced ties are included in the database without modification. Note that the
Dynamic Database does not include any columns with s>d, hence the impact of the penalty for such large spacing
cannot be evaluated here. It is considered prudent to maintain this limit considering the potential for a diagonal
crack not crossing any hoops for such wide spacing.

Secondly, it is found that most of the flexure-shear-critical columns that are misclassified as Condition iii
have ratio Vp /V0 only slightly over unity. If the criterion for flexure-shear-critical columns is modified to
1.1≥ Vp /V0 >0.6 instead of 1.0≥ Vp /V0 >0.6, the classification improves substantially. This relaxation of the upper
limit on the shear ratio for Condition ii reflects the fact that columns with plastic shear demand very close to the
shear capacity are expected to exhibit some limited plastic deformation capacity even if post yield deformations are
dominated by shear deformations. The relaxation of the upper limit is also consistent with the observation noted
above that columns subjected to dynamic loads tend to exhibit slightly higher shear strengths (Su 2007).

As shown in Table 5, by using the unreduced column shear strength and the modified upper limit on Vp /V0
for condition ii, flexure-shear-critical columns could be properly identified, resulting in more accurate modeling
parameters and acceptance criteria, and subsequently, a more economical retrofit solution. The modified
classification criteria are used to classify columns and evaluate the recommended modeling parameters in the
following section.

MODELING PARAMETERS FOR CONCRETE COLUMNS

By classifying columns into the aforementioned three conditions based on failure modes, ASCE/SEI 41
provides recommendations on modeling parameters and allowable plastic rotations for columns in each condition.
As shown in Figure 1, the modeling parameters a and b are provided to represent the plastic rotation at significant
loss of lateral resistance and loss of axial load resistance, respectively. Plastic Rotation Ratio (PRR), calculated as
the ratio of measured plastic rotation to the modeling parameter, is adopted here to assess the accuracy of the
ASCE/SEI 41 modeling parameters.

2.9
@Seismicisolation
@Seismicisolation
Y. Li et al.

Assessment of Modeling Parameter a: loss of significant lateral resistance

Non-ductile columns with inadequate transverse reinforcement are vulnerable to damage associated with
shear failure when subjected to seismic loads. Consistent with the definition of modeling parameter a in ASCE/SEI
41, the column shear failure point is taken as the point that column lateral resistance degrades to 80% of its peak
lateral resistance. Most drift capacity models available also have been developed to provide similar estimations of
column drift at 80% of maximum effective force (Elwood and Moehle 2005; Kato and Ohnishi 2002, etc.).

The measured plastic rotation capacity was taken equal to the measured drift ratio corresponding to a 20%
reduction in the maximum measured shear resistancel) minus the calculated yield drift ratio (y) using the
recommended effective stiffness from ASCE/SEI 41 and the plastic shear demand, Vp, as defined previously. Hence,
the plastic rotation ratio at loss of significant lateral resistance is defined as:
Plastic Rotation Ratio (PRR) = (l - y)/a Equation 5

Figure 7 shows the PRR at loss of significant lateral resistance for columns in the Dynamic Database. Data in
solid markers provide the PRR for columns that experienced failure (20% loss in lateral resistance) during the test,
while the hollow markers represent the lower bound of PRR based on the maximum recorded drifts for columns that
did not fail in the shake table test. It is noted that all flexure-critical columns (Condition i) did not experience
significant loss of lateral strength during the shaking table tests. The red solid diamond marker (Condition ii)
represents the PRR for flexure-shear critical columns that were connected by a rigid beam in the reinforced concrete
frame specimens. The PRR of flexure-shear-critical columns that were connected by flexible beam are represented
by purple solid diamond marker, namely “Condition ii (Beam)”. In such cases the measured drift ratio is influenced
by the flexibility of the beams and joints, indicating that the resulting PRR is likely higher than the true value for the
column alone and should be evaluated with caution.

To compare the PPR for columns subjected to dynamic loads and static cyclic loads, 50 non-ductile columns
subjected to quasi-static cyclic lateral loads are selected from the PEER Structural Performance Database (Berry et
al. 2004) to form a Static Dataset A (detailed column properties could be found in Appendix B). Figure 8 plots the
PRR at loss of significant lateral resistance for columns in Static Dataset A.
Condition i(lower) Condition ii Condition ii (Beam) Condition ii (lower)
5
PPR ( Plastic Rotation Ratio)

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Initial Axial Load Ratio (Pini /Agfc')
Figure 7 Plastic rotation ratio at significant loss of lateral resistance for Dynamic Database
Condition ii
5
PPR (Plastic Rotation Ratio)

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Axial load ratio (P/Agfc ')
Figure 8 Plastic rotation ratio at significant loss of lateral resistant force for Static Dataset A

2.10
@Seismicisolation
@Seismicisolation
Assessment of ASCE/SEI 41 Concrete Column Provisions using Shaking Table Tests

Significant scatter exists among the data for PRR in Figure 7 and Figure 8, emphasizing the need to
determine the level of safety provided by ASCE/SEI 41 provisions. Results of plastic rotation ratio can be
represented by a lognormal cumulative distribution, commonly referred to as a fragility curve. The fragility curve in
Figure 9 focuses on the probability of failure for flexure-shear critical columns (i.e. Condition ii). In Figure 9(a), the
fragility curve is constructed based on the test data shown in red diamonds in Figure 7 (Condition ii columns
connected to rigid beams and observed to experience flexure-shear failures). The fragility curves for flexure-shear
critical columns in the Dynamic Database and Static Dataset A are compared in Figure 9 (b).

1 1

0.9 0.9

(Loss of Significant Lateral Resistance)


0.8 0.8

0.7 0.7
Probability of Failure (Pf)

Probability of Failure
0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2
From Test Data Dynamic Database
0.1 0.1 Static Dataset A
Lognormal CDF
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Plastic rotation ratio Plastic Rotation Ratio
(a) (b)
Figure 9 Lognomal fragility curve for plastic rotation at significant loss of lateral resistance in
Dynamic Database and Static Dataset A

According to the target safety levels selected in the development of the ASCE/SEI 41 tables (as reported in
Elwood et al 2007), at a plastic rotation demand equal to parameter a the probability of failure, Pf, should be less
than 15% for columns vulnerable to shear failure, (i.e. columns in condition ii and iii); while for flexure-critical
columns in condition i, Pf up to 35% is acceptable since flexure failure is less sudden and brittle. (The “failure” here
refers to the point at which the column has lost more than 20% of its lateral resistance.) Based on Figure 9 (a), there
is a probability of failure of 3% for “Condition ii” columns in the Dynamic Database when subjected to plastic
rotation demand equal to parameter a specified in ASCE/SEI 41 (PRR=1), and thus, is lower than the ASCE/SEI 41
targets. As shown in Figure 9 (b), the probability of failure at PRR =1 for columns in the Static Dataset A is more
than four times that of the Dynamic Database, but still consistent with the ASCE/SEI 41 targets.

Most quasi-static cyclic tests are displacement-controlled (column specimens are typically subjected to two
or three reversed lateral force-deformation cycles at each horizontal drift ratio level until failure) and commonly
include more cycles at high displacement demands than expected from real ground motions used in the shaking table
tests. During multiple cycles in quasi-static cyclic tests, diagonal shear cracks open and close repeatedly, increasing
the column shear damage and thus resulting in faster degradation of column lateral resistance. This provides a
possible explanation for why the fragility curve for columns subjected to shaking table tests shifts to lower
probabilities of failure compared to that from quasi-static cyclic tests.

For columns expected to experience shear failure prior to flexural yielding (Condition iii), plastic deformations
cannot be relied upon prior to shear failure according to ASCE/SEI 41; and hence, parameter a is set to zero and the
drift ratio at lateral load failure is taken as V0/keff. Hence, the Total Rotation Ratio (TRR, Equation 6) is used here in
place of the Plastic Rotation Ratio to assess the level of conservatism provided by ASCE/SEI 41 for Condition iii
columns.
Total Rotation Ratio = l / (V0/keff Equation 6

2.11
@Seismicisolation
@Seismicisolation
Y. Li et al.

It is observed that all Condition iii columns exhibited TRR larger than unity, indicating that the ASCE/SEI 41
limits for shear-critical columns are conservative. Since only five shear-critical columns are included in the database,
the data are not sufficient to construct a fragility curve or recommend revisions to ASCE/SEI 41.

Condition iii
5
TPR ( Total Rotation Ratio)

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Initial Axial Load Ratio (Pini /Agfc')

Figure 10 Total Rotation Ratio at significant loss of lateral resistance for Condition iii columns in Dynamic
Database

Assessment of Modeling Parameter b: loss of axial-load support

Modeling parameter b in ASCE/SEI 41 represents the plastic rotation at the point that a column experiences
axial-load failure and loses axial-load support. While it is intuitive to define the axial-load failure based on the axial
load response, load redistribution between columns in frame systems and column elongation/shortening due to
imposed lateral deformations can significantly influence the axial load data, rendering it difficult to capture the true
point of column axial-load failure from frame tests such as those in the Dynamic Database. For consistency among
all tests, this research adopts the method used in Yavari (2011) to define the column axial-load failure point as “the
point at which the peak vertical elongation of the column is recorded immediately prior to a sudden shortening of the
column”. Yavari (2011) found that this definition resulted in a failure point corresponding to a sudden drop in axial
load and visual failure of the column in videos. An example of the application of this definition is shown in Figure
11.

Figure 11 Interpretation of column axial-load failure

2.12
@Seismicisolation
@Seismicisolation
Assessment of ASCE/SEI 41 Concrete Column Provisions using Shaking Table Tests

The measured plastic rotation capacity at axial-load failure was taken equal to the measured drift ratio at loss
of axial-load support a) minus the calculated yield drift ratio using the recommended effective stiffness from
ASCE/SEI 41 and the plastic shear demand, Vp, as defined previously. Hence, the plastic rotation ratio at loss of
axial load support is defined as:
Plastic Rotation Ratio (PRR) = (a - y)/b Equation 7

To compare the plastic rotation at axial-load failure for columns subjected to quasi-static cyclic tests and
shaking table tests, Static Dataset B consisting of 21 Condition ii columns is introduced and summarized in
Appendix C. These columns are subjected to unidirectional displacement cycles and constant axial load and
experience axial-load failure after shear failure. Static Dataset A could not be used for this assessment since only a
small subset of these tests experienced axial load failure. Figure 12 and Figure 13 plot the PRR (Plastic Rotation
Ratio) at loss of axial-load support for columns in the Dynamic Database and Static Dataset B, respectively.

Condition i(lower) Condition ii Condition ii (Beam) Condition ii (lower) Condition iii


5
PPR ( Plastic Rotation Ratio)

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Initial Axial Load Ratio (Pini /Agfc ')
Figure 12 Plastic rotation ratio at loss of axial-load support for Dynamic Database

Condition ii
5
PPR (Plastic Rotation Ratio)

0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Axial load ratio (P/Agfc ')
Figure 13 Plastic rotation ratio at loss of axial-load support for Static Dataset B

Similar to the previous section for modeling parameter a, results of plastic rotation ratio for Condition ii
columns in Dynamic Database are represented by a fragility curve, as shown in Figure 14 (a). The fragility curve is
constructed based on the test data of flexure-shear-critical columns that experienced axial-load failure during the
tests (data in red diamond marker in Figure 12).

2.13
@Seismicisolation
@Seismicisolation
Y. Li et al.

1 1

0.9 0.9

0.8 0.8

0.7 0.7

(Loss of Axial Load Support)


Probability of Failure (Pf)

Probability of Failure
0.6 0.6

0.5 0.5

0.4 0.4

0.3 0.3

0.2 0.2
Test Data Static Dataset B
0.1 0.1
Fragility curve Dynamic Database
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Plastic rotation ratio Plastic Rotation Ratio
Figure 14 Lognomal fragility curve for plastic rotation at loss of axial load support in Dynamic
Database and Static Dataset B

According to the target safety levels selected in the development of the ASCE/SEI 41 tables, when columns
experience plastic rotation demand equal to the specified parameter b, the probability of failure, Pf, should be less
than 15% (Elwood et al. 2007). (“Failure” here refers to the point that the column loses axial-load support.) Figure
14 (a) shows a 9% probability of failure for PRR=1.0, indicating that the probability of failure at loss of axial-load
support for flexure-shear-critical columns in the Dynamic Database is consistent with ASCE/SEI 41 targets. The
probability of failure for PRR = 1.0 for Static Dataset B is slightly higher, at 11%, as shown in Figure 14 (b). The
results indicate that the level of conservatism against axial load failure in ASCE/SEI 41 is consistent regardless of
the dataset used in the assessment, suggesting that axial load failure is not sensitive to the form of load application
(dynamic or static) or the number of cycles.

Considering the conclusions for the drift at lateral load failure (drifts from dynamic tests greater than drifts
from static tests) and axial load failure (drifts from dynamic tests similar to drifts from static tests) together, it is
noted that the gap between shear failure and axial load failure for columns subjected to shaking table tests is smaller
than that for columns in static cyclic tests. Indeed 56% of the flexure-shear-critical columns in the Dynamic
Database experience axial-load failure at a drift ratio within +/-0.1% of the drift at 20% drop in the lateral resistance.
This limited “buffer” between lateral and axial load failure for the Dynamic Database columns may be a result of
dynamic amplification of drift demands with the development of a negative stiffness (softening) and a rapid
acceleration of response after lateral load failure. Caution should be exercised in interpreting this observation,
however, since the specimens in the database included only a limited number of columns and the dynamic response
characteristics were heavily influenced by strength degradation in any one column. In a larger building structure,
particularly one with walls, failure of one column will not typically result in the same degree of softening and
associated dynamic amplification of drift demands.

CONCLUSIONS

Acceptance criteria in codes and standards (e.g. ASCE 41-13) have generally been developed and validated
based on comparisons with test data from quasi-static tests. To investigate the influence of real earthquake loading
on the performance of concrete columns a database consisting of 59 large-scale reinforced concrete columns tested
on a shaking table is compiled. This “Dynamic Database” is used to evaluate the accuracy and conservatism of the
concrete column provisions in ASCE/SEI 41-13.

The conclusions of this study include the following:

 ASCE/SEI 41 overestimates the effective stiffness ratio for columns in Dynamic Database. The mean
value of EIMeas/EIASCE for flexural-shear-critical columns connected with rigid beam in Dynamic

2.14
@Seismicisolation
@Seismicisolation
Assessment of ASCE/SEI 41 Concrete Column Provisions using Shaking Table Tests

Database is 0.65, with coefficient of variation of 24%. This result is consistent with the observation
for columns subjected to quasi-static cyclic test.

 Based on the current criteria of ASCE/SEI 41, most of the flexure-shear-critical columns in the
Dynamic Database are categorized into Condition iii (pure shear failure). This is too conservative,
since Condition iii columns are assumed to have zero plastic rotation capacity prior to shear failure.
The following modifications to the ASCE/SEI 41 column classification criteria are recommended: (1)
remove the penalty on column shear strength (V0) for columns transverse reinforcement spaced at
greater than d/2, while maintaining the penalty (Vs = 0) for s>d; (2) change the criteria for Condition ii
from 0.6<Vp/V0≤1.0 to 0.6<Vp/V0≤1.1.

 Flexure-shear critical columns tend to develop larger plastic drifts at loss of significant lateral
resistance in shaking table tests, compared with columns subjected to quasi-static cyclic loads. Based
on Condition ii columns in the Dynamic Database, modeling parameter a in ASCE/SEI 41 provided a
3% probability of failure at significant loss of lateral resistance, which is significantly lower than the
probability of failure assessed based on columns subjected to quasi-static cyclic loads.

 Flexure-shear critical columns tend to develop similar plastic drifts at loss of axial-load support in
shaking table tests, compared with columns subjected to quasi-static cyclic loads. Based on Condition
ii columns in the Dynamic Database, modeling parameter b in ASCE/SEI 41 provided a similar
probability of failure (9%) at loss of axial-load support compared to that assessed using columns
subjected to quasi-static cyclic loads.

 Dynamic tests tend to result in a smaller difference between the drifts at lateral-load and axial-load
failure compared with static cyclic tests. For systems where strength deterioration of columns may
dominate the dynamic response of the structure, this observation suggests that real earthquakes may
result in more rapid degradation of lateral resistance for flexure-shear columns after shear failure than
had been previously assumed based on quasi-static cyclic tests.

REFERENCES

ASCE 2008. Seismic Rehabilitation of Existing Buildings. ASCE/SEI 41, Supplement 1. Reston,VA: American
Society of Civil Engineers.

ASCE 2013. Seismic Evaluation and Retrofit of Existing Buildings. ASCE/SEI 41. Reston,VA: American Society of
Civil Engineers.

Berry, M. Parrish, M. and Eberhard, M. 2004. PEER Structural Performance Database User’s Manual, Pacific
Earthquake Engineering Research Center, University of California, Berkeley. [available at:
www.ce.washington.edu/~peera1].

EN 1998-3. 2005. Eurocode 8: Design of structures for earthquake resistance – Part 3: Assessment and retrofitting
of buildings, European Committee for Standardization, Brussels.

Elwood, K. 2002. Shaking Table Tests and Analytical Studies on the Gravity Load Collapse of Reinforced Concrete
Frames. PhD Dissertaiton, University of California, Berkeley.

Elwood, K., Matamoros, A., Wallace, J., Lehman, D., Heintz, J., Mitchell, A., Moore, M., Valley, M. and Lowes, L.
2007. Update to ASCE/SEI 41 Concrete Provisions. Earthquake Spectra, 23(3), 493-523.

Elwood, K. J. and Eberhard, M. O. 2009. Effective Stiffness of Reinforced Concrete Columns. ACI Structural
Journal, 106 (4), 476-484.

2.15
@Seismicisolation
@Seismicisolation
Y. Li et al.

Elwood, K. J. and Moehle, J. P. 2005.Drift Capacity of Reinforced Concrete Columns with Light Transverse
Reinforcement. Earthquake Spectra, 21(1), 71-89.

Elwood, K.J., and Moehle, J.P., 2006. “Idealized backbone model for existing reinforced concrete columns and
comparisons with FEMA 356 criteria”, Structural Design of Tall and Special Buildings, 15(5), 553-569.

Ghannoum, W. M. 2007. Experimental and Analytical Dynamic Collapse Study of a Reinforced Concrete Frame
with Light Transverse Reinforcement. Ph.D. Dissertation, University of California, Berkeley.

Haselton, C. B., Liel, A. B., Taylor Lange, S., and Deierlein, G. G. (2008). Beam-Column Element Model
Calibrated for Predicting Flexural Response Leading to Global Collapse of RC Frame Buildings. Pacific
Earthquake Engineering Research Center, Berkeley, CA.

Kato, D., and Ohnishi, K. 2002. Axial Load Carrying Capacity of R/C Columns under Lateral Load Reversals. Third
US-Japan Workshop on Performance-Based Earthquake Engineering Methodology for Reinforced concrete
Building Structures, Seattle, WA, PEER Report 2002/02.

Kuo, W. W. 2008. Study on the collapse behavior of nonductile reinforced concrete frames subjected to earthquake
loading. Ph.D. Dissertation, Department of Construction Engineering, National Taiwan University of
Science and Technology, Taipei, Taiwan, 542pp.

Li, Y. 2012. The shaking table tests column database and evaluation of drift capacity models for non-ductile
columns. Master Thesis, University of British Columbia.

Lynn, A.C. 2001. Seismic evaluation of existing reinforced concrete building columns. Ph.D. Dissertation,
Department of civil and Environmental Engineering, University of California, Berkeley.

Panagiotakos T.B. and Fardis, M.N., 2001 "Deformations of RC Members at Yielding and Ultimate", ACI Structural
Journal, 98(2), 135-148.

Sezen, H. 2002. Seismic response and modeling of reinforced concrete building columns, Ph.D. Dissertation,
Department of Civil and Environmental Engineering, University of California, Berkeley.

Sezen, H. and Moehle, J. 2004. Shear strength model for lightly reinforced concrete columns. ASCE Journal of
Structural Engineering, 130(11), 1692-1703.

Shin, Y. B. 2007. Dynamic Response of Ductile and Non-Ductile Reinforced Concrete Columns. Ph.D. Dissertation,
University of California, Berkeley.

Su, R.S. 2007. Shake table tests on reinforced concrete short columns failed in shear. Master thesis, Department of
Construction Engineering, National Taiwan University of Science and Technology, Taipei, Taiwan, 195pp.

Wu C.L., Yang Y.S. and Loh C.H. 2006. Dynamic Gravity Load Collapse of Non-ductile RC Frames I :
Experimental Approach. 8th US National Conference on Earthquake Engineering - the 100th Anniversary
Earthquake Conference Commemorating the 1906 San Francisco Earthquake, San Francisco, Moscone
Center, April 18-22.

Yavari, S., Elwood, K. J. and Wu, C. L. 2009. Collapse of a non-ductile concrete frame: Evaluation of analytical
models. Earthquake Engineering and Structural Dynamics, 38(2), 225-241.

Yavari, S. 2011. Shaking table tests on the response of reinforced concrete frames with non-seismic detailing. Ph.D.
Dissertation, University of British Columbia.

2.16
@Seismicisolation
@Seismicisolation
Assessment of ASCE/SEI 41 Concrete Column Provisions using Shaking Table Tests

APPENDIX A: PROPERTIES OF THE SHAKING TABLE TESTS COLUMN DATABASE

Column Information Geometry Long. Rein. Trans. Rein. Column Drift*


Failure
D b L f c' dl c fyl dt Tie s fyt ρt ∆80%/L ∆a/L
Specimen Column Scale ρl Mode
  (mm) (mm) (mm) (MPa) (mm) (mm) (MPa) (mm) Type (mm) (MPa) (%) (%) (%)
NCREE 2009 Tests (Yavari 2011)
1 MCFS A1 1/2.25 200 200 1400 34.0 12.7 17 439.0 2.53% 5 r90 120 469.0 0.16% 3.59 3.59 FS
2 MCFS B1 1/2.25 200 200 1400 34.0 12.7 17 439.0 2.53% 5 r90 120 469.0 0.16% 2.15 2.15 FS
3 MCFS C1 1/2.25 200 200 1400 34.0 12.7 17 439.0 2.53% 5 r90 120 469.0 0.16% 2.39 2.39 FS
4 HCFS A1 1/2.25 200 200 1400 34.4 12.7 17 439.0 2.53% 5 r90 120 469.0 0.16% N/A 3.87 FS
5 HCFS B1 1/2.25 200 200 1400 34.4 12.7 17 439.0 2.53% 5 r90 120 469.0 0.16% 2.59 2.59 FS
6 HCFS C1 1/2.25 200 200 1400 34.4 12.7 17 439.0 2.53% 5 r90 120 469.0 0.16% 4.09 4.10 FS
7 MUF A1 1/2.25 200 200 1400 35.8 12.7 17 467.0 2.53% 5 r135 40 475.0 0.49% N/A 2.97 F
8 MUF B1 1/2.25 200 200 1400 35.8 12.7 17 467.0 2.53% 5 r135 40 475.0 0.49% N/A 2.80 F
9 MUF C1 1/2.25 200 200 1400 35.8 12.7 17 467.0 2.53% 5 r135 40 475.0 0.49% N/A 2.82 F
10 MUFS A1 1/2.25 200 200 1400 36.5 12.7 17 467.0 2.53% 5 r90 120 475.0 0.16% N/A 2.96 FS
11 MUFS B1 1/2.25 200 200 1400 36.5 12.7 17 467.0 2.53% 5 r90 120 475.0 0.16% N/A 2.75 FS
12 MUFS C1 1/2.25 200 200 1400 36.5 12.7 17 467.0 2.53% 5 r90 120 475.0 0.16% N/A 2.80 FS
NCREE 2007 Tests (Su 2007)
13 S1 C1 1/2 250 250 1000 29.9 12.7 16 436.4 1.62% 4 r135 50 643.3 0.30% N/A 5.93 FS
14 S1 C2 1/2 250 250 1000 29.9 12.7 16 436.4 3.24% 4 r90 150 643.3 0.07% 1.41 3.35 S
15 S1 C3 1/2 250 250 1000 29.9 12.7 16 436.4 1.62% 4 r135 50 643.3 0.30% N/A 3.37 FS
16 S3 C1 1/2 250 250 750 29.9 12.7 16 436.4 3.24% 4 r90 150 643.3 0.07% 1.19 1.28 S
17 S3 C2 1/2 250 250 750 29.9 12.7 16 436.4 3.24% 4 r90 150 643.3 0.07% 0.76 1.44 S
18 S4 C1 1/2 250 250 1000 29.9 12.7 16 436.4 3.24% 4 r90 150 643.3 0.07% 0.91 0.91 S
19 S4 C2 1/2 250 250 1000 29.9 12.7 16 436.4 3.24% 4 r90 150 643.3 0.07% 1.22 1.22 S
NCREE 2005 Tests (Kuo 2008)
20 P2 C1 1/3 150 150 1000 33.8 9.53 14 470.9 2.53% 3.2 r90 100 548.1 0.16% 4.17 4.98 FS
21 P2 C2 1/3 150 150 1000 33.8 9.53 14 470.9 2.53% 3.2 r90 100 548.1 0.16% 4.22 4.98 FS
22 P2 C3 1/3 150 150 1000 33.8 6 12 231.7 1.39% 5 r135 33 661.9 1.19% N/A 11.00 F
23 P2 C4 1/3 150 150 1000 33.8 6 12 231.7 1.39% 5 r135 33 661.9 1.19% N/A 11.00 F
24 L C1 1/3 150 150 1000 32.3 9.53 14 470.9 2.53% 3.2 r90 100 548.1 0.16% 3.34 8.50 FS
25 L C2 1/3 150 150 1000 32.3 9.53 14 470.9 2.53% 3.2 r90 100 548.1 0.16% 3.18 8.50 FS

2.17
@Seismicisolation
@Seismicisolation
Y. Li et al.

Column Information Geometry Long. Rein. Trans. Rein. Column Drift*


Failure
D b L f c' dl c fyl dt Tie s fyt ρt ∆80%/L ∆a/L
Specimen Column Scale ρl Mode
  (mm) (mm) (mm) (MPa) (mm) (mm) (MPa) (mm) Type (mm) (MPa) (%) (%) (%)
26 L C3 1/3 150 150 1000 32.3 6 12 231.7 1.39% 5 r135 33 661.9 1.19% N/A 8.50 F
27 L C4 1/3 150 150 1000 32.3 6 12 231.7 1.39% 5 r135 33 661.9 1.19% N/A 8.50 F
NCREE 2004 Tests (Wu et al. 2006)
28 S2 C1 1/2 200 200 1730 31.8 9.53 9.5 385.9 1.43% 4 r90 125 385.9 0.10% 2.83 4.31 FS
29 S2 C2 1/2 200 200 1730 31.8 9.53 9.5 385.9 1.43% 4 r90 125 385.9 0.10% 4.30 4.31 FS
30 S3 C1 1/2 200 200 1730 31.8 9.53 9.5 385.9 1.43% 4 r90 125 385.9 0.10% 4.39 4.49 FS
31 S3 C2 1/2 200 200 1730 31.8 9.53 9.5 385.9 1.43% 4 r90 125 385.9 0.10% 4.49 4.49 FS
Ghannoum 2006 Tests (Ghannoum 2007)
32 S1 A1 1/3 152 152 991 24.6 9.53 17 441.3 2.45% 3.18 r90 102 655.0 0.15% 4.16 4.66 FS
33 S1 B1 1/3 152 152 991 24.6 9.53 17 441.3 2.45% 3.18 r90 102 655.0 0.15% 4.72 5.11 FS
34 S1 C1 1/3 152 152 991 24.6 6.35 17 482.6 1.09% 4.76 r135 31.8 556.3 1.10% N/A 5.05 F
35 S1 D1 1/3 152 152 991 24.6 6.35 17 482.6 1.09% 4.76 r135 31.8 556.3 1.10% N/A 5.58 F
Elwood 2002 Tests (Elwood 2003)
36 S1 CenC 1/2 229 229 1473 24.6 12.7 25 479.3 2.48% 4.76 d90 152 689.7 0.18% 5.27 6.07 FS
37 S2 CenC 1/2 229 229 1473 23.9 12.7 25 479.3 2.48% 4.76 d90 152 689.7 0.18% 3.43 3.81 FS
Shin 2005 Tests (Shin 2007)
38 Setup I EastC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 5.39 F
39 Setup I WestC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 5.39 F
40 Setup I EastC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 8.07 F
41 Setup I WestC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 8.07 F
42 Setup I EastC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 4.84 F
43 Setup I WestC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 4.84 F
44 Setup I EastC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 5.95 F
45 Setup I WestC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 5.95 F
46 Setup II EastC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 6.94 F
47 Setup II WestC 1/3 152 152 991 25.6 9.53 13.8 444.1 2.46% 3.17 r90 102 554.5 0.15% 2.46 4.26 FS
48 Setup II EastC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 8.93 F
49 Setup II WestC 1/3 152 152 991 25.6 9.53 13.8 444.1 2.46% 3.17 r90 102 554.5 0.15% 4.36 4.36 FS
50 Setup II EastC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 6.93 F

2.18
@Seismicisolation
@Seismicisolation
Assessment of ASCE/SEI 41 Concrete Column Provisions using Shaking Table Tests

Column Information Geometry Long. Rein. Trans. Rein. Column Drift*


Failure
D b L f c' dl c fyl dt Tie s fyt ρt ∆80%/L ∆a/L
Specimen Column Scale ρl Mode
  (mm) (mm) (mm) (MPa) (mm) (mm) (MPa) (mm) Type (mm) (MPa) (%) (%) (%)
51 Setup II WestC 1/3 152 152 991 25.6 9.53 13.8 444.1 2.46% 3.17 r90 102 554.5 0.15% 3.40 3.40 FS
52 Setup II EastC 1/3 152 152 991 25.6 9.53 12.2 444.1 2.46% 4.76 r135 38.1 549.0 0.92% N/A 6.42 F
53 Setup II WestC 1/3 152 152 991 25.6 9.53 13.8 444.1 2.46% 3.17 r90 102 554.5 0.15% 3.41 3.41 FS
54 Setup III EastC 1/3 152 152 991 25.6 9.53 13.8 444.1 2.46% 3.17 r90 102 554.5 0.15% 2.47 2.47 FS
55 Setup III WestC 1/3 152 152 991 25.6 9.53 13.8 444.1 2.46% 3.17 r90 102 554.5 0.15% 2.26 3.83 FS
56 Setup III EastC 1/3 152 152 991 25.6 9.53 13.8 444.1 2.46% 3.17 r90 102 554.5 0.15% 3.16 3.35 FS
57 Setup III WestC 1/3 152 152 991 25.6 9.53 13.8 444.1 2.46% 3.17 r90 102 554.5 0.15% 4.28 4.28 FS
58 Setup III EastC 1/3 152 152 991 25.6 9.53 13.8 444.1 2.46% 3.17 r90 102 554.5 0.15% 3.16 3.16 FS
59 Setup III WestC 1/3 152 152 991 25.6 9.53 13.8 444.1 2.46% 3.17 r90 102 554.5 0.15% 2.43 2.48 FS
* For flexure-critical columns (Type “F” columns) and flexure-shear-critical columns (Type “FS” columns) that did not experience 20% degradation in lateral
resistance during the test, ∆80%/L is noted as “N/A” and ∆a/L is the maximum horizontal drift measured during the tests.

Note: a = the shear span (equals to L/2 for fixed-fixed columns); b = width of column section; c = clear cover (measured from the outer face of the concrete to
outer face of longitudinal reinforcement); D = height of column section; dl = the diameter of longitudinal reinforcement; dt = the diameter of transverse
reinforcement; fc’= concrete compressive stress at test day; fyl = yield stress of longitudinal reinforcement; fyt = yield stress of transverse reinforcement; L= clear
height of the column; s = spacing of transverse reinforcement; ρl = longitudinal reinforcement ratio; ρt = volume transverse reinforcement ratio; Tie Type = the
type of the transverse reinforcement, “r135” means rectangular-shaped hoops with 135-degree hooks and “d90” means diamond-shaped transverse reinforcement
with 90-degree hooks; s = spacing of transverse reinforcement; ∆80% = column displacement at 80% of peak lateral resistance.

2.19
@Seismicisolation
@Seismicisolation
Y. Li et al.

APPENDIX B: CONDITION II COLUMNS IN STATIC DATASET A

Aspect
b D L f’c fyl s fyt Δ80%/L
Specimens Ratio ρl ρt P/Agf’c
(×10-2)
(mm) (mm) (mm) (MPa) (MPa) (mm) (MPa)
(a/D)
Bechtoula, Kono, Arai and Watanabe, 2002,
D1N30
250 250 625 37.6 2.50 0.0243 461 40 0.0050 485 0.30 3.96
Esaki, 1996 H-2-1/3 200 200 400 23.0 2.00 0.0253 363 40 0.0065 364 0.33 1.99
Esaki, 1996 H-2-1/5 200 200 400 23.0 2.00 0.0253 363 50 0.0052 364 0.20 2.52
Esaki, 1996 HT-2-1/3 200 200 400 20.2 2.00 0.0253 363 60 0.0065 364 0.33 2.49
Esaki, 1996 HT-2-1/5 200 200 400 20.2 2.00 0.0253 363 75 0.0052 364 0.20 2.94
Galeota et al. 1996, AB2 250 250 1140 80.0 4.56 0.0603 579 150 0.0054 579 0.30 4.02
Galeota et al. 1996, AB3 250 250 1140 80.0 4.56 0.0603 579 150 0.0054 579 0.30 3.74
Galeota et al. 1996, AB4 250 250 1140 80.0 4.56 0.0603 579 150 0.0054 579 0.20 4.06
Lynn et al. 1998, 3CMD12 457.2 457.2 1473 27.6 3.22 0.0303 331 305 0.0017 400 0.26 1.74
Nosho et al. 1996, No. 1 279.4 279.4 2134 40.6 7.64 0.0101 407 229 0.0010 351 0.34 1.61
Ohue et al. 1985, 2D16RS 200 200 400 32.0 2.00 0.0201 369 50 0.0048 316 0.14 3.83
Ohue et al. 1985, 4D13RS 200 200 400 29.9 2.00 0.0265 370 50 0.0048 316 0.15 1.78
Pujol 2002, No. 10-1-2.25N 152.4 304.8 686 36.5 2.25 0.0245 453 57 0.0073 411 0.08 3.21
Pujol 2002, No. 10-1-2.25S 152.4 304.8 686 36.5 2.25 0.0245 453 57 0.0073 411 0.08 3.14
Pujol 2002, No. 10-2-2.25N 152.4 304.8 686 34.9 2.25 0.0245 453 57 0.0073 411 0.08 3.21
Pujol 2002, No. 10-2-2.25S 152.4 304.8 686 34.9 2.25 0.0245 453 57 0.0073 411 0.08 3.17
Pujol 2002, No. 10-2-3N 152.4 304.8 686 33.7 2.25 0.0245 453 76 0.0055 411 0.09 3.19
Pujol 2002, No. 10-2-3S 152.4 304.8 686 33.7 2.25 0.0245 453 76 0.0055 411 0.09 3.05
Pujol 2002, No. 10-3-2.25N 152.4 304.8 686 27.4 2.25 0.0245 453 57 0.0073 411 0.10 3.05
Pujol 2002, No. 10-3-2.25S 152.4 304.8 686 27.4 2.25 0.0245 453 57 0.0073 411 0.10 3.22
Pujol 2002, No. 10-3-3N 152.4 304.8 686 29.9 2.25 0.0245 453 76 0.0055 411 0.10 3.13
Pujol 2002, No. 10-3-3S 152.4 304.8 686 29.9 2.25 0.0245 453 76 0.0055 411 0.10 3.15
Pujol 2002, No. 20-3-3N 152.4 304.8 686 36.4 2.25 0.0245 453 76 0.0055 411 0.16 3.33
Pujol 2002, No. 20-3-3S 152.4 304.8 686 36.4 2.25 0.0245 453 76 0.0055 411 0.16 3.36
Ramirez and Jirsa, 1980, 00-U 305 305 458 34.5 1.50 0.0245 374 65 0.0032 455 0.00 3.71
Saatcioglu and Ozcebe 1989, U1 350 350 1000 43.6 2.86 0.0321 430 150 0.0030 470 0.00 4.87
Saatcioglu and Ozcebe 1989, U2 350 350 1000 30.2 2.86 0.0321 453 150 0.0030 470 0.16 4.20
Saatcioglu and Ozcebe 1989, U3 350 350 1000 34.8 2.86 0.0321 430 75 0.0060 470 0.14 5.11
Sezen and Moehle No. 1 457.2 457.2 1473 21.1 3.22 0.0247 434 305 0.0017 476 0.15 2.56
Sezen and Moehle No. 2 457.2 457.2 1473 21.1 3.22 0.0247 434 305 0.0017 476 0.60 1.33

2.20
@Seismicisolation
@Seismicisolation
Assessment of ASCE/SEI 41 Concrete Column Provisions using Shaking Table Tests

Aspect
b D L f’c fyl s fyt Δ80%/L
Specimens Ratio ρl ρt P/Agf’c
(×10-2)
(mm) (mm) (mm) (MPa) (MPa) (mm) (MPa)
(a/D)
Sezen and Moehle No. 4 457.2 457.2 1473 21.8 3.22 0.0247 434 305 0.0017 476 0.15 3.49
Soesianawati et al. 1986, No. 4 400 400 1600 40.0 4.00 0.0151 446 94 0.0030 255 0.30 2.74
Takemura and Kawashima, 1997, Test 1 (JSCE-4) 400 400 1245 35.9 3.11 0.0158 363 70 0.0020 368 0.03 3.51
Takemura and Kawashima, 1997, Test 2 (JSCE-5) 400 400 1245 35.7 3.11 0.0158 363 70 0.0020 368 0.03 3.90
Takemura and Kawashima, 1997, Test 3 (JSCE-6) 400 400 1245 34.3 3.11 0.0158 363 70 0.0020 368 0.03 5.96
Takemura and Kawashima, 1997, Test 4 (JSCE-7) 400 400 1245 33.2 3.11 0.0158 363 70 0.0020 368 0.03 8.15
Takemura and Kawashima, 1997, Test 5 (JSCE-8) 400 400 1245 36.8 3.11 0.0158 363 70 0.0020 368 0.03 6.79
Umehara and Jirsa 1982, 2CUS 230 410 455 42.0 1.11 0.0301 441 89 0.0055 414 0.27 1.04
Wehbe et al. 1998, A1 380 610 2335 27.2 3.83 0.0222 448 110 0.0027 428 0.10 5.23
Wehbe et al. 1998, A2 380 610 2335 27.2 3.83 0.0222 448 110 0.0027 428 0.24 4.38
Wight and Sozen 1973, No. 00.033(East) 152.4 304.8 876 32.0 2.88 0.0245 496 127 0.0032 345 0.00 3.00
Wight and Sozen 1973, No. 00.033(West) 152.4 304.8 876 32.0 2.88 0.0245 496 127 0.0032 345 0.00 5.50
Wight and Sozen 1973, No. 25.033(East) 152.4 304.8 876 33.6 2.88 0.0245 496 127 0.0032 345 0.07 3.59
Wight and Sozen 1973, No. 25.033(West) 152.4 304.8 876 33.6 2.88 0.0245 496 127 0.0032 345 0.07 4.80
Wight and Sozen 1973, No. 40.048(East) 152.4 304.8 876 26.1 2.88 0.0245 496 89 0.0046 345 0.15 4.88
Wight and Sozen 1973, No. 40.048(West) 152.4 304.8 876 26.1 2.88 0.0245 496 89 0.0046 345 0.15 5.49
Wight and Sozen 1973, No. 40.067(East) 152.4 304.8 876 33.4 2.88 0.0245 496 64 0.0064 345 0.11 6.85
Wight and Sozen 1973, No. 40.067(West) 152.4 304.8 876 33.4 2.88 0.0245 496 64 0.0064 345 0.11 6.88
Xiao and Martirossyan 1998, HC4-8L16-T6-0.1P 254 254 508 86.0 2.00 0.0246 510 51 0.0075 449 0.10 6.40
Xiao and Martirossyan 1998, HC4-8L16-T6-0.2P 254 254 508 86.0 2.00 0.0246 510 51 0.0075 449 0.19 4.26
Note: a = the shear span (equals to L/2 for fixed-fixed columns); b = width of column section; D = height of column section; fc’= concrete compressive stress at
test day; fyl = yield stress of longitudinal reinforcement; fyt = yield stress of transverse reinforcement; L= clear height of the column; s = spacing of transverse
reinforcement; ρl = longitudinal reinforcement ratio; ρt = volume transverse reinforcement ratio; P = column axial load; Ag = area of column cross section; ∆80% =
column displacement at 80% of peak lateral resistance.

2.21
@Seismicisolation
@Seismicisolation
Y. Li et al.

APPENDIX C: CONDITION II COLUMNS IN STATIC DATASET B

b D L f'c fyl fyt Δa/L


Specimen Name (a/D) ρl s (mm) ρt P/(f'cAg)
(mm) (mm) (mm) (MPa) (Mpa) (MPa) (×10-2)
Kato et al. 2000, No.2 350 350 875 1.3 40.5 0.0328 404 45 0.01016 309 0.65 2.00
Kato et al. 2002, LC 250 250 1000 2.0 27.9 0.0127 362 60 0.00420 390 0.43 2.40
Sezen 2002, 2CHD12 457.2 457.2 2946 3.2 21.1 0.0247 434.4 304.8 0.00176 476 0.61 1.90
Kabeyasawa et al. 2002, A-1 420 150 1260 1.5 18.3 0.0100 349 200 0.00133 289 0.29 1.48
Ousalem et al.2004, C4 300 300 900 1.5 13.5 0.0169 340 75 0.00280 384 0.30 4.03
Ousalem et al.2004, C8 300 300 900 1.5 18.0 0.0169 340 75 0.00280 384 0.30 1.52
Ousalem et al.2004, C12 300 300 900 1.5 18.0 0.0169 340 75 0.00280 384 0.20 8.07
Ousalem et al.2004, D16 300 300 600 1.0 26.1 0.0169 447 50 0.00420 398 0.23 4.00
Ousalem et al.2004, D1 300 300 600 1.0 27.7 0.0169 447 50 0.00420 398 0.22 4.06
Ousalem et al.2004, D13 300 300 900 1.5 26.1 0.0226 447 50 0.00420 398 0.23 3.50
Ousalem et al.2004, D14 300 300 900 1.5 26.1 0.0226 447 50 0.00420 398 0.23 10.06
Ousalem et al.2004, C1 300 300 900 1.5 13.5 0.0169 340 160 0.00083 587 0.30 1.01
Sezen 2002, 2CLD12 457.2 457.2 2946 3.2 21.1 0.0247 434.4 304.8 0.00176 476 0.15 4.99
Sezen 2002, 2CLD12M 457.2 457.2 2946 3.2 21.8 0.0247 434.4 304.8 0.00176 476 0.15 5.09
Lynn et al. 1992, 3CMD12 457.2 457.2 2946 3.2 27.6 0.0312 331 304.8 0.00176 400 0.26 2.07
Yoshimura et al. 2002, 2C13 300 300 600 1.0 25.2 0.0169 350 100 0.00210 392 0.19 3.00
Yoshimura et al. 2002, 2M13 300 300 600 1.0 25.2 0.0169 350 100 0.00210 392 0.19 4.10
Lynn et al. 1992, 2SLH18 457.2 457.2 2946 3.2 33.1 0.0194 331 457.2 0.00068 400 0.07 3.63
Yoshimura et al. 2003, No.7 300 300 1200 2.0 30.7 0.0169 409 150 0.00140 392 0.20 2.00
Yoshimura et al. 2003, No.5 300 300 1200 2.0 30.7 0.0268 402 100 0.00210 392 0.35 2.00
Yoshimura et al. 2003, No.4 300 300 1200 2.0 30.7 0.0268 402 100 0.00210 392 0.30 2.00
Note: b = width of column section; D = height of column section; fc’= concrete compressive stress at test day; fyl = yield stress of longitudinal reinforcement; fyt =
yield stress of transverse reinforcement; L= clear height of the column; s = spacing of transverse reinforcement; ρl = longitudinal reinforcement ratio; ρt = volume
transverse reinforcement ratio; P = column axial load; Ag = area of column cross section; ∆a = column displacement at axial load failure.

2.22
@Seismicisolation
@Seismicisolation
SP-297—3

Numerical Models for Beam‐Column Joints in Reinforced Concrete Building 
Frames 
Jong‐Su Jeon, Laura N. Lowes, Reginald DesRoches 

 
 
 

Synopsis: The  results  of  laboratory  testing  and  earthquake  reconnaissance  studies  of  reinforced 
concrete frames indicate that beam‐column joint deformation can determine total frame deformation 
and  that  for  older  buildings  joint  failure  can  result  in  frames  losing  lateral  and  gravity  load  carrying 
capacity.  Given  the  impact  of  joints  on  frame  response,  numerical  models  used  to  evaluate  the 
earthquake performance of reinforced concrete frames must include nonlinear joint models. This paper 
reviews  previously  proposed  models  for  simulating  joint  response  with  the  objective  of  identifying 
models that provide i) accurate simulation of response to earthquake loading, ii) simple implementation 
in  nonlinear  analysis software, iii) numerical robustness, iv) computational  efficiency, and v) objective 
calibration procedures. Ultimately, no set of models was identified that met all of these requirements 
for the range of geometric and design parameters found in reinforced concrete buildings in the United 
States. With the objective of extending current modeling capabilities for interior joints, an experimental 
data set was assembled. The data set was used to evaluate existing envelope response models and used 
to  calibrate  cyclic  response  parameters  for  use  with  the  preferred  existing  model.  A  new  response 
model for interior beam‐column joints is presented that meets the above requirements for the range of 
geometric and design parameters found in reinforced concrete buildings in the United States.        
 

Key words: Reinforced  concrete,  beam‐column  joint,  earthquake,  seismic  behavior,  model, 
simulation, bond, shear strength     

3.1
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

Biographical Sketches of Authors


Jong‐Su Jeon, is a post‐doctoral researcher in the School of Civil and Environmental Engineering at the 
Georgia  Institute  of  Technology.  His  research  addresses  nonlinear  analysis  of  reinforced  concrete 
structures and seismic risk assessment.  

Laura N. Lowes, FACI, is an Associate Professor of Civil and Environmental Engineering at the University 
of  Washington,  Seattle.  Her  research  and  teaching  focusses  on  nonlinear  analysis  of  civil  structures 
subjected  to  severe  loading  including  earthquake  loading.  She  is  a  member  of  ACI  Committees  369, 
Seismic Repair and Rehabilitation, and 447, Finite Element Analysis of Reinforced Concrete Structures.  

Reginald DesRoches is the Karen and John Huff School Chair, and Professor of Civil and Environmental 
Engineering at the Georgia Institute of Technology. His research addresses earthquake engineering and 
risk  assessment  of  civil  infrastructure.  He  is  a  member  of  the  executive  committee  of  the  National 
Academy of Sciences Disasters Roundtable, the Board of the Earthquake Engineering Research Institute, 
and  the  advisory  board  for  the  Natural  Disasters,  Coastal  Infrastructure  and  Emergency  Management 
Research Center. 

Introduction
Numerical  simulation  to  evaluate  the  earthquake  performance  of  reinforced  concrete  frames  requires 
simulation  of  the  nonlinear  response  of  beam‐column  joints.  Laboratory  testing  of  building  frame 
subassemblages  shows  that  beam‐column  joints  with  design  details  typical  of  pre‐19671 construction 
exhibit severe stiffness and strength loss (Walker 2001; Leon 1990; Meinheit and Jirsa 1981; Pantelides 
et al. 2002). Significant strength and stiffness loss have been observed also for joints designed to achieve 
superior performance, such as those compliant with ACI 318 Code requirements (e.g., Park and Ruitong 
1988;  Durrani  and  Wight  1985).  Finally,  earthquake  damage  to  reinforced  concrete  frames  observed 
during post‐earthquake reconnaissance efforts (e.g., Hall 1996; Holmes and Somers 1996) suggests also 
that  joint  failure  may  result  in  structural  collapse.  Given  the  potential  for  beam‐column  joints  to 
determine  frame  performance,  simulation  of  nonlinear  joint  response  is  required  for  accurate 
evaluation of the earthquake performance of older and modern reinforced concrete frame buildings.   

Experimental investigation of the earthquake response of reinforced concrete beam‐column joints has 
been the focus of numerous research studies during the last 40 years. Experimental testing has resulted 
in  data  characterizing  the  response  of  frame  subassemblages,  which  comprise  column  and  beam 
segments  as  well  as  joints  and,  in  some  cases,  characterizing  the  local  response  of  the  beam‐column 
joint. Experimental tests have addressed a range of joint configurations and designs, including interior 
and exterior joints from two‐dimensional and three‐dimensional frames with detailing representing pre‐
1967  and  modern  construction.  However,  the  objective  of  many  of  these  laboratory  tests  was  to 
establish that modern building  code requirements result in acceptable  earthquake  performance; thus, 
many of these tests employed joint designs representative of modern construction, including transverse 
reinforcement,  moderate  joint  shear  stress  demands,  and  moderate  bond  stress  demands  for  beam 
                                                            
1
 In 1967 UBC provisions changed to include requirements that i) joint shear strength defined using the column shear strength 
equation exceed joint shear demand, ii) transverse reinforcement be provided in the joint to achieve required shear strength, 
iii)  the  sum  of  column  flexural  strengths  exceed  the  sum  of  beam  flexural  strengths,  and  iv)  splices  in  beam  and  column 
longitudinal reinforcement be located away from the joint. Similar revisions were made to the ACI Code in 1971. These Code 
changes are reflected in drawings for buildings designed for construction on the West Coast from 1923 – 1979 (Mosier 2000).     

3.2
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

longitudinal reinforcement. Fewer tests have investigated the earthquake response of older joints with 
no  transverse  reinforcement,  high  shear  stress  demands  and  high  bond  stress  for  beam  longitudinal 
reinforcement  anchored  in  the  joint.  Additionally,  most  experimental  investigations  have  employed 
interior  beam‐column  joint  subassemblages  from  two‐dimensional  frames  (the  test  specimen  is 
cruciform‐shaped  and  comprises  a  column  that  is  continuous  through  the  joint,  a  beam  that  is 
continuous  through  the  joint,  and  the  joint  region  with  no  out‐of‐plane  beam  segments).  Fewer 
investigations  have  employed  interior  joints  from  three‐dimensional  frames  (the  test  specimen 
comprises a column that is continuous through the joint, a beam that is continuous through the joint, a 
second  orthogonal  beam  that  is  continuous  through  the  joint  or  terminates  in  the  joint,  and  the  joint 
region).  A  very  few  tests  have  considered  the  behavior  of  exterior  corner  joints  in  three‐dimensional 
frames  (the  test  specimen  comprises  a  column  that  is  continuous  through  the  joint,  two  orthogonal 
beam  segments  that  terminate  in  the  joint,  and  the  joint  region).  Despite  the  fact  that  a  plethora  of 
experimental  data  do  not  exist  for  all  joint  geometries  and  designs,  sufficient  data  exist  to  enable 
calibration of numerical models for use in evaluating the earthquake performance of concrete frames.  

Given  the  impact  on  frame  response  of  joint  flexibility  and  strength  loss,  many  previous  studies  have 
addressed the development of beam‐column joint models for use in assessing the earthquake behavior 
and performance of concrete frames. Previously proposed models have ranged in sophistication from i) 
“super‐elements”  comprising  multi‐dimensional  continuum  elements  in  which  model  response  is 
defined  purely  by  fundamental  material  properties  and  joint  geometry  (Pantazopoulou  and  Bonacci 
1994;  Mitra  2007;  Baglin  and  Scott  2000),  to  ii)  “super‐elements”  comprising  one‐dimensional  springs 
and employing simplifying assumptions about joint response mechanisms (Lowes and Altoontash 2003; 
Altoontash 2004; Mitra and Lowes 2007; Tajiri et al. 2006), to iii) rotational hinge models calibrated to fit 
experimental  data  (Otani  1974;  Anderson  and  Townsend  1977).  In  recent  years,  validation  and 
calibration efforts have employed larger and larger data sets; for example for interior joints, Mitra and 
Lowes  (2007),  Shin  and  LaFave  (2004),  and  Jeon  (2013)  employed  data  from  57,  26  and  124  tests, 
respectively.  Unfortunately,  however,  previous  research  has  not  resulted  in  a  model  or  set  of  models 
that fully meets the needs of the earthquake engineering community with respect to modeling accuracy 
and  precision,  computational  efficiency  and  robustness,  and  ease  of  use  for  the  full  range  of  beam‐
column joint configurations found in the building inventory and for the case of dynamic cyclic as well as 
quasi‐static monotonic loading.           

Modeling Objectives and Research Process


The objective of the research presented here was the development of joint modeling recommendations 
for evaluation of the earthquake performance of concrete frames. Specifically, joint models were sought 
that met the following criteria: 

1. Accurate simulation of the fundamental characteristics of joint response to earthquake loading. 
Multiple  models may be  required  to provide accurate response simulation for a range of joint 
designs  and  configurations.  Simulation  of  response  through  significant  loss  of  joint  shear 
strength was considered to be an important attribute; simulation of the loss of joint axial load 
carrying  capacity  was  not  considered  to  be  an  important  attribute.  In  the  laboratory,  beam‐
column  joint  subassemblages  have  typically  maintained  moderate  axial  loads  following 
significant loss of joint shear strength. A recent study by Hassan (2011) found that for joints with 
beam‐to‐column  depth  ratios  less  than  2.5,  joint  axial  load  failure  was  unlikely  to  precede 
column axial load failure. 
2. Simple  implementation  of  the  model  in  OpenSees  using  existing  element  formulations  and 
material models. The OpenSees software platform was chosen for use because it is free, widely 

3.3
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

used  within  the  earthquake  engineering  research  community,  includes  element  and  material 
models  that  can  readily  be  employed  to  simulate  joint  response,  and  provides  efficient  and 
robust solution algorithms to support analysis of building frames subjected to quasi‐static and 
dynamic loads. In particular, OpenSees includes joint elements and 1D spring elements that can 
be  used  to  simulate  joint  kinematics  and  material  models  that  are  capable  of  simulating  the 
stiffness  and  strength  loss  and  pinched  hysteretic  response  exhibited  by  joints  subjected  to 
cyclic loading.    
3. Numerical robustness and stability, so that analyses are not hampered by “failures to converge”, 
and computational efficiency.  
4. Objective  calibration  procedures,  so  that  the  model‐building  process  is  as  efficient  as  possible 
and modeling parameters do not need to be adjusted to define the model for a specific beam‐
column joint with a specific set of design characteristics. 

To  accomplish  the  above  objective,  previously  proposed  models  were  reviewed  and  evaluated  with 
respect to the above criteria. As needed, model development activities were accomplished to fill gaps in 
the capabilities of existing models.  

Modeling Joint Geometry


Previous  research  has  resulted  in  several  approaches  for  modeling  the  geometry  of  the  beam‐column 
joint region for frame analysis. This research has primarily addressed analysis of 2D frames, with analysis 
of  3D  systems  accomplished  by  duplicating  the  2D  modeling  approach  in  the  two  orthogonal  framing 
directions.  Few  studies  have  addressed  modeling  of  the  3D  joint  volume  to  support  modeling  of  joint 
behavior  under  3D  loading  scenarios;  few  3D  joint  elements  exist.  Most  research  addressing  joint 
modeling  in  2D  has  employed  either  a  four‐node,  finite‐area  joint  element  or  a  zero‐length  rotational 
spring element. Zero‐length elements have been used at the beam‐joint and column‐joint interface to 
simulate stiffness and strength loss associated with bond‐slip within the joint.   

Using a four‐node, finite‐area joint element, joint behavior is typically simulated through the action of 
multiple 1D spring elements that compose the joint element (Altoontash 2004; Youssef and Ghobarah 
2001;  Shin  and  LaFave  2004;  Lowes  and  Altoontash  2003;  Mitra  and  Lowes  2007).  However,  the  joint 
element may comprise only a single rotational spring (Celik 2007; Park 2010; Jeon 2013) or a mesh of 
continuum elements (Baglin and Scott 2000; Hegger et al. 2004; Deaton 2013; Mitra 2007). Joint super‐
elements that comprise more than one spring or continuum element typically require an intra‐element 
solution2. Typically the user has minimal control of the intra‐element solution algorithm; as such, the use 
of super‐elements introduces the potential for numerical problems and reduced numerical robustness 
and efficiency. For 2D analysis, element nodes are typically located at the joint‐beam and joint‐column 
interface; thus, no additional modeling effort is required to ensure that inelastic flexural action in beams 
and columns occurs at the beam‐joint / column‐joint interface and that beam and column elements do 
not  contribute  flexural  deformation  within  the  joint  region.  Appropriate  element  formulations  are 
required to enable simulation of geometric nonlinearity.  

                                                            
2
 The  OpenSees  joint  elements  are  essentially  small  structural  subassemblages  within  the  global  structural  model.  These 
subassemblages  comprise  multiple  nodes  and  multiple  nonlinear  springs  that  are  not  (explicitly)  part  of  the  global  structural 
model. For each iteration to establish equilibrium of the global structural model, it is necessary to solve for equilibrium of the 
joint  subassemblage  model.  While  the  user  has  control  over  the  global  solution  algorithm,  with  the  ability  to  change  the 
tolerance or step size to improve the rate of convergence, the user has no control over the solution algorithm used to solve the 
joint  subassemblage  model.  Thus,  the  user  has  limited  ability  to  continue  the  analysis  when  a  “failure  to  converge”  occurs 
within the joint element.        

3.4
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

Using  a  two‐node,  zero‐length  rotational  joint  spring  (Figure  1  and  2),  joint  response  is  simulated 
through the action of a single rotational spring located at the centerline intersection of the beam and 
column. Since the beam and column elements extend to the center of the joint, it is necessary to add 
rigid links or rigid‐end‐offsets to the ends of beams and columns, within the joint volume, to ensure that 
inelastic flexural action in beams and columns occurs at the beam‐joint / column‐joint interface and that 
beam  and  column  elements  do  not  contribute  flexural  deformation  within  the  joint  region.  The 
rotational  spring  element  is  appropriate  for  use  with  linear  and  nonlinear  geometric  transformations. 
The  rotational  joint  spring  requires  a  1D  joint  moment  versus  joint  rotation  response  model.  Joint 
response models are discussed below; however, most of these models define joint shear stress versus 
strain. Since joints are not subjected to pure shear loading at the geometric perimeter of the joint, Celik 
(2007)  provides  equations  relating  shear  stress  to  joint  moment.  Joint  shear  deformation  may  be 
assumed equal to joint rotation.    

Two‐node, zero‐length slip elements (Figure 2) may be employed in combination with joint elements to 
simulate  slip  of  longitudinal  reinforcement  in  the  joint  region.  Typically  the  joint  element  is  used  to 
simulate  shear  deformation  of  the  finite‐area  joint  region  and  the  slip  element  is  used  to  simulate 
stiffness and strength loss due to bond failure for beam or column reinforcement anchored in the joint. 
At one location, multiple independent zero‐length elements or a single section element may be used to 
simulate load‐deformation response for transfer of moment, axial load and shear.  

Two nodes
connected by Steel
Concrete
zero-length stress-slip
stress-slip
Nonlinear fiber- rotational spring model
model
Rotational spring type beam-column Slip section
elements

Rigid links (i.e. elastic


Rotational spring Two nodes connected by
Two nodes beam-column elements zero-length slip section
connected by with relatively high element
zero-length stiffness) Nonlinear fiber-
rotational spring type beam-column Rigid links (i.e. elastic beam-
elements column elements with
relatively high stiffness)

   
Figure 1: OpenSees Model of a 2D Interior Joint   Figure 2: OpenSees Model of a 2D Exterior Joint 
Subassemblage    Subassemblage 

Figure  2  shows  a  zero‐length  barslip  fiber‐section  model  located  at  the  beam‐joint  interface  of  an 
exterior  joint.  These  models  have  been  used  successfully  to  simulate  rotation  due  to  slip  of 
reinforcement anchored in footings (Berry 2006; Oyen 2006) and are consistent with the barslip springs 
introduced into joint models by Youssef and Ghobarah (2001) and Mitra and Lowes (2007). The barslip 
section  model  may  be  used  to  simulate  deformation  due  to  barslip  for  interior  joints  with  continuous 
reinforcement,  but  is  likely  most  relevant  for  joints  in  which  beam  longitudinal  reinforcement 
terminates within the joint with a short straight anchorage. Berry (2006) provides general guidance on 
defining the barslip model; Mitra and Lowes (2007) provides recommendations for bond strength within 
joints and cyclic response parameters. 

3.5
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

Review of Existing Joint Response Models


The  following  sections  provide  a  review  of  the  joint  core  constitutive  models  and  barslip  models 
available  in  the  literature  that  offer  the  greatest  potential  for  meeting  the  modeling  objectives 
presented above. It is expected that these constitutive models will be used to define input parameters 
for  the  1D  OpenSees  Pinching4  material  model;  though  other  OpenSees  material  models  may  be 
adequate. The Pinching4 model allows for definition of a multi‐linear response envelope, simulates the 
“pinched” response histories that are typical of beam‐column joint stress versus shear strain data, and 
provides a mechanism for simulating strength and stiffness loss under cyclic loading.  

Interior Beam‐Column Joints in 2D Frames: Earthquake Behavior


An  interior  joint  in  a  2D  frame  (Figure  1)  represents  the  intersection  of  a  continuous  beam  and 
continuous  column.  For  these  joints,  accurate  simulation  of  the  following  response  characteristics  is 
required: 1) joint stiffness and strength, 2) subassemblage drift capacity, as limited by the joint, where 
drift capacity is defined as the drift at which significant (in excess of 20% of maximum strength) loss in 
lateral  load  carrying  capacity  is  observed,  and  3)  beam‐controlled  versus  joint‐controlled  response, 
where  specimens  exhibiting  beam‐controlled  response  develop  the  yield  strength  of  the  beams  (or 
columns)  framing  into  the  joint  and  exhibit  significant  drift  capacity  while  specimens  exhibiting  joint‐
controlled response do not develop the yield strength of the beams (or columns) framing into the joint 
and  exhibit  limited  drift  capacity.  Figures  3  and  4  show  column  shear  versus  subassemblage  drift  for 
joint  test  specimens  exhibiting  joint‐  and  beam‐controlled  response;  specimens  have  no  transverse 
reinforcement and do not satisfy the requirements of ACI 318‐11 for a joint in a special moment frame 
(SMF). Note that shear and bond stress demands defined per Eqs. 1 and 2 are provided.  
150 100 Extreme
Yield of Beam Reinforcement Center Joint Spalling
Extreme Spalling Spalling
80 Yielding
100 Initial Spalling
60
Column Shear (kip)

50 40 Center Joint
Center Joint Cracking Cracking
Column Shear (kips)

20
0
0
-50 -20
-40
-100 -60
-80
-150
-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 -100
Drift (%) -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6
     Drift (%)

Figure  3:  Column  shear  versus  drift  for  a  joint‐ Figure  4:  Column  shear  versus  drift  for  a  beam‐
controlled  specimen  (PEER  4150  from  Walker  controlled specimen (PEER 0995 from Alire (2002) 
(2001) joint = 0%,  = 47 ,  = 46 ).  joint = 0%,  = 8 ,  = 32 ). 

Numerous  studies  have  investigated  the  parameters  that  affect  joint  response  to  earthquake  loading 
(Kim and LaFave 2009; Mitra 2007). Kim and LaFave (2009), using a data set comprising 212 interior 2D 
and 3D joint test specimens (3D specimens were subjected to unidirectional lateral loading), identifies 
the  following  as  the  primary  parameters  controlling  joint  strength:  presence  of  out‐of‐plane  beams, 
beam  eccentricity,  normalized  joint  transverse  reinforcement  ratio,  normalized  beam  reinforcement 
ratio,  and  concrete  compressive  strength.  Birely  et  al.  (2012),  considering  a  data  set  comprising  45 
interior 2D joint test specimens, proposes that joint versus beam controlled response is determined by 
joint shear stress demand per ACI 352R‐02, , and joint bond stress demand assuming beam bars yield 
on both sides of the joint,  (Figure 5). In Figure 5, joint demands are defined as follows:  

3.6
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

  Eq. 1 

and 

   Eq. 2 

where   is the nominal concrete compressive strength;   is  Beam‐controlled response


the depth of the column;   is the out‐of‐plane dimension of   
Beam‐controlled response with ductility < 4 
   
the  joint;   is  the  nominal  yield  strength  of  beam  Joint‐controlled response
Proposed limits  
longitudinal  steel;   and   are  the  longitudinal  steel  ACI limits  
areas in the top and bottom of the beam, respectively;   is 
the column shear corresponding to the development of the 
nominal  flexural  strength  of  the  beams  framing  into  the 
joint;  and   accounts  for  the  actual‐versus‐nominal  yield 
strength  of  the  reinforcing  steel  and  the  hardening  of  the 
steel under loading. Eqs. 1 and 2 employ measured concrete 
and steel strengths and   = 1.25/1.1, where the 1.1 factor is 
used  to  approximate  the  design  strengths  from  the 
measured  properties,  and  the  1.25  factor  is  used  to 
approximate  the  effect  of  strain  hardening  and  over‐
strength.  

It should be noted that the above discussion of interior joint 
Figure  5:  Relationship  of  ductility 
behavior  does  not  consider  the  case  of  interior  joints  for 
classification  to  design  shear  stress 
which  beam  longitudinal  reinforcement  terminates  within 
and bond stress.  
the  joint.  Prior  to  19671,  anchorage  for  beam  longitudinal 
reinforcement  was  often  determined  on  the  basis  of  gravity  loads  with  the  result  that  beam  bottom 
reinforcement  often  terminated  within  the  joint  with  short  (6  in.  typ.)  embedment  lengths. 
Experimental testing (e.g., Pessiki et al. 1990) shows that these joints typically exhibit strength loss prior 
to  beam  yielding  due  to  pull‐out  of  the  reinforcement.  Exterior  joints  have  similar  design  details  and 
simulation of this type of response is discussed below.  

Interior Beam‐Column Joints in 2D Frames: A Review of Existing Response Models


Many models have been proposed for simulating the earthquake response of interior joints. In general, 
models  define  response  on  the  basis  of  joint  geometry,  design  configuration,  and  material  properties. 
Some models are highly calibrated such that the user is required to input minimal information but the 
accuracy  of  predicted  response  for  a  particular  beam‐column  joint  depends  on  the  similarity  between 
the  design  parameters  of  the  particular  joint  and  the  data  set  used  for  model  calibration.  Table  1 
describes  models  by  Birely  et  al.  (2012)  and  Anderson  et  al.  (2008),  which  are  examples  of  highly 
calibrated  models.  Some  models  employ  simplified  representations  of  the  joint  load‐transfer 
mechanism; for these models, the accuracy of predicted response for a particular beam‐column joint is 
less dependent on the similarity between the design parameters of the particular joint and the data set 
used for model calibration. The model by Mitra and Lowes (2007) is one such model. Table 1 provides 
information  about  those  models  considered  to  offer  the  greatest  potential  for  meeting  the  modeling 
objectives  identified  above.  Information  included  in  the  Defining  Characteristics  column  of  Table  1  is 
intended  to  differentiate  the  model  from  others;  information  included  in  the  Advantages  and 
Limitations columns relates to the modeling objectives identified above.      

3.7
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

Table 1: Response Models for Interior Beam‐Column Joints in 2D Frames 

Elem. 
Reference  C/M1  Defining Characteristics  Strengths  Limitations 
Type2 
Mitra and  C  FA  ‐ Model includes  vs.  for joint  ‐ Validated using a large data set  ‐ Complicated to define model 
Lowes  core derived from concrete strut  (57 specimens).  parameters for specific set of joint 
(2007)  model as well as bar‐slip springs  ‐ Developed for OpenSees.  characteristics. 
derived from bond stress    ‐ Requires intra‐element solution. 
capacity. 
Andreson et  C  RS   vs. model only.  ‐ Developed using data from test  ‐ Model definition is relatively 
al. (2008)  ‐ Developed using experimental  specimens with no trans.  complicated. 
 vs. data.  reinforcement.  ‐ May not possible to implement 
‐ Joint strength is 0.95ACI.  ‐ Demonstrated to provide  model in OpenSees without 
‐ Joint strength degrades with  accurate response of specimen  approximations.   
cyclic loading once maximum  with limited transverse  ‐ Because beam strength 
strength is reached; rate of  reinforcement.  determines joint strength, 
degradation depends on ACI.   accurate simulation of post‐peak 
response and impact of ACI on this 
response is critical. 
‐ Model was validated using joint  
vs. data, not subassemblage 
response data. 
Kim and  M  RS   vs. model.  ‐ Developed using large data set  ‐ Implementation in OpenSees 
LaFave  ‐ Developed using experimental  (212 specimens). ‐ Defines  requires definition of cyclic 
(2009)   vs. data.  response for joints with a wide  response quantities. 
‐ Quadrilinear envelope, trilinear  range of design parameters.  ‐ Model calibration, for individual 
to max. strength and then    joint design, requires calculation 
descending branch to 90% max.  of a number of design parameters. 
strength.  ‐ Model only predicts response 
‐ Envelope points are functions  through 10% strength loss.  
of multiple design parameters 
(geometry, joint reinforcement 
ratio, concrete strength, etc.).   
Birely et al.  M  RS  Joint moment versus joint  ‐ Developed using large data set  ‐ Requires definition of post‐peak 
(2012)  deformation model (i.e. includes  (45 specimens).  response. 
all non‐frame member  ‐ Simple.  ‐ Requires definition of cyclic 
deformation).  ‐ Joint model includes all non‐ response quantities. 
‐ Bilinear to max.  frame member deformation, so no  ‐ Model calibration, for individual 
strength/failure; does not  barslip model required; model was  joint design, requires calculation 
include post‐peak response.    developed using subassemblage  of beam yield moments. 
test data and OpenSees lumped‐
plasticity elements used to model 
beams and columns. 
‐ Model validation demonstrates 
that model can predict joint failure 
versus beam‐yielding; specimens 
exhibiting limited ductility are 
predicted to exhibit joint failure. 
Jeon (2013)  C  RS  ‐ vs. model.  ‐ Developed using large data set  ‐ Joint shear strain at maximum 
‐ Developed using experimental  (375 specimens).  strength and residual strength are 
 vs. data.  ‐ Validated using large data set  not defined. 
‐ Mult‐linear envelope, trilinear  (124 specimens). 
to max. strength, descending  ‐ Defines response for joints with a 
branch to 20% max. strength,  wide range of design parameters. 
residual strength of 20% max.   
strength.  
‐ Envelope points are functions 
of multiple design parameters 
(geometry, joint reinforcement 
ratio, concrete strength, etc.).
Notes: 1. Model defines cyclic response history and is appropriate for cyclic analysis, C, or model defines envelope to cyclic response history 
and is appropriate for analysis under monotonic loading. 2. Model is appropriate for use with a 4‐node, finite‐area element, FA, or zero‐length 
rotational spring element, RS. 

3.8
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

On  the  basis  of  the  information  provided  in  Table  1,  the  models  by  Birely  et  al.  (2012)  and  Kim  and 
LaFave  (2009)  were  identified  as  warranting  further  evaluation  and,  potentially,  development.  Both 
models provide a relatively simple approach for defining a shear strength versus deformation envelope 
that is appropriate for use for a range of design configurations and that can provide simulation of ductile 
as well as non‐ductile response mechanisms. However, both models require further evaluation. The Kim 
and LaFave model was developed using experimental joint shear stress versus strain data and was not 
validated  using  subassemblage  data;  thus,  there  is  the  potential  for  the  model  to  neglect  bar  slip  or 
other  deformation  not  captured  by  laboratory  instrumentation  of  the  joint‐core  region.  The  Birely 
model  was  developed  and  calibrated  using  beam‐column  joint  subassemblage  data  and  a  numerical 
model  such  as  shown  in  Figure  1;  thus,  simulated  joint  deformation  includes  all  frame  subassemblage 
deformation not attributed to beams or columns. However, the Birely model defines a hardening type 
response  to  the  point  of  maximum  strength  with  definition  of  post‐peak  response  left  to  the  user; 
because initial strength loss may not be extremely rapid (Figure 3 and 4), there is the potential for the 
model to neglect and/or provide highly inaccurate results for the post‐peak response history. To enable 
use  of  the  models  for  analysis  of  frames  subjected  to  dynamic  earthquake  excitation,  both  models 
require definition of the post‐peak response envelope to the point of residual joint strength and of cyclic 
response quantities.  

Exterior Beam‐Column Joints in 2D Frames: A Review of Existing Response Models


A number of models have been proposed or are appropriate for simulating the earthquake response of 
exterior joints. In general, models define response on the basis of joint geometry, design configuration, 
and  material  properties.  A  defining  characteristic  for  the  models  is  simulation  of  anchorage  failure. 
Figures 6 and 7 show the response of two exterior beam‐column joint subassemblages from 2D frames. 
Figure  6  shows  response  for  a  joint  for  which  strength  under  positive  drift  demand  was  controlled  by 
anchorage failure for beam bottom reinforcement that was anchored in the joint with a short anchorage 
length. Figure 7 shows the response an exterior joint subassemblage for which response was controlled 
by shear failure of the joint. Table 2 provides information about those models considered to offer the 
greatest potential for simulating response histories such as shown in Figures 6 and 7 and meeting the 
modeling objectives identified above. 

Strength  under  pos.  drift 


demands  controlled  by 
pullout of beam bottom bar 

   
Figure  6:  Column  shear  versus  drift  for  an  Figure  7:  Column  shear  versus  drift  for  a  shear‐
anchorage‐controlled exterior joint specimen (Unit  controlled  exterior  joint  specimen  (Unit  6  from 
2  from  Pantelides  et  al.  (2002),  joint  =  0%,  max  Pantelides et al. (2002)  joint = 0%, max = 11.3  
=7    (up)  =  10.6    (down),  beam  ,  beam  bar  anchorage  is  180‐degree  hooks, 
bottom bar anchorage length = 6 in., column axial  column  axial  load  =  0.25Agfc.  Joint  design  is  not 
load  =  0.25Agfc. Joint  design  is  not  compliant  with  compliant with ACI Code requirements for SMF). 
ACI Code requirements for SMF). 
   

3.9
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

Table 2: Response Models for Exterior Beam‐Column Joints  

Elem. 
Reference  C/M1  Defining Characteristics  Advantages  Limitations 
Type2 
Sharma et al.  M  RS  ‐ Model includes  vs.  for joint  ‐ Validated using  ‐ Model cannot be directly applied to joints with 
(2011)  core derived from Priestley  data from 12 joint  straight anchorage lengths other than 6 in.  
(1997) principal tensile stress  subassemblage tests;  ‐ Proposed implementation / spring model is 
model.  specimens have  different from that shown in Figure 2. 
‐ Max tensile stress is empirically  range of design  ‐ Validation was not in OpenSees, used zero‐
calibrated; one strength for well‐ characteristics and  length plastic hinges for beams and columns and 
anchored bars and one strength  exhibit anchorage  used a spring configuration different from that 
for bars with 6 in. straight  and joint failure.  in Figure 2. 
anchorage length.   ‐ Simple.  ‐ Requires definition of cyclic response 
‐ Does not require  quantities; achieving accurate simulation may be 
barslip section.  challenging given different strengths in +/‐ 
directions.  
Kim and  M  RS &  ‐ Defines response for joints with  ‐ Developed using  ‐ Implementation requires definition and use of 
LaFave (2009)  BSS  well‐anchored beam longitudinal  large data set (111  barslip section model to simulate strength loss 
reinforcement (90‐ or 180‐degree  specimens).  due to anchorage failure. 
hooks).  ‐ Includes  ‐ Requires definition of cyclic response 
‐  vs. model.  descending branch  quantities. 
‐ Developed using experimental   of response curve, so  ‐ Model calibration, for individual joint design, 
vs. data.  no additional  requires calculation of a number of design 
‐ Quadrilinear envelope, trilinear  modeling is required  parameters. 
to max. strength, descending  to simulate strength  ‐ Requires definition of post‐peak response; 
branch to 90% max. strength.  loss.    model only predicts response through 10% 
‐ Envelope points are functions of    strength loss. 
multiple design parameters 
(geometry, joint reinforcement 
ratio, concrete strength, etc.).   
Mitra and  M,C  RS &  ‐  vs. models for interior beam‐ ‐ Simple. ‐ Requires validation. 
Lowes (2007),  BSS column joints.  ‐ Allows for use of  ‐ Requires definition and use of barslip section 
Anderson et  ‐ Assume that these can be  same model for  model to simulate strength loss due to 
al. (2008),  applied to exterior joints.  interior and exterior  anchorage failure. 
Birely et al.  ‐ Employ barslip section model to  joints.  ‐ Requires definition of cyclic response 
(2012)   simulate pullout of beam bars  quantities. 
with short, straight anchorage. 
Hassan  C  RS &  ‐  vs. model to simulate joint  ‐ Simple. ‐ Not validated for exterior joints in which beam 
(2011)   BSS core response combined with M‐ ‐ Developed for use  bars have inadequate straight anchorage. 
model to simulate barslip.   in OpenSees.   ‐ Validation effort for 2D joints is limited. 
developed for exterior joints in  ‐ Validated using 
3D frames and applied to 2D  subassemblage 
joints.   model and test data. 
Park (2010)  C  RS &  ‐ M vs.  model to simulate joint  ‐ Simple. ‐ Does not include simulation of barslip and 
BSS  response.   ‐ Developed for use  cannot simulate failure due to pullout of beam 
developed for exterior joints in  in OpenSees.   bars with short anchorage lengths. 
3D frames and applied to 2D  ‐ Validated using  ‐ Validation effort for 2D joints is limited. 
joints.  subassemblage 
model and test data. 
Jeon (2013)  C  RS  ‐ vs. model.  ‐ Developed using  ‐ Joint shear strain at maximum strength and 
‐ Developed using experimental   large data set (340  residual strength are not defined. 
vs. data.  specimens). 
‐ Mult‐linear envelope, trilinear  ‐ Validated using 
to max. strength, descending  large data set (72 
branch to 20% max. strength,  specimens). 
residual strength of 20% max.  ‐ Defines response 
strength.   for joints with a wide 
‐ Envelope points are functions of  range of design 
multiple design parameters  parameters. 
(geometry, joint reinforcement   
ratio, concrete strength, etc.). 
Notes: 1. Model defines cyclic response history and is appropriate for cyclic analysis, C, or model defines envelope to cyclic response history 
and is appropriate for analysis under monotonic loading, M. 2. Model is appropriate for use with a zero‐length rotational spring element, RS, or 
requires use of a rotational spring element in combination with one or more bar‐slip springs, RS & BSS. 

3.10
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

Knee‐Joints in 2D Frames
A knee joint in a 2D frame represents the intersection of an exterior column and the beam at the top of 
a frame. The knee joint on the left side of a frame is modeled in OpenSees by removing the top column 
and its associated nodes from the model shown in Figure 2.  

Given that the top story of a frame is typically not a critical region of the frame, with respect to loss of 
lateral load carrying capacity or frame collapse, accurate simulation of response for these components 
should  be  considered  less  important  than  for  interior  or  exterior  joints.  Relatively  little  research  has 
addressed the earthquake response of knee‐joints in older or modern buildings frames. For knee‐joints, 
joint  shear  demands  are  likely  similar  in  magnitude  to  those  in  exterior  joints;  for  older  joints, 
inadequate  anchorage  detailing  for  beam  and/or  column  bars  terminating  in  the  knee  joint  likely 
controls  response.  Kim  and  LaFave  (2009)  provides  a    vs.  model  for  the  envelope  to  the  cyclic 
response history for knee joints (18 specimens) with transverse reinforcement and adequate anchorage 
detailing for beam and column reinforcement; recommendations are provided for extending this model 
to  the  case  of  no  transverse  reinforcement  with  adequate  anchorage  detailing,  but  no  data  exist  for 
validation. Barslip section models at the beam‐joint and column‐joint interfaces are recommended for 
simulating the response of older knee‐joints with inadequate straight bar anchorages for beam and/or 
column  longitudinal  reinforcement  anchored  in  the  joint.  Development  activities  must  address 
validation  of  models  using  subassemblage  data,  validation  of  bar‐slip  models  and  definition  of  cyclic 
response parameters.    

Joints in 3D Frames
In 3D frames, the interior, exterior and knee joints described above are extended to include continuous 
or  terminating  beams  in  the  out‐of‐plane  direction  and  are  subjected  to  loading  in  the  out‐of‐plane 
direction.  The  addition  of  beams  framing  into  the  joint  is  typically  considered  to  improve  joint  core 
response  by  providing  additional  confinement  of  the  core  (Kitayama  et  al.  1991;  Oka  and  Shiorhara 
1992);  however,  3D  loading  may  increase  joint  damage  and  reduced  joint  deformation  capacity  (Leon 
and  Jirsa  1986).  Relatively  few  experimental  studies  have  employed  joint  subassemblages  from  3D 
frames  subjected  to  bidirectional  lateral.  Model  development  activities  have  employed  data  from  3D 
joint  tests  to  calibrate  2D  joint  response  models  (i.e.  models  for  use  in  2D  frames  subjected  to 
unidirectional  lateral  loading);  these  models  are  included  in  Tables  1  and  2.  Few  models  for  use  in 
simulating  the  response  of  joints  in  3D  frames  subjected  to  bidirectional  lateral  loading  for  which 
response in orthogonal lateral directions is coupled do not exist.    

To  enable  accurate  simulation  of  the  response  of  beam‐column  joints  in  3D  frames,  extensive 
experimental testing and model development are required. Existing models are reviewed below.  

Interior Joints in 3D Frames


Interior joints in 3D frames represent the intersection of a continuous column, a continuous beam, and 
either a second continuous beam or a beam that terminates at the joint. Relatively few experimental or 
numerical  studies  have  addressed  these  joint  configurations;  no  experimental  studies  have  addressed 
interior joints in 3D frames without transverse reinforcement. Beams and columns framing into all (all 
but one) side of the joint could be expected to provide some additional confinement of the joint core 
and  improve  performance;  however,  some  interior  joint  subassemblages  from  3D  frames,  with 
transverse  reinforcement,  might  have  exhibited  joint  failure  prior  to  beam  yielding  in  the  laboratory 
(Meinheit and Jirsa 1981; Teraoka 1997). Given the potential for joint failure, simulation of the nonlinear 
response of interior joints, including reduced stiffness as well as strength loss, is required for the current 
study. The model developed by Kim and LaFave (2009) can be extended to interior joints in 3D frames; 

3.11
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

only a few data exist for validation of the model. For the case of older interior beam‐column joints with 
three  beam  segments  framing  into  the  joint  (i.e.  an  interior  joint  in  an  exterior  frame),  inadequate 
anchorage  of  beam  reinforcement  may  control  response.  In  this  case,  the  model  by  Kim  and  LaFave 
(2009) can be supplemented by a barslip section model at the joint‐beam interface.  

Exterior Corner Joints in 3D Frames


Recent  research  at  the  University  of  California,  Berkeley  (Park  2010;  Hassan  2011)  has  addressed  the 
earthquake response of older exterior joints in 3D frames and, in particular, older exterior corner joints 
in  3D  frames.  This  research  included  experimental  tests  of  exterior  corner  joints  to  investigate  the 
impact  on  response  of  column  axial  load,  beam‐to‐column  depth  ratio,  joint  shear  demand  /  beam 
longitudinal  reinforcement  ratio,  and  load  history.  Modeling  efforts  addressed  definition  of  joint 
strength,  as  well  as  development  of  joint  cyclic  response  models.  Models  by  Park  (2010)  and  Hassan 
(2011)  both  i)  were  developed  using  OpenSees,  ii)  were  validated  using  OpenSees  models  of 
subassemblages and experimental data for exterior joints and exterior corner joints in 3D frames , and 
iii) provided acceptably accurate simulation of measured response. The model by Park does not account 
for explicitly separate deformation due to joint shear response and bar slip, while the model by Hassan 
does. Neither model addresses simulation of response for joints in which beam bars have inadequate, 
straight  anchorage  within  the  joint;  in  this  case,  the  models  should  be  combined  with  a  barslip  fiber 
section model at the joint‐beam interface.   

Evaluation and Advancement of Existing Joint Response Models


The above review of existing response models for beam‐column joints identifies a number of limitations 
with  respect  to  the  modeling  objectives.  To  enable  assessment  of  the  seismic  vulnerability  of  existing 
reinforced  concrete  frame  buildings  (Jeon  2013),  a  research  effort  was  undertaken  to  evaluate  and 
address  limitations  of  existing  beam‐column  joint  models.  Initial  activities,  the  results  of  which  are 
presented here, focused on evaluation and advancement of response models for interior joints from 2D 
and 3D frames subjected to gravity and unidirectional lateral loading.  

Evaluation of Existing Response Models for Interior Joints


The review of existing interior joints identified models by Kim and LaFave (2009) and Birely et al. (2012) 
as offering the greatest potential for meeting the modeling objectives. These models as well as the Jeon 
(2013) model were evaluated using an experimental data set comprising 58 interior beam‐column joint 
tests for which subassemblage load‐displacement as well as joint shear stress‐strain response data were 
available (Table 3). Specimens in Table 3 represent a wide range of joint configurations (cruciforms from 
2D frames as well as joints from 3D frames for which 3 or 4 beam segments frame into the joint) and 
design  parameters,  have  joint  transverse  reinforcement  that  is  non‐compliant  (specimens  1‐16)  and 
compliant  (specimens  17‐58)  per  ASCE  41‐06  (Elwood  et  al.  2007),  and  exhibited  beam‐  and  joint 
controlled response in the laboratory. All specimens in Table 3 were subjected to unidirectional lateral 
loading with or without gravity loading. The data set presented in Table 3 is a subset of that assembled 
by Jeon (2013); additional information about the test specimens is provided in Jeon (2013). 

The  response  of  the  joint  subassemblages  was  simulated  using  OpenSees  and  a  model  configuration 
such  as  shown  in  Figure  1.  Jeon  (2013)  provides  details  of  the  model.  A  zero‐length  rotational  spring 
element  and  the  Pinching4  material  model  were  used  to  model  the  joint.  Lateral  and  vertical 
displacements  were  constrained  to  be  equal  for  duplicate  nodes  located  at  the  center  of  the  joint. 
Elastic  beam‐column  elements  with  relatively  large  stiffnessess  were  introduced  within  the  joint  area. 
The  beamWithHinges  element  was  used  to  model  beams  and  columns.  This  element  is  a  force‐based 
element  for  which  a  linear  moment  distribution  and  constant  shear  and  axial  load  distribution  are

3.12
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

 Table 3: Data set used in this study 
  Test program  Specimen     (√psi)  Error estimate (Eτ) 
  (√psi)  Jeon Kim and LaFave Birely Jeon Kim and LaFave  Birely
1  Alire (2002)  PEER0995  11.1  13.0 12.8 12.4 1.18 1.16  1.12
2     PEER1595  16.5  15.7 14.6 17.2 0.95 0.88  1.04
3     PEER4150  25.4  19.7 17.9 22.0 0.78 0.71  0.86
4  Goto and Joh (1996)  J‐OH  18.3  18.5 17.3 22.0 1.01 0.94  1.20
5  Ohwada (1984)  LJO‐6  20.1  14.6 15.0 22.0 0.73 0.74  1.09
6     LJXY‐6  19.1  16.8 17.7 22.0 0.88 0.92  1.15
7     LJXY‐8  18.2  16.8 17.7 22.0 0.93 0.97  1.21
8    JO‐5  17.6  14.9 14.7 21.5 0.85 0.83  1.22
9     JXY‐3  21.8  16.4 17.3 21.5 0.75 0.80  0.99
10  Walker (2001)  PEER14  10.9  11.0 11.8 11.6 1.02 1.09  1.07
11     CD1514  11.6  11.1 11.8 12.0 0.96 1.03  1.04
12    CD3014  11.2  10.8 11.6 10.2 0.97 1.04  0.91
13    PADH14  10.8  10.8 11.6 10.2 1.00 1.07  0.94
14    PEER22  14.6  15.2 14.3 15.5 1.04 0.99  1.07
15    CD3022  15.5  14.9 14.2 15.7 0.96 0.92  1.01
16    PADH22  15.7  15.1 14.4 15.8 0.96 0.92  1.00
17  Endoh et al. (1991)  LA1  20.0  22.5 22.2 20.1 1.12 1.11  1.00
18     A1  20.1  22.6 22.7 20.5 1.12 1.13  1.02
19  Fujii and Morita   A1  18.7  22.4 23.2 22.0 1.20 1.24  1.18
20  (1991)  A3  18.7  22.4 23.2 22.0 1.20 1.24  1.18
21     A4  19.0  24.8 27.8 22.3 1.31 1.46  1.17
22  Goto and Joh (1996)  J‐HH  20.3  21.1 23.7 30.1 1.04 1.17  1.49
23  Goto and Joh (2003)  LM‐60  16.1  14.5 15.6 16.7 0.90 0.97  1.04
24     LM‐125  12.7  12.5 12.5 13.1 0.98 0.99  1.03
25     HM‐60  20.1  15.9 17.2 21.6 0.79 0.85  1.07
26     HM‐125  14.4  13.9 13.6 16.3 0.97 0.95  1.14
27     HH‐125  14.5  14.3 14.4 15.8 0.98 0.99  1.09
28  Kurose et al. (1991)  J1  21.3  17.2 19.8 16.9 0.81 0.93  0.80
29  Matsumoto et al.   J‐5  22.0  21.1 23.5 21.9 0.96 1.07  1.00
30  (2010)  J‐10  20.4  20.3 21.8 22.0 0.99 1.07  1.08
31  Noguchi and   OKJ‐1  20.4  22.8 24.9 21.2 1.12 1.22  1.04
32  Kashiwazaki (1992)  OKJ‐4  21.6  21.0 23.5 21.2 0.98 1.09  0.98
33     OKJ‐5  21.2  23.8 27.1 21.9 1.12 1.28  1.03
34     OKJ‐6  21.5  22.7 25.7 21.9 1.05 1.20  1.02
35  Oka and Shiohara   J‐5  22.9  22.8 23.6 21.9 0.99 1.03  0.96
36   (1992)  J‐6  20.4  18.7 17.9 18.6 0.92 0.88  0.91
37  Ozaki et al.   NO1  18.1  17.4 19.7 18.2 0.96 1.09  1.00
38   (2010)  NO2  16.9  16.7 17.1 17.7 0.99 1.01  1.05
39  Raffaelle and Wight   SP1  13.5  14.6 15.4 13.8 1.08 1.14  1.03
40   (1992)  SP2  10.2  11.3 14.3 14.7 1.11 1.40  1.44
41     SP3  9.6  10.8 13.1 11.0 1.13 1.36  1.14
42     SP4  11.5  12.1 13.0 12.1 1.06 1.13  1.06
43  Shin and LaFave   SL1 11.6  13.4 13.3 12.0 1.15 1.15  1.03
44   (2004)  SL2 12.4  13.9 13.3 12.4 1.12 1.08  1.00
45     SL4 16.5  16.0 16.6 13.0 0.97 1.00  0.78
46  Takamori et al.   NO1  14.4  14.4 14.4 14.1 1.00 1.00  0.98
47   (2006)  NO2  15.1  15.1 15.3 15.0 0.99 1.01  0.99
48  Teng and Zhou   S1  18.0  16.7 18.5 16.7 0.93 1.03  0.93
49   (2003)  S2  17.7  16.5 17.3 16.6 0.93 0.98  0.94
50     S5  14.5  14.5 15.2 14.7 1.00 1.05  1.02
51  Teraoka (1997)  NO01  22.8  20.0 23.1 22.2 0.88 1.01  0.97
52     NO04  25.5  25.3 26.1 22.2 0.99 1.03  0.87
53     NO07  24.2  24.7 27.7 22.2 1.02 1.14  0.92
54     NO08  24.2  24.6 29.1 22.2 1.02 1.20  0.92
55     NO09  21.2  22.5 23.5 22.2 1.06 1.11  1.05
56     NO10  20.9  24.8 28.3 22.2 1.19 1.36  1.07
57     NO35  17.0  18.6 19.4 15.5 1.10 1.14  0.91
58     NO36  17.3  20.7 23.6 17.8 1.20 1.37  1.03
  Mean      1.01 1.06  1.04
  COV      0.12 0.15  0.12

3.13
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

assumed along the length of the element. Inelastic flexural response is assumed to occur only in “plastic‐
hinge  regions”  at  the  ends  of  the  element;  for  the  system  shown  in  Figure  1,  hinges  were  introduced 
only where the beam (column) framed into the joint. A fiber‐type discretization of the beam (column) 
cross section and typical material response models were used to simulate inelastic response within the 
hinge  regions;  a  plastic  hinge  length  equal  to  half  the  member  depth  was  used.  Outside  the  hinge 
region,  beam  and  column  elastic  flexural  stiffness  was  defined  per  ASCE  41‐06  (Elwood  et  al.  2007). 
Lateral load was applied via displacement control to the top of the column. A constant gravity load was 
applied at the top of the column as required to simulate laboratory testing. 

Figure 8 shows the response models by Jeon (2013), Kim and LaFave (2009) and Birely et al. (2012). The 
Jeon  model  includes  parameters  defining  cyclic  response  (not  shown);  the  Kim  and  LaFave  and  Birely 
models define only the envelope of the response and are not directly applicable for quasi‐static cyclic or 
dynamic analysis. The Jeon model (Figure 8a) represents a simplification of the model by Anderson et al. 
(2008);  experimental  data  characterizing  the  response  of  interior  joints  from  2D  and  3D  frames  were 
used to develop an expression defining maximum joint shear strength, max, as a function of joint design 
parameters and post‐peak stiffnesses, kdeg, as a function of joint transverse reinforcement configuration 
(kdeg = ‐2max and kdeg = ‐1.2max for joints that are ASCE 41‐06 non‐compliant and compliant, respectively). 
An  expression  was  not  developed  for  joint  shear  strain  at  maximum  strength;  instead  measured 
quantities  were  used  in  evaluating  the  model.  Thus,  the  Jeon  model  is  not  appropriate  for  predictive 
analysis. Figure 8b shows the Kim and LaFave model, which was developed using statistical observation 
methods and an extensive experimental data base comprising interior joint specimens from 2D and 3D 
frames.  The  database  included  relatively  few  joints  without  joint  transverse  reinforcement,  and  the 
model may not provide accurate response for this class of joints. The response envelope for the Kim and 
LaFave  model  is  defined  by  maximum  joint  shear  stress,  max,  and  the  joint  shear  strain  at  maximum 
strength,  γmax;  expressions  are  provided  defining  these  quantities  as  a  function  of  joint  material, 
geometric and reinforcement parameters. The Birely model shown in Figure 8c. It was developed using 
data  from  tests  of  interior  joints  from  2D  frames;  it  does  not  simulate  the  impact  of  confinement 
provided by out‐of‐plane beam segments framing into the joint as the Jeon and Kim and LaFave models 
do. The Birely model defines a joint shear stress‐strain response envelope that is bilinear to maximum 
strength;  brittle  post‐peak  response  is  conservatively  assumed  and  response  is  not  defined  beyond 
maximum  strength.  Initial  stiffnesses  are  a  function  of  the  concrete  shear  modulus,  G;  the  transition 
from initial to second stiffness occurs once beam longitudinal reinforcement yields and analysis of the 
beams framing into the joint is required to determine this point. For all three models, equilibrium of the 
joint is considered to convert joint shear stress versus strain response to joint moment versus rotation 
response for use with the OpenSees model shown in Figure 1.  

(γexp, τmax) (γmax, τmax)


Joint shear stress

Joint shear stress

Joint shear stress

(0.006, 0.95τmax) (0.362γmax, 0.890τmax) τyield,min


k2= 0.038G
kdeg (2.02γmax, 0.900τmax)
(0.00043, 0.48fc ) 0.5
(0.0197γmax, 0.442τmax)
k1= 0.14G
(γres, 0.2τmax)
γlimit = 0.0069
Joint shear strain Joint shear strain Joint shear strain
(a) Jeon model  (b) Kim and LaFave model  (c) Birely model 
Figure 8: Joint shear stress‐strain response envelope models 

3.14
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

Models  were  evaluated  on  the  basis  of  the  accuracy  with  which  subassemblage  and  joint  response 
quantities  were  simulated.  Table  3  lists,  for  each  specimen,  the  ratio  of  simulated  to  measured  joint 
shear  strength,  ,  as  defined  below.  Data  in  Table  4  and  Table  5,  respectively,  quantify  the  accuracy 
with which subassemblage and joint response are simulated; the mean and coefficient of variation (COV) 
of  various  error  measures  defined  below  are  provided  for  all  specimens  in  the  data  set.  Note  that 
statistics  provided  in  Tables  4  and  5  are  similar  when  specimens  are  grouped  as  compliant  and  non‐
compliant  per  ASCE  41‐06  (Elwood  et  al.  2007).  Figures  9  and  10Error!  Reference  source  not  found. 
show, respectively, measured and simulated subassemblage and joint response envelopes for specimens 
exhibiting  joint‐  and  beam‐controlled  response.  The  following  response  quantities  were  considered  in 
model evaluation and are listed in Tables 3, 4 and 5: 

1. Ratio of simulated to measured maximum joint shear stress, as controlled by either joint shear 
,
failure or beam flexural strength. For a test specimen, this ratio is defined as   where 
,
τ ,  and τ ,  are the simulated and measured maximum joint shear stresses, respectively.  
2. Ratio  of  simulated  to  measured  drift  at  computed  yield  strength.  This  is  used  to  evaluate  the 
accuracy  with  which  initial  stiffness  is  computed.  For  a  test  specimen,  the  ratio  is  defined  as 
,
 where Δ ,  is  the  simulated  drift  at  the  computed  initial  yield  of  beam  (or  column) 
,
longitudinal steel or, for specimens that did not reach the computed yield strength of the beams 
(columns), to 90% of maximum strength and Δ ,  is the measured drift at the same load level.  
3. Ratio  of  simulated  to  measured  maximum  column  shear,  as  controlled  by  either  joint  shear 
,
failure or beam flexural strength. For a test specimen, this ratio is defined as   where 
,
V ,  is the simulated maximum shear load carried by the column and V ,  is the measured 
maximum load. Note that  and that both are provided for completeness.  
4. Ratio of simulated to measured drift at maximum subassemblage strength. For a test specimen, 
,
this ratio is defined as  ,  where Δ ,  is the simulated drift at V ,  and Δ ,  is 
,
the measured drift at  V , . Note that  ∆,  is not provided for the Jeon model as the model 
does not predict joint shear strain at maximum shear strength; thus, measured shear strain at 
maximum  shear  strength  was  used  to  define  the  model  and  ,  is  approximately  1.0  for  this 
model.  
5. Ratio  of  simulated  to  measured  drift  at  10%  loss  of  subassemblage  strength.  For  a  test 
,
specimen,  this  ratio  is  defined  as  ,  where Δ ,  is  the  simulated  and Δ ,  is  the 
,
measured drift at 10% loss of subassemblage strength.  
6. Normalized error in the simulated response envelope. For a given specimens, this is computed 

as  ∑  where  n  is  the  number  of  drift  demands  for  which  the 
,
difference between simulated,  , and measured,  , column shear is computed and  ,  is 
the measured maximum column shear. For the Birely model and subassemblages predicted to 
exhibit  joint‐controlled  response,  this  quantity  was  computed  to  the  point  of  maximum 
strength; for all for models and subassemglages, this quantity was computed to the point of 10% 
strength loss. 
7. Ratio  of  simulated  to  measured  initial  joint  shear  stiffness.  For  a  test  specimen,  this  ratio  is 
,
defined  as   where γ ,  and γ ,  are,  respectively,  the  simulated  and  measured  shear 
,
strain at 90% of computed joint shear strength.  

3.15
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

8. Ratio  of  simulated  to  measured  joint  shear  strain  at  maximum  joint  shear  stress.  For  a  test 
,
specimen, this ratio is defined as  ,  where γ ,  is the simulated joint shear strain 
,
at simulated maximum shear stress, τ , , and γ ,  is the measured joint shear strain at the 
measured maximum shear stress,  τ , . Note that γ ,  is not predicted by the Jeon model 
and  ,  is not provided for this model.  
9. Ratio of simulated to measured joint shear strain at 10% loss of joint shear strength. For a test 
,
specimen,  this  ratio  is  defined  as  ,  where γ ,  is  the  simulated  and γ ,  is  the 
,
measured joint shear strain at 10% loss of maximum joint shear strength. 
10. Normalized  error  in  the  simulated  joint  response  envelope.  For  a  given  specimen,  this  is 

computed as  , ∑  where n is the number of joint shear strains for 
,
which the difference between simulated,  , and measured,  , joint shear stress is computed 
and  ,  is  the  measured  maximum  joint  shear  stress.  For  the  Birely  model  and 
subassemblages  predicted  to  exhibit  joint‐controlled  response,  this  quantity  was  computed  to 
the  point  of  maximum  strength;  for  all  for  models  and  subassemglages,  this  quantity  was 
computed to the point of 10% strength loss. 

150 70

125 60
Column shear (kips)

Column shear (kips)

50
100
40
75
30
50 Experiment 20 Experiment
Jeon (2013) Jeon (2013)
25 Kim and LaFave (2009) 10 Kim and LaFave (2009)
Birely et al. (2012) Birely et al. (2012)
0 0
0 1 2 3 4 5 0 1 2 3 4 5
 
Drift (%) Drift (%)
 
a) PEER4150 (Alire 2002), joint = 0%, max =25.4   b) PEER14 (Walker 2001), joint = 0%, max=10.9  
 and  λ=45.6 .  Strength  is  controlled  by   and  λ=28.1 .  Strength  is  controlled  by 
joint failure.   beam yielding.  
60 400

50
Column shear (kips)

Column shear (kips)

300
40

30 200

20 Experiment Experiment
Jeon (2013) 100 Jeon (2013)
10 Kim and LaFave (2009) Kim and LaFave (2009)
Birely et al. (2012) Birely et al. (2012)
0 0
0 1 2 3 4 0 1 2 3 4 5
Drift (%)
  Drift (%)
 
c)  NO01  (Teraoka  1997),  joint  =  1.09%,  max  d) NO02 (Takamori et al. 2006), joint = 0.60%, max 
=22.8    and  λ  =  33.0 .  Strength  is  =14.9    and  λ  =  19.7 .  Strength  is 
controlled by joint failure.   controlled by beam yielding. 

3.16
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

Figure 9: Observed and simulated response envelopes for interior joint subassemblages.  

2 1

0.8
Joint shear stress (ksi)

Joint shear stress (ksi)


1.5

0.6
1
0.4
Experiment Experiment
0.5 Jeon (2013) Jeon (2013)
0.2
Kim and LaFave (2009) Kim and LaFave (2009)
Birely et al. (2012) Birely et al. (2012)
0 0
0 0.005 0.01 0.015 0.02 0.025 0.03 0 0.005 0.01 0.015 0.02 0.025 0.03
Joint shear strain (rad)    
Joint shear strain (rad)
b) PEER14 (Walker 2001), joint = 0%, max =10.9  
a) PEER4150 (Alire 2002), joint = 0%, max =25.4    and  λ=28.1 .  Strength  is  controlled  by 
 and  λ=45.6 .  Strength  is  controlled  by  beam yielding.  
joint failure.   1.4
2
1.2
Joint shear stress (ksi)

1
Joint shear stress (ksi)

1.5
0.8

1 0.6

0.4 Experiment
Experiment Jeon (2013)
0.5 Jeon (2013) 0.2 Kim and LaFave (2009)
Kim and LaFave (2009) Birely et al. (2012)
Birely et al. (2012) 0
0 0.005 0.01 0.015 0.02 0.025 0.03
0   Joint shear strain (rad)
0 0.005 0.01 0.015 0.02 0.025 0.03
  d) NO02 (Takamori et al. 2006), joint = 0.60%, max 
Joint shear strain (rad)
c)  NO01  (Teraoka  1997),  joint  =  1.09%,  max  =14.9    and  λ  =  19.7 .  Strength  is 
=22.8    and  λ  =  33.0 .  Strength  is  controlled by beam yielding. 
controlled by joint failure.  
Figure 10: Observed and simulated joint shear stress versus strain response envelopes for interior joints. 

Table 4. Comparison of subassemblage load versus drift error values for the three models  

  Initial  Strength,         Drift at  Drift Capacity,  Response 


  Stiffness,      Strength,  ,   ,   Envelope,   
  Mean  COV  Mean  COV  Mean  COV  Mean  COV  Mean  COV 
Birely   1.27  0.20  1.04  0.13  0.59  0.30  0.64  0.79  0.031  0.52 
Kim and LaFave  1.12  0.27  1.07  0.14  1.29  0.62  1.18  0.41  0.017  0.69 
Jeon  1.23  0.23  1.01  0.11  NA  NA  1.07  0.28  0.015  0.67 
 
Table 5. Comparison of joint shear stress versus strain error values for the three models  

  Initial  Strength,         Shear Strain at  Shear Strain  Shear Response 


  Stiffness,      Strength,  ,   Capacity,  ,   Envelope,  ,  

3.17
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

  Mean  COV  Mean  COV  Mean  COV  Mean  COV  Mean  COV 
Birely   1.12  0.41  1.04  0.12  0.48  0.48  0.29  0.50  0.077  0.46 
Kim and LaFave  0.96  0.45  1.06  0.15  1.28  0.51  1.53  0.50  0.011  0.77 
Jeon  1.12  0.35  1.01  0.12  NA  NA  1.43  0.37  0.009  0.79 
 

Simulated and observed response envelopes for the entire data set, such as  those shown in Figures 9 
and 10, and the data in Table 4 and Table 5 support the following observations and conclusions:  

1. The Kim and LaFave and Jeon models, on average, provide acceptable accuracy for simulation of 
subassemblage and joint response; though both models provide poor prediction of joint shear 
strain  at  failure.  Both  the  Kim  and  LaFave  models  provide  relatively  imprecise  simulation  of 
response,  with  relatively  large  COVs  for  most  response  quantities  (initial  joint  stiffness,  drift 
capacity, etc.).  
2. The  Birely  model,  on  average,  provides  acceptable  accuracy  for  simulation  of  subassemblage 
and  joint  response  up  to  maximum  strength.  The  model  provides  acceptable  accuracy  of 
simulation  of  response  beyond  maximum  strength  for  beam‐controlled  specimens.  For  joint‐
controlled specimens, the model makes the conservative assumption of a relatively brittle joint 
failure  mode  and  does  not  define  response  beyond  the  point  of  maximum  joint  strength.  As 
shown  in  Figures  9  and  10,  many  specimens  exhibiting  joint‐controlled  failure  do  not  exhibit 
classical  brittle  response.  Thus,  on  average,  the  model  provides  poor  prediction  of  response 
beyond maximum strength.  
3. For  all  models,  corresponding  subassemblage  and  joint  response  quantities  (e.g.  initial 
subassemblage  stiffness  and  initial  joint  shear  stiffness)  are  simulated  with  approximately  the 
same  accuracy  and  precision;  the  COV  for  simulation  of  a  particular  joint  response  quantity  is 
typically equal to or slightly larger than the COV for the corresponding subassemblage response 
quantity.    
4. The initial stiffness of the subassemblage and joint region are accurately simulated by the Kim 
and LaFave model (average error of 12% and ‐4% for the subassemblage and joint, respectively) 
and are simulated with acceptable accuracy by the Birely and Jeon models (average errors of 12% 
for  the  subassemblage  and  approximately  25%  for  the  joint).  For  all  of  the  models, 
subassemblage stiffness is simulated with greater precision (COVs of 0.20, 0.23 and 0.27) than 
joint  stiffness  (COVs  of  0.35,  0.41  and  0.45).  On  the  basis  of  these  data,  the  Kim  and  LaFave 
model is considered to be the best modeling approach to estimate the initial stiffness.  
5. Maximum  column  (joint)  shear  strength  is  simulated  accurately  and  precisely  for  all  three 
models.  For  simulation  of  column  strength  the  Birely,  Kim  and  LaFave,  and  Jeon  models, 
respectively, provide average errors of 4%, 7%, and 1% with respective COVs of 0.13, 0.14, and 
0.11. Statistics are essentially the same for joint shear strength. 
6. Subassemblage  drift  and  joint  shear  strain  at  maximum  strength  are  simulated  poorly  by  the 
Birely model (average errors of ‐41% and ‐52% for these two quantities) and relatively poorly by 
the  Kim  and  LaFave  model  (average  errors  of  29%  and  28%  for  these  two  quantities).  These 
relatively large errors are attributed to the relatively low stiffness of the system in the vicinity of 
maximum  strength,  which  results  in  significant  variation  in  both  measured  and  simulated 
deformation  at  maximum  strength.  The  Jeon  model  does  not  predict  this  quantity,  and 
measured quantities were used to complete definition of the response envelope; thus no error 
value is provided for this model. 
7. Drift capacity of the subassemblage is simulated with acceptable accuracy by both the Joen and 
the Kim and LaFave models (average errors of 7% and 18% for the two models); however joint 

3.18
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

shear strain at subassemblage drift capacity is poorly predicted by the models (average errors of 
43%  and  53%).  In  all  cases  COVs  are  large  (greater  than  0.28).  The  Birely  model  does  not 
simulate  response  beyond  the  point  of  maximum  joint  strength;  thus,  drift  capacity  and  joint 
shear strain at drift capacity are, on average, relatively poorly.  

Given the above evaluation of the Birely, Kim and LaFave, and Jeon models, it was decided that further 
model  development  activities  would  employ  the  Kim  and  LaFave  model.  The  Kim  and  LaFave  model 
supports predictive modeling, while the Jeon model, which employs measured joint shear strain values, 
does not. The Kim and LaFave model also provides definition of the response envelope beyond the point 
of maximum strength, while the Birely model does not. 

Development of Cyclic Response Parameters for the Kim and LaFave Model for Interior Joints
Model development activities employed the Kim and LaFave model and focused on defining parameters 
to support cyclic analysis using the OpenSees Pinching4 material model. Figure 11 shows the response 
envelope  and  response  path  under  cyclic  loading  as  well  as  key  parameters  used  in  defining  the 
Pinching4  material  model.  For  each  test  specimen,  the  Kim  and  LaFave  model  was  used  to  define  the 
envelope  points  (ePd*,  ePf*,  eNd*,  dNf*)  in  Figure  11,  and  cyclic  response  parameters  (rDispP,  rDispN, 
rForceP and rForceN) in Figure 11 were determined that provided a least‐squares, best‐fit to the cyclic 
column shear versus story drift data. Average values of the cyclic response parameters were computed 
for the entire data set (124 specimens) used in Jeon (2013), for specimen group according to whether 
joint  transverse  reinforcement  was  conforming  or  non‐conforming  per  ASCE  41‐06,  for  specimens 
grouped according to whether or not joint transverse reinforcement was provided, and for specimens 
grouped  according  to  joint‐  or  beam‐controlled  response  as  predicted  using  Birely  et  al.  (2012).  In  all 
cases, the average value of the cyclic response parameters ranged from 0.19 to 0.27, with the averages 
for the entire data set being rDispP = rDispN = rForceN = 0.24 and rForceP = 0.23. On the basis of these 
results, rDispP = rDispN = rForceN = rForceP = 0.25 was chosen as the preferred values for use. Figures 
12 and 13 show simulated and observed cyclic response histories for column shear versus specimen drift 
(Figure  12)  and  joint  shear  versus  joint  strain  (Figure  13).  In  these  figures,  simulated  response  was 
generated  using  the  previously  described  OpenSees  modeling  approach  with  the  Pinching4  material 
model used to define the cyclic response of the joint rotational spring, the Kim and LaFave model used 
to  define  the  envelope  to  the  joint  moment‐rotation  history,  and  the  Pinching4  cyclic  response 
parameters defined equal to 0.25. Pinching4 model parameters related to cyclic stiffness and strength 
deterioration were not considered in this study. Comparison of simulated and observed cyclic response 
histories, such as those shown in Figures 12 and 13, shows that the proposed Pinching4 cyclic response 
parameters provide relatively accurate simulation of the observed “pinched” behavior of the specimens. 
However, the data in Figures 12 and 13 show the potential for poor simulation of joint strain demands 
and  thus  joint  response  as  well  as  the  potential  for  relatively  accurate  simulation  of  subassemblage 
response to be achieved even when joint response is poorly simulated.  

3.19
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

(dmax,f(dmax))

rDispP  dmax , rForceP  f (dmax )


, uForceP  ePf3 

, uForceN  eNf3 

rDispN  dmin , rForceN  f (dmin )

(dmin,f(dmin))

 
Figure 11: OpenSees Pinching4 uniaxial material model (Lowes and Altoontash 2003). 

150 80
60
100
Column shear (kips)

Column shear (kips)


40
50
20
0 0

−50 −20
−40
−100 Experiment
−60 Experiment
Simulation Simulation
−150 −80
−6 −4 −2 0 2 4 6 −6 −4 −2 0 2 4 6
 
Drift (%)
  Drift (%)
a) PEER4150 (Alire 2002), joint = 0%, max =25.4   b) PEER14 (Walker 2001), joint = 0%, max =10.9  
 and  λ  =  45.6 .  Strength  is  controlled  by   and  λ  =  28.1 .  Strength  is  controlled  by 
joint failure.   beam yielding.  
60 400

40
Column shear (kips)

Column shear (kips)

200
20

0 0

−20
−200
−40 Experiment Experiment
Simulation Simulation
−60 −400
−8 −6 −4 −2 0 2 4 6 8 −6 −4 −2 0 2 4 6
 
Drift (%) Drift (%)  
c)  NO01  (Teraoka  1997),  joint  =  1.09%,  max  d) NO02 (Takamori et al. 2006), joint = 0.60%, max 
=22.8    and  λ  =  33.0 .  Strength  is  =14.9    and  λ  =  19.7 .  Strength  is 
controlled by joint failure.   controlled by beam yielding. 
Figure 12: Simulated and observed column shear versus story drift response histories for interior joint 
subassemblage test specimens. 

3.20
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

Summary and Conclusions


Existing beam‐column joint response models were reviewed with the objective of identifying models for 
use  in  dynamic  analysis  of  reinforced  concrete  frames.  Models  were  sought  that  meet  the  following 
criteria:  accurate  simulation  of  response,  computational  efficiency  and  robustness,  relatively  easy 
implementation within the OpenSees platform, and objective calibration procedures. Ultimately, no set 
of models was identified that met the above criteria and was appropriate for use for the full range of 
geometric and design configurations found in the US building inventory.  

With  the  objective  of  advancing  current  modeling  capabilities,  three  recently  proposed  models  for 
interior beam‐column joints (Jeon 2013; Kim and LaFave 2009; Birely et al. 2012) were evaluated using 
an  experimental  data  set  comprising  58  subassemblage  test  specimens.  Test  specimens  had  a  wide 
range of material properties, geometric configurations and design parameters and were representative 
of  modern  and  older  construction  in  the  United  States.  The  results  of  the  evaluation  include  1)  all 
models provide accurate and precise simulation of strength, 2) all models provide acceptably accurate 
simulation  of  initial  stiffness,  but  the  Kim  and  LaFave  provides  relatively  accurate  simulation  of  initial 
stiffness, and 3) the Jeon model provides accurate and the Kim and LaFave model provides acceptably 
accurate simulation of drift capacity, but all models provide poor simulation of joint shear strain demand 
at drift capacity (90% strength loss).   

2 2
Joint shear stress (ksi)

Joint shear stress (ksi)

1 1

0 0

−1 −1
Experiment Experiment
Simulation Simulation
−2 −2
−0.06 −0.04 −0.02 0 0.02 0.04 0.06 −0.09 −0.06 −0.03 0 0.03 0.06 0.09
Joint shear strain (rad)
  Joint shear strain (rad)
 
a) PEER4150 (Alire 2002), joint = 0%, max =25.4   c)  NO01  (Teraoka  1997),  joint  =  1.09%,  max 
 and  λ  =  45.6 .  Strength  is  controlled  by  =22.8    and  λ  =  33.0 .  Strength  is 
joint failure.   controlled by joint failure.  
1 12

8
Joint shear stress (ksi)
Joint shear stress (ksi)

0.5
4

0 0

−4
−0.5
Experiment −8 Experiment
Simulation Simulation
−1 −12
−0.04 −0.03 −0.02 −0.01 0 0.01 0.02 0.03 0.04 −0.06 −0.04 −0.02 0 0.02 0.04 0.06
 
Joint shear strain (rad) Joint shear strain (rad)
 
b) PEER14 (Walker 2001), joint = 0%, max =10.9   d) NO02 (Takamori et al. 2006), joint = 0.60%, max 
 and  λ  =  28.1 .  Strength  is  controlled  by  =14.9    and  λ  =  19.7 .  Strength  is 
beam yielding.   controlled by beam yielding. 

3.21
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

Figure 13: Simulated and observed joint shear stress versus shear strain response histories for interior 
joint subassemblage test specimens. 

The Kim and LaFave model was chosen for use in development of a cyclic response model for interior 
joints; this decision was based on the results of the model evaluation as well limitations of the Jeon and 
Birely  models  (the  Jeon  model  does  not  provide  prediction  of  joint  shear  strain  at  maximum  strength 
and  the  Birely  model  does  not  predict  joint  response  beyond  maximum  strength).  Cyclic  response 
parameters for the OpenSees Pinching4 material model were determined using a least‐squares, best‐fit 
method and cyclic column shear versus story drift data for the 58 specimens in the assembled data set. 
Results  show  that  the  proposed  cyclic  response  parameters  provide  accurate  simulation  of  the 
hysteretic response exhibited by beam‐column joint subassemblages tested in the laboratory. However, 
results  show  also  the  potential  for  relatively  poor  simulation  of  joint  shear  strain  demands  and  that 
relatively  accurate  simulation  of  subassemblage  response  can  be  achieved  with  relatively  poor 
simulation of joint shear response. 

The results of this study suggest that additional research is need to develop a suite of numerical models 
that  enable  accurate  simulation  of  the  nonlinear  response  of  beam‐column  joints  with  a  range  of 
geometric and design configurations found in the US building inventory.   

Acknowledgements
This  material  is  based  in  part  upon  work  supported  by  the  National  Science  Foundation  under  Grant 
Number  1000700.  Any  opinions,  findings,  and  conclusions  or  recommendations  expressed  in  this 
material  are  those  of  the  authors  and  do  not  necessarily  reflect  the  views  of  the  National  Science 
Foundation.  

References
Alire, D.A. (2002). “Seismic evaluation of existing unconfined reinforced concrete beam‐column joints,” 
MS Thesis. University of Washington, Seattle. 

Altoontash,  A.  (2004).  “Simulation  and  damage  models  for  performance  assessment  of  reinforced 
concrete beam‐column joints,” PhD Dissertation. Stanford University, CA. 

Anderson,  M.,  Lehman,  D.E.,  Stanton,  J.  (2008).  “A  cyclic  shear  stress–strain  model  for  joints  without 
transverse reinforcement.” Engineering Structures, 30: 941‐954. 

Anderson,  J.C.  and  Townsend,  W.H.  (1977).  “Models  for  reinforced  concrete  frames  with  degrading 
stiffness,” Journal of the Structural Division, ASCE, 103 (ST12): 1433‐1449. 

Baglin,  P.S.  and  Scott,  R.H.  (2000).  “Finite  Element  Modeling  of  Reinforced  Concrete  Beam‐Column 
Connections,” ACI Structural Journal, 97 (6): 886‐894. 

Berry, M. (2006). “Performance Modeling Strategies for Modern Reinforced Concrete Bridge Columns,” 
PhD Dissertation. University of Washington, Seattle.  

Birely, A., Lowes, L.N., Lehman, D.E. (2012). “A model for the practical nonlinear analysis of reinforced 
concrete frames including joint flexibility,” Engineering Structures 34: 455–465. 

Celik, O.C. (2007). “Probabilistic Assessment of Non‐Ductile Reinforced Concrete Frames Susceptible to 
Mid‐America Ground Motions,” PhD Dissertation. Georgia Institute of Technology. 

3.22
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

Deaton,  J.B.  (2013).  “Nonlinear  finite  element  analysis  of  reinforced  concrete  exterior  beam‐column 
joints with nonseismic detailing,” PhD Dissertation. Georgia Institute of Technology. 

Durrani,  A.J.  and  Wight,  J.K.  (1985).  “Behavior  of  interior  beam‐to‐column  connections  under 
earthquake‐type loading,” ACI Journal 82 (3): 343‐349. 

Elwood,  K.J.,  Matamoros,  A.B.,  Wallace,  J.W.,  Lehman,  D.,  Heintz,  J.A.,  Mitchell,  A.D.,  Moore,  M.A., 
Valley,  M.T.,  Lowes,  L.N.,  Comartin,  C.D.,  and  Moehle,  J.P.  (2007).  “Update  to  ASCE/SEI  41  concrete 
provisions,” Earthquake Spectra, 23 (3): 493‐523. 

Hall,  J.,  ed.  (1996).  Northridge  Earthquake  of  January  17,  1994  Reconnaissance  Report  Vol  1  Oakland: 
Earthquake Engineering Research Institute. 

Hassan,  W.  (2011).  “Analytical  and  Experimental  Assessment  of  Seismic  Vulnerability  of  Beam‐Column 
Joints  without  Transverse  Reinforcement  in  Concrete  Buildings,”  PhD  Dissertation.  University  of 
California, Berkeley. 

Hegger, J., Sherif, A., and Roeser, W. (2004). “Nonlinear Finite Element Analysis of Reinforced Concrete 
Beam‐Column Connections,” ACI Structural Journal, 101 (5), 604‐614. 

Holmes,  W.  and  Somers,  P.,  eds.  (1996).  Northridge  Earthquake  of  January  17,  1994  Reconnaissance 
Report Vol 2. Oakland: Earthquake Engineering Research Institute. 

Jeon,  J.‐S.  (2013).  “Aftershock  Vulnerability  Assessment  of  Damaged  Reinforced  Concrete  Buildings  in 
California,” PhD dissertation. Georgia Institute of Technology. 

Kim, J. and LaFave, J. (2009). “Joint Shear Behavior of Reinforced Concrete Beam‐Column Connections 
Subjected  to  Seismic  Lateral  Loading,”  NSEL  Report  NSEL‐020,  University  of  Illinois  at  Urbana‐
Champaign. 

Kitayama,  K.,  Otani,  S.,  and  Aoyama,  H.  (1991).  “Development  of  design  criteria  for  RC  interior  beam‐
column  joints,”  Design  of  Beam‐Column  Joints  for  Seismic  Resistance  (SP‐123),  American  Concrete 
Institute, Detroit, MI, 97‐113. 

Leon,  R.T.  (1990).  “Shear  strength  and  hysteretic  behavior  of  interior  beam‐column  joints,”  ACI 
Structural Journal, 87(1): 3‐11.  

Leon, R. and Jirsa, J.O. (1986). “Bidirectional loading of R.C. Beam‐Column Joints,” Earthquake Spectra, 2 
(3): 537‐564. 

Lowes, L.N. and Altoontash, A. (2003). “Modeling Reinforced Concrete Beam‐Column Joints Subjected to 
Cyclic Loading,” Journal of Structural Engineering, ASCE, 129 (12): 1686‐1697.  

Meinheit,  D.F.  and  Jirsa,  J.O.  (1981).  “Shear  strength  of  RC  beam‐column  connections,”  Journal  of  the 
Structural Division, ASCE, 107 (ST11): 2227‐2244.  

Mitra, N. (2007). “An analytical study of reinforced concrete beam‐column joint behavior under seismic 
loading,” PhD Dissertation. University of Washington, Seattle.  

Mitra,  N.  and  Lowes,  L.N.  (2007).  “Evaluation,  Calibration,  and  Verification  of  a  Reinforced  Concrete 
Beam–Column Joint Model,” Journal of Structural Engineering, ASCE, 133 (1): 105‐120.  

3.23
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

Moiser,  W.G.  (2000).  “Seismic  assessment  of  reinforced  concrete  beam‐column  joints,”  MS  Thesis. 
University of Washington, Seattle. 

Oka,  K.  and  Shiohara,  H.  (1992).  “Tests  of  high‐strength  concrete  interior  beam‐column‐joint 
subassemblages,” 10th World Conference on Earthquake Engineering, Madrid, Spain, 3211‐3217. 

Otani,  S.  (1974).  “Inelastic  analysis  of  reinforced‐concrete  frame  structures,”  Journal  of  the  Structural 
Division, ASCE, 100 (ST7): 1422‐1449. 

Oyen,  P.  (2006).  “Evaluation  of  Analytical  Tools  for  Determining  the  Seismic  Response  of  Reinforced 
Concrete Shear Walls,” PhD Dissertation. University of Washington, Seattle. 

Pantazopoulou,  S.J.  and  Bonacci,  J.F.  (1994).  “On  earthquake  resistant  reinforced  concrete  frame 
connections,” Canadian Journal of Civil Engineering, 21 (2): 307–328. 

Pantelides,  C.P.,  Hansen,  J.,  Nadauld,  J.,  Reaveley,  L.D.  (2002).  “Assessment  of  reinforced  concrete 
building exterior joints with substandard details,” PEER Report 2002/18, Pacific Earthquake Engineering 
Center, University of California, Berkeley, CA. 

Park,  S.  (2010).  “Experimental  and  Analytical  Studies  on  Old  Reinforced  Concrete  Buildings  with 
Seismically Vulnerable Beam‐Column Joints,” PhD Dissertation. University of California, Berkeley. 

Park,  R.  and  Ruitong,  D.  (1988).  “A  Comparison  of  the  behavior  of  reinforced  concrete  beam‐column 
joints  designed  for  ductility  and  limited  ductility,”  Bulletin  of  the  New  Zealand  National  Society  of 
Earthquake Engineering, 21 (4): 255‐278. 

Pessiki,  S.P.,  Conley,  C.H.,  Gergely,  P.,  and  White,  R.N.  (1990).  “Seismic  behavior  of  lightly‐reinforced 
concrete column and beam‐column joint details,” Technical Report NCEER‐90‐0014, National Center for 
Earthquake Engineering Research, State University of New York, Buffalo.  

Sharma, A., Eligehausen, R., and Reddy, G.R. (2011). “A new model to simulate joint shear behavior of 
poorly detailed beam‐column connections in RC structures under seismic loads, Part I: Exterior joints,” 
Engineering Structures, 33 (3): 1034‐1051.  

Shin, M. and LaFave, J.M. (2004). “Modeling of Cyclic Joint Shear Deformation Contributions in RC Beam‐
Column connections to overall frame behavior,” Structural Engineering and Mechanics, 18 (5): 645‐669. 

Tajiri,  S.,  Shiohara,  H.,  and  Kusuhara,  F.  (2006).  “A  new  macroelement  of  reinforced  concrete  beam‐
column joint for elasto‐plastic plane frame analysis,” Proceedings of the 8th U.S. National Conference on 
Earthquake Engineering, San Francisco, California.  

Takamori,  N.,  Hayashi,  K.,  Sasaki,  S.,  and  Teraoka,  M.  (2006).  “Experimental  study  on  full‐scale  R/C 
interior beam‐column joints (in Japanese),” Proceedings of the Japan Concrete Institute, 28 (2): 283‐288. 

Teraoka, M. (1997). “A study on seismic design of R/C beam‐column joint in high rise frame structure (in 
Japanese),” Research Report, Extra Issue No. 5, Fujita Institute of Technology, Tokyo, Japan. 

Walker, S. (2001). "Seismic performance of existing reinforced concrete beam‐column joints," MS Thesis. 
University of Washington. 

Youssef,  M.  and  Ghobarah,  A.  (2001).  “Modeling  of  RC  Beam—Column  Joints  and  Structural  Walls,” 
Journal of Earthquake Engineering, 5 (1):93‐111. 

3.24
@Seismicisolation
@Seismicisolation
Numerical Models for Beam‐Column Joints in Reinforced Concrete Building

References Used in the Database 

Alire, D. 2002. "Seismic evaluation of existing unconfined reinforced concrete beam‐column joint," MS 
Thesis. University of Washington. 

Endoh,  Y.,  Kamura,  T.,  Otani,  S.,  and  Aoyama,  H.  (1991).  “Behavior  of  RC  beam‐column  connections 
using light‐weight concrete,” Transactions of the Japan Concrete Institute, 13: 319‐326. 

Fujii,  S.  and  Morita,  S.  (1991).  “Comparison  between  interior  and  exterior  RC  beam‐column  joint 
behavior,”  Design  of  Beam‐Column  Joints  for  Seismic  Resistance  (SP123),  American  Concrete  Institute, 
Detroit, MI, 145‐165. 

Goto,  Y.  and  Joh,  O.  (1996).  “An  experimental  study  on  shear  failure  mechanism  of  RC  interior  beam‐
column joints,” 11th World Conference on Earthquake Engineering, Paper No. 1194, Acapulco, Mexico. 

Goto,  Y.  and  Joh,  O.  (2003).  “Experimental  study  on  shear  resistance  of  RC  interior  eccentric  beam‐
column  joints,”  9th  East  Asia‐Pacific  Conference  on  Structural  Engineering  and  Construction,  Bali, 
Indonesia, RSC‐170. 

Kurose,  Y.,  Guimaraes,  G.N.,  Zuhua,  L.,  Kreger,  M.E.,  and  Jirsa,  J.O.  (1991).  “Evaluation  of  slab‐beam‐
column  connections  subjected  to  bi‐directional  loading,”  Design  of  Beam‐Column  Joints  for  Seismic 
Resistance (SP123), American Concrete Institute, Detroit, MI: 39‐67. 

Matsumoto,  T.,  Nishihara,  H.,  Nakao,  M.,  and  Castro,  J.J.  (2010).  “Performance  on  shear  strength  of 
reinforced concrete eccentric beam‐column joints subjected to seismic loading,” Proceedings of the 9th 
U.S.  National  and  10th  Canadian  Conference  on  Earthquake  Engineering,  Paper  No.  1457,  Toronto, 
Canada. 

Noguchi,  H.  and  Kashiwazaki,  T.  (1992).  “Experimental  studies  on  shear  performances  of  RC  interior 
column‐beam joints with high‐strength materials,” 10th World Conference on Earthquake Engineering, 
Madrid, Spain, 3163‐3168. 

Ohwada, Y. (1984). “Studies on reinforced concrete beam‐column joints under biaxial seismic load (2) (in 
Japanese),”  Summaries  of  Technical  Papers  of  Annual  Meeting  Architectural  Institute  of  Japan,  1869‐
1870. 

Oka,  K.  and  Shiohara,  H.  (1992).  “Tests  of  high‐strength  concrete  interior  beam‐column‐joint 
subassemblages,” 10th World Conference on Earthquake Engineering, Madrid, Spain, 3211‐3217. 

Ozaki,  E.,  Nishio,  A.,  Tajima,  K.,  and  Shirai,  N.  (2010).  “Damage  evaluation  of  RC  beam  column  joints 
based  on  the  image  measurement  (in  Japanese),”  Proceedings  of  the  Japan  Concrete  Institute,  32  (2): 
301‐306. 

Raffaelle,  G.S.  and  Wight,  J.K.  (1995).  “Reinforced  concrete  eccentric  beam‐column  connections 
subjected to earthquake‐type loading,” ACI Structural Journal, 92 (1): 45‐55. 

Shin, M. and LaFave, J.M. (2004) “Modeling of cyclic joint shear deformation contributions in RC beam‐
column connections to overall frame behavior,” Structural Engineering and Mechanics, 18 (5): 645‐669. 

3.25
@Seismicisolation
@Seismicisolation
Jong-Su Jeon et al.

Takamori,  N.,  Hayashi,  K.,  Sasaki,  S.,  and  Teraoka,  M.  (2006).  “Experimental  study  on  full‐scale  R/C 
interior beam‐column joints (in Japanese),” Proceedings of the Japan Concrete Institute, 28 (2): 283‐288. 

Teng,  S.  and  Zhou,  H.  (2003).  “Eccentric  reinforced  concrete  beam‐column  joints  subjected  to  cyclic 
loading,” ACI Structural Journal, 100 (2): 139‐148. 

Teraoka, M. (1997). “A study on seismic design of R/C beam‐column joint in high rise frame structure (in 
Japanese),” Research Report, Extra Issue No. 5, Fujita Institute of Technology, Tokyo, Japan. 

Walker,  S.G.  (2001).  “Seismic  performance  of  existing  reinforced  concrete  beam‐column  joints,”  MS 
Thesis. University of Washington, Seattle. 

3.26
@Seismicisolation
@Seismicisolation
SP-297—4

 
Evaluation of ASCE 41 Modeling Parameters for Slender Reinforced Concrete Structural Walls

Anna C. Birely, Laura N. Lowes, and Dawn E. Lehman

Synopsis: ASCE/SEI 41-06 provides guidelines for evaluating the seismic adequacy of existing buildings. For
nonlinear dynamic analysis of a building, ASCE 41 provides modeling parameters to define the backbone curve for
the response of structural components. Seismic adequacy is then determined by comparing simulated response to
predetermined acceptance criteria. In the reinforced concrete (RC) community, there is interest in evaluating the
modeling parameters and acceptance criteria for RC components, and if deemed necessary, developing updated
values that reflect the current state of understanding of the seismic performance of RC components. For some
structural components (e.g. columns), large databases of experimental data can be used to develop empirical
acceptance criteria that reflect the behavior of the component. In the case of slender structural walls, relatively
limited tests have been conducted such that sufficient variation in critical design and loading characteristics
including shape, aspect ratio (elevation and cross-sectional), confinement, and axial load are not represented by
experimental data to justify use of an experimental database to develop acceptance criteria. Evaluation of this
limited set of experimental data indicates current ASCE 41 modeling parameters and acceptance criteria for flexure-
controlled walls is inappropriate, generally resulting in overprediction of wall deformation capacity at high axial
load ratios and underprediction at low axial load ratios and low shear demands. Although suitable for evaluation of
criteria, the data set is not sufficiently varied such that revised provisions can be developed. To overcome the lack of
sufficient experimental data, a parameter study was conducted to provide data to support development of updated
acceptance criteria. The parameter study was conducted using a modeling approach validated to provide accurate
simulation of flexural failures in slender reinforced concrete walls. Simulation results were used to develop
preliminary recommendations for revised modeling parameters for slender RC walls. An evaluation of these
simulation results and preliminary recommendations for revised flexure-controlled RC wall modeling parameters are
presented in this paper.

Keywords: ASCE 41; reinforced concrete; walls; flexure-control; acceptance criteria; plastic hinge rotation

4.1
@Seismicisolation
@Seismicisolation
A. C. Birely et al.

Anna C. Birely is an assistant professor in the Zachry Department of Civil Engineering at Texas A&M University.
She received her BS from the University of Colorado and her MS and PhD in civil engineering from the University
of Washington. She is a member of ACI Committees 369, Seismic Repair and Rehabilitation; 374, Performance-
Based Seismic Design of Concrete Buildings; and 133, Disaster Reconnaissance. Her research interests include the
performance of reinforced concrete structures subjected to hazardous loads, including earthquakes.

Laura N. Lowes is an associate professor in the Department of Civil and Environmental Engineering at the
University of Washington. She received her BS from the University of Washington and her MS and PhD from the
University of California, Berkeley. She is a member of ACI Committee 369, Seismic Repair and Rehabilitation, and
Joint ACI-ASCE Committees 445, Shear and Torsions and 447, Finite Element Analysis of Reinforced Concrete
Structures. Her research interests include numerical modeling of concrete structures and earthquake engineering.

Dawn E. Lehman is an associate professor in the Department of Civil and Environmental Engineering at the
University of Washington. She received her BS from Tufts University and her MEng and PhD from the University
of California, Berkeley. She is a member of ACI Committees 341, Earthquake Resistant Concrete Bridges; 374,
Performance-Based Seismic Design of Concrete Buildings; and Joint ACI-ASCE Committee 352, Joints and
Connections in Monolithic Concrete Structures. Her research interests include seismic design of reinforced concrete
and composite structures.

1. INTRODUCTION

ASCE/SEI 41-06 Seismic Rehabilitation of Existing Buildings (ASCE-41) provides provisions to guide in the
seismic evaluation of existing buildings. The first generation of ASCE 41 was based extensively on the precursor
document FEMA 356 Prestandard and Commentary for the Seismic Rehabilitation of Buildings (FEMA-356, 2000).
A supplement was developed in 2007 (ASCE-41, 2007) to improve the accuracy of modeling parameters and
acceptance criteria for structural components. Ideally, such values are based on extensive experimental evidence
such that empirical models can be established for the range of component characteristics that may be found in
existing buildings. In anticipation of future editions of ASCE 41, modeling parameters and acceptance criteria for
some reinforced concrete components have been reevaluated. Evaluation of modeling parameters and acceptance
criteria for slender reinforced concrete walls are addressed in this paper. ASCE 41 classifies walls based on the
expected behavior being either shear-controlled and flexure-controlled. Provisions for shear-dominated walls were
updated in the most recent supplement; however, provisions for flexure controlled walls have remained largely
unchanged since the original FEMA 356 document, for which experimental justification of the modeling parameters
and acceptance criteria is not available. An evaluation of the provisions for slender (flexure-controlled) reinforced
concrete structural walls and recommendations for modifications to future versions of ASCE 41 are presented here.
The current ASCE 41 provisions classifies structural walls as being either flexure-controlled or shear-
controlled. While no explicit distinction is made between the two, commentary in Section C.7.1 indicates that
flexure-controlled walls should be taken as those with an aspect ratio (AR = height/length) of 3.0 or greater, shear-
controlled walls should be taken as those with an aspect ratio of 1.5 or less, and walls with intermediate response
should be considered to have response controlled by both flexure and shear. This assumption for wall behavior
based on the aspect ratio is commonly used to define walls as squat (shear-controlled) or slender, with an aspect
ratio of 1.5 or 2.0 typically used to indicate the transition between the two wall types. This aspect ratio has, for
experimental subassemblage tests, proved to be an accurate indicator of whether the failure mode of the wall is shear
(AR < 2.0) or flexure (AR > 2.0) dominated. Most such specimens are loaded with a single lateral force at the top of
the wall, and thus, the aspect ratio equal to the shear span ratio ((effective height of lateral load)/length), which
essentially describes the relationship between shear and flexure base reactions and is a more appropriate descriptive
quality to consider in a multi-story building. Thus, flexure-controlled walls in this study are considered to have a
shear span ratio, M/Vlw, greater than 1.5. In the case of investigating multistory walls, the shear span ratio is
determined by assuming a uniform distribution of lateral forces at each floor.
Section 2 provides an overview of the existing provisions for flexure-controlled walls. Section 3 provides
an evaluation of these provisions based on a database of experimental data. Due to a fairly limited number of tests
available from which to develop comprehensive empirically based recommendations for future versions of the
ASCE 41 provisions, a parameter study was conducted to provide simulation data to support the development new
recommendations. Section 4 provides an overview of the parameter study, which was conducted using numerical
modeling techniques validated using much of the experimental data used to evaluate existing provisions. Section 5
provides an analysis of trends in the experimental and simulation data to identify the parameters that should be used

4.2
@Seismicisolation
@Seismicisolation
Evaluation of ASCE 41 Modeling Parameters for
Slender Reinforced Concrete Structural Walls

identifying modeling parameters and acceptance criteria. Finally, Section 6 provides recommendations for updated
values.

2. ASCE 41 PROVISIONS FOR FLEXURE-CONTROLLED WALLS

ASCE 41 provides provisions for both linear and nonlinear analysis of existing structures. Linear procedures
determine the adequacy of a structural component based on the ratio of the simulated demand to the strength, with
consideration for over strength in linear models accounted for by `m-factors’, where m-factor values are provided in
tabulated form based on key characteristics of a component. Linear procedures are not evaluated here.
For nonlinear analysis of a structure, the response is represented using a load-deformation response such as
the one shown in Figure 1. For structural walls controlled by the flexural response, the load-deformation response is
the moment-rotation response, where the rotation is the rotation in the plastic hinge region, Lp, at the base of the wall
and Lp is defined as equal to one-half the wall length, but not greater than the floor height. Point ‘B’ in Figure 1 is
the yield point of the wall. Points ‘C’, ‘D’, and ‘E’ are determined by the values ‘a’, ‘b’, and ‘c’, defined by
tabulated values in ASCE 41 (see Figure 2). For flexure-controlled walls, these parameters are specified as a
function of the axial load ratio, the shear demand, and confinement of the boundary regions, with linear interpolation
permitted between the tabulated values.
In addition to specifying modeling parameters for use in nonlinear models, ASCE 41 provides acceptance
criteria. For structural walls controlled by the flexural response, the acceptance criteria specify the maximum plastic
hinge rotations that shall not be exceeded. The maximum plastic hinge rotations are specified as a function of the
axial load ratio, the shear demand, and confinement in the boundary regions (see Figure 2). Six sets of acceptance
criteria are provided; selection of the appropriate set is driven by the desired performance level (Immediate
Occupancy (IO), Life Safety (LS), or Collapse Prevention (CP)) and if the component studied is a primary
component or a secondary component.

Figure 1 - Generic load-deformation response for flexural members (from ASCE/SEI 41 supplement 1).

3. EVALUATION OF CURRENT PROVISIONS USING EXPERIMENTAL DATA

The current ASCE/SEI 41-06 provisions were evaluated by comparing modeling parameters and acceptance criteria
to the equivalent response from experimental tests. The modeling parameters that are provided are the plastic hinge
rotations ‘a’ and ‘b’, and the residual strength ratio ‘c’ (see Figure 1). The plastic hinge rotation ‘a’ is the change in
rotation from yield of the component to loss of lateral load carrying capacity. The plastic hinge rotation ‘b’ is the
change in rotation from the loss of lateral load carrying capacity to the rotation at which the wall lateral capacity is
reduced. For flexure-controlled walls, the critical value to be evaluated is ‘a’; the additional plastic hinge rotation
following loss of lateral load carrying capacity and the residual strength ratio are secondary response characteristics
that do not have dramatic effects on determining whether or not a structural component is seismically adequate.
Additionally, available experimental data typically does not include the residual strength and/or the decay from
initial failure to the residual strength. Thus, the discussion presented here will be limited to a discussion of the
plastic hinge rotation value ‘a’.
ASCE 41 modeling parameters and acceptance criteria for flexure-controlled walls were evaluated using a
database of experimental wall tests consisting of 66 walls with shear span ratios equal to or greater than 2.0. Details
of the database are provided by Birely (2012). Not all experimental test programs reported measurements that could

 
4.3
@Seismicisolation
@Seismicisolation
A. C. Birely et al.

Figure 2 - Acceptance criteria for reinforced concrete walls controlled by flexure (from ASCE/SEI 41
supplement 1). Shear stress demands, V/twlw√f’c, are for concrete strengths in psi; for MPa, multiply by
1/12.043.

be used to calculate rotation and, for those that did, the measurements were not consistent between test programs.
Thus, to provide a consistent measure for the rotation in the walls, an equivalent plastic hinge rotation was
calculated from the reported load-displacement histories and geometric properties of the walls. The equivalent
plastic hinge rotation was defined as the rotation in a lumped-plasticity model that would result in the experimental
load-displacement history. The lumped-plasticity model comprised a rotational hinge at the base of the wall with a
hinge length of one-half the length of the wall. The region outside the hinge was modeled as elastic with an effective
flexural stiffness of 0.5EcIg and an effective shear stiffness of 0.4EcIg.
Experimental ‘a’ values were calculated by subtracting the equivalent plastic hinge rotation at yield from the
equivalent plastic hinge rotation at the drift capacity. To ensure consistency between the experimental specimens,
for which varying levels of details are available in the literature, the yield points were defined as the moment at
which the extreme reinforcing bar in tension yielded in a moment-curvature analysis of the cross-section (note that
this is different than the calculation of the yield point specified in ASCE 41). The drift capacity was defined as either
the maximum experimental drift or the drift at which a 10% loss of lateral load carrying capacity occurred.
The literature for walls supports the general understanding by the engineering community that the performance
of structural walls, particularly when measured by the drift capacity of the wall, is greatly impacted by the axial load
ratio, the shear demand, and the confinement of boundary regions. Other factors have also been shown to impact the
performance, but it is these three that are the biggest factors, and it is these on which the current ASCE 41
provisions are based.

4.4
@Seismicisolation
@Seismicisolation
Evaluation of ASCE 41 Modeling Parameters for
Slender Reinforced Concrete Structural Walls

The experimental values are compared to the ASCE 41 values in Figures 3 and 4. Figure 3 compares the values
for the modeling parameter ‘a’. The values are shown versus the experimental values for axial load ratio (left
column) and maximum shear stress demand (right column). The first row shows all test specimens (rectangular,
barbell, C-shape, H-shape, and T-shape walls). The second and third rows show the rectangular and flanged walls
(C-, H-, and T-shape), respectively. For tests with bi-directional loading, reported data is for strong-axis bending.
Figure 4 provides the same set of comparisons but for the acceptance criteria (maximum plastic hinge rotation). The
following general observations can be made:

1. Overall, the modeling parameter ‘a’ is conservative; however, for the limited data points at higher axial
load ratio in rectangular walls, the ASCE 41 values are unconservative.
2. There is a lack of experimental data in the range of 3-4√f’c psi (0.25-0.33√f’c MPa), a reasonable shear
demand in many walls, particularly rectangular walls.
3. For rectangular walls, there is a general decrease in the experimental values as the axial load ratio and the
maximum shear stress demand increase, confirming observations in the literature that these factors impact
the performance of structural walls and justifying the use of these values as selection criteria for the ASCE
41 modeling and acceptance criteria.
4. For flanged walls, there are no distinct relationships between the experimental values for ‘a’ and the axial
load ratio or shear stress demand.
5. Relative to the magnitude of the experimental values, the distinction between acceptable plastic hinge
rotations for Life Safety (LS) and Collapse Prevention (CP) are minor. In applying these limits, it seems
unlikely to achieve LS without achieving CP, an observation that is consistent with analysis of buildings
damaged in the 2010 Maule earthquake using the ASCE 41 acceptance criteria (Birely et al. 2013).

Although the experimental data indicates trends, there is limited data, particularly for walls with axial load
ratios in excess of 10% and flanged walls. The top-left plot in Figure 3 clearly shows that many experimental tests
have no axial load, an unrealistic condition in buildings. Figure 5 shows, for rectangular and flanged walls, the
relationship between ‘a’ and the shear stress demand for walls with axial loads, illustrating the limited data for use in
developing recommendations for revised ASCE 41 modeling parameters and acceptance criteria. Additionally,
despite visible trends in Figures 3-5, other parameters may also be crucial to the performance of walls. Birely (2012)
investigated the impact of a wide range of wall characteristics on the drift capacity and damage progression for the
same set of data presented here and found that other characteristics may impact the performance of walls as well, yet
there is insufficient data to state this with certainty.

4. PARAMETER STUDY

The experimental data for slender RC walls is sufficiently limited that the empirical calibration of
recommendations for revised acceptance criteria is not practical for slender walls as it may be for other components
for which hundreds of experimental data points are available. To provide a more diverse set of data for use in
developing proposed revisions to ASCE 41, a parameter study was conducted to generate additional data. The
parameter study was conducted using OpenSees following modeling recommendations by Pugh (2012). The
following sections provide an overview of the models used and the characteristics of the parameter study.

4.1 OpenSees numerical model

The parameter study was conducted using OpenSees nonlinear force-based beam-column elements. One element
was used per floor, with five integration points per element. Each integration point consisted of an aggregated
section made up of independent responses of the wall to i) shear demand, and ii) flexural and axial demands. Out-of-
plane displacements were restrained, thus, the torsional stiffness of the walls was not considered.
Shear was model using an elastic shear stiffness equal to the gross shear stiffness of the wall in the
direction of bending. It should be noted that the other more sophisticated models for the shear response, including
those that couple the shear and flexure response of the walls, would likely result in more accurate models of the
cyclic response of the walls; however, as the focus of the developed acceptance criteria is on the ability to capture
the flexural response of the wall, the approach used for modeling the shear is reasonable, is the simplest method
available for including shear, and is consistent with the shear stiffness in the current version of ASCE 41.
The flexural and axial response of the walls was modeled using fiber sections followed the
recommendations of Pugh (2012), which were validated for simulation of planar walls using an extended data set;
 
4.5
@Seismicisolation
@Seismicisolation
A. C. Birely et al.

results show accurate simulation of stiffness to yield, strength, and drift capacity. Initial evaluation of the model for
non-planar walls using limited experimental data suggests that the model is appropriate for C-shaped walls but not
for T-shaped walls. For T-shaped walls, the assumptions of a linear strain field on the wall cross-section, which is
inherent to fiber-based beam-column elements, can result in significant error in simulated response. The basis for
Pugh’s recommendations is that distributed plasticity elements have the potential for localization of damage at the
critical section and that the global response is sensitive to the number of integration points used. To achieve an
objective response regardless of mesh size, the proposed recommendations use post-peak regularization of the
stress-strain response of the material models.

Axial Load Ratio Shear Stress Demand


All Walls
Rectangular Walls
Flanged Walls (H-, T-, C-shape)

Figure 3 - Comparison of experimental values for the parameter ‘a’ and the ASCE 41 values for ‘a’;
experimental specimens used are detailed in Birely (2012). Shear stress demands shown are for concrete
strengths in psi; for MPa, multiply by 1/12.043.

4.6
@Seismicisolation
@Seismicisolation
Evaluation of ASCE 41 Modeling Parameters for
Slender Reinforced Concrete Structural Walls

All Walls
Rectangular Walls
Flanged Walls (H-, T-, C-shape) Axial Load Ratio Shear Stress Demand

Figure 4 - Comparison of experimental rotation at failure and the ASCE 41 values for acceptable plastic
hinge rotation; experimental specimens used are detailed in Birely (2012). Shear stress demands shown are
for concrete strengths in psi; for MPa, multiply by 1/12.043.

 
4.7
@Seismicisolation
@Seismicisolation
A. C. Birely et al.

(a) Rectangular Walls (b) Flanged Walls

Figure 5 - Relationship between shear and experimental values for the parameter ‘a’ and the ASCE 41 values
for ‘a’ for experimental specimens with axial load. Shear stress demands shown are for concrete strengths in
psi; for MPa, multiply by 1/12.043.

4.2 Wall characteristics

The geometry of the parameter study walls was selected to provide a range of demands while falling within a range
representative of building characteristics for low- to mid-rise buildings. Two wall heights, 4 and 8 floors (10 ft. floor
heights), were considered. Wall thicknesses of 8 in. (200 mm) and 12 in. (300 mm) were used in combination with
cross-sectional aspect ratios of 10, 15, and 20. This resulted in aspect ratios (wall height/wall length) ranging from
2-12. If a uniform distribution of lateral loads is assumed, the resulting shear span ratios could range from 1.25-6.75.
To ensure that only flexure dominated walls were considered in the study, only shear span ratios equal to or greater
than 1.5 were used in developing acceptance criteria. The cross-sectional aspect ratios used are representative of the
range found in most building. The thicknesses used are at the lower range of those found in mid-rise buildings,
however, it is worth noting that thicker walls would typically be found in taller walls and that the corresponding
range of shear span ratio and cross-sectional aspect ratio is accounted for in the limited parameter study presented
here.
Planar and symmetric flanged walls were considered. Symmetric walls were defined to have two identical
flanges at both ends of the walls. The lengths of the flanges were defined as 10%, 25%, or 50% of the length of the
wall. For the range of wall heights and lengths used in the parameter studies, no flanges exceeded the effective
flange width specified by ACI 318.
The distribution of longitudinal reinforcing bars was considered to follow two general layouts. The first
longitudinal reinforcement is a uniform distribution of the bars throughout the wall. For these walls, reinforcement
ratios of 0.25%, 0.5%, 1.0%, and 1.5% were used. Walls with 0.25% reinforcement meet the minimum required by
the present ACI 318 code (ACI-318 2011). For older walls that may be evaluated with ASCE 41, the gross
reinforcement ratio may be less than 0.25%; however, such walls are essentially unreinforced and therefore not
consider here. The second longitudinal reinforcement layout is that commonly used in modern walls, with most
reinforcement concentrated in the end region (boundary element). For the planar and symmetric flanged walls, a
reinforcement ratio of 0.25% was used in web (the region between the boundary elements) and the reinforcement in
the boundary elements was set such that the gross reinforcement ratio of the wall was 0.5%, 1.0%, or 1.5%. In the
walls with boundary elements, 0.25% was not considered as adding boundary elements would not impact the
distribution of the steel. For unconfined walls, this would result in walls similar to the uniform distribution of steel
and confined boundary elements are unlikely to be found in such lightly reinforced walls. In the planar walls, the
boundary element length was considered to be 10% or 20% of the total wall length. For the symmetric flanged walls,
each flange was considered to be the boundary element.
The material properties were kept consistent for all walls considered in the parameter study. Lower-bound
values were adopted from ASCE 41 Tables 6-2 and 6-3 for the most recent time periods; for concrete, compressive
strength of 4,000 psi (28 MPa) and for reinforcing steel, a yield strength of 60 ksi (420 MPa). The lower-bound
properties were multiplied by the factors in ASCE 41 Table 6-4 (1.5 for concrete and 1.25 for reinforcing steel) to

4.8
@Seismicisolation
@Seismicisolation
Evaluation of ASCE 41 Modeling Parameters for
Slender Reinforced Concrete Structural Walls

convert to expected material properties. Concrete was modeled using the OpenSees Concrete02 uniaxial material,
with a residual concrete strength of 20% of the peak concrete strength. Steel was modeled using the OpenSees
Steel02 uniaxial material with strain limits included to indicate failure of the specimens. The compressive strain
limit was set equal to the concrete compressive strain when the residual strength is first reached, essentially
indicating that concrete crushing and buckling will occur simultaneously; such an approach was deemed reasonable
by Pugh (2012) for modeling the response of experimental test specimens. The tension strain limit was set to be
30%, ensuring a very ductile steel response; no considerations were made for fracture due to low-cycle fatigue.
Three levels of confinement were considered for the boundary elements: unconfined, confined, and well-
confined. The distinction between well-confined and confined was intended to capture the difference between
boundary elements that, essentially, meet the ACI 318 provisions for special boundary elements (well-confined) and
boundary elements for which some level of confinement is provided to delay bar bucking but which cannot be
considered as special boundary elements. The distinction between the two levels of confinement is intended to
provide data to support maintaining the current ASCE 41 practice of not only defining values for confined and
unconfined boundary elements, but also reduced confined values for boundary elements that are confined but do not
fully meet the ACI 318 provisions for special boundary elements.

4.3 Loading

For each wall design analyzed in the parameter studied, six separate analyses were conducted, each with a different
axial load ratio. The lowest axial load ratio used was 2% of the gross axial capacity of the wall, which was
considered to be a lower bound on the axial load ratio that would account for the self-weight of the wall and a very
small tributary area of load. Other axial load ratios used ranged from 5-25% of the gross-capacity axial capacity of
the wall, in increments of 5%. An axial load ratio of 25% of the gross-capacity would be considered to be a very
large axial load ratio in a structural wall; however, coupled walls, particularly planar walls, have been estimated to
sustain axial compressive demands as large as 40% of the gross-capacity (Lehman et al. 2013) due to the axial
demands introduced by coupling beam shear.
For all walls, lateral loads were distributed uniformly at all floors. Taking into account the cross-sectional
geometry of the walls studied and the use of four- and eight-story walls, the resulting shear span ratios ranged from
1.25-6.75; however, only walls with shear span ratios greater than 1.5 were used.
The same reverse-cyclic drift history was used for each wall to ensure consistency of the compared data;
this meant using different displacement histories for the four- and eight-story walls. Two full cycles were completed
for each peak drift. The drifts used were intended to capture the cyclic response of the walls at both pre- and post-
yield levels. Drifts of 0.01%, 0.02%, 0.05%, 0.1%, 0.2%, 0.3%, 0.4%, 0.5%, 0.75%, 1.0%, 1.5%, 2.0%, 2.5%, 3.0%,
3.5%, 4.0%, 4.5%, and 5.0% were used. If failure had not been simulated at 5% drift, additional cycles were
completed to peak drifts increasing by increments of 1.0%.

4.4 Post-processing

Results of the parameter study were evaluated and any walls with maximum shear stress demand in excess
of 8√f’c psi (0.66√f’c MPa), the limiting shear stress in ACI 318, were removed. Yield of the walls was defined as
when the extreme tension reinforcing bar reached the yield strain. Failure of the walls was defined to occur when a
10% loss of shear occurred.
Although the use of the distributed plasticity beam-column elements would permit the calculation of the
wall rotation at yield and failure, it was desired to provide values that were consistent with the lumped-plasticity
approach in ASCE 41 and that were compatible with the experimental rotations considered. Thus, the equivalent
plastic hinge rotation approach was used, in which the rotation reported is the plastic hinge rotation back-calculated
from the simulated lateral drift of the wall.

5. RESULTS

The simulation results were evaluated to determine the appropriate parameters to use in providing recommendations
for revised modeling parameters and acceptance criteria for ASCE 41. While all modeling parameters (‘a’, ‘b’, and
‘c’) and acceptance criteria (for three performance levels and two component importance categories) matter, the
most critical is the parameter for ‘a’. The value ‘a’, a key aspect of the backbone curve shown in Figure 1, defines
the plastic hinge rotation from initial yield to loss of significant loss of load carrying capacity. The other modeling
parameters define the response after loss of load carrying capacity and the acceptance criteria provide acceptable
 
4.9
@Seismicisolation
@Seismicisolation
A. C. Birely et al.

total plastic hinge rotations for specified performance levels. These values can be defined in relationship to the
parameter ‘a’. In ASCE/SEI 41-06, the acceptance criteria for the collapse prevention performance level of primary
components are defined as equal to ‘a’ and all other acceptance criteria are a percentage of this value. Thus, the
evaluation of the simulation data presented here focuses solely on the ‘a’ parameter, with the objective of achieving
the following:

1. Evaluate the expectation that shear and axial load ratio are the primary variables that impact the
performance of walls and thus should be the characteristics upon which the modeling parameter ‘a’ is
defined.
2. Establish appropriate limits, if any, for shear and axial demands if tabulated values of modeling parameters
and acceptance criteria are desired.
3. Evaluate the impact of confinement on the value of ‘a’, including varied levels of confinement (i.e. well
confined versus poorly confined), and establishing how, if at all, confinement should be accounted for in
determining modeling and acceptance criteria.
4. Determine if other wall characteristics should impact the definition of modeling and acceptance criteria.

Values for ‘a’ from the parameter study walls are shown against the shear stress demand in Figure 6. The
simulation data are indicated by open black circles. For comparison, values for experimental data with axial load are
shown by solid blue diamonds. It is evident that the simulated results are reasonable in comparison to the
experimental data. The ‘a’ parameter values for each simulated wall as specified by ASCE 41 are indicated by red
lines. A detailed examination of the simulated data is presented in the sections that follow.

Figure 6 - Relationship between shear stress demand and simulated values for the parameter ‘a’ and the
ASCE 41 values for ‘a’. Shear stress demands shown are for concrete strengths in psi; for MPa, multiply by
1/12.043.

5.1 Axial load ratio and shear demand limits

A cursory evaluation of the wall design and loading characteristics impacting the value of ‘a’ indicates that,
as anticipated based on experimental data, shear stress demand and axial load ratio have a significant relationship
with the value of ‘a’. These two characteristics are highlighted in Figure 7, where ‘a’ is shown as a function of the
shear demand, with separate plots for each axial load ratio considered. Planar and flanged walls are shown
separately (Figure 7a and 7b, respectively). Simulated data is indicated by black circles. ASCE/SEI 41-06 values are
indicated by red lines; the three distinct lines shown are a result of the three confinement levels considered.
Key observations from evaluation of the data presented in Figure 7 include:
1. The impact of the axial load ratio is significantly greater in the simulated values for ‘a’ than is provided for
in the ASCE 41 modeling parameters. At low axial load ratios (2-5% of the gross axial capacity) the ASCE

4.10
@Seismicisolation
@Seismicisolation
Evaluation of ASCE 41 Modeling Parameters for
Slender Reinforced Concrete Structural Walls

41 values fall within the broad range of ‘a’ values. At axial load ratios of 10% or greater, the ASCE 41
values are unconservative, with the values for confined and well confined walls falling above the maximum
simulated values.

(a) Planar Walls

(b) Symmetric Flanged Walls


Figure 7 - Relationship between shear stress demand and simulated values for the parameter ‘a’ and the
ASCE 41 values for ‘a’ for varying levels of axial load ratio, λN = P/(Agf’c). Shear stress demands shown are
for concrete strengths in psi; for MPa, multiply by 1/12.043.

 
4.11
@Seismicisolation
@Seismicisolation
A. C. Birely et al.

2. Current ASCE 41 values have lower limits at 10% axial load ratio and upper limits at 25% axial load ratio,
with interpolation permitted in between the two limits. Figure 7 indicates that such limits are inappropriate.
With increasing axial load ratio from 2% to 5% to 10%, the range of values for ‘a’ decreases significantly.
At axial load ratios greater than 10%, the range of values for ‘a’ continues to decrease, but the drop is less
dramatic than at smaller axial loads. Consequently, revised limits would be warranted.
3. At high axial load ratios, there is limited variation in the values for ‘a’ as a function of the shear stress
demand. At lower axial load ratios, there is a clear decrease in the value for ‘a’ as the shear stress demands
increases, however, this trend levels out at shear stress demands of approximately 3-5√f’c psi (0.25-0.42√f’c
MPa), indicating that the current lower and upper bound limits for shear demands may not be the
appropriate limits to use.
4. The planar and flanged walls are shown separately to illustrate the observation that, for a specified shear
demand, particularly at lower axial load ratios, the value ‘a’ can be expected to be much larger in flanged
walls. Thus, it may be necessary to provide separate modeling parameters and acceptance criteria for planar
and flanged walls, or to define these values based on characteristics that will account for these differences.

5.2 Confinement

Current ASCE 41 modeling parameters and acceptance criteria for flexure-controlled walls provide
different sets of values for walls with unconfined and confined boundary elements, and permits a reduction in
confined values for poorly confined walls. It is well accepted by structural engineers that confinement may increase
the lateral capacity of walls. Evaluation of the impact of confinement based on the parameter study was done to
establish the most appropriate method for accounting for the impact of confinement. The parameter study was
structured to allow for a direct method of considering the confinement. For each wall geometry and reinforcement
ratio for walls with boundary element, three levels of confinement were considered, with the all other wall
characteristic remaining unchanged. The levels of confinement were achieved by adjusting the crushing energy of
the concrete and thus increasing the strain at which crushing occurred. For the intermediate level of confinement
(referred to as confined), the crushing strain was adjusted but the confined concrete was not considered to have an
increase in compressive strength. For the highest level of confinement (referred to as well confined), the crushing
strain was adjusted using a different crushing energy and the confined concrete was considered to have a 40%
increase in compressive strength. Although all other characteristics of the walls stayed the same, the increased
compressive strength results in higher nominal moment strength and corresponding higher based shear demand
relative to unconfined concrete. Thus, evaluation of the impact of confinement on ‘a’ should account for the increase
in shear strength; Figure 8 shows the ratio of the change in ‘a’ due to the addition of confinement relative to the
change in the shear stress demand for planar walls with a 10% axial load ratio. For confined walls, there is
essentially no increase in the shear demand and an increase in ‘a’ of up to 10% from that of the unconfined walls.
For well confined walls, the value for ‘a’ increases as much as 80% of the confined value and there is a distinct
positive slope for the increase in ‘a’ relative to the increase in the shear demand. These observations from Figure 8
clearly indicate that the level of confinement is an appropriate parameter to consider in defining modeling
parameters for flexure-controlled walls. It is not clear, however, that the method used in the current ASCE 41
provisions is appropriate. Rather, it may be more appropriate to define values for walls with unconfined and well
confined boundary regions, with an increase in the unconfined values permitted for poorly confined boundary
regions. Additional analysis and comparison to experimental data is necessary to provide improved
recommendations.

5.3 Geometric characteristics of symmetric flanged walls

For planar walls, Figure 7a indicates that consideration for shear and axial demands is capable of capturing
distinct trends; other parameters may affect the modeling parameters, but the impact is expected to be relatively
minimal. For symmetric flanged walls, the trends are less distinct and it is necessary to further explore relationships
to wall characteristics to determine impact on the modeling parameter ‘a’.
A key characteristic of the symmetric flanged walls is the length of the flange (bf) relative to the length of
the wall (lw), referred to here as the flange aspect ratio (FAR). Three FAR values (10%, 20%, and 50%) were
considered. For larger FAR values, the moment capacity is increased, and since the load distribution is held
constant, there is an increased shear demand at the base of the wall; this impact of the FAR on the shear demand
should be accounted for in evaluating the impact of shear demand on the modeling parameter ‘a’ for the symmetric

4.12
@Seismicisolation
@Seismicisolation
Evaluation of ASCE 41 Modeling Parameters for
Slender Reinforced Concrete Structural Walls

flanged walls. Figure 9 shows, for symmetric flanged walls with an axial load demand of 10% of the gross capacity,
the relationship between ‘a’ and the shear demand for each of the three FAR values considered.
Figure 9a, b, and c individually indicate less variation in ‘a’ relative to shear than for the same axial load in
Figure 7b, which makes no distinction for the FAR. From Figure 9, it is evident that the FAR has an impact on the
value of ‘a’. For relatively short flanges (FAR = 0.10, Figure 9a), the relationship between ‘a’ and the shear demand
is very similar to that for planar walls. For wider flanges (FAR = 0.25 and FAR = 0.50), the median values for ‘a’
are larger. Figure 9b and 9c show more scatter in the value for ‘a’, but the lower-bound and median values are larger
for FAR = 0.50 than for FAR = 0.25. Collectively, this suggests the need to provide modeling parameter
recommendations for symmetric flanged walls that are dependent on the flange aspect ratio (FAR).
100

90

80

70
 a / aunconfined

60

50

40

30

20

10

0
-0.5 0 0.5 1 1.5 2 2.5 3 3.5
 v / vconfined

Figure 8 - Ratio of the increase in the modeling parameter ‘a’ relative to the change in the shear demand
when the level of confinement is increased from unconfined to confined (gray circles) and unconfined to well
confined (black circles). Only planar walls with axial load of 10% are shown.

(a) FAR = 0.10 (b) FAR = 0.25 (c) FAR = 0.50


Figure 9 - Impact of flange aspect ratio (FAR = bf/lw) on the modeling parameter ‘a’ for symmetric flanged
walls with axial load ratio of 10%. Shear stress demands shown are for concrete strengths in psi; for MPa,
multiply by 1/12.043.

6. PRELIMINARY RECOMMENDATIONS

Evaluation of the simulated values of the modeling parameter ‘a’ confirms that it is appropriate to use the axial load,
shear demand, and level of confinement to define modeling parameters and acceptance criteria for flexure-controlled
walls. This is consistent with the current provisions of ASCE 41; however, the current values do not predict
simulation and experimental values well, thus, it is necessary to provide recommendations for revised provisions.
Additionally, the presence of flanges and the length of the flanges relative to the wall length should be accounted
for. Using the simulated data, linear regression was used to provide preliminary recommendations for updated
modeling parameters. Two sets of recommendations were developed: i) for planar walls, with ‘a’ defined as a
function of axial load ratio and the shear stress demand, and ii) for flanged walls, with ‘a’ defined as a function of
 
4.13
@Seismicisolation
@Seismicisolation
A. C. Birely et al.

axial load ratio, shear stress demand, and flange aspect ratio (FAR). For each set of recommendations,
recommendations are provided for the three levels of confinement considered: unconfined, confined, and well
confined.
Evaluation of the simulated results indicates that it may be appropriate to a) specify upper and lower bound
values with interpolation between the limits or b) define a bilinear model that is steep at low shear and axial load
demands but nearly flat at high demands. However, it was desired to provide proposed revisions that are simple to
implement and multi-parameter linear regression was used to develop proposed recommendations. Thus, the
recommendations for the planar and flanged walls are defined by Equations (1) and (2), respectively:

a = α + αλλN + αvV/(Acv√f’c) (1)

a = α + αλλN + αvV/(Acv√f’c) + αFARbf/lw (2)

The regression coefficients for each geometry (planar and flanged) and confinement level (unconfined, confined,
and well confined) are provided in Table 1, along with the R2 values for the regression plane. Although it was
desired to have a consistent linear expression for all conditions, at high axial loads and high shear demands,
application of the linear regression equations to the simulation data resulted in negative values for ‘a’ at high shear
stress demands and high axial demands. To provide realistic conditions, an lower bound limit was specified at an
axial load ratio of 20% and shear stress demand of 4√f’c psi (0.33√f’c MPa). The value predicted at this point is set
as the lower bound for all walls with higher axial load or for the same axial load and greater shear stress demand.
This point was found by inspection and found to produce a reasonable model for both planar walls and symmetric
flanged walls. Revision may be appropriate upon revising recommendations to include asymmetric flanged walls
and greater diversity of confinement values. Figure 10 compares the proposed modifications of Equations (1) and
(2) to the current values for ‘a’ in ASCE 41.

Table 1 - Regression coefficients and R2 values for proposed regression equation for revised modeling
parameter ‘a’.
Shape Confinement Level α αλ αv1 αFAR R2

Planar Unconfined 0.011 -0.038 -0.0006 - 0.60


Confined 0.012 -0.041 -0.0006 - 0.65
Well Confined 0.014 -0.044 -0.0007 - 0.65
Flanged Unconfined 0.012 -0.055 -0.0006 0.020 0.55
Confined 0.015 -0.073 -0.0006 0.022 0.59
Well Confined 0.017 -0.080 -0.0005 0.027 0.51
1
For SI units (MPa), multiply by 1/12.043.

To evaluate the proposed regression expressions for the modeling parameter ‘a’, an error measure was
specified as the ratio of the ‘a’ parameter from the simulation results, asim, to the ‘a’ parameter predicted by the
regression expressions, apredict. Using this error measure, a value of 1.0 indicates exact prediction of the simulated
parameter; a value less than 1.0 indicates an over-prediction of the simulated parameter (unconservative); and a
value greater than 1.0 indicates that an under-prediction of the simulated parameter (conservative). Figure 11a and
11b summarize this error value for the planar and flanged walls, respectively, by showing the cumulative
distribution of the error as a solid black line. For comparison, the same error measure using the ‘a’ value predicted
by ASCE/SEI 41-06 are shown in as red lines.
For planar walls at low axial loads (2-5%), the proposed regression equation for the modeling parameter ‘a’
presents approximately the same cumulative error distribution as does the current ASCE 41 modeling. For these
walls, approximately 50% of the ‘a’ values are over-predicted and 50% are under-predicted. Given that the error is
largely unchanged for low axial load in planar walls, the proposed equation for the modeling parameter ‘a’ offers a
benefit to the current as it is continuous with the values for higher axial loads. For planar walls with axial loads from
10-20%, the median error for the proposed regression equation again is such that 50% of the ‘a’ values are over-
predicted and 50% are under-predicted. At these higher demands, this is a significant improvement over the current
ASCE 41 model parameter, for which the modeling parameter is generally much larger than the ‘a’ value from the
simulation results. At an axial load of 25%, the median for proposed regression equations is significantly less than
1.0; although this is unconservative, it does offer a slight improvement over the ASCE 41 prediction.

4.14
@Seismicisolation
@Seismicisolation
Evaluation of ASCE 41 Modeling Parameters for
Slender Reinforced Concrete Structural Walls

(a) Planar Walls

(b) Symmetric Flanged Walls


Figure 10 – Comparison of proposed values for ‘a’ (Equations (1) and (2) with the current ASCE 41 values.
Shear stress demands shown are for concrete strengths in psi; for MPa, multiply by 1/12.043.

For flanged walls, the proposed regression equations, which take into account the flange aspect ratio,
provide an overall improvement in the prediction of the ‘a’ value. This is evident in Figure 11b, which shows the
propsed equations shifting the median error measure closer to 1.0 relative to the median error measure for the ASCE
41 values, which overall provided unconservative values for ‘a’. As is the case for the planar walls, the prediction of
‘a’ at high axial demands is significantly unconservative.

 
4.15
@Seismicisolation
@Seismicisolation
A. C. Birely et al.

Figure 11 presents the evaluation of the recommended modeling parameter ‘a’ for the simulation results
used in establishing the recommendations. The recommended values were then used to model the experimental data
set described in the evaluation of the current ASCE 41 provisions. Figure 12 shows a comparison of the current and
recommended values for rectangular and flanged walls. For rectangular walls, the recommended values are slightly
more conservative than the current values. The largest impact is for flanged walls, for which the prediction of the
experimental values is significantly improved over that of the current ASCE 41 modeling parameters which
significantly underestimated the value for ‘a’.

(c) Planar Walls

(d) Symmetric Flanged Walls


Figure 11 - Error between simulated ‘a’ parameters and those i) predicted by the current ASCE/SEI 41-06
provisions, and ii) predicted by the proposed modifications.

4.16
@Seismicisolation
@Seismicisolation
Evaluation of ASCE 41 Modeling Parameters for
Slender Reinforced Concrete Structural Walls

(a) Rectangular Walls (b) Flanged Walls

Figure 12 – Comparison of existing ASCE 41 modeling parameter ‘a’ values and proposed values to values
from experimental tests. Shear stress demands shown are for concrete strengths in psi; for MPa, multiply by
1/12.043.

7. SUMMARY AND CONCLUSIONS

ASCE/SEI 4106 Seismic Rehabilitation of Existing Buildings is a guideline used by engineers to determine the
seismic adequacy of buildings based on the response of individual building components. The modeling parameters
for flexure-controlled reinforced concrete walls were evaluated using a database of experimental tests and were
found to be conservative, particularly for flanged walls; however, for the limited data points at higher axial loads
and for flanged walls, the values were found to be unconservative. In the study presented here, recommendations for
improved modeling parameters for flexure-controlled walls were developed. To provide a diverse range of data from
which to develop the recommendations, a parameter study was conducted using models validated using
experimental data. Both rectangular and symmetrically-flanged walls were considered, with a variety of geometry,
reinforcement ratios (0.25-1.5%), confinement levels, and axial load ratios (0.02-0.25Agf’c). Simulation results were
used to develop regression equations for the recommended modeling parameters.
Separate regression equations for the recommended modeling parameters were developed for rectangular
walls and for symmetric flanged walls. The regression equations and associated coefficients are provided in
Equations 1 and 2 and Table 1. For both geometries, three sets of equations are provided based on the level of
confinement in the boundary regions (unconfined, confined, and well confined). All equations are a function of the
axial load ratio and the shear stress demand in the walls. The confinement, axial, and shear dependence are
consistent with the current ASCE 41 modeling parameters, thereby providing a direct comparison and allowing for
familiarity in implementing the recommended modeling parameters. For symmetric flanged walls, the recommended
modeling parameters introduce the consideration of the flange aspect ratio (ratio of the flange length to the wall
length). Another deviation the recommended modeling parameters make from the current ASCE 41 modeling
parameters is the elimination of tabulated values in favor of linear regression equations, with the exception of lower
bound limits provided on walls with high axial load and high shear demands.
The proposed regression equations were developed from a preliminary parameter study for which further
analysis is intended. There remains a need for additional simulations and evaluations prior to providing finalized
recommendations. Needs for further investigation include:

1. Observations for the simulated data have focused on the modeling parameter ‘a’ with the assumption that
relating all other modeling parameters & acceptance criteria to this would be appropriate. Further
evaluation of the data is necessary to determine if this is true and to establish relationships between ‘a’ and
all other modeling parameters and acceptance criteria.
2. There is a need to expand evaluation of the parameter study to consider asymmetric flanged walls. For
asymmetric flanged walls, it is necessary to not only consider the ratio of the flange length to the web
length, but the length of the boundary region at the tip of the flange and the unbalanced reinforcement of
the two boundary regions (web and flange). The current modeling parameters of ASCE 41 are based on
 
4.17
@Seismicisolation
@Seismicisolation
A. C. Birely et al.

axial load ratios that include consideration for unbalanced reinforcement in the tension and compression
regions, which along with the recommended inclusion of the flange aspect ratio, would likely provide a
simple method for addressing the difference in response between rectangular, symmetric flanged, and
asymmetric flanged walls. For asymmetric flanged walls, the modeling parameters would necessarily be
different for bending with the flange in compression verses the flange in tension, therefore resulting in
modeling parameters that will would likely be bound by the recommendations for the rectangular and
symmetric flanged walls.
3. Expand the parameter study to further evaluate not just the impact of confinement, but also the degree of
confinement. The data presented here was for unconfined, confined, and well-confined boundary regions.
In the case of well-confined boundary regions, the modification to the confined concrete properties was an
upper bound for reasonable values. Additional refinement could provide useful data for establishing
recommendations for the impact of confinement on the modeling and acceptance criteria.
4. Expand the parameter study to include more walls with shear stress demands greater than 6√f’c psi (0.5√f’c
MPa). The data for such data points is sparse relative to the data for low shear demands.
5. For flanged walls, investigate the impact of varied reinforcement configurations in the flange (discrete
confined elements in the toes of the flanges as opposed to the uniform distribution of reinforcement and a
fully confined flange (for well-confined walls, this may lead to simulated data in excess of reasonable
characteristics for buildings))
6. Results presented here are based on the assumption that the failure mode is flexure-compression, where
failure is defined in the models by crushing of the concrete. Flexure-tension failure should also be
considered, where failure of the walls is defined by fracture of the reinforcing bars in tension. It is expected
that such consideration would result in a decrease in the simulated ‘a’ parameter for walls with low axial
demands and low shear stress demands. When using the recommended modeling parameters, engineers
should make consideration for whether or not strain demands are large enough to cause fracture in
reinforcing bars, taking into consideration of the ductility of the steel in the wall analyzed.
7. The experimental data used in validating the modeling technique used and evaluating the proposed
recommendations has upper bound axial load ratios of approximately 10% for flanged walls and 15% for
planar walls. Where walls have axial loads in excess of this, such as in coupled wall systems, the proposed
modeling parameters should be used with caution due to the lack of validating data. Further improvement
of the prediction at high axial loads may be practical when a larger experimental data set is available.

8. REFERENCES

ASCE-41, 2006. Seismic Rehabilitation of Existing Buildings (ASCE/SEI Standard 41-06), Structural Engineering
Institute, American Society of Civil Engineers, Reston, Virginia.

ASCE-41, 2007. Supplement No. 1, Seismic Rehabilitation of Existing Buildings (ASCE/SEI Standard 41-06),
Structural Engineering Institute, American Society of Civil Engineers, Reston, Virginia.

Birely, A.C., 2012, “Seismic Performance of Slender Reinforced Concrete Structural Walls”, PhD Dissertation,
University of Washington, Seattle, WA.

Birely, A.C., Lowes, L.N., Lehman, D.E., Aviram, A., Kelly, D.J., 2013. ASCE 31/41 Evaluation of Damaged
Chilean Walled Buildings. Proceedings of ASCE Structures Congress 2013, Pittsburg, PA.

FEMA-356, 2000, Prestandard and Commentary for the Seismic Rehabilitation of Buildings, American Society of
Civil Engineers (ASCE) for the Federal Emergency Management Association (FEMA), Washington, D.C.

Pugh, J.S., 2012 “Numerical Simulation of Walls and Capacity Design Recommendations for Walled Buildings”,
PhD Dissertation, University of Washington, Seattle, WA.

Lehman, D.E., Turgeon, J.A., Birely, A.C., Hart, C.R., Marley, K.P., Kuchma, D.A., Lowes, L.N., 2013. Seismic
Behavior of a Modern Concrete Coupled Wall. ASCE Journal of Structural Engineering, 139(8)

4.18
@Seismicisolation
@Seismicisolation
SP-297—5

Modeling Parameters for Reinforced Concrete Slab-Column Connections

Amy C. Hufnagel, YeongAe Heo, and Thomas H.-K. Kang

Synopsis: Over the past few decades, flat plate concrete building systems have been widely adopted in the United
States and other countries because it enables not only to save construction time and cost but also to make better use
of interior spaces. However, it has been observed that such buildings whose columns are cast into the concrete flat
plate are highly vulnerable to collapse. Such integrated slab-column frames have suffered severe damage or
completely collapsed during the past earthquake events. The main failure mode of these structures is punching shear.
Although a general guideline for seismic design and a seismic rehabilitation design guideline for both existing and
new reinforced concrete frame structures are specified in ACI 318 and ACI 369R, respectively, many uncertainties
have still resided in the modeling parameters to accurately predict seismic behavior of the flat plate concrete frame
structures. Therefore, a new chart of allowable plastic rotation values for correlating values of gravity shear ratio is
presented in this study with the objective of updating the modeling parameters of ACI 369R. The major issues which
lead to errors in evaluating flat plate concrete structural behavior under seismic loads are thoroughly investigated.
Also, dominant parameters from previous and recent experiments on integrated slab-column connections subjected
to seismic loading are utilized to assess the estimation of seismic performance of the flat plate system based on ACI
369R. Consequently, although it is confirmed that there is a trend in correlation between the allowable plastic
rotation and gravity shear ratio while almost no correlation is observed with reinforcement ratio, more experimental
data are necessary to enhance this correlation study. It is also noticed that the current ACI 369 recommendations for
allowable plastic rotation values for slab-column connection under seismic and gravity loading are unconservative.

Keywords: reinforced concrete; slab-column connections; modeling; plastic rotation; seismic; evaluation.

5.1
@Seismicisolation
@Seismicisolation
A. C. Hufnagel et al.

Amy C. Hufnagel is a Structural Engineer in the Walter P. Moore, Houston. She received her BS and MS from the
University of Oklahoma, Norman. Her research interests include the design and analysis of reinforced and
prestressed concrete structures.

YeongAe Heo, PhD, is a Post-doctoral Researcher in the Department of Architecture and Architectural Engineering,
Seoul National University, Korea. She received her BS and MS from Dong-A University, Korea, and her PhD from
the University of California at Davis. Her research interests include the seismic evaluation and nonlinear dynamic
modeling of reinforced concrete structures.

Thomas H.-K. Kang, PhD, PE, FACI, FPTI, is an Associate Professor in the Department of Architecture and
Architectural Engineering, Seoul National University, Korea. He received his BS from Seoul National University,
Korea, his MS from Michigan State University, and his PhD from the University of California at Los Angeles. His
research interests include the design and behavior of reinforced and prestressed concrete structures.

INTRODUCTION

Flat plate concrete building systems consist of monolithically cast slab and column frames, and are widely used
with and without shear walls in the United States and other countries for several advantages, including economy and
building efficiency. This setup reduces story heights, which allows for lower costs for building material, cladding,
electrical and mechanical ductwork, and annual heating and air-conditioning. This lowers both dead loads and lateral
loads on the structure. Despite these many advantages, there are also many issues such as complexity in modeling
and punching shear failure with flat plate building systems.
Because of the many benefits they provide, there is high motivation for the implementation of flat plate
building systems in every type of environment and loading situation. This includes areas of moderate to high
seismicity, despite the additional risks of increased lateral loads during earthquake events. Over the years, this has
resulted in a high percentage of flat plate buildings that have suffered large amounts of damage or even completely
collapsed during past seismic events, like the 1971 San Fernando earthquake, which initially called attention to the
weaknesses in flat plate building design. During the 1985 Mexico City earthquake, many non-ductile flat plate slabs
suffered severe damage caused by punching shear failure (Figure 1). During this event, 91 buildings completely
collapsed, while 44 also suffered severe damage (Robertson and Johnson, 2006). This triggered increased scrutiny
into the behavior and the code design methods for these structures in both the United States and Mexico. One source
of weakness in these buildings is the slab-column connections and the possibility of punching shear failure, which
occurs when the slab area around the column fails and collapses. This failure is caused by excessive shear stress
applied to the connection, as well as poor design for lateral loads, including non-ductile connections. It is important
that these connections withstand certain levels of lateral drift, while maintaining their gravity load carrying capacity
during an earthquake. There is a need to better understand the behavior of slab-column connections in flat plate
structures in order to improve design methods, as well as to suggest methods for investigation and retrofit of older
buildings in seismic regions.

Figure 1 — Progressive collapse of flat plate buildings in Virginia and Mexico City.

5.2
@Seismicisolation
@Seismicisolation
Modeling Parameters for Reinforced Concrete Slab-Column Connections

The purposes of this research are: 1) to investigate several issues with flat plate concrete structures in order to
better understand their behavior under gravity and lateral loads; 2) to compile past research and experiments
completed on slab-column connections in which specimens were subjected to seismic loading, and record relevant
information in a database; and 3) to examine the existing ACI 369 modeling parameters and provide a new chart of
modeling parameters of allowable plastic rotation values for correlating values of gravity shear ratio. The research
consists of a review of previous large-scale experiments, as well as a series of analyses on the data contained therein,
and is concluded by presenting experimentally suggested modeling parameters of allowable plastic rotation for slab-
column connections. This study is applicable to both pure flat plate systems and flat plate systems with walls and/or
moment frames, though slab-column connections in buildings with walls and/or frames may not be modeled. In such
cases, the acceptance criteria of plastic rotation become very important, which will be investigated in depth in the
next study of Slab-Column Connections Task Group of ACI Committee 369.

ISSUES IN FLAT PLATE DESIGN AND ACI 369R-11

Complexity in Modeling

Despite the many advantages for flat plate design and construction, the nonlinear behavior of these structures is
difficult to predict (Farhey et al., 1993). The complexity of the nonlinear behavior of slab-column connections is the
result of several contributing factors. The method of load transfer between the slab and column is complicated and
difficult to calculate, because of the combination of torsional and flexural moments and the three-dimensional stress
distribution as well as shear forces concentrated on the connection. Several analytical modeling methods have been
developed for slab-column frames (e.g., Kang et al., 2009; see Fig. C4.2 of ACI 369R-11). Modeling is especially
important in gaining a higher understanding of the complicated behavior of structural components, like slab-column
connections.

Punching Shear Failure

A major issue in flat plate concrete structures is the possibility of punching shear failure at the slab-column
connections, especially when subjected to cyclic lateral loading combined with gravity loading. This is a result of
several contributing factors, and is specifically problematic for those built in the 1950s and 1960s. These older
buildings were designed with non-ductile connections to resist gravity loads only, without shear walls or frames to
handle the applied lateral loads. The slab-column connection is responsible for transferring the unbalanced moment
and the shear forces. If shear stress due to these applied forces and moments surpass the shear capacity of the slab at
the face of the column, punching shear failure is a risk.
Another factor contributing to punching shear failure in the design of these older buildings is the lack of
continuous bottom reinforcement through the interior columns, which prevents the rebar from developing ultimate
strength at the location at the face of the column by acting as “a membrane to suspend the slab following failure of
the concrete” (Moehle et al., 1988). This lack of continuity in the connection makes the structure more vulnerable
and can lead to sudden punching shear failure or possibly a progressive collapse of the entire building. This collapse
may be the result of excessive lateral loading or the redistribution of gravity loads following lateral deformations.
When punching shear failure initially takes place, the slab drops and collapses onto the floor below. This, in turn,
overloads the floor below, eventually leading to a progressive collapse of the entire building. Punching shear failure
can happen for several reasons, including poor construction practices or premature removal of shoring, like the
collapse of the Skyline Plaza in Bailey’s Crossroads, Virginia (Figure 1); however, the main focus of this research is
the punching shear failure of flat plate systems subjected to lateral loading during seismic events. Even in flat plate
buildings that are designed with shear walls, the slab-column connections still need to have the capability of
maintaining gravity load carrying capacity under lateral load deformation. In order to prevent punching failure, the
connections must be capable of exhibiting ductility in the inelastic range, while still carrying the design loads (Pan
and Moehle, 1989). Non-ductile detailing in these connections does not provide enough anchorage between the
reinforcement and the concrete at the interior columns, which can cause a sudden brittle pull out and failure (Dovich
and Wight, 1996).
As such, several contributing factors aside from overly intensive ground motion may have led to the failure of
so many existing flat plate structures. Throughout the years, it is common for the function of certain buildings to
change. When these changes call for building modifications, this can sometimes cause an increase in gravity loading
that was not considered in the original design (Tian et al., 2008). Any building modifications that cause an increase

5.3
@Seismicisolation
@Seismicisolation
A. C. Hufnagel et al.

in loading must be accounted for by checking the capacity of the existing structure prior to use. Another source of
weakness is that the connections are assumed to take the majority of the shear resistance from the concrete, but
damage to the slab during a seismic event may reduce connection strength and shear capacity, thereby significantly
overestimating the shear strength of the connection.

ACI 369R-11

ACI 369R-11 is the American Concrete Institute’s Guide for Seismic Rehabilitation of Existing Concrete
Frame Buildings and Commentary. The guide was based on ASCE/SEI 41-06 and is intended to be used in
combination with the code in analyzing the behavior of existing and new concrete elements in buildings subjected to
seismic loading. The purpose of ACI 369R-11 is to provide recommendations for acceptance criteria and modeling
parameters that are necessary for the linear and nonlinear analysis of concrete components in the seismic
rehabilitation design process. This research focuses on Chapter 4, Section 4.4 of the guide, which discusses slab-
column moment frames and the analytical processes used in assessment. Section 4.1.3 of ACI 369R-11 defines these
frames as monolithically cast concrete structures with slabs, columns, and connections as the main structural
framing components, and with non-prestressed or prestressed reinforcement to serve as the primary reinforcement in
the slab. General considerations outlined in Section 4.4.1 of ACI 369R-11 state that stiffness, strength, and
deformation capacity of the building components should all be represented in analytical models for these frames.
This section also warns against the possibility of punching shear failure as a result of a combination of flexural,
shear, and torsional transfer at the connections.
Section 4.4.3 of ACI 369R-11 outlines the method for determining the flexural strength of a slab to resist
moment resulting from lateral deformations using Eq. (1).

M nCS  M gCS (1)


where MnCS is the nominal flexural strength of the column strip and MgCS is the column strip moment due to gravity
loads (ACI 369, 2011). This section goes on to describe the possibility of failure for slab-column connections under
shear and moment transfer. These connections should be investigated, because of the combined action of torsion,
flexure, and shear that all act in the slab at these locations.
The current code gives recommendations for modeling parameters a and b in Table 4.9 of ACI 369R-11. These
parameters are the allowable inelastic deformations for slab-column frames. In the table, these parameters vary
based on the continuity of bottom reinforcement through the interior columns, as well as on the gravity shear ratio.
The table contains data for reinforced concrete and post-tensioned concrete slab-column connections; however, this
portion of the research focuses solely on reinforced concrete components.
Larger allowable deformations are permissible in the event that there is experimental data to support the use of
alternative values. Otherwise, slab-column moment frame components that do not meet the acceptance criteria must
be rehabilitated as per the Section 3.7 of ACI 369R-11. These suggested seismic rehabilitation methods include
deficient component removal and replacement, or modification of the structure in order for the existing portions to
meet the criteria.

REVIEW OF PREVIOUS EXPERIMENTS

Reinforced Concrete Slab-Column Connections

A major portion of the research consisted of a review of past studies on interior reinforced concrete slab-
column connections subjected to combined lateral and gravity loading. This included individual interior connections,
as well as two-bay frames with interior connections. Other qualifications for relevant specimens include absence of
shear reinforcement in the slab in the form of either rebar reinforcement, like stirrups, shearbands or headed shear
studs, or a shear capital or drop panel.
In these past experiments, reinforced concrete slab-column connections composed of a square or rectangular
concrete slab with both an upper and lower column, most commonly measuring half a floor height each, were placed
in a tested rig and lateral loading was applied. In most cases, the connections were loaded cyclically and uniaxially,
in which a load was applied to one face at the top of the column up to a certain drift ratio, and then the process was
reversed in the opposite direction, simulating loads from a seismic event. This uni-directional cycle is typically
repeated at increasing levels of drift until failure of the connection. For specimens tested biaxially, force is applied

5.4
@Seismicisolation
@Seismicisolation
Modeling Parameters for Reinforced Concrete Slab-Column Connections

to the column in two principal directions. The order and configuration of the load applications are discussed with
each relevant specimen (Hufnagel, 2011).
The critical section perimeter of the interior connection (bo) was determined based on the information in the
text or figure of the literature. The critical section perimeter affects the shear capacity of the concrete (Vo) at the
connection, which is defined in ACI 369R-11 (and ACI 318-11) and typically determined using Eq. (2).

Vo  4 fc' (psi)bo d or 0.33 fc' (MPa)bo d (2)


where f’c is the compressive strength of the concrete and d is the effective depth. The shear capacity can then be
compared to the applied gravity shear (Vg) to determine the gravity shear ratio (Vg/Vo). The applied gravity shear was
taken from the literature, and should be the value at the peak lateral load so as to determine the most extreme
loading case. The gravity shear ratio is thought to significantly affect the overall strength and deformation capacity
of the connection.
The top and bottom reinforcement ratios of the slabs were determined at the location at the face of the column
using provided rebar layout diagrams for several different slab widths, including c2 + 3h, c2 + 5h, the column strip,
and the full width of the slab, where c2 is the column dimension in the direction perpendicular to the lateral loading,
and h is the thickness of the slab. The different widths were calculated by the provided dimensions, and the amount
of steel was counted for each width. The reinforcement ratios were then calculated using the thickness of the slab.
The moment capacities (Mn) were calculated for the same slab widths as the reinforcement ratios discussed
above. The total unbalanced moment capacity of an interior connection was calculated as a combination of the
positive and negative moment capacities. The total unbalanced moment capacity for each slab width for the frame
members with two exterior connections and one interior connection is doubled, because there are two additional
contributing connections. Peak loading values were determined either from the literature or from provided hysteretic
curves, and were used to calculate the unbalanced moment. Most of the specimens were tested cyclically in unaxial
direction, so the calculations include both a positive and negative unbalanced moment.
Provided hysteretic curves for each specimen were analyzed to collect information on drift capacities. First a
backbone curve was established to find the upper bounds of the testing cycles, making it easier to identify the
location of the peak load, as well as the drift percentage at that load. Next the unbalanced moment capacities were
plotted on the curve and drift percentages were recorded. Then the points at which the loads drop to 95%, 80%, and
20% of the peak loading were determined and plotted on the decreasing portion of the curve (see Figure 2). The
corresponding drift values for the points were recorded. These specific values are important in determining the
modeling parameter a, a value of plastic deformation from the yielding point, and modeling parameter b, a value of
plastic deformation to the point of residual strength. Although not all of the tests were conducted to the point where
the hysteretic curves provided information for all of these values, the lower bound values for b were obtained. Some
of the specimens showed unreasonable unbalanced moment capacities possibly because of an improper testing setup
or data reading, so this drift information could not be collected for those curves. For those specimens that were
tested biaxially, the hysteretic curves for each testing direction were analyzed, and the vector drift data was used.

Figure 2 — Selected lateral load vs. drift curves for recently tested slab-column connections (ND4LL – Robertson
and Johnson, 2006; CO – Kang and Wallace, 2008).

5.5
@Seismicisolation
@Seismicisolation
A. C. Hufnagel et al.

Punching Shear Capacity

Each specimen is investigated so as to determine whether or not punching shear failure could be predicted. The
connections are analyzed using the eccentric shear stress model for a slab width of c2 + 5h, the center of the column
strip, which should have the highest reinforcement ratio. The nominal moment of the slab over this width is
currently the representative recommendation from ACI 369R as to the yield moment of the slab. This width should
have the strength to transfer the unbalanced moment by flexure. If the applied shear stress exceeds the nominal shear
stress capacity of the concrete, a brittle failure may take place; otherwise, a flexural failure or flexure yielding
followed by drift-induced punching failure can be expected (Kang and Wallace, 2006). The eccentric shear stress
model was used in order to compare the nominal shear capacity of each connection to the applied shear stresses
[Eqs. (3) and (4)].

vc  vu (3)
Vg M c
vu   v u AB (4)
bo d Jc
where vc is the nominal shear stress capacity provided by the concrete as per ACI 318-11, vu is the total applied
shear stress, Vg is the applied gravity shear, v is the factor to determine the portion of unbalanced moment
transferred by shear, Mu is the unbalanced moment (M+n_c+5h + M–n_c+5h was used for this research), cAB is the
distance between the centroid of the column critical section to the edge of the critical section, and Jc is the property
of the assumed critical section analogous to the polar moment of inertia.

ACI 369 Modeling Parameters

From the information collected during the review of previous experiments, the data are analyzed to estimate the
approximate value of allowable plastic rotation for each specimen. This includes the determination of the parameters
a and b, which, as defined in ACI 369R, are used to measure deformation capacity in component load-deformation
curves. These parameters are defined in Figure 3. Current specifications for these modeling parameters can be found
in Table 4.9 in ACI 369R-11 (Table 1 of this paper).

Figure 3 — Generalized force-deformation relations for concrete elements or components (ACI 369R-11).

In this paper, the modeling parameters a and b are calculated using Eqs. (5) and (6), respectively; however, the
method to determine the yield point should be discussed further in the ACI Committee 369.

a   0.95 Peak   (5)


 M n _ c 5h
b   0.2 Peak   (6)
 M n _ c 5h

where 0.95Peak and 0.2Peak are the drifts at a 5% drop from the peak (green dots in Figure 2) and at a 80% drop from
the peak (purple dots in Figure 2), respectively, and  M is the drift at which the lateral load corresponds to
 n _ c 5h
the sum of M+n_c+5h and M–n_c+5h during the ascending branch of the backbone curve (blue dots in Figure 2). Here,
M+n_c+5h and M–n_c+5h are design flexural strengths of the c2 + 5h width for positive and negative bending,
respectively.

5.6
@Seismicisolation
@Seismicisolation
Modeling Parameters for Reinforced Concrete Slab-Column Connections

In this paper, these parameters are compared to the reinforcement ratios (ρc+5h) for the width of c2 + 5h, as well
as to the gravity shear ratios (Vg/Vo) in order to develop a chart for recommendations to ACI 369R-11.

Table 1 — Modeling parameters and numerical acceptance criteria for nonlinear procedures – Two-way slabs and
slab-column connections (ACI 369R-11)

*
Values between those listed in the table should be determined by linear interpolation.

Primary and secondary component demands should be within secondary component acceptance criteria where the full
backbone curve is explicitly modeled, including strength degradation and residual strength, in accordance with Section 3.4.3.2
of ASCE/SEI 41-06.

Where more than Condition i occur for a given component, use the minimum value from Table 4.9 of ACI 369R-11.
#
Action should be treated as force-controlled

RESULTS FROM REVIEW OF PREVIOUS EXPERIMENTS

Summary of Results from Assessment of Data

A summary of the data recorded in the database during the literature review can be found in Table 2 for the
specimens with and without continuous bottom reinforcement. The results for predicted punching shear failure can
be found in Table 3. This table includes the values for nominal shear stress capacity (vc) and total applied shear
stress (vu_c+5h) calculated as per Eq. (4) and using unbalanced moment values of (M+n_c+5h + M–n_c+5h) at an interior
connection and M+n_c+5h or M–n_c+5h for an exterior connection depending on the direction of slab bending. The width
of c2 + 5h was evaluated, because this width typically falls inside the column strip or close to the width of the
column strip, and therefore has the highest amount of slab reinforcement. The stress applied at the column is
transferred away from the joint as the area weakens, so the region closest to the column is the most crucial. These
results do not take into consideration the presence or absence of continuous bottom reinforcement through the
column. While many of these specimens seemed to exhibit enough strength to avoid stress-induced punching shear
failure, the applied cyclic loading at higher drift levels resulted in drift-induced punching failure anyway, despite the
presence of continuity in reinforcement (Kang and Wallace, 2006). While connections are not expected to withstand
drift levels beyond a certain level, the maximum allowable inelastic drift needs to be quantified.

5.7
@Seismicisolation
@Seismicisolation
A. C. Hufnagel et al.

Table 2 — Summary of database for specimens with and without continuous bottom reinforcement

d bo Failure Vg/Vo Munb 0.95Peak


Authors Specimen Loading (kip-ft) (%)
(in.) (in.) Pattern
+ - + - + -
AP1 Biaxially 0.37 39.1 36.1 1.6 1.8
Pan and
Cyclic;
Moehle (1992) AP3 4.1 59.4 Punching 59.7 56.8 3.6 3.4
Applied to 0.18
: Interior AP4 70.1 59.0 3.7 2.5
Column
Robertson and 2C Cyclic; Flexure 0.2 99.8 95.3 4 4.3
Durrani (1991) Applied to 3.6 54.3
: Frame 7L Column Punching 0.4 63.0 66.0 1.7 1.7

Farhey et al. 3 Cyclic; 50.8 0.2 13.3 13.3 3.7 2.6


One-sided
(1993) Applied to 2.8
punching
: Interior 4 Column 44.5 0.2 11.1 11.1 2.9 2.6
Robertson et Cyclic;
al. (2002) 1C Applied to 3.7 39.4 Punching 0.15 42.8 37.6 3.5 3.1
: Interior Column
Zee and Cyclic;
Moehle (1984) Interior Applied to 2.1 29.9 Punching 0.14 7.4 7.4 3.9 4.0
: Interior Column
Wey and Cyclic;
Durrani (1992) SC0 Applied to 3.8 55.1 Punching 0.05 0.06 45.7 45.0 3.6 4.0
: Interior Column
Kang and Cyclic;
Wallace (2008) CO Applied to 5.1 60.6 Punching 0.3 76.7 60.5 2.9 2.0
: Interior Column
Flexure/
ND1C 0.2 28.8 31.0 5.1 3.3
Punching
Flexure/
ND4LL 0.3 31.7 32.5 3.2 3.1
Robertson and Punching
Cyclic;
Johnson
ND5XL Applied to 3.9 55.7 Punching 0.4 22.9 23.6 2.0 2.0
(2006)
Column
: Interior†
ND6HR Punching 0.3 41.3 43.5 3.1 3.2
Flexure/
ND7LR 0.2 19.2 22.1 3.3 3.3
Punching
L0.5 Cyclic 95.9 99.6 1.6 1.8
Tian et al. Followed by
(2008) LG0.5 Monotonic; 5.0 83.9 Punching 0.2 90.7 89.2 1.2 1.2
: Interior† Applied to
LG1.0 Column 121 117 1.2 1.2
Stark et al. Cyclic;
(2005) C-63 Applied to 3.5 61.9 Punching 0.1 30.8 27.7 1.8 2.4
: Interior† Column
Durrani et al. Cyclic;
(1995) DNY_3 Applied to 3.9 55.1 Flexure 0.22 51.6 46.5 3.2 3.5
: Frame† Column
Luo et al. Cyclic;
(1994) II Applied to 3.9 55.1 Flexure 0.0 28.8 25.8 5.0 5.0
: Interior† Column

= without continuous bottom reinforcement; + = positive drift direction; – = negative drift direction
bo = perimeter of critical section; d= effective depth of section
Vg = applied direct shear; Vo = concrete punching shear capacity [see Eq. (2)];
Munb = measured peak unbalanced moment; 0.95Peak = drift at a 5% drop from the peak;
Conversion: 1 in. = 25.4 mm; 1 ft = 305 mm; 1 kip = 4.45 kN.

5.8
@Seismicisolation
@Seismicisolation
Modeling Parameters for Reinforced Concrete Slab-Column Connections

Based on the data and the results in Table 3, the nominal shear stress capacity provided by the concrete (vc)
appears to be quite conservative in terms of the stress-induced punching failure (that is, brittle punching shear failure
with limited slab bar yielding and ductility); therefore, the nominal stress capacity of 4√f’c (psi) or 0.33√f’c (MPa) is
recommended to be increased to some degree [e.g., 5√f’c (psi) or 0.42√f’c (MPa)] for more realistic modeling of
stress-induced punching and reasonable seismic evaluation of reinforced concrete interior slab-column connections
with typical slab dimension and reinforcement detailing.
Tables 3 and 4 show the calculations for modeling parameters a and b, respectively. The tables provide
information for specimens with and without continuous bottom reinforcement. The positive/negative sign represents
just the direction of lateral drift. Table 4 includes data for the modeling parameters with respect to gravity shear ratio
(Vg/Vo) only, like in the existing Table 4.9 in ACI 369R-11 (or Table 1 of this paper). Note that for interior slab-
column connection subassemblies with a pin condition at the column below, story drift due to column elastic
bending is typically negligible and no plastic deformation exists in the column (Pan and Moehle, 1992). Thus, the
plastic drift is assumed to be equal to the plastic rotation of the slab on either side of an interior slab-column
connection. The information shows the relationship between the measured modeling parameter and gravity shear
ratio. The values shown in Table 4 are mean values for the specimens with and without continuous bottom
reinforcement under variable gravity levels. However, because the data used in this study did not apply to each
gravity shear ratio, the data is lacking and should be expanded in future research studies. The relevant data including
the mean values are analyzed in this study to choose the best recommendation for each range.

Table 3 — Failure mode and modeling parameters (Gray shades: b is based on the maximum drift during testing)
C v/ f c a (%) b (%) Stress-Induced
Authors Specimen Vg/Vo Punching Expected? /
v=vc v=vu + - + - Occurred?
AP1 4 3.6 0.37 0.8 0.8 0.8 1.9 No / No
Pan and Moehle (1992) AP3 4 2.7 0.18 3 2.9 4.6 3.0 No / No
AP4 4 2.7 0.19 2.7 1.4 4.0 0.9 No / No
Robertson and Durrani 2C 4 5.6 0.20 3.4 3.4 4.4 4.1 Yes / No
(1991) 7L 4 6.7 0.40 0.5 0.7 1.8 2.0 Yes / Yes
3 4 3.0 0.20 3.0 2.0 4.4 2.0 No / No
Farhey et al. (1993)
4 4 4.9 0.20 2.2 1.8 4.9 1.8 Yes / No
Robertson et al. (2002) 1C 4 2.4 0.17 2.4 1.9 2.9 2.8 No / No
Zee and Moehle (1984) Interior 4 4.1 0.14 2.9 2.6 2.9 2.6 Yes / No
Wey and Durrani SC0 4
0.05 2.3 2.6 4.2 4.2 No / No
(1992) 2.5
Kang and Wallace CO 4
0.30 2.3 1.2 3.9 3.4 No / No
(2008) 3.1
ND1C 4 2.7 0.20 3.5 2.5 7.1 4.4 No / No
ND4LL 4 2.8 0.30 1.9 1.8 3.5 3.4 No / No
Robertson and Johnson
ND5XL 4 3.7 0.50 0.6 0.6 0.6 0.6 No / No
(2006)
ND6HR 4 4.2 0.30 1.1 1.4 3.2 3.3 Yes / No
ND7LR 4 2.5 0.30 2.0 2.2 3.7 3.9 No / No
L0.5 4 1.9 0.23 1.3 1.5 1.7 1.7 No / No
Tian et al. (2008) LG0.5 4 1.9 0.23 1.0 1.0 1.0 1.0 No / No
LG1.0 4 2.5 0.23 1.0 0.9 1.1 1.0 No / No
Stark et al. (2005) C-63 4 2.5 0.10 0.6 1.0 1.6 2.0 No / No
Durrani et al. (1995) DNY_3 4 4.0 0.22 2.6 2.4 3.7 3.2 Yes / No
Luo et al. (1994) II 4 1.6 0 3.8 3.9 3.8 3.9 No / No

5.9
@Seismicisolation
@Seismicisolation
A. C. Hufnagel et al.

Table 4 — Comparison of modeling parameters obtained from database to ACI 369R-11

Modeling Parameters
Conditions
Plastic Rotation, Radians
Continuity a b
x = Vg/Vo
Reinforcement ACI 369R-11 Mean Proposed ACI 369R-11 Mean Proposed
0 ≤ x < 0.2 0.035 0.025 0.025 0.050 0.032 0.035
0.2 ≤ x < 0.4 0.030 0.021 0.020 0.040 0.032 0.03
Yes
0.4 ≤ x < 0.6 0.020 0.006 0.015 0.030 0.019 0.025
x ≥ 0.6 0 - - 0 - -
0 ≤ x < 0.2 0.025 0.023 0.023 0.025 0.028 0.025
0.2 ≤ x < 0.4 0.020 0.018 0.018 0.020 0.029 0.020
No
0.4 ≤ x < 0.6 0.010 0.006 0.008 0.010 0.006 0.010
x ≥ 0.6 0 - - 0 - -

Discussion of Results from Assessment of Data

The review of previous experiments on reinforced concrete slab-column connections subjected to lateral
loading resulted in a database of test details including dimensions, detailing, and calculated moment strength. This
database will be useful as this research continues to expand to include more specimens that are relevant to the study.
It also enabled the comparison of several different factors in order to determine which factors may contribute to the
punching shear strength and allowable plastic drift capacity for certain connections.
Many comparisons were made between the different variables of interest in this study in order to determine
trends in the data (Figures 4 to 7). Both positive and negative story drifts are plotted in Figures 4 to 7. The
correlation between gravity shear ratio and allowable plastic drift levels was already known, so it was analyzed in
order to provide recommendations for ACI 369R, which are discussed below. Figure 4 shows the relationship
between the gravity shear ratio and the measured modeling parameter a. As the applied gravity shear ratio increases,
the modeling parameter a tends to decrease for both the specimens with and without continuous bottom
reinforcement. Similar comparisons were made for the modeling parameter b. A trend of a positive correlation
between gravity shear ratio and the measured modeling parameter b was noticed (Figure 5). The scatter was
particularly large for the parameter b. Note that the “maximum” drift value was used to determine b for more than
half of the specimens, because the testing was stopped without significant loss of strength in order to avoid the
complete collapse of the specimen.
The comparisons also include an analysis between the reinforcement ratios and the modeling parameters, which
showed little trend (Figures 6 and 7). Figures 6 and 7 show the modeling parameters with respect to reinforcement
ratio (ρc+5h), in order to determine the relationship between allowable plastic drift and the level of reinforcing steel at
the face of the interior column. The total reinforcement ratio, top and bottom included, for the slab width of c2 + 5h
was used, and the reinforcement ratio is shown as a percentage (%). These figures also have some missing
information for lower reinforcement levels, but this is mainly due to minimum code requirements for reinforcement
amounts. The critical section area was also compared to the modeling parameters with the thought that larger joint
proportions would allow increased levels of plastic deformation; however, no correlation was noted in the data. This
may be due to many different variables that are taken into account in testing a slab-column specimen, so the direct
comparison of two joints differing only in critical section area was not possible at this time.

5.10
@Seismicisolation
@Seismicisolation
Modeling Parameters for Reinforced Concrete Slab-Column Connections

(a) with continuous bottom reinforcement (b) without continuous bottom reinforcement
Figure 4. Gravity shear ratio vs. measured modeling parameter a.

(a) with continuous bottom reinforcement (b) without continuous bottom reinforcement
Figure 5. Gravity shear ratio vs. measured modeling parameter b.

(a) with continuous bottom reinforcement (b) without continuous bottom reinforcement
Figure 6. Reinforcement ratio vs. measured modeling parameter a.

5.11
@Seismicisolation
@Seismicisolation
A. C. Hufnagel et al.

(a) with continuous bottom reinforcement (b) without continuous bottom reinforcement
Figure 7. Reinforcement ratio vs. measured modeling parameter b.

(a) with continuous bottom reinforcement (b) without continuous bottom reinforcement
Figure 8. Gravity shear ratio vs. proposed modeling parameter a.

(a) with continuous bottom reinforcement (b) without continuous bottom reinforcement
Figure 9. Gravity shear ratio vs. proposed modeling parameter b.

5.12
@Seismicisolation
@Seismicisolation
Modeling Parameters for Reinforced Concrete Slab-Column Connections

The majority of the specimens also had continuous bottom reinforcement, which should be used to prevent
collapse and improve punching shear capacity (by providing dowel action and connection integrity and by
increasing flexural transfer capacity of unbalanced moment). The eccentric shear stress model does not take into
account the lack or presence of continuity in bottom reinforcement through the column. This is important, because
an insufficient development length may lead to brittle and unexpected failure by pullout of the rebar, which in turn
could lead to a progressive collapse of an entire structure. This emphasizes the importance of redundancy in
structural design. It is also important to know the varying strength of the slab with increasing flexural transfer
widths, in order to better understand the transfer mechanism and failure behavior of a structure. It is important that a
slab region around a column has the ability to transfer the load to the column, rather than punch suddenly and
catastrophically under load application. ACI 369R recommends the use of a slab width of c2 + 5h for flexural
transfer width, unlike ACI 318 where a width of c2 + 3h is specified. This study follows the current recommendation
of ACI 369R, because the database suggests that the c2 + 5h width is more reasonable.
Figures 8 and 9 compare the values preliminarily proposed from this research to the current values for the
modeling parameters a and b that are recommended in ACI 369R. The values in Figures 8 and 9 and Table 4 are
mainly based on the mean values for both specimens with and without continuity in reinforcement, with some
engineering judgment applied to address cases where limited data and significant scatter exist (see Figures 4 and 5).
The standard deviations of the errors between the experimental values and estimated modeling parameters are also
provided (Table 5). The proposed values follow the same basic trends as the current numbers, in that with increasing
gravity shear ratio, the modeling parameters decrease. However, due to unconservative estimates of the plastic
rotation capacity, smaller values are proposed for both the modeling parameters a and b. The median values are
quite close to the mean values (Table 6), so the use of the median would not affect the proposed values for
reinforced concrete slab-column connections. For specimens with a gravity shear ratio of 0.6 or greater, ACI 369R
typically recommends zero allowable plastic rotation angles for all categories. The research results did not include
any specimens that fell in this category, but with such a high amount of gravity loading, it appears that zero would
be an appropriate and conservative recommendation. In general, the updated numbers are smaller than those
currently recommended, meaning that the ACI 369 recommendations may have areas that are unconservative, and
therefore may not be safe for reinforced concrete slab-column connections.
In terms of modeling parameter c, residual strength ratio (see Table 1), at this point no updates are made to the
current values. Using the compiled database, acceptance criteria can also be assessed. ASCE/SEI 41-06 provides
procedures to define acceptance criteria for the Immediate Occupancy (IO), Life Safety (LS) and Collapse
Prevention (CP). For primary members, acceptance criteria for LS and CP are defined as 0.75(a) and the larger of
1.0(a) and 0.75(b), respectively. For secondary members, acceptance criteria for LS and CP are defined as 0.75(b)
and 1.0(b), respectively. However, to have the same (or similar) probability of exceeding a modeling parameter a or
b for various members/connections that meet acceptance criteria, ACI Committee 369 is proposing a new procedure
utilizing specific percentiles. While specific numbers are still being discussed, the same or similar procedure will be
applicable to the assessment of slab-column connections’ acceptance criteria. In Table 6, a few selected percentiles
are presented to provide a general idea.

Table 5 — Standard deviations of errors between experimental values and estimated modeling parameters (proposed
and ACI 369R-11) for a and b in %
Continuity
Vg/Vo STDa_proposed STDa_ACI369 STDb-proposed STDb_ACI369
Reinforcement
0 ≤ x < 0.2 0.47 1.13 1.07 2.07
0.2 ≤ x < 0.4 0.93 1.30 1.35 1.58
Yes
0.4 ≤ x < 0.6 1.17 1.40 0.61 1.10
x ≥ 0.6 - - - -
0 ≤ x < 0.2 1.53 1.54 1.09 1.09
0.2 ≤ x < 0.4 0.72 0.76 1.84 1.84
No
0.4 ≤ x < 0.6 0.20 0.40 0.40 0.40
x ≥ 0.6 - - - -

5.13
@Seismicisolation
@Seismicisolation
A. C. Hufnagel et al.

Table 6 — Data analysis of modeling parameters of a and b

Continuity a b Percentile of a Percentile of b


Vg/Vo
Reinf. mean median mean median 20th 35th 10th 25th
0 ≤ x < 0.2 0.025 0.026 0.032 0.030 0.022 0.024 0.024 0.028
0.2 ≤ x < 0.4 0.021 0.021 0.032 0.037 0.011 0.017 0.017 0.019
Yes
0.4 ≤ x < 0.6 0.006 0.006 0.019 0.019 0.005 0.006 0.018 0.019
x ≥ 0.6 - - - - - - - -
0 ≤ x < 0.2 0.023 0.024 0.028 0.029 0.008 0.011 0.015 0.018
0.2 ≤ x < 0.4 0.018 0.017 0.029 0.033 0.010 0.013 0.010 0.013
No
0.4 ≤ x < 0.6 0.006 0.006 0.006 0.006 0.006 0.006 0.006 0.006
x ≥ 0.6 - - - - - - - -

CONCLUSION

The primary purpose of the research was to review the experimental data of previously tested reinforced
concrete slab-column connections and to recommend potential updates to ACI 369R concerning modeling
parameters for allowable plastic rotations based on the investigation. The thorough review of previous and recent
experiments resulted in a database of useful information, which was used in further calculations and analysis of the
punching shear strength of reinforced concrete slab-column connections in flat plate systems. This study showed
that the nominal punching shear strength of connections subjected to combined seismic and gravity loading provides
a conservative estimate and that conversely the current modeling parameters for plastic deformation outlined in ACI
369R are unconservative. This would help provide more reasonable and safer guidelines for seismic evaluation and
rehabilitation, and may work to limit the damage caused to rehabilitated flat plate systems by strong seismic events.
This research was also successful in confirming that flat plate concrete structures without continuous
reinforcement are more susceptible to punching shear failure, and therefore have lower recommended allowable
plastic deformation values. The modeling parameters correlate with different values of gravity shear ratio and
reinforcement ratio; however, the trends suggest that the main variable of interest is the gravity shear ratio. Little
trend was noticed in the data between the reinforcement ratio and the modeling parameters. Given that the current
ACI 369 recommendations for plastic rotation values for slab-column connections were found to be unconservative,
the authors have made preliminary proposals on the values of a and b based on limited experimental data.
This study was especially important in pointing out the lack of information available for this area of research.
While many studies have taken place, the large amount of different variables among the specimens allows for less
comparable data between the specimens, especially for those without continuity in bottom reinforcement. Although
buildings are not designed this way anymore, it is important to determine if those that currently exist with non-
ductile connections have the strength to survive a seismic event or if retrofit methods are necessary. Further analysis
and specimen testing would be useful in expanding this study to provide more accurate recommendations to ACI
369R.

ACKNOWLEDGEMENTS

Financial support from Seoul National University is greatly appreciated. The authors acknowledge members of
ACI Committee 369 who provided constructive comments during the committee meetings, particularly John
Wallace, Ying Tian, Ken Elwood, Wassim Ghannoum, Roberto Stark, Murat Melek, and Mary Beth Hueste.

REFERENCES

ACI Committee 318 (2011), Building Code Requirements for Structural Concrete (ACI 318-11) and Commentary,
American Concrete Institute, Farmington Hills, MI.

5.14
@Seismicisolation
@Seismicisolation
Modeling Parameters for Reinforced Concrete Slab-Column Connections

ACI Committee 369 (2011), Guide for Seismic Rehabilitation of Existing Concrete Frame Buildings and
Commentary (ACI 369R-11), American Concrete Institute, Farmington Hills, MI.

American Society of Civil Engineers/Structural Engineering Institute (ASCE/SEI) Committee 41, Seismic
Rehabilitation of Existing Structures, in ASCE Standard 2007, Reston, VA.

American Society of Civil Engineers/Structural Engineering Institute (ASCE/SEI) Committee 41, Supplement 1 to
ASCE 41, in ASCE Standard 2007, Reston, VA.

Dovich, L. and Wight, J. K. (1996), “Lateral Response of Older Flat Slab Frames and the Economic Effect on
Retrofit,” Earthquake Spectra, Vol. 12, No. 4, pp. 667-691.

Durrani, A. J., Du, Y., and Luo, Y. H. (1995), “Seismic Resistance of Nonductile Slab-Column Connections in
Existing Flat-Slab Buildings,” ACI Structural Journal, Vol. 92, No. 4, pp. 479-487.

Farhey, D. N., Adin, M. A., and Yankelevsky, D. Z. (1993), “RC Flat Slab-Column Subassemblages Under Lateral
Loading,” ASCE Journal of Structural Engineering, Vol. 119, No. 6, pp. 1903-1916.

Hufnagel, A. C. (2011), “Analytical Studies of Reinforced and Post-Tensioned Concrete Flat-Plate Systems,” MS
thesis, School of Civil Engineering and Environmental Science, University of Oklahoma, Norman, OK.

Kang, T. H.-K. and Wallace, J. W. (2006), “Punching of Reinforced and Post-Tensioned Concrete Slab-Column
Connections,” ACI Structural Journal, V. 103, No. 4, pp. 531-540.

Kang, T. H.-K. and Wallace, J. W. (2008), “Seismic Performance of Reinforced Concrete Slab-Column Connections
with Thin Plate Stirrups,” ACI Structural Journal, Vol. 105, No. 5, pp. 617-625.

Kang, T. H.-K., Wallace, J. W., and Elwood, K. J. (2009), “Nonlinear Modeling of Flat-Plate Systems,” ASCE
Journal of Structural Engineering, V. 135, No. 2, pp. 147-158.

Luo, Y. H., Durrani, A. J., and Conte, J. P. (1994), “Equivalent Frame Analysis of Flat Plate Buildings for Seismic
Loading,” ASCE Journal of Structural Engineering, Vol. 120, No. 7, pp. 2137-2155.

Moehle, J. P., Kreger, M. E., and Leon, R. (1988), “Background to Recommendations for Design of Reinforced
Concrete Slab-Column Connections,” ACI Structural Journal, Vol. 85, No. 6, pp. 636-644.

Pan, A. and Moehle, J. P. (1989), “Lateral Displacement Ductility of Reinforced Concrete Flat Plates,” ACI
Structural Journal, Vol. 86, No. 3, pp. 250-258.

Pan, A. and Moehle, J. P. (1992), “An Experimental Study of Slab-Column Connections,” ACI Structural Journal,
Vol. 89, No. 6, pp. 626-638.

Robertson, I. N. and Durrani, A. J. (1991), “Gravity Load Effect on Seismic Behavior of Exterior Slab-Column
Connections,” ACI Structural Journal, Vol. 88, No. 3, pp. 255-267.

Robertson, I. and Johnson, G. (2006), “Cyclic Lateral Loading of Nonductile Slab-Column Connections,” ACI
Structural Journal, Vol. 103, No. 3, pp. 356-364.

Robertson, I. N., Kawai, T., Lee, J. and Enomoto, B. (2002), “Cyclic Testing of Slab-Column Connections with
Shear Reinforcement,” ACI Structural Journal, V. 99, No. 5, pp. 605-613.

Stark, A., Binici, B., and Bayrak, O. (2005), “Seismic Upgrade of Reinforced Concrete Slab-Column Connections
Using Carbon Fiber-Reinforced Polymers,” ACI Structural Journal, Vol. 102, No. 2, pp. 324-333.

5.15
@Seismicisolation
@Seismicisolation
A. C. Hufnagel et al.

Tian, Y., Jirsa, J. O., Bayrak, O., Widianto, and Argudo, J. F. (2008), “Behavior of Slab-Column Connections of
Existing Flat-Plate Structures,” ACI Structural Journal, Vol. 105, No. 5, pp. 561-569.

Wey, E. H. and Durrani, A. J. (1992), “Seismic Response of Interior Slab-Column Connections with Shear
Capitals,” ACI Structural Journal, Vol. 89, No. 6, pp. 682-691.

Zee, H. L. and Moehle, J. P. (1984), “Behavior of Interior and Exterior Flat Plate Connections Subjected to Inelastic
Load Reversals,” Report No. UCB/EERC-84/07, Earthquake Engineering Research Center, University of
California, Berkeley, CA.

5.16
@Seismicisolation
@Seismicisolation
SP-297—6

Non-linear Modeling Parameters for Jacketed Columns Used in Seismic


Rehabilitation of RC Buildings

José C. Alvarez and Sergio F. Breña


Department of Civil and Environmental Engineering
University of Massachusetts Amherst

Abstract
Rehabilitation of structures requires knowledge on the anticipated nonlinear
behavior of building components. Column retrofitting using jackets made of
different materials, designed to counter deficiencies in the original design, can be
used to improve the overall performance of an existing building. This paper
presents a procedure to obtain the backbone nonlinear response of jacketed
columns based on existing experimental results for this class of elements. Column
jacketing has been used in the past to increase ductility and strength of reinforced
concrete columns. Jacket designs can be varied to selectively improve the lateral-
load response of columns depending on the original design deficiency. A uniform
methodology or specific recommendations on how to estimate the response of
jacketed columns with different jacket materials types do not exist. Existing
experimental data are used to determine the backbone nonlinear response of
jacketed columns. These data are then used to develop recommendations to
determine non-linear force and deformation parameters of jacketed columns.
Although the most common types of jackets used are made from: concrete, steel,
and FRP, only the last two are included in this paper. The backbone curves thus
generated can then be used in performance assessment of reinforced concrete
buildings where columns have been jacketed.

Introduction
Performance-based design and seismic rehabilitation of structures requires a thorough
understanding of the expected response of a structure under various levels of seismic demand.
The elastic and inelastic cyclic behavior of structural components needs to be well understood in
order to be able to get better estimates of overall structure response under different ground
shaking scenarios. Current practice is to approximate the nonlinear cyclic response of structural
components by using force-deformation curves that bound the hysteretic response of these
components (backbone curves). To model the nonlinear response of components, average
backbone curves are simplified by determining only those points that impact the response of the
component importantly. These simplified backbone curves are constructed using values that

6.1
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña

represent the mean response of components tested in the laboratory. Modeling parameters in
ASCE/SEI Standard 41-06 (2006), for example, allow users to determine force and deformation
values at yield, at peak strength, and at residual strength. Multi-linear backbone curves are
constructed by joining these three key points. Backbone curves for positive and negative
directions of loading are assumed identical for components with symmetrical cross-section and
reinforcement.

Modeling parameters included in ASCE/SEI 41-06 and ACI 369R-11 were developed for
existing structural components. In a seismic rehabilitation project, the ability to estimate the
response of the retrofitted structure is important to assess the success of a proposed intervention.
Nonlinear modeling of rehabilitated components is therefore needed. A common feature of
existing frame structures that predate modern seismic procedures is the presence of columns with
details that make them vulnerable to collapse during earthquakes. Columns in these frames may
contain short lap splices of longitudinal reinforcement, inadequate shear strength, and/or
insufficient confinement to resist the anticipated seismic demands. Column jacketing is a
technique that has successfully been used in the past, as shown by experiments, to correct these
design and detailing deficiencies. There is little guidance, however, on procedures to determine
nonlinear modeling parameters for jacketed columns so the seismic response of retrofitted frames
cannot be determined consistently. This paper provides recommendations for construction of
backbone curves by determining modeling parameters of columns retrofitted using fiber-
reinforced polymer (FRP) jackets or steel jackets. Backbone curves of jacketed columns are first
constructed using existing experimental data to determine force-deformation parameters at yield
and peak strength. These data are then used for comparison with values determined from models
proposed to determine relevant force-deformation parameters of jacketed columns.

Description of Jacketed Column Database


A database of existing laboratory experiments of jacketed reinforced concrete columns
was compiled to study the force-deformation response characteristics of these elements. These
data were used to construct backbone curves from experimental data, as discussed in the
following section, which could be used to estimate yield and ultimate force-deformation values
for comparison with proposed models to generate these parameters for other jacketed columns.

All columns in the database were tested under cyclic static loading applied incrementally
following a prescribed protocol. Columns were tested in original (un-retrofitted) condition and
retrofitted using jackets made with fiber-reinforced polymer materials or steel. The concrete
compressive strength of columns in the database ranged between 1300 and 8000 psi (9 and 55
MPa). Columns were tested under either single curvature bending (cantilever) or double
curvature bending (fixed-fixed). Although some tests were intended to simulate bridge column
behavior, data are still included in the database for these specimens.

The database contains 146 columns of which 52 are circular columns and 94 are
rectangular columns. Figure 1 summarizes the classification of columns in the database

6.2
@Seismicisolation
@Seismicisolation
Non-linear Modeling Parameters for Jacketed Columns
Used in Seismic Rehabilitation of RC Buildings

according to jacket type and type of deficiency encountered in the original design. Figure 2
illustrates the distribution of key parameters of the columns in the database. The parameters
presented in this figure are spacing of transverse reinforcement (s) normalized by effective depth
to the tension force resultant, d, assumed equal to 0.8 dc; axial force ratio, defined as the axial
load divided by the nominal compressive strength of the concrete (f’c) and the gross cross-
sectional area (Ag); type of jacket used to retrofit each column (steel or FRP); the ratio between
shear at plastic hinging (Vo) and nominal shear strength (Vn). Nominal shear strength was
calculated using the shear strength equation 4-1 in ACI 369R-11 using a ductility of 8.

Figure 1 – Classification of columns in database

Figure 2 – Distribution of selected parameters of columns in database

6.3
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña

A short lap splice condition refers to splicing longitudinal reinforcement within the
plastic hinge region of the column over a length of 20 to 36 longitudinal bar diameters. Columns
with inadequate shear strength are those that fail in shear prior to developing the plastic hinge
capacity of the column at high displacement demands. Inadequate confinement of longitudinal
reinforcement is caused by a low volume of transverse reinforcement because of either small
diameter hoops and/or hoops at a large spacing. Poor seismic detailing of hoops with 90° hooks
at their ends also leads to inadequate confinement after cover concrete spalls at moderate
displacement demands.

The column database includes information about the material properties and geometry of
the jacket. The diameter of circular columns are between 9.5 and 30 in. (241 and 762 mm), while
the width of rectangular columns is between 6 and 36 in. (152 and 914 mm) and the depth is
between 7.5 and 36 in. (191 and 914 mm). Key jacket properties include jacket material (steel or
FRP), jacket thickness, length of jacket, and jacket mechanical properties. The properties of the
jacket vary widely depending on the retrofitting objective and the original column deficiency.

Appendix A includes tables that list force-deformation parameters of individual columns in the
database. Force-deformation values are provided at three levels: yield force, peak force, and
maximum force and deformation. Force is given as applied lateral load, V, at these different
levels. Peak values are those measured at the highest force applied to the specimens. Maximum
values are those measured at the maximum deformation imposed to the specimens. The
deformation parameter used in the tables and throughout this paper was drift, given as lateral
displacement divided by specimen height.

Construction of Backbone Curves from Experimental Data


Data available on jacketed columns were used to construct simplified backbone curves
that could be used for nonlinear analysis. Results of cyclic tests of jacketed columns are
typically presented by different researchers in plots of lateral-load (shear) as a function of
displacement (top displacement, drift, or displacement ductility demand). Cyclic force-
displacement is available in the literature either in the form of hysteresis curves or response
envelopes. Multi-linear backbone curves were extracted using reported hysteresis plots or, if
available, force-displacement envelopes for each specimen. These curves were further
simplified by identifying and plotting the coordinates of three key points (yield, peak strength,
and residual strength) to obtain simplified backbone curves consistent with modeling parameters
of ASCE/SEI 41-06.

Backbone curves were constructed by digitizing data from figures of hysteresis plots
included in the original test references. The data were digitized by obtaining the force and
displacement (or drift) at each displacement level applied to the columns. Force and

6.4
@Seismicisolation
@Seismicisolation
Non-linear Modeling Parameters for Jacketed Columns
Used in Seismic Rehabilitation of RC Buildings

displacement measured during the first cycle at each displacement level were used to construct
these curves. If authors directly reported backbone curves, then these plots were digitized to
obtain the force-deformation relationships instead of extracting them from hysteresis curves.
Generation of a backbone curve from the hysteresis response of one column in the database is
illustrated in Figure 3. The key points needed to construct simplified backbone curves were
obtained from these curves as illustrated in the figure; they are listed by column specimen in
Appendix A.

Figure 3 – Backbone curve generated from experimental data


Force and deformation at yield (Vy, Δy) were defined at the point of intersection between
two secant lines. The first line passes through the origin with the same slope as the experimental
force-deformation curve. The force at yield, Vy, was defined in two different ways depending on
the drift measured at peak strength. A value Vy = 0.8Vpeak was used for columns with peak
strength at a drift at or below 2% followed by rapid degradation. A value Vy = 0.7 Vpeak was used
for columns with peak strength at a drift exceeding 3%. This approach was needed to include
retrofitted columns where the jacket enhanced the lateral strength but did not affect the yield
capacity significantly.

The coordinates at peak force (VPeak, ΔPeak) were determined by using the average of the
peaks measured in the positive and negative directions of loading. The residual force (Vres) and
maximum displacement (max) are not frequently reported in the literature. Therefore Vres was
defined as a fraction (20%) of the peak force (Vpeak) and max was defined conservatively as the
maximum displacement reported for each tested column. The definitions of nonlinear modeling
parameters (a, b, and c) in ASCE/SEI 41-06 were used, wherever possible, to construct the
simplified backbone curves of jacketed columns. Parameter a is defined in ASCE/SEI 41-06 as
the difference between the deformation at lateral strength degradation of 20% and the
deformation at yield. In some tests the columns were not displaced to cause a 20% strength
degradation. In these cases parameter a was defined using the maximum deformation imposed
during the test. Parameter b is defined in ASCE/SEI 41-06 as the deformation corresponding to
loss of axial capacity, accompanied by a significant drop in lateral-load strength. Most of the
jacketed columns were not tested to this level of strength degradation, so parameter b was

6.5
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña

determined as the deformation corresponding to a degradation of 25% from Vpeak. In cases where
columns did not exhibit a strength degradation level of 20% or more, parameter b could not be
determined and is not reported in Appendix A.

To allow comparison among columns with different geometric characteristics,


reinforcement contents, and different material properties, the simplified backbone curves were
normalized by dividing all force values by the force at yield. Lateral deflection of the columns
was divided by column height (/H) to consistently plot drift for all the simplified backbone
curves from the database. The mean simplified backbone curves of circular and rectangular
columns jacketed using steel or FRP jackets in the database are shown in Figure 4.

Figure 4 – Mean force-deformation relationships of jacketed columns from database

Effect of Column Jacketing on Yield Force, Peak Strength, and Flexural Stiffness
The key parameters that define simplified backbone curves of jacketed columns need to
be calculated reliably for use in nonlinear analyses. Techniques used to determine these
parameters for jacketed columns must be developed to allow evaluation of the response of
retrofitted frames. Models that can be used to determine the yield force and peak strength of
jacketed columns are presented in this section. Because of lack of accurate deformation models,
drift at these key points may be based on results collected from the column database, so no
model was developed.

Depending on the jacket material used and the way it is connected to the existing
concrete column, jackets may also contribute to the flexural and shear strength of the column.
The two types of jackets discussed in this paper, steel and FRP, contribute in different degrees to
confinement, shear, and flexural strength. Steel jackets may contribute to the shear strength and
flexural strength of the columns since steel mechanical properties can be considered as isotropic.
FRP jackets, on the other hand, primarily contribute to strength and stiffness in the direction of

6.6
@Seismicisolation
@Seismicisolation
Non-linear Modeling Parameters for Jacketed Columns
Used in Seismic Rehabilitation of RC Buildings

fiber orientation. Fibers in FRP jackets used for confinement and shear strength increase are
applied transversely to the column axis so their contribution to flexural strength is negligible.
Only the effects of jacket confinement are considered here; contribution of jackets to shear
strength is not discussed in detail.

Steel jackets contribute to bending stiffness of the columns due to the increase in cross-
sectional dimensions (grout between original column and steel jacket) and the contribution of the
external steel jacket to bending stiffness. The contribution of FRP jackets to bending stiffness is
negligible so it was therefore not considered. Stiffness reductions to account for concrete
cracking (ACI 369R-11, Table 3.1) were used to modify the gross cross-sectional stiffness of
jacketed columns.

Confining Effects of Column Jacketing


The main effect of column jacketing is to provide confinement to the concrete in columns
with insufficient transverse reinforcement because of the presence of small diameter or widely
spaced hoops. Jacket confinement is activated after the concrete expands because of micro-
cracking. Increased concrete confinement will allow development of higher stresses and strains
in the confined concrete resulting in an increase in flexural strength. If inadequately short lap
splices exist within the plastic hinge region, the concrete splitting stresses that generate along the
splice at large displacement demands cause the concrete to expand. This expansion causes the
column jackets to develop lateral confining stresses (clamping stresses) that prevent failure at
low ductility because of splitting along the lap splice. Therefore, models intended to capture
jacketed column behavior must consider the confining effects that the jackets provide to the
column cross section.

Short lap splices or inadequate confinement of existing older columns may cause
localized failures primarily within the plastic hinge region at low ductility demands. The global
response of a column with these deficiencies is dominated by the behavior within the plastic
hinge region. In these cases it is appropriate to use a sectional model within the plastic hinge
zone to capture the global column behavior. The goal of localized column jacketing is to
eliminate a brittle splice failure from occurring within the plastic hinge region so that the global
response of the column becomes ductile. The jackets, however, may not necessarily provide the
level of clamping force necessary to develop yielding of longitudinal bars in the spliced region.
The stress developed in the spliced reinforcement was therefore reduced using equation 3-2 of
the ACI 369R-11, which considers the effect of bar slip in stress reduction. This modification
resulted in better agreement with the measured backbone behavior.

Column yield and peak force were estimated by including the effects of concrete
confinement using existing models with some modification to account for the properties of the
jackets. The cross–sectional behavior of a jacketed column was calculated by dividing the
section into fibers. The contribution of each fiber to flexural capacity at various deformations
(curvatures) was calculated using the uniaxial stress-strain behavior of the material composing

6.7
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña

the fiber. For a jacketed column fibers representing reinforcing steel, confined concrete, and
jacket were used. Unconfined concrete fibers were not included since jackets are applied
externally so all the concrete in the cross section is confined.

Similar to confinement provided by closely spaced hoops, jacket confinement efficiency


depends on the jacket and column cross-sectional properties. The lateral confining stress (fl)
generated by steel or FRP jackets for circular columns was estimated using equations proposed
by Priestley et al. (1994) and ACI 440.2R-08 (2008), respectively. For circular steel jackets the
lateral confining stress was estimated using:

(1)

and for circular FRP jackets, the lateral confining stress was calculated as:

(2)

where,

fyj = yield stress of the steel jacket


tj = steel jacket thickness
Dj = diameter of column to outside face of jacket
EFRP = modulus of FRP jacket
n = number of FRP plies (layers)
tf = thickness of FRP jacket
fe = effective strain in FRP jacket, assumed equal to 0.57fu
fu = maximum strain at failure of FRP jacket material
D = diameter of existing column

To allow use of the popular confinement model developed by Mander et al. (1988),
lateral confining stressed developed by steel or FRP jackets was converted to an equivalent
quantity of transverse reinforcement that would result in the same lateral confining stress. The
jacket confining stresses were equated to the lateral confining stress equation in the Mander et al.
(1988) model to determine the spacing required of No. 3 bar hoops that would result in the same
confining stress using:
∙ ∙ (3)

where ke equals

(4)

ke represents the confinement efficiency factor that accounts for the reduced confining
effect of discretely spaced transverse hoops at spacing s. The spacing s' refers to the clear
spacing between transverse hoops. The ratio between the area of longitudinal reinforcement (As)
to concrete area (Ac), denoted as cc, may be taken equal to zero in the case of jacketed columns.

6.8
@Seismicisolation
@Seismicisolation
Non-linear Modeling Parameters for Jacketed Columns
Used in Seismic Rehabilitation of RC Buildings

Equations (3) and (4) are solved for s for an assumed area of transverse reinforcement Atr, in this
case taken equal to 0.11 in2 for No. 3 transverse reinforcing bars.

(a) Circular columns (b) Rectangular columns


Figure 5 – Consideration of jacket confinement as equivalent hoop confinement
Confining efficiency of jackets on rectangular columns decreases significantly as the
ratio of long-to-short cross sectional dimensions increases. The confining stress generated by
rectangular steel or FRP jackets was estimated using an equivalent substitution depending on the
type of jacket. For rectangular steel jackets, the cross section of the column was assumed to
deform into an elliptical shape as the concrete expands due to micro-cracking. This approach
resulted in reasonable estimates of the global behavior of jacketed columns. The amount of
concrete expansion that would be required to deform a rectangular jacket into an ellipse is very
large, so this approximation was not intended to represent sectional behavior. Using this
approximation, the original rectangular column cross section of depth dc and width bc was
assumed to deform into an equivalent ellipse with major and minor axis dimensions determined
from:

2 (5)

2 (6)

The lateral confining stress of the equivalent elliptical jacket is calculated using the
equations proposed by Priestley et al. (1994), for confining stresses in the strong direction:

(7),

and for confining stresses in the weak direction:

(8)

6.9
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña

For rectangular FRP jackets, the column cross section was transformed into an equivalent
circular cross section instead of an ellipse given the higher flexibility of this material. The
diameter of the equivalent section was calculated equal to the diagonal dimension of the column
cross section, or:

(9)

Following the same philosophy as for circular columns, the confining stress generated by
the steel or FRP jackets was equated to the confining stress provided by an equivalent amount of
transverse hoops determined using the model by Mander et al. (1988). The difference in cross-
sectional dimensions, and the difference in longitudinal and transverse reinforcement
arrangement in the two directions of rectangular columns result in different confining stresses in
each direction. The effective confining stress provided by transverse reinforcement in the x and
y directions of a rectangular column is given by:

(10)

(11)

where x and y represent the transverse reinforcement content in the x (Atr-x/s·dc) and y
(Atr-y/s·bc) directions, respectively; fyh is the yield stress of transverse hoops; and ke is an
effectiveness factor to account for unconfined concrete zones in the vertical direction between
hoops and in the horizontal direction between longitudinal reinforcing bar positions. This
effectiveness factor may be estimated using:

∑ ′ ′
1 1 1 (12)

where bc and dc are the width and depth of the rectangular column. The term Σwi2 is used
to evaluate the reduction in confinement efficiency between longitudinal bars of the confined
core. Rectangular jackets were assumed restrained only at the column corners so the length of
each side of the column cross-section was used to estimate the ∑wi2 variable, resulting in a value
equal to 2 2 . Substituting this value into equation

′ ′
1 1 1 (13)

Using the area of No. 3 hoops for Atr-x and Atr-y the spacing required to give an equivalent
confining stress as generated by steel or FRP jackets was determined from equations (10), (11),
and (13).

Once the confining stress acting on the concrete was estimated, the confined concrete
material properties were determined using the Mander et al. (1988) model. Confined concrete
properties were used throughout the cross section to compute the moment-curvature response. It

6.10
@Seismicisolation
@Seismicisolation
Non-linear Modeling Parameters for Jacketed Columns
Used in Seismic Rehabilitation of RC Buildings

was assumed that the response of each column was governed by the sectional behavior within the
plastic hinge region. Using this assumption, the force-displacement response was calculated for
each column in the database where the plastic deformation concentrated in the column plastic
hinge region.

Comparison between Calculated and Measured Column Yield and Peak Capacities
The calculated forces at yield and at peak of the columns in the database were compared
with measured values. An assessment of the proposed force models was conducted to verify the
assumptions that were made and propose modifications if necessary. The proposed models only
incorporate the effects of confinement generated by the column jackets. Future efforts will
concentrate on determining the contribution of jackets to column shear strength.

A comparison between calculated and experimental yield force of the columns in the
database is shown in Figure 6. A similar comparison is given in Figure 7 for the peak force in
the columns. The data are divided into circular and rectangular columns because the confining
effects are markedly different between these two cross-sectional shapes. The data shown
indicate that the proposed models match the experimental data much better in the case of circular
columns than rectangular columns, but significant scatter is observed. Bending flexibility of the
jackets in rectangular columns, which is not captured adequately by the proposed models, may
be partially responsible for a better fit of model predictions to experimental values. The
proposed model overestimates the effect of confinement leading to higher calculated forces at
yield and peak. In some of the columns tested with partial jackets, failure moved to a section
above the jacket causing the differences observed between calculated and measured force values.

The columns in the database include details leading to three different types of
vulnerabilities and various jacket configurations. The lack of uniformity in the original column
designs and geometry and configuration of jackets used to correct the column vulnerabilities
introduces variability into the database that is difficult to capture with a single model. However,
a modification to the model by a reduction in the jacket confining efficiency seems warranted.

Summaries of the mean and standard deviation of the ratio of model predictions to
experimental values of force at yield and peak are given in Table 1 and Table 2 for circular and
rectangular columns, respectively. The values in these tables show that the discrepancies
between measured and calculated values vary significantly by jacket type and type of original
column vulnerability. Much better correlation in calculated to measured values is obtained if the
entire column database is considered regardless of vulnerability in the original design (Table 3).

6.11
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña

(a) Circular columns (b) Rectangular columns

Figure 6 – Comparison between calculated and test yield force

(a) Circular columns (b) Rectangular columns

Figure 7 – Comparison between calculated and test peak force

Table 1 – Comparison of circular column data by jacketing type

Columns with Steel Jackets  Columns with FRP Jackets 
Circular Columns 
Lap Splice  Shear  Confinement  Lap Splice  Shear  Confinement 
   Mean  STD  Mean  STD  Mean  STD  Mean  STD  Mean  STD  Mean  STD 
Vy calc/Vy test ‐  ‐  0.95  0.02  1.17  0.11  1.43  0.33  1.24  0.06  ‐  ‐ 
Vpeak calc/Vpeak test ‐  ‐  0.98  0.04  1.13  0.17  1.23  0.24  0.99  0.03  ‐  ‐ 

6.12
@Seismicisolation
@Seismicisolation
Non-linear Modeling Parameters for Jacketed Columns
Used in Seismic Rehabilitation of RC Buildings

Table 2 – Comparison of rectangular column data by jacketing type

Rectangular  Columns with Steel Jackets  Columns with FRP Jackets 


Columns  Lap Splice  Shear  Confinement  Lap Splice  Shear  Confinement 
   Mean  STD  Mean  STD  Mean  STD  Mean  STD  Mean  STD  Mean  STD 
Vy calc/Vy test 1.33  0.25  0.95  0.04  1.47  0.43  2.17  0.73  1.39  0.19  1.24  0.24
Vpeak/Vpeak test 1.09  0.15  0.98  0.04  1.30  0.17  1.23  0.12  1.09  0.16  1.11  0.24

Table 3 – Overall summary of statistical data for all columns in database

Columns with Steel Jackets  Columns with FRP Jackets 
Vy calc/Vy test Vpeak calc/Vpeak testt Vy calc/Vy test Vpeak calc/Vpeak test
Mean  1.09  1.08  1.38  1.16 
All Circular Specimens 
STD  0.14  0.15  0.30  0.23 
Mean  1.25  1.12  1.56  1.14 
All Rectangular Specimens 
STD  0.35  0.18  0.61  0.20 

Summary and Conclusions


Nonlinear modeling techniques of structural components in existing rehabilitation
standards and reports only include recommendations for existing elements. There is a lack of
information on approaches to follow for rehabilitated components. To help fill that void in
information a database was created to include jacketed columns because of the importance of
these elements in certain classes of rehabilitation projects.

The database includes steel and FRP jacketed columns of circular or rectangular cross-
section. A large variety of jacket configurations and design parameters were found in the
literature because of the different design vulnerabilities that original columns had. Existing
models intended to predict the load capacity of columns were modified to include the effects of
the various jacket designs that may be encountered in practice. In this paper, focus is given to
the effect of confinement on column strength and ductility. Shear strength enhancement was not
included in the current discussion. Following are the main observations that can be drawn from
the work conducted to date:

1. Normalized yield force is not affected significantly by jacket type or cross-sectional


shape. However, the deformation at yield may change depending on jacket type.
2. The ratio of peak strength to yield force of FRP jacketed columns is higher than for steel
jacketed columns. This result is due to the linear behavior of FRP materials to failure
compared with a loss in stiffness of steel after reaching the yield point.

6.13
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña

3. The ratio of peak strength to yield force of jacketed circular columns is on average higher
than for rectangular columns. This is caused by the higher confining efficiency of
circular jackets.
4. Drifts at yield and peak varied depending on jacket type and original column deficiency.
Columns with short lap splices within the plastic hinge region had higher deformations at
the same loads as those with continuous bars because of slippage of spliced bars.
5. The maximum deformation of columns varied significantly, but no firm conclusion can
be reached regarding this parameter. Most columns found in the literature were not
tested to displacements corresponding to significant loss of lateral force capacity. Many
tests were stopped prior to or when strength had degraded to 80% of peak because of
equipment limitations.
6. The proposed models used to predict strength of jacketed columns were able to estimate
the yield force and strength of columns in the database with different degrees of
accuracy. Models for jacketed circular columns are more accurate than those used for
jacketed rectangular columns. Strength was better predicted than yield with the models
proposed in this paper. More work is needed to develop physical models that produce
more accurate results.

The number of columns found in the literature for each design category and jacket type was
in many cases too small to allow making firm conclusions. However, the observations given
above may serve as a guide on the anticipated performance of jacketed columns and allow
development of consistent modeling parameters for different types of jacketed columns. It is
clear that more tests are necessary to develop comprehensive design recommendations on this
important class of rehabilitated component.

6.14
@Seismicisolation
@Seismicisolation
Non-linear Modeling Parameters for Jacketed Columns
Used in Seismic Rehabilitation of RC Buildings

References
Aboutaha, R. S., Engelhardt, M. D., Jirsa, J. O., and Kreger, M. E. (1999). "Rehabilitation of
Shear Critical Concrete Columns by Use of Rectangular Steel Jackets". ACI Structural
Journal. V. 96, No. 1, 68-78.

Aboutaha, R. S., Engelhardt, M. D., Jirsa, J. O., and Kreger, M. E. (November 1996). “Retrofit
of Concrete Columns with Inadequate Lap Splices by the Use of Rectangular Steel Jackets”.
Earthquake Spectra, V. 12, No. 4 , 693-714.

Aboutaha, R. S., Engelhardt, M. D., Jirsa, J. O., & Kreger, M. E. (1999). “Experimental
Investigation of Seismic Repair of Lap Splice Failures in Damaged Concrete Columns”. ACI
Structural Journal, V. 96, No. 2 , 297-307.

ACI Committee 369. (2011). “Guide for Seismic Rehabilitation of Existing Concrete Frame
Buildings”, ACI 369R-11, American Concrete Institute, Farmington Hills, MI.

ACI Committee 440. (2008). “Guide for the Design and Construction of Externally Bonded FRP
Systems for Strengthening Concrete Structures”, ACI 440.2R-08, American Concrete
Institute, Farmington Hills, MI.

Alcocer, S. M., and Durán-Hernández, R. Seismic Performance of a RC Building with Columns


Rehabilitated with Steel Angles and Straps. Innovations in Design with Emphasis on Seismic,
Wind, and Environmental Loading; Quality Control and Innovations in Materials/Hot
Weather Concreting. 531-552.

ASCE. (2006). “Seismic Rehabilitation of Existing Buildings”, ASCE/SEI 41-06, American


Society of Civil Engineers, Reston, VA.

Breña, S. F., and Schlick, B. M. (2007). "Hysteretic Behavior of Bridge Columns with FRP-
Jacketed Lap Splices Designed for Moderate Ductility Enhancement".Journal of Composites
for Construction. V. 11, No. 6, 565-574.

Chai, Y. H., Nigel Priestley, M. J. N., Seible, F., California., and University of California, San
Diego. (September – October 1991). “Seismic Retrofit of Circular Bridge Columns for
Enhanced Flexural Performance”. ACI Structural Journal, V. 88, No. 5, 572-584.

Esmaeily, A., and Xiao, Y. (2002). “Seismic Behavior of Bridge Columns Subjected to Various
Loading Patterns”. California: Pacific Earthquake Engineering Research Center (PEER)
2002/15.

Esmaeily, A., and Xiao, Y. (2005). “Behavior of Reinforced Concrete Columns Under Variable
Axial Loads: Analysis”. ACI Structural Journal ,V. 102, No. 5 , 736-744.

6.15
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña

Galal, K., Arafa, A., and Ghobarah, A. (2005). "Retrofit of RC Square Short Columns".
Engineering Structures. V. 27, no. 5, 801-813.

Harajli, M. H. (2005). "Behavior of Gravity Load-Designed Rectangular Concrete Columns


Confined with Fiber Reinforced Polymer Sheets". Journal of Composites for Construction. V.
9, no. 1, 4-14.

Harajli, M. H., and Rteil, A. A. (2004). "Effect of Confinement Using Fiber-Reinforced Polymer
or Fiber-Reinforced Concrete on Seismic Performance of Gravity Load-Designed
Columns". ACI Structural Journal. V. 101, No. 1, 47-56.

Harajli M.H., and Dagher F. (2008). "Seismic Strengthening of Bond-Critical Regions in


Rectangular Reinforced Concrete Columns Using Fiber-Reinforced Polymer Wraps". ACI
Structural Journal. V. 105, No. 1, 68-77.

Haroun, M. A., and Elsanadedy, H. M. (2005). "Fiber-Reinforced Plastic Jackets for Ductility
Enhancement of Reinforced Concrete Bridge Columns with Poor Lap-Splice
Detailing". Journal of Bridge Engineering. V. 10, No. 6, 749-757.

Haroun, M. A., and Elsanadedy, H. M. (2005). "Behavior of Cyclically Loaded Squat Reinforced
Concrete Bridge Columns Upgraded with Advanced Composite-Material Jackets". Journal of
Bridge Engineering. V. 10, No. 6, 741-748.

Iacobucci, R. D., Sheikh, S. A., and Bayrak, O. (2003). "Retrofit of Square Concrete Columns
with Carbon Fiber-Reinforced Polymer for Seismic Resistance". ACI Structural Journal. V.
100, No. 6, 785-794.

Li, Y.-F., Hwang, J.-S., Chen, S.-H., and Hsieh, Y.-M. (2005). “A Study of Reinforced Concrete
Bridge Columns Retrofitted by Steel Jackets”. Journal of the Chinese Institute of Engineers,
V. 28, No. 2, 319-328.

Mander, J. B., Nigel Priestley, M. J., and Park, R. et al. (1988). “Theoretical Stress-Strain Model
for Confined Concrete”. Journal of Structural Engineering, V. 114, No. 8 , 1804-1826.

Memon, M. S., and Sheikh, S. A. (2005). "Seismic Resistance of Square Concrete Columns
Retrofitted with Glass Fiber-Reinforced Polymer". ACI Structural Journal, V. 102, no. 5,
774-783.

Nigel Priestley, M. J., Seible, F., Xiao, Y., and Verma R. (July–August 1994). “Steel Jacket
Retrofitting of Reinforced Concrete Bridge Columns for Enhanced Shear Strength Part 1:
Theoretical Considerations and Test Design”. ACI Structural Journal , V. 91, No.4 , 394-405.

Nigel Priestley, M. J., Seible, F., Xiao, Y., and Verma, R. (September–October 1994). “Steel
Jacket Retrofitting of Reinforced Concrete Bridge Columns for Enhanced Shear Strength Part

6.16
@Seismicisolation
@Seismicisolation
Non-linear Modeling Parameters for Jacketed Columns
Used in Seismic Rehabilitation of RC Buildings

2: Test Results and Comparison with Theory”. ACI Structural Journal, V. 91, No. 5 , 537-
551.

Sezen, H. and Setzler, E. J. (May-June 2008). “Reinforcement Slip in Reinforced Concrete


Columns”. ACI Struuctural Journal, V. 105, No. 3, 280-289.

Seible, F., Priestley, M. J. N., Hegemier, G. A., and Innamorato, D. (1997). "Seismic Retrofit of
RC Columns with Continuous Carbon Fiber Jackets". Journal of Composites for
Construction. V. 1, No. 2, 52-62.

Wu, Y.-F., Wang, L., and Liu, T. (2008). "Experimental Investigation on Seismic Retrofitting of
Square RC Columns by Carbon FRP Sheet Confinement Combined with Transverse Short
Glass FRP Bars in Bored Holes". Journal of Composites for Construction. V. 12, No. 1, 53-
60.

Xiao, Y., and Wu., H. (2003). "Retrofit of Reinforced Concrete Columns Using Partially
Stiffened Steel Jackets". Journal of Structural Engineering. V. 129, No. 6, 725-732.

Xiao, Y., and Ma, R. (1997). "Seismic Retrofit of RC Circular Columns Using Prefabricated
Composite Jacketing". Journal of Structural Engineering. V. 123, No. 10, 1357-1364.

Youm, K.-S., Lee, H.-E., and Choi, S. (2006). "Seismic Performance of Repaired RC Columns".
Magazine of Concrete Research. V. 58, No. 5, 267-276.

6.17
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña

Appendix A: Summary of Jacketed Column Database Parameters


Specimen Author Vy (kip) Δy/H (%) Vpeak/Vy Δp/H (%) Vmax/Vy Δmax/H (%) a b
FC6ₒ Aboutaha 35.0 1.0 1.4 2.5 0.6 4.6 1.9 3.6
FC7ₒ Aboutaha 52.0 1.5 1.4 3.9 * * 2.4
FC10 Aboutaha 37.3 0.8 1.4 2.4 1.2 3.4 2.6
FC13ₒ Aboutaha 50.1 0.4 1.4 3.6 1.4 5.0 4.6
Lap Splice

FC9 Aboutaha 38.4 0.5 1.4 2.7 0.9 4.9 3.5 4.3
FC11 Aboutaha 45.4 0.8 1.4 2.4 0.5 5.5 2.6 4.7
FC12 Aboutaha 43.1 0.6 1.4 2.7 0.9 5.5 3.4 4.9
FC17 Aboutaha 45.6 0.2 1.4 2.4 1.3 5.4 5.2 5.2
Mean 43.4 0.7 1.4 2.8 1.0 4.9 3.3 4.5
Steel Jacketed Rectangular Columns

STD 5.7 0.4 0.0 0.6 0.3 0.7 1.1 0.5


SC6 Aboutaha 111.2 1.0 1.4 3.3 0.9 5.1 3.8 4.1
SC7 Aboutaha 101.3 0.5 1.4 4.2 1.2 6.3 5.8 5.8
SC10 Aboutaha 205.3 0.6 1.4 4.0 1.1 5.3 4.7 4.7
R2R Priestly 104.4 0.3 1.4 3.6 † † 3.3
Shear

R4R Priestly 154.8 0.3 1.4 3.8 † † 3.5


R6R Priestly 205.8 0.4 1.4 3.7 † † 3.3
SC8ₒ Priestly 109.8 0.7 1.4 3.9 1.1 7.0 6.0 6.3
Mean 141.8 0.6 1.4 3.8 1.1 5.9 4.3 5.2
STD 43.6 0.2 0.0 0.3 0.1 0.8 1.1 0.9
C-66-R Alcocer 42.0 1.0 1.4 2.5 0.8 3.0 2.0 2.0
C-66-S Alcocer 64.7 1.0 1.4 2.7 ** ** 1.7
Confinement

RC-2R Wu 45.8 0.4 1.4 2.1 0.6 6.0 3.0 5.6


RC-3R Wu 51.5 0.5 1.4 3.0 1.3 8.0 7.3 7.5
RC-4R Wu 51.0 0.4 1.4 3.1 1.1 8.0 7.6 7.6
RC-5R Wu 52.7 0.4 1.4 3.1 1.1 8.5 8.1 8.1
Mean 51.3 0.6 1.4 2.8 1.0 6.7 4.9 6.2
STD 7.0 0.3 0.0 0.4 0.2 2.0 2.8 2.3
All Mean 83.0 0.6 1.4 3.2 1.0 5.5 4.1 5.3
All STD 54.5 0.3 0.0 0.6 0.3 1.3 1.9 1.6

6.18
@Seismicisolation
@Seismicisolation
Non-linear Modeling Parameters for Jacketed Columns
Used in Seismic Rehabilitation of RC Buildings

Specimen Author Vy (kip) Δy/H (%) Vpeak/Vy Δp/H (%) Vmax/Vy Δmax/H (%) a b
C2R Priestly 115.5 0.3 1.4 4.4 † † 4.1
C4R Priestly 150.5 0.3 1.4 4.1 † † 3.8
Shear
C6R Priestly 161.0 0.4 1.4 5.5 † † 5.1
C8R Priestly 193.2 0.3 1.4 5.2 1.0 ‡ 4.9
Steel Jacketed Circular Columns

Mean 173.3 0.4 1.3 4.8 0.9 - 4.5


STD 35.3 0.1 0.0 0.6 0.0 - 0.6
1-R Chai 38.0 0.9 1.4 2.7 1.1 5.1 4.2 4.2
2 Chai 38.9 0.8 1.4 2.5 0.8 2.5 1.7 1.7
4 Chai 48.1 0.4 1.4 6.0 * * 5.6
Confinement

6 Chai 49.2 0.5 1.4 4.6 1.1 6.1 5.6 5.6


5 Chaiₒ 36.9 0.4 1.3 1.1 0.7 6.0 1.4 5.6
SC1 Hwang 54.9 0.4 1.4 3.8 0.9 5.8 4.8 5.4
SC2 Hwang 51.3 0.3 1.4 3.8 1.2 5.8 5.0 5.5
Mean 45.6 0.7 1.4 3.5 0.9 5.2 4.0 4.6
STD 7.3 0.2 0.2 1.5 0.2 1.2 1.6 1.4
All Mean 97.9 0.4 1.4 4.1 1.0 5.4 4.2 4.6
All STD 57.4 0.2 0.0 1.2 0.2 1.1 1.4 1.4

*Test was stopped at the peak load.


**Test was stopped at a load corresponding to the force capacity of the actuator.
†Tests stopped at the maximum displacement capacity of the actuator.
‡The strength of Specimen C8R degraded during cycling at the peak load. The degraded strength is taken as the value for Vmax.
ₒ These columns were not fully jacketed, and may have experienced failures outside the jacketed section.

6.19
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña
Specimen Author Vy (kip) Δy/H (%) Vpeak/Vy Δp/H (%) Vmax/Vy Δmax/H (%) a b
RF-R1 Haroun 59.7 1.3 1.3 2.4 0.9 2.8 1.2 1.5
RF-R2 Haroun 58.4 1.2 1.3 4.5 0.9 1.9 0.3 0.7
RF-R3 Haroun Lap Splice
54.4 0.9 1.4 3.6 1.2 3.6 1.8 2.7
RF-R4 Haroun 59.1 1.2 1.3 3.5 1.0 2.7 1.3 1.5
Mean 57.9 1.2 1.3 3.5 1.0 2.7 1.1 1.6
STD 2.1 0.1 0.1 0.7 0.1 0.6 0.6 0.7
RS-R1 Haroun 107.1 0.2 1.3 1.7 0.9 4.7 3.4 4.5
RS-R2 Haroun 104.7 0.3 1.3 1.1 0.9 4.7 2.7 4.4
RS-R3 Haroun 105.2 0.2 1.3 1.5 1.1 4.4 4.1
RS-R4 Haroun 95.9 0.1 1.4 2.1 1.2 4.6 4.5
FRP Jacketed Rectangular Columns

RS-R5 Haroun 97.3 0.2 1.4 2.5 1.3 4.2 4.0


RS-R6 Haroun 107.1 0.2 1.3 1.8 1.1 4.9 4.7
Shear R Seible 98.0 0.1 1.2 1.5 1.1 2.3 2.1
Shear

SC2 Galal 76.7 0.1 1.3 1.8 1.1 5.7 5.6


SC1R Galal 72.9 0.3 1.2 1.8 0.8 5.8 2.8
SC2R Galal 76.4 0.3 1.3 1.8 0.6 4.2 2.4
SC1U Galal 78.9 0.1 1.3 0.9 1.3 5.6 5.5 5.5
SC3 Galal 78.5 0.2 1.3 1.8 0.8 5.7 3.2 5.5
SC3R Galal 54.6 0.3 1.3 0.6 0.5 3.6 1.5 3.3
Mean 89.1 0.2 1.3 1.6 1.0 4.6 3.6 4.6
STD 16.2 0.1 0.2 0.5 0.2 1.0 1.2 0.8
S2 Ozcan 8.6 0.2 1.2 3.0 0.4 7.2 3.4 7.0
S3 Ozcan 9.4 0.3 1.3 2.0 0.5 4.8 3.5 4.6
S4 Ozcan 7.7 0.2 1.4 2.4 0.7 4.6 3.4 4.3
Confinement

S5 Ozcan 14.1 0.3 1.2 1.0 0.6 6.0 3.9 5.6


C1FP1 Harajli 10.9 0.9 1.4 3.0 0.7 6.0 3.3 5.2
C1FP2 Harajli 5.4 0.6 1.4 3.0 2.1 5.1 4.5
C1F1 Harajli 12.5 0.9 1.3 2.0 0.7 5.0 3.5 4.1
C1F2 Harajli 11.2 1.0 1.4 3.0 1.0 5.1 2.8 4.1
C2FP1 Harajli 15.2 0.8 1.4 3.0 0.6 5.1 2.7 4.3

6.20
@Seismicisolation
@Seismicisolation
Non-linear Modeling Parameters for Jacketed Columns
Used in Seismic Rehabilitation of RC Buildings
Δp/H
Specimen Author Vy (kip) Δy/H (%) Vpeak/Vy (%) Vmax/Vy Δmax/H (%) a b
C2FP2 Harajli 15.1 0.7 1.4 3.0 0.8 5.1 4.0 4.3
C2F1 Harajli 15.6 0.8 1.4 3.1 0.9 5.0 3.7 4.3
C2F2 Harajli 15.3 0.8 1.4 3.0 0.9 5.1 3.4 4.3
FRP Jacketed Rectangular Columns

Confinement R Seible 107.3 0.4 1.4 2.7 1.1 3.3 6.0


C3 Wu 10.8 1.0 1.3 2.1 1.2 7.2 3.3 6.2
C4 Wu 11.8 1.1 1.3 2.4 1.2 8.0 2.4 6.9
Confinement

C5 Wu 11.1 1.0 1.3 4.0 1.3 7.2 6.2


C6 Wu 11.0 1.0 1.3 2.0 1.0 7.3 6.3
ASG-2NSS Memon 24.6 0.7 1.4 4.6 0.3 23.7 11.6 23.0
ASG-3NSS Memon 24.9 0.2 1.4 6.0 1.2 12.0 9.7 11.8
ASG-4NSS Memon 24.2 0.3 1.4 2.4 0.9 11.4 11.2
ASG-5NSS Memon 24.4 0.6 1.4 4.3 0.8 11.2 8.9 10.6
ASG-6NSS Memon 29.8 0.9 1.4 11.8 1.1 20.7 11.4 19.8
ASGR-7NSS Memon 24.2 0.5 1.4 5.8 0.7 14.9 7.5 14.4
ASGR-8NSS Memon 25.8 1.0 1.4 4.7 1.2 12.2 11.3
Mean 19.6 0.7 1.4 3.5 0.9 8.5 5.7 8.0
STD 19.5 0.3 0.4 2.1 0.4 5.1 3.1 5.6
All Mean 44.2 0.5 1.4 2.9 0.9 6.9 4.6 6.5
All STD 36.9 0.3 0.3 1.9 0.3 4.5 2.9 5.2
CFRP-05 Brena 9.4 1.9 1.4 6.0 1.0 9.2 6.8 7.3
FRP Jacketed Circular

KFRP-05 Brena 9.7 1.4 1.4 6.3 0.9 9.5 7.5 8.1
CFRP-15 Brena 9.1 1.3 1.8 6.0 1.5 9.2 7.9
Lap Splice
Columns

KFRP-15 Brena 9.5 1.1 1.4 6.2 1.1 10.0 8.9


CF-R1 Haroun 25.4 0.4 1.4 5.3 1.1 6.4 5.2 6.0
CF-R2 Haroun 28.3 0.4 1.4 4.4 1.1 5.9 4.9 5.5
CF-R3 Haroun 31.3 0.5 1.4 3.8 1.1 5.6 4.3 5.2
CF-R4 Haroun 30.8 0.4 1.4 3.9 1.1 5.9 5.5
CF-R5 Haroun 31.2 0.5 1.3 3.9 1.1 5.9 4.5 5.4

6.21
@Seismicisolation
@Seismicisolation
J. C. Alvarez and S. F. Breña

Specimen Author Vy (kip) Δy/H (%) Vpeak/Vy Δp/H (%) Vmax/Vy Δmax/H (%) a b
CF-R6 Haroun 30.3 0.4 1.4 4.8 1.1 7.0 6.2 6.6
Lap Splice Lap Splice Rₒ Seible 42.7 0.1 1.4 2.4 1.1 3.8 7.1
FRP Jacketed Circular Columns

C2-RT4 Xiao 53.3 0.4 1.3 1.4 0.8 5.1 4.7


C3-RT5 Xiao 52.8 0.4 1.4 2.5 1.0 5.0 4.5
Mean 28.0 0.7 1.4 4.4 1.1 6.8 6.0 6.3
STD 14.9 0.5 0.1 1.5 0.2 1.9 1.4 1.0
CS-R1 Haroun 80.3 0.2 1.4 3.1 1.3 3.6 3.4
CS-R2 Haroun 83.2 0.2 1.4 3.0 1.4 3.7 3.6
CS-R3 Haroun 116.2 0.5 1.4 3.8 1.4 4.3 3.8
Shear

CS-R4 Haroun 112.2 0.5 1.4 2.1 1.3 3.1 2.5


CS-P1 Haroun 119.9 0.3 1.4 3.9 1.3 4.1 3.7
Mean 102.4 0.4 1.4 3.2 1.3 3.7 3.4
STD 17.0 0.2 0.0 0.6 0.0 0.4 0.4
All Mean 48.6 0.6 1.4 4.0 1.2 6.0 5.3 6.3
All STD 36.8 0.5 0.1 1.4 0.2 2.1 1.7 1.0
ₒ These columns were not fully jacketed, and may have experienced failures outside the jacketed section.

6.22
@Seismicisolation
@Seismicisolation
SP-297—7
 

Analysis of Seismic Response of Masonry‐Infilled RC Frames through Collapse 
 
P. Benson Shing and Andreas Stavridis 
 
 
 
Synopsis: 
The assessment of the seismic vulnerability and collapse potential of masonry‐infilled RC frame buildings 
presents a significant challenge because of the complicated failure mechanisms they could exhibit and 
the  number  of  factors  that  could  affect  their  behavior.  In  general,  there  are  two  types  of  analysis 
methods that can be used to simulate the inelastic behavior of infilled frames. One is to use simplified 
frame models in which infill walls are represented by equivalent diagonal struts, and the other is to use 
refined  finite  element  models  that  can  capture  the  failure  behavior  of  RC  frames  and  infill  walls  in  a 
detailed  manner.  However,  both  types  of  models  have  limitations  in  simulating  structural  response 
through  collapse.  While  refined  finite  element  models  are  not  computationally  efficient,  simplified 
models  are  less  accurate  because  of  their  inability  to  represent  some  failure  mechanisms  that  could 
occur in an infilled frame. In this paper, possible failure mechanisms and causes of collapse of masonry‐
infilled RC structures are discussed, and both simplified and refined finite element analysis methods that 
can  be  used  to  simulate  the  inelastic  response  of  these  structures  and  assess  their  vulnerability  to 
collapse are presented with numerical examples. Additional research and development work needed to 
improve collapse simulations is discussed. 
 
 
Keywords:  masonry;  infill  walls;  reinforced  concrete  frames;  earthquakes;  seismic;  collapse;  finite 
element method; equivalent strut models. 
   

7.1
@Seismicisolation
@Seismicisolation
P. B. Shing and A. Stavridis

P.  Benson  Shing  is  a  Professor  of  Structural  Engineering  at  the  University  of  California,  San  Diego.  He 
received his BS, MS, and PhD from the University of California, Berkeley. He is a member of ACI and his 
main research interests are in seismic design and analysis, and in the inelastic behavior and performance 
of concrete and masonry structures, including large‐scale testing and analytical modeling.  
 
Andreas Stavridis is an Assistant Professor of Civil Engineering at the University at Buffalo. He received 
his  BS  degree  from  the  National  Technical  University  of  Athens,  Greece,  and  MS  and  PhD  from  the 
University of California, San Diego. His main research interests include the experimental and analytical 
evaluation of the seismic performance of structures, and the condition assessment and rehabilitation of 
existing concrete and masonry buildings. 
 
 
 
INTRODUCTION 
Masonry‐infilled RC frames can be frequently found in seismically active regions around the world. They 
represent  a  common  form  of  construction  in  many  countries.  Many  older  buildings  in  the  western 
coastal regions of the United States have unreinforced masonry infill walls, and infill walls are frequently 
used  in  newer  buildings  in  the  mid‐western  and  eastern  parts  of  the  country.  Masonry  infill  walls  are 
normally  neither  reinforced  nor  tied  to  the  RC  frames.  In  spite  of  this,  infill  walls  often  enhance  the 
seismic  resistance  of  non‐ductile  RC  frames,  which  would  otherwise  not  be  able  to  withstand  a  major 
earthquake.  However,  the  partial  or  total  collapse  of  an  infilled  RC  frame  could  be  sudden  as 
demonstrated in past earthquakes. The assessment of their seismic vulnerability and collapse potential 
presents a significant challenge because of the complicated failure mechanisms they can exhibit and the 
number  of  factors  that  can  affect  their  behavior.  In  general,  there  are  two  types  of  analysis  methods 
that  can  be  used  to  simulate  the  inelastic  behavior  of  infilled  frames.  One  is  to  use  simplified  frame 
models  in  which  infill  walls  are  represented  by  equivalent  compression‐only  diagonal  struts,  and  the 
other is to use refined finite element models that can capture the failure behavior of RC frames and infill 
walls  in  a  detailed  manner.  However,  both  types  of  models  have  limitations  in  simulating  structural 
response  through  collapse.  While  refined  finite  element  models  are  not  computationally  efficient, 
simplified models are less accurate because of their inability to represent some main failure mechanisms 
that could occur in an infilled frame.  In this paper, possible failure mechanisms and causes of collapse of 
masonry‐infilled RC structures are first discussed, and both simplified and refined finite element analysis 
methods  that  can  be  used  to  simulate  the  inelastic  response  of  these  structures  and  assess  their 
vulnerability to collapse are presented with numerical examples. Additional research and development 
work needed to improve collapse simulations is discussed. 
  
FAILURE MODES AND CAUSES OF COLLAPSE 
Failure modes, load resistance, and ductility 
When built with high quality masonry materials and workmanship, infill walls can significantly enhance 
the earthquake resistance of RC frames. However, under severe seismic loads, their interaction with the 
bounding  frames  can  result  in  undesired  failure  mechanisms,  which  may  lead  to  the  collapse  of  the 
structure.  Hence,  to  evaluate  the  ability  of  this  type  of  structures  to  survive  an  earthquake  without 
collapsing,  the  frame‐wall  interaction  and  the  resulting  failure  mechanism  have  to  be  accurately 
simulated  in  the  analytical  model.  Under  earthquake  loading,  the  interaction  of  an  RC  frame  with 
masonry infill walls can result in one of several possible failure mechanisms depending on the strength 
and stiffness of the walls as compared to those of the frame. Figure 1 shows three failure mechanisms 
that have been frequently observed in laboratories and in the field (Mehrabi et al. 1994, 1996).  

7.2
@Seismicisolation
@Seismicisolation
Analysis of Seismic Response of Masonry-Infilled RC Frames through Collapse

Figure 1 – Failure mechanisms of infilled frames  Figure 2 – Failure Mechanism (b) exhibited in a 
(Mehrabi et al. 1996).  quasi‐static cyclic loading test (Blackard et al. 
2009). 
 
Failure Mechanism (a) shown in Figure 1 may occur when infill walls develop profuse horizontal sliding 
shear cracks because of weak mortar joints. This mechanism can result in flexural hinges at the top and 
bottom of the columns. However, because of the relatively small horizontal forces exerted by the wall 
against the columns in this case, the columns can avoid severe damage as long as the story drift is not 
excessive.  Mechanism  (b)  can  occur  in  non‐ductile  RC  frames  infilled  with  relatively  strong  masonry, 
which tends to develop dominant horizontal and/or diagonal cracks as shown by a laboratory specimen 
in  Figure  2.  In  this  case,  shear  failures  can  occur  in  the  columns  as  a  result  of  a  large  horizontal  force 
exerted by the cracked wall against the top or bottom of a column. Sometimes, shear failure can also 
occur at the mid‐height of a column when the masonry bed‐joint at that level develops a through crack 
leading  to  a  short‐column  effect.  This  can  be  the  case  when  a  weak  horizontal  sliding  plane  has  been 
introduced by a window opening. Mechanism (c) is associated with corner crushing in a wall as a result 
of the racking force from the frame. This can be expected for walls constructed of weak masonry units, 
such as hollow clay tile, which is vulnerable to compressive failure. This mechanism will result in flexural 
hinges in the columns and may also introduce a short‐column effect. 
 
In  addition  to  the  aforementioned  failure  modes,  older  multi‐story  RC  frames  with  strong  infill  walls 
could  develop  lap‐splice  failures  at  the  base  of  the  columns.  In  such  a  case,  the  infill  walls  and  the 
boundary  columns  act  together  to  function  like  a  monolithic  shear  wall  with  severe  tensile  and 
compressive forces developing at the base of the columns. 
 
Hence, the lateral load resistance contributed by the frame and the wall in an infilled frame is governed 
by  the  frame‐wall  interaction  and  the  resulting  failure  modes.  As  a  consequence,  the  resistance  of  an 
infilled  frame  is  not  a  simple  sum  of  that  of  a  bare  frame  and  that  of  a  masonry  wall  considered 
independently. 
 
The ability of an infilled frame to resist earthquake loads depends not only on its strength but also on its 
ductility.  As  shown  in  Figure  3,  a  frame  exhibiting  Failure  Mechanism  (a)  discussed  above  tends  to  be 
more  ductile  than  a  frame  developing  Mechanism  (b)  because  the  latter  has  the  shear  failure  of  the 
columns. Furthermore, as shown in Figure 4, increasing the axial load on an infilled frame governed by 
Failure  Mechanism  (b)  can  increase  its  lateral  load  resistance  but  reduces  its  ductility.  This  can  be 
explained by the fact that the masonry infill is the main load resisting element of an infilled frame, and 

7.3
  @Seismicisolation
@Seismicisolation
P. B. Shing and A. Stavridis

the  lateral  resistance  of  a  masonry  wall  is  largely  governed  by  the  shear  resistance  of  the  bed  joints, 
which largely depends on the compressive stress in the joints. Nevertheless, masonry units will be more 
susceptible  to  crushing  failure  when  subjected  to  higher  axial  and  shear  stresses.  Such  crushing  can 
occur in the interior of a wall (which is different from Mechanism (c)), and will lead to a rapid strength 
degradation. Such behavior can also be observed in frames exhibiting Mechanism (a). 
 
 

Mechanism (a)  Mechanism (b) 

Figure 3 – Non‐ductile RC frames of the same design exhibiting Failure Mechanism (a) and Failure 
Mechanism (b) due to different infill walls (Mehrabi et al. 1996). 
   

 
 

Figure 4 – Influence of vertical load in non‐ductile RC frames exhibiting Failure Mechanism (b) 
(Mehrabi et al. 1996). 
 
Hence, to assess the shear strength of a masonry wall, it is important to know the amount of axial load 
carried by the wall. The proportion of the gravity load initially carried by an infill wall depends on the 
ratio of the elastic axial stiffness of the wall to that of the RC columns and the construction sequence, 
i.e.,  whether  the  walls  were  constructed  before  the  upper  stories  were  built.  If  the  infill  walls  were 
constructed  after  the  entire  RC  frame  had  been  built,  then  these  walls  would  experience  almost  no 
gravity  load  right  after  construction.  However,  part  of  the  sustained  gravity  load  may  be  gradually 
shifted to the walls as creep and shrinkage occurs in the RC columns and the clay units expand in time 
due to moisture absorption. Therefore, one can expect that infill walls will eventually carry a portion of 
the gravity loads unless gaps are introduced between the walls and the beams above. The vertical load 
carried  by  a  wall  can  be  measured  by  in‐situ  tests  using  apparatus  like  flat‐jacks.  However,  in  the 

7.4
@Seismicisolation
@Seismicisolation
Analysis of Seismic Response of Masonry-Infilled RC Frames through Collapse

absence of such data, it has to be estimated analytically. An analytical method to calculate the gravity 
load distribution in an infilled frame considering the creep of the RC columns and masonry wall can be 
found in Koutromanos (2011). 
 
As compared to a bare frame, the ductility of an infilled frame is normally much lower. For a non‐ductile 
frame with a strong infill wall, diagonal/sliding shear cracks usually occur in the infill at a story‐drift ratio 
of 0.2 to 0.3%. This will be quickly followed by the shear failure of the RC columns, after which the in‐
plane lateral resistance starts to decrease rapidly with increasing drift. Normally, severe damage will be 
inflicted in an infilled frame when the story‐drift ratio exceeds 1%. 
 
In  summary,  the  performance  of  an  infilled  frame  and  its  ability  to  survive  an  earthquake  depend  on 
many  factors,  including  the  strength  and  quality  of  the  masonry  infill,  the  reinforcing  details  of  the 
frame,  and  the  resistance  of  the  frame‐wall  system  as  compared  to  the  seismic  load  demand.  In 
addition,  the  strength  and  ductility  of  an  infilled  frame  depend  on  the  shapes,  sizes,  and  locations  of 
openings in the walls (Stavridis 2009), and the locations of infill walls in the frame. 
 
Causes of collapse 
Even  though  an  infilled  frame  may  exhibit  an  undesired  failure  mode,  such  as  Failure  Mechanism  (b) 
shown in Figure 1, and the falling debris of damaged masonry can be a major life‐safety concern, both 
field  observations  from  past  earthquakes  and  laboratory  studies  (e.g.,  see  Stavridis  et  al.  2012)  have 
shown that masonry infill walls can significantly enhance the lateral resistance of a non‐ductile RC frame 
and  protect  it  from  major  damage  or  collapse  in  the  event  of  a  severe  earthquake.  The  collapse  of 
infilled RC frames in past earthquakes was often associated with a weak story mechanism, which could 
be attributed to the lack of infill walls in the bottom story of a building for certain practical reasons or to 
the severe damage and subsequent loss of infill during strong shaking. Studies have shown that intact, 
well‐built,  multi‐wythe,  infill  walls  can  develop  significant  resistance  to  out‐of‐plane  loads  because  of 
the  arching  mechanism  that  can  develop  with  the  bounding  frame  (Dawe  and  Seah  1989,  Angel  et  al. 
1994,  Mander  et  al.  1993,  Bashandy  et  al.  1995,  and  Flanagan  and  Bennett  1999).  However,  the 
effectiveness  of  the  arching  mechanism  and  the  stability  of  a  wall  depend  on  its  height‐to‐thickness 
(slenderness) ratio, and such mechanism can be jeopardized when the masonry infill has been damaged 
by the frame‐panel interaction due to in‐plane shaking. Single‐wythe infill walls are especially vulnerable 
to out‐of‐plane collapse but their contribution to the structural resistance is small to begin with. Figure 
5(a) shows the collapse of a single‐wythe clay tile wall during an earthquake while the building remained 
standing. Figure 5(b) shows the collapse of an infilled‐frame building in the same earthquake, which can 
be attributed to the lack of sufficient walls in the first story. One can see from the picture that significant 
drift  occurred  in  the  first  story  of  the  building  leading  to  the  separation  of  the  RC  frame  from  the 
exterior wall, which together with the damage inflicted on the columns might have contributed to the 
loss  of  the  vertical  load  carrying  capacity  of  the  structure.  Walls  having  window  openings  are  more 
susceptible to collapse, which was observed in a shaking‐table test as shown in Figure 6. 
 
Computational models 
To  simulate  the  earthquake  response  of  a  masonry  infilled  RC  frame  through  collapse,  it  is  important 
that the model can directly or indirectly account for the aforementioned failure mechanisms including 
the loss of stability of infill walls due to in‐plane and out‐of‐plane loads. In particular, it is important that 
the model be able to account for the following possible local and global behavior of a structure. 
 

7.5
  @Seismicisolation
@Seismicisolation
P. B. Shing and A. Stavridis

1. The  inelastic  behavior  of  RC  members,  including  flexural  hinging,  diagonal  shear  failure,  axial 
load failure of RC columns, lap‐slip failures, and the failure of beam‐to‐column joints. 
 
2. The inelastic behavior of masonry walls, including the cracking and shear sliding of mortar joints 
and the crushing failure of masonry units. 
 
3. The  loss  of  stability  and  out‐of‐plane  collapse  of  infill  walls,  considering  the  influence  of  the 
arching mechanism, wall damage induced by in‐plane loads, and wall openings. 
 
4. The frame‐wall interaction and the resulting failure mechanism. 
 
5. The P‐ effect on the RC frame. 
 
6. The three‐dimensional response of a building to multi‐axial ground motions. 
 
 
 

(a) Collapse of infill walls  (b) Collapse of a building (Courtesy of Bin Wu, 


Harbin Institute of Technology) 

Figure 5 – Damage of buildings with masonry infill walls in 2008 Wenchuan Earthquake in China. 
 
 
 
It  is  also  important  that  the  computational  models  should  be  general  enough  to  account  for  the 
construction quality. However, a model that has all the aforementioned simulation capabilities can be 
computationally prohibitive, and simplifications will be normally needed. In the following sections, two 
modeling  methods  are  discussed.  One  is  the  use  of  detailed  finite  element  models  using  continuum 
elements and the other is the use of simplified models with frame elements. 
 

7.6
@Seismicisolation
@Seismicisolation
Analysis of Seismic Response of Masonry-Infilled RC Frames through Collapse

 
 
Figure 6 – Collapse of masonry infill with a window opening in a shaking‐table test (Stavridis et al. 2012). 
 
 
NONLINEAR FINITE ELEMENT MODELS 
A detailed nonlinear finite element model can capture the global as well as local failure behavior of an 
infilled frame. However, these models require a significant computational effort. There are two general 
finite element modeling approaches to simulate the fracture behavior of quasi‐brittle materials, such as 
concrete  and  masonry.  They  are  the  smeared  and  discrete  crack  approaches.  The  smeared‐crack 
approach  has  been  often  used  to  model  diffused  tensile  cracks  as  well  as  the  compressive  failure  of 
concrete  in  reinforced  concrete  structures.  However,  this  approach  suffers  from  several  inherent 
limitations, including stress locking (Rots 1991, Lotfi and Shing 1991) that limits its ability to simulate the 
brittle  behavior  of  an  RC  member  failing  in  diagonal  shear  and  the  shear  sliding  behavior  of  masonry 
joints. This limitation can be overcome by representing cracks in a discrete manner using zero‐thickness 
interface elements. Hence, combining the smeared and discrete approaches is necessary to model the 
failure behavior of infilled frames in a realistic way (Stavridis and Shing 2010, Koutromanos et al. 2011). 
An example of this is shown in Figure 7, in which zero‐thickness interface elements were used to model 
flexural  and  diagonal  shear  cracks  in  RC  columns  and  cracks  in  masonry  joints,  while  smeared  crack 
elements were used to model diffused damage, including compressive failure in the concrete members 
and  brick  units.  In  this  model,  each  reinforcing  bar  in  the  columns  was  divided  into  multiple  truss 
elements  (Stavridis  and  Shing  2010)  so  that  each  discrete  crack  was  resisted  by  the  right  quantity  of 
reinforcement. This also improved the robustness of the numerical solution by having as many smeared‐
crack elements connected to a truss bar as possible. 
 
Constitutive  laws  that  can  be  used  for  the  smeared‐crack  and  zero‐thickness  interface  elements  are 
described  in  Koutromanos  (2011)  and  Koutromanos  et  al.  (2011).  A  smeared‐crack  model  represents 
cracks in a distributed fashion and can also simulate the compressive failure behavior of concrete and 
masonry. The model used by Koutromanos et al. (2011) has an uncracked material represented by a J2‐
plasticity constitutive law. As shown in Figure 8(a), it has a von Mises failure surface with a tension cut‐
off. When the maximum principal stress reaches the tensile strength of the material,  cracks initiate in a 
direction  normal  to  the  direction  of  the  maximum  principal  stress.  For  a  cracked  material,  it  adopts  a 
nonlinear orthotropic law with the axes of orthotropy normal and parallel to the crack, whose direction 
is assumed fixed. At each point, two orthogonal cracks are allowed to occur. The uniaxial stress‐strain 
relation for each direction of the orthotropic law is shown in Figure 8(b). 

7.7
  @Seismicisolation
@Seismicisolation
P. B. Shing and A. Stavridis

 
Figure 7 – Finite element modeling of a masonry infilled RC frame. 
 
 
INITIAL YIELD SURFACE σ2

-f ’m -fo f ’t ft  
Initial stiffness σ Exponential
unloading/reloading softening
σ1 ft
f ’ t ε2 ε1
LINEAR ε
ELASTIC -fo
Secant stiffness
unloading/reloading
ELASTOPLASTIC Exponential
softening
-f ’m f΄c
fc   parabola

VON MISES
FAILURE SURFACE

(a) Failure surface  (b) Uniaxial stress‐strain law 

Figure 8 – Smeared‐crack constitutive model. 
 
The  shear  failure  of  RC  columns  and  the  sliding  shear  behavior  of  mortar  joints  are  best  captured  by 
zero‐thickness interface elements with a cohesive crack constitutive law. A cohesive crack elastic‐plastic 
constitutive  model  has  been  proposed  by  Koutromanos  and  Shing  (2012)  to  capture  the  mix‐mode 
fracture as well as the cyclic crack opening‐closing and sliding shear behavior of concrete and masonry. 

7.8
@Seismicisolation
@Seismicisolation
Analysis of Seismic Response of Masonry-Infilled RC Frames through Collapse

It is implemented in a 4‐node, zero‐thickness, isoparametric, line element. The model can simulate the 
initiation  and  propagation  of  cracks  under  combined  normal  and  shear  stresses.  It  also  accounts  for 
reversible shear dilatation induced by the roughness of the crack surface, which can have a significant 
effect on the response of a confined crack. As shown in Figure 9(a), the yield surface for the model is 
hyperbolic  and  its  evolution  is  controlled  with  internal  variables  that  are  functions  of  the  mode‐I  and 
mode‐II fracture energies as well as the frictional work. Figure 9(b) shows an example on the modeling 
of the shear behavior of a mortar joint. As shown, the model accounts for the shear dilatation and joint 
compaction due to damage. 
 
Shear Displacement (in)
-0.4 -0.2 0.0 0.2 0.4
n 1.5 218

1.0 146

Shear Stress (MPa)


dn,σ

Shear Stress (psi)


0.5 73
dt,τ
0.0 0
t -0.5 -73
 
-1.0 experiment -146
 
analysis
  -1.5 -218

  -10 -5 0 5 10
Shear Displacement (mm)

Shear Displacement (in)


-0.4 -0.2 0.0 0.2 0.4

Normal Displacement (in)


Normal Displacement (mm)

0.5 0.02
0.0 0.00
-0.5 -0.02
-1.0 -0.04
-1.5 -0.06
-2.0 experiment -0.08
-2.5 analysis -0.10
-3.0 -0.12
  -10 -5 0 5 10
Shear Displacement (mm)

(a) Constitutive law  (b) Shear test of a mortar joint 


(0.687‐MPa [100‐psi] normal stress) 

Figure 9 – Cohesive crack interface model. 
 
It  is  well  known  that  the  interaction  between  the  mortar  joints  and  the  masonry  units  due  to  the 
different moduli of elasticity and Poisson effects of the two materials has a significant influence on the 
strength  and  failure  mechanism  of  a  masonry  assembly.  However,  when  a  zero‐thickness  interface  is 
used to represent the behavior of a mortar joint, the 3‐D behavior of a mortar layer and its interaction 
with  the  adjacent  masonry  units  cannot  be  simulated.  In  this  case,  the  influence  of  the  brick‐mortar 
interaction  on  the  masonry  strength  has  to  be  accounted  for  in  an  indirect  way  by  assuming  that  the 
compressive strength of the masonry units is equal to the compressive strength of the masonry prisms. 
For  brick  masonry,  prism  strengths  are  normally  lower  than  the  brick  strengths  due  to  the  tensile 
splitting  stress  introduced  by  the  softer  mortar  to  the  stiffer  brick.  In  addition,  the  stiffness  of  the 
interface  elements  representing  mortar  joints  has  also  to  be  so  determined  that  the  stiffness  of  the 
masonry assembly is properly represented. A procedure to calibrate these constitutive models is given 
in Stavridis and Shing (2010), and Koutromanos et al. (2011). 

7.9
  @Seismicisolation
@Seismicisolation
P. B. Shing and A. Stavridis

 
This  modeling  approach  has  been  proven  to  be  able  to  accurately  capture  the  failure  mode  and  load‐
displacement  response  of  an  infilled  frame  up  to  a  severe  damage  state,  at  which  the  lateral  load 
resistance  decreases  to  as  much  as  50%  of  the  peak  resistance  and  the  story‐drift  ratio  exceeds  1%. 
Figure 10 shows the simulation of a quasi‐static test conducted by Blackard et al. (2009) on a masonry‐
infilled non‐ductile RC frame. A picture of the damaged specimen is shown in Figure 2. It can be seen 
that the model was able to capture the crack pattern as well as the load‐displacement response of the 
test specimen up to a state with severe load degradation. 
 

800

400

force (kN)
0
-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5

-400 Experiment
Analysis
  -800
drift ratio (%)
 
Figure 10 – Analysis of an infilled frame subjected to quasi‐static cyclic loading. (Koutromanos et al. 
2011). 
 
 
This  modeling  method  has  also  been  successfully  applied  to  simulate  the  response  of  a  three‐story, 
masonry‐infilled, non‐ductile, RC frame tested on a shaking table by Stavridis et al. (2012). As shown in 
Figure 11, the model is able to accurately capture the failure mechanism, response time histories, and 
base shear‐vs.‐bottom story drift relations of the structure through the second last test on the shaking 
table that had the Gilroy motion (from the 1989 Loma Prieta Earthquake) scaled to 120%. The specimen 
collapsed  out‐of‐plane  in  the  subsequent  test  that  had  the  El  Centro  record  from  the  1940  Imperial 
Valley Earthquake scaled to 250%. This occurred after severe diagonal shear failures had developed in 
the bottom‐story columns and the bottom‐story infill wall that had a window had partially collapsed (as 
shown in Figure 6). However, neither the out‐of‐plane collapse of the frame nor the collapse of the wall 
could  be  simulated  by  the  model.  This  is  because  of  the  plane‐stress  and  small  displacement 
assumptions adopted in the model formulation. 
 
 

7.10
@Seismicisolation
@Seismicisolation
Analysis of Seismic Response of Masonry-Infilled RC Frames through Collapse

0.4
0.3
0.2

drift ratio (%)


0.1
0.0
-0.1 3.0 3.5 4.0 4.5 5.0 5.5 6.0
-0.2
Experiment
-0.3 Analysis
-0.4
time (sec)
 
   
(a) Crack Pattern after 120% Gilroy  (b) Bottom‐story drift time history for Gilroy 
motion scaled to 67% 

1.2 2.0
1.5
0.8
1.0
drift ratio (%)

0.4 0.5

Vb/W
0.0 0.0
123.0 123.5 124.0 124.5 125.0 125.5 126.0 -1.2 -0.8 -0.4
-0.5 0.0 0.4 0.8 1.2
-0.4
-1.0
-0.8 Experiment Experiment
-1.5
Analysis Analysis
-1.2 -2.0
time (sec) drift ratio (%)
   
(c) Bottom‐story drift time history for Gilroy  (d) Base shear vs. story drift for Gilroy motion 
motion scaled to 120%  scaled to 120% 

Figure 11 – Analysis of a three‐story infilled frame tested on a shaking table (Koutromanos et al. 2011). 
 
 
It should be noted that even though all the validation studies were based on laboratory specimens that 
had  good  quality  materials  and  workmanship,  the  material  parameters,  such  as  the  mortar  joint 
properties, used in the models can be calibrated to account for the deterioration due to aging and for 
the less‐than‐ideal construction quality that may be encountered in real structures. To model the failure 
of lap splices in RC columns, a bond‐slip model can be used to connect reinforcing bars to smeared‐crack 
elements. Such a model has been developed by Murcia‐Delso (2013) to simulate the bond‐deterioration 
between the reinforcing steel and concrete, and can be used to simulate the failure of bar anchorages 
and lap‐splices. 
 
The constitutive models presented here can be extended to 3 dimensions to simulate the out‐of‐plane 
collapse  of  an  RC  frame  or  that  of  a  masonry  infill  wall.  While  this  extension  is  conceptually 
straightforward,  it  can  significantly  increase  the  computational  effort.  Furthermore,  to  simulate  the 
collapse  of  a  masonry  wall  due  to  brick  dislocation  and  the  collapse  of  an  RC  column  with  dominant 
shear  cracks  under  an  axial  load,  the  cohesive  crack  interface  model  presented  here  needs  to  be 
extended to account for the loss of contact due to shear sliding. This can be achieved with a contact law. 

7.11
  @Seismicisolation
@Seismicisolation
P. B. Shing and A. Stavridis

However, combining a cohesive crack model with a contact law will introduce additional computational 
overheads. 
 
SIMPLIFIED ANALYSIS METHODS 
A simple and efficient approach to model infill walls in frame structures is to use the equivalent diagonal 
strut concept as proposed by Holmes (1961), Stafford Smith (1967), and Mainstone and Weeks (1970). 
This concept has been investigated and extended by others in numerous studies. However, the original 
equivalent  strut  theory  was  based  on  the  observed  behavior  of  small‐scale  steel  frames  infilled  with 
either  brickwork  or  concrete.  For  masonry  infilled  RC  frames,  the  equivalent  strut  representation 
appears  to  be  an  over‐simplification  of  the  actual  behavior  and  fails  to  capture  some  main  failure 
mechanisms,  such  as  Mechanisms  (a)  and  (b)  depicted  in  Figure  1.  A  study  by  Stavridis  (2009)  using 
detailed  nonlinear  finite  element  models  has  demonstrated  that  the  compressive  stress  field  in  a 
masonry infill wall may not be accurately represented by a single diagonal strut. As shown in Figure 12, 
the distribution and orientation of compressive forces developed in an infill wall can be quite different 
from that introduced by a single diagonal strut. Hence, replacing a wall by a diagonal strut will not lead 
to a realistic representation of the frame‐wall interaction. Furthermore, it will not simulate the possible 
shear failure of a column that could be induced by the frame‐wall interaction. Models to simulate the 
shear  and  axial  load  failures  of  columns  using  zero‐length  springs  have  been  proposed  (e.g.,  Elwood 
2004). The zero‐length spring concepts can be combined with a strut model in the way shown in Figure 
13  to  account  for  the  axial  and  shear  failures  of  a  column  as  a  result  of  the  frame‐wall  interaction. 
However, three issues are associated with such a model. One is that it ignores the shear transfer that 
develops  between  the  beam  and  the  wall,  as  shown  in  Figure  12,  and  it  will,  thereby,  induce 
unrealistically large shear demands on the columns causing their premature shear failure. Second, such 
a  model  ignores  the  variation  of  the  axial  force  along  the  height  of  each  column  due  to  the  interface 
stresses introduced by the infill. Finally, in such a model, the shear failure of a column may result in the 
immediate  collapse  of  the  column  under  the  axial  load,  which  will  lead  to  the  collapse  of  the  frame 
without  being  able  to  account  for  the  vertical  load  carrying  capacity  of  the  wall.  Moreover,  the 
orientation  of  the  compressive  forces  in  a  wall  can  change  as  damage  evolves,  while  the  angle  of  a 
diagonal  strut  is  fixed.  Multiple‐strut  models  as  proposed  in  some  studies  might  partially  overcome 
these problems. Nevertheless, this will increase the complexity of a model and its benefits remain to be 
demonstrated. 
 
 

Vlc = 209 kN Vrc = -10 kN


Nlc = 27 kN Vw = 393 kN Nw = -368 kN Nrc = 3 kN

Vlc = -25 kN Vrc = -22 kN


Nlc = 234 kN Vw = 638 kN Nw = -551 kN Nrc = -21 kN

Vlc = 7 kN Vrc = 272 kN


Nlc = 214 kN Vw = 311 kN Nw = -246 kN Nrc = -306 kN

Vtot = 591 kN
 
 
Figure 12 – Strut model vs. forces on an infill from  Figure 13 – Strut model with zero zero‐thickness 
finite element analysis.  springs. 
 

7.12
@Seismicisolation
@Seismicisolation
Analysis of Seismic Response of Masonry-Infilled RC Frames through Collapse

Lateral Load
Qmax
Qy K2
1

Qres
K
1

y Q
max
Q Story Drift
res

(a) Load‐vs.‐displacement envelop for an infilled  (b) Comparison of the simplified curve with the 
fame  experimental result for the structure shown in 
Figure 10 
 
Figure 14 ‐ Comparison of a simplified curve with experimental result. 
 
 
In  spite  of  the  aforementioned  issues,  the  use  of  strut  models  is  probably  the  most  efficient  way  for 
Monte‐Carlo‐type  simulations.  For  this  purpose,  one  can  treat  diagonal  struts  as  purely 
phenomenological  models.  However,  the  struts  should  be  calibrated  in  such  a  way  that  they  not  only 
reflect  the  behavior  of  the  infill  walls  but  also  the  response  of  the  frame  members  associated  with 
mechanisms  that  are  not  represented  in  the  model,  such  as  the  shear  failure  of  the  RC  columns.  Two 
steps  are  needed  for  such  calibration.  The  first  is  to  divide  a  multi‐bay,  multi‐story,  structure  into 
multiple  single‐bay,  single‐story  frames,  and  the  second  is  to  derive  the  lateral  load‐vs.‐displacement 
curve  for  each  of  these  frames.  The  second  step  can  be  accomplished  by  creating  a  detailed  finite 
element model for each bay of the structure. However, this can be impractical for structures with many 
bays  and  stories.  To  circumvent  this  difficulty,  Stavridis  (2009)  has  used  experimental  data  and  finite 
element  analysis  results  to  derive  a  set  of  simple  rules  to  define  ASCE  41‐type  pushover  curves  for 
infilled  frames.  His  study  focused  on  non‐ductile  RC  frames  with  strong  infill  walls  consisting  of  multi‐
wythe  solid  clay  brick  units,  which  typically  lead  to  Failure  Mechanism  (b)  shown  in  Figure  1.  The 
idealized  lateral‐vs.‐displacement  curve  adopted  is  shown  in  Figure  14  and  it  is  calibrated  with  the 
following steps. 
 
1. The initial stiffness of an uncracked infilled frame is calculated with a shear beam model as proposed 
by Fiorato et al. (1970). 
 
1
  K    (1) 
1 1

K fl K shl
 
in which  K fl and K sh represent the flexural and shear stiffness of an uncracked cantilever wall. With this 
approach, the structure is assumed to be a composite beam with the RC columns being the flanges and 
the  masonry  wall  being  the  web  of  the  beam.  Hence,  for  the  flexural  stiffness,  K fl ,  the  equivalent 
properties of the composite beam should be used, but for the shear stiffness, only the contribution of 

7.13
  @Seismicisolation
@Seismicisolation
P. B. Shing and A. Stavridis

the wall needs to  be considered. The flexural stiffness can be calculated as  K fl  3Ec I ce hb3  , in which 


hb is the height of the composite wall measured from the top of the foundation to the mid‐height of the 
RC  beam,  E c is  the  modulus  of  elasticity  of  concrete,  and  I ce is  the  moment  of  inertia  of  the 
transformed section with the masonry wall section replaced by an equivalent concrete section. With the 
assumption  that  the  shear  stress  is  uniform  across  the  wall,  the  shear  stiffness  can  be  calculated  as
K sh  AwGw hw , in which Aw , G w , and h w are the cross‐sectional area, shear modulus, and height of the 
infill wall.  
 
2.  The  yield  strength, Qy ,  is  the  point  at  which  there  is  a  distinct  reduction  in  stiffness  due  to  the 
separation between the infill and the RC frame. The study by Stavridis (2009) has indicated that the yield 
force is typically between 65 and 80% of the peak strength. Hence, a conservative estimate of the yield 
force is as follows. 
 
  Q y  2 Qmax   (2) 
3
 
3.  The  peak  strength,  Q max ,  for  a  non‐ductile  frame  with  a  strong  infill  occurs  at  the  instant  a  major 
diagonal/shear sliding crack develops in the infill but prior to the shear failure of the columns. Hence, 
this strength can be calculated as  
 
  Q max   Qlc  Q rc   c o Aw  Pw   (3) 
 
in which  Qlc  and  Q rc are the shear forces carried by the two columns,  co is the cohesive strength of the 
mortar joints,  Aw is the cross‐sectional area of the infill wall,   is the friction coefficient, which is equal 
to one assuming a friction angle of 45o,  Pw is the portion of the gravity load that is carried by the infill, 
and ψ is a reduction factor between 0 and 1 to account for the fact that the two columns and the infill 
may not develop their peak resistances at the same time. One can conservatively assume that ψ is equal 
to 0. The axial load, Pw , on the infill can be estimated based on the elastic axial stiffnesses of the infill 
and  columns.  This  is  a  simplifying  assumption  but  a  more  exact  assessment  of  the  force  is  not  simple 
because  this  force  changes  along  the  height  and  also  as  the  lateral  displacement  increases.  The  shear 
resistance of the columns can be estimated based on the formulas provided in ACI 318.  
 
4. The story drift at the peak strength,   Q , depends on the aspect ratio of the wall, ARw. The following 
max

empirical equations have been proposed by Stavridis (2009) based on the parametric study. 
 
1
  Q  0.86  AR w   for  AR w  2.15   (4a) 
max
3
 
  Q max
 0.15   for  AR w  2.15   (4b) 
 

7.14
@Seismicisolation
@Seismicisolation
Analysis of Seismic Response of Masonry-Infilled RC Frames through Collapse

5. The residual strength,  Q res , can be calculated as the sum of the residual strength of the infill due to 


friction along the bed joints,  Qw,res ,and the residual shear resistance of the columns,  Qc ,res , after the 
development  of  dominant  shear  cracks.  For  the  latter,  only  the  resistance  of  the  shear  reinforcement 
crossed by a diagonal shear crack should be considered.  
 
  Qres  2Qc ,res  Qw,res   (5) 
 
6.  Based  on  the  parametric  study  of  Stavridis  (2009),  the  value  of  the  story‐drift  ratio  at  which  the 
residual strength,   Qres , is reached can be taken to be 1.40  Q . This is a conservative estimate and it 
max

corresponds to a brittle behavior with a high post‐peak negative stiffness. 
 
The  aforementioned  procedure  has  been  calibrated  only  for  non‐ductile  RC  frames  infilled  with  solid 
masonry panels. All these frames exhibit Failure Mechanism (b) in Figure 1, with shear cracks developed 
in the infill and the columns. If a different failure mechanism develops due to different infill and frame 
properties, the pushover curve has to be recalibrated. It is also important to point out that this study has 
considered in‐plane seismic loads only with the out‐of‐plane loads ignored. 
 
The load‐vs.‐displacement relation of an infilled frame is also affected by the existence of window and 
door  openings  in  the  infill  walls.  The  openings  reduce  the  strength  and  stiffness  of  an  infilled  frame. 
Stavridis (2009) has examined the behavior of 30 infilled frames with various opening shapes, sizes, and 
locations in the infill using nonlinear finite element models. In his study, the ratio of the opening area to 
the  gross  surface  area  of  the  wall  is  between  7  and  19%.    Based  on  this  study,  he  has  proposed  the 
following recommendations to derive a conservative backbone curve for infilled frames with openings. 
 
1. The initial stiffness of a frame with an opening is to be estimated with the following formula. 
   
  K
 1   RA   (6) 
K solid
 
in  which  K solid   is  the  initial  stiffness  of  the  same  frame  with  a  solid  infill  calculated  with  Equation  1, 
RA  Aop AWtot , which is the ratio of the opening area,  Aop , to the area of the solid infill wall,  AWtot , and 
α is a factor depending on the shape of the opening. Based on the results of the parametric study, it has 
been suggested that    2  for infills with a window and    1.6 for infills with a door. 
 
2. The yield force,  Q y , corresponds to the load at which cracks initiate at the corners of the openings. 
Based  on  the  parametric  study,  the  ratio,  Qy / Qmax ,  is  between  65  and  80%,  similar  to  the  case  with 
solid infill walls. Hence, this ratio can be assumed to be 2/3 for infilled frames with openings as well.  
 
3. For RA between 7 and 19%,  Qmax can be conservatively estimated as follows. 
 
  Qmax  0.8  Qmax
solid
  (7) 
 
Solid
in which  Qmax  is the peak strength of the same frame with a solid infill.  

7.15
  @Seismicisolation
@Seismicisolation
P. B. Shing and A. Stavridis

 
4. The drift at the peak load,   Qmax , can be conservatively assumed to be equal to that for the frame with 
a  solid  masonry  panel,  which  can  be  estimated  with  Equation  4.  Alternatively,  the  drift,  Qmax ,  can  be 
determined with the post‐yield hardening stiffness,  K 2  , as shown in Figure 14(a). This stiffness can be 
estimated as a fraction of the initial stiffness. 
 
  K2   K   (8) 
 
with   recommended to be 0.10 for infills with a window and 0.08 for infills with a door. 
 
5. The residual strength  Qres can be determined as a fraction of the peak strength. 
 
  Qres  0.5  Qmax   (9) 
 
6. The drift   Qres , can be assumed to be equal to 1.40  Q , which is the same as that for frames with 
max

solid infills.  
 
Once the backbone curve for an infilled frame has been defined with the procedure described above or 
with a finite element analysis, the curve can be used to determine the load‐displacement relation for an 
equivalent  diagonal  strut.  It  should  be  emphasized  that  this  curve  accounts  for  the  load  degradation 
exhibited  by  an  infilled  frame  due  to  the  shear  and/or  flexural  failure  of  the  RC  columns  as  well  as 
damage in the infill. For a multi‐bay, multi‐story, structure, the calibration can follow these steps: 
 
Step 1: Develop the backbone curve for the lateral load‐vs.‐drift relation for each bay and each story of 
the structure using the aforementioned procedure or a finite element model.  
 
Step 2: Construct a bare RC frame model for each bay and each story of the structure, and determine 
the lateral load‐vs.‐displacement backbone curve for the frame considering the appropriate gravity load. 
 
Step 3: Calibrate the strut model so that when it is added to the bare frame considered in Step 2, one 
can obtain the behavior of the infilled frame estimated in Step 1. 
 
Step 4: Develop a model for the entire structure consisting of the frame elements determined in Step 2 
and the diagonal struts calibrated in Step 3. 
 
The  above  method  was  applied  to  model  the  response  of  the  three‐story  infilled  frame  tested  on  a 
shaking table by Stavridis et al. (2012), with the structural configuration shown in Figure 11. The frame 
model  with  equivalent  diagonal  struts  is  shown  in  Figure  15(a).  The  analysis  was  carried  out  with  the 
software platform OpenSees. The axial behavior of the diagonal struts was modeled with a constitutive 
law  for  concrete,  which  was  calibrated  with  simplified  pushover  curves  derived  with  the  procedure 
described  above.  The  strut  for  each  bay  was  individually  calibrated.  The  load‐displacement  curves  for 
the infilled frame models representing the two bottom‐story bays are shown in Figures 15(b) and 15(c). 
One was for the solid infill while the other was for the infill with a window opening. In both cases, the 
numerical model in OpenSees deviated slightly from the idealized (simplified) pushover curve because of 
the specific constitutive law used. In Figure 15(d), the base shear‐vs.‐bottom story drift hysteresis curves 

7.16
@Seismicisolation
@Seismicisolation
Analysis of Seismic Response of Masonry-Infilled RC Frames through Collapse

obtained from the dynamic analysis of the three‐story frame with the strut‐based model, for a sequence 
of ground motion records up to the 120% Gilroy motion, are compared to the test results. The model 
shows  more  rapid  load  degradation  than  the  test  results  due  to  the  conservatism  introduced  in  the 
simplified pushover curves. The quality of the simulation could probably be improved if a detailed finite 
element model was used to calibrate the model in place of the simple rules. 
 
This type of phenomenological model can be extended to simulate the response of an infilled frame to 
out‐of‐plane  loads  by  using  two  bi‐axial  fiber‐section  beam‐column  elements  to  represent  a  diagonal 
strut.  A  fiber‐section  beam‐column  model  can  simulate  both  the  arching  mechanism  and  the  out‐of‐
plane failure of an infill wall. This approach has been considered by Kadysiewski and Mosalam (2008), 
who  have  proposed  an  interaction  curve  to  estimate  the  capacity  of  an  infill  wall  subjected  to 
simultaneous  in‐plane  and  out‐of‐plane  loads.  The  proposed  interaction  curve  is  calibrated  with  finite 
element models. This approach deserves to be further explored and evaluated with experimental data 
(e.g., the data of Dawe and Seah 1989, Angel et al. 1994, Mander et al. 1993, Bashandy et al. 1995, and 
Flanagan and Bennett 1999). 
 
 
700

600 OpenSEES model


Simplified curve 120
500 Bare Frame

Lateral force, kips


Lateral force, kN

400
80
300

200
40

100

0 0
0 0.2 0.4 0.6 0.8 1 1.2
  Drift ratio, %
   
(a) Diagonal strut model  (b) Pushover analysis of a bay with solid infill 

1800 400
700
1350 Shake-Table Tests 300
600 OpenSEES model
Simplified curve 900 Strut model 200
120
Base shear, kips
Base shear, kN

500 Bare Frame


Lateral force, kips

450 100
Lateral force, kN

400 0 0
80
300 -450 -100
-900 -200
200
40
-1350 -300
100
-1800 -400
0 0 -1.5 -1 -0.5 0 0.5 1
0 0.2 0.4 0.6 0.8 1 1.2 st
1 Story drift, %
Drift ratio, %
 
(c) Pushover  analysis of an infilled bay 
(d) Hysteretic curves from dynamic analysis 
with a window 
to up 120% Gilroy 
Figure 15 – Frame analysis using equivalent diagonal struts. 
 

7.17
  @Seismicisolation
@Seismicisolation
P. B. Shing and A. Stavridis

COLLAPSE CRITERION 
The detailed finite element modeling and simplified analysis methods presented above can be used to 
simulate the response of an infilled frame up to a severe damage stage close to collapse. Nevertheless, 
neither  method  can  simulate  the  actual  collapse  of  an  infilled  frame  in  a  satisfactory  manner  without 
further enhancements. For Monte‐Carlo‐type simulations, the use of detailed finite element models is 
not practical and simplified strut‐based models should be used. 
 
Regardless of the type of model being used, one critical issue in a collapse simulation of any structure is 
the definition and identification of collapse. While collapse can be defined as the loss of the vertical load 
carrying capacity of a structure, the distinction between the partial and total collapse needs to be made. 
The simulation of either the partial or the total collapse of an infilled frame is a big challenge because of 
the interaction and possible changes of the load transfer mechanisms among the columns, beams, and 
walls. Once the RC columns in an infilled frame have suffered severe damage such as shear failure, the 
gravity load can be shifted to the infill walls as long as these walls remain standing. A severely cracked 
masonry wall could still carry significant gravity load even though it might have lost a large portion of its 
in‐plane  and  out‐of‐plane  lateral  load  resisting  capacity.  This  can  be  perceived  as  an  unstable 
equilibrium state, which has been often observed in in‐plane quasi‐static tests (see Figure 4) as well as 
nonlinear finite element analyses (Koutromanos et al. 2011). Hence, the collapse of an infilled frame has 
to be preceded by the collapse of the infill walls. The collapse of infill walls could occur before or after 
the RC frame has suffered severe damage. The walls could collapse prior to the development of severe 
damage  in  the  frame  if  the  masonry  is  weak  or  large  openings  exist  in  the  walls,  and  this  will 
immediately result in a weak‐story mechanism. For this situation, simple models like those proposed by 
Elwood (2004) can be used to simulate the subsequent collapse of the RC columns.  
 
In  view  of  the  aforementioned  modeling  challenges  and  the  large  uncertainties  in  the  collapse 
prediction for an infilled frame, it is prudent to treat the first attainment of an unstable equilibrium state 
of the structure as the total collapse  condition.  This can be considered as a  state in which the infilled 
frame has lost a significant portion of its lateral in‐plane load carrying capacity due to the failure of the 
walls at one or more stories. Until more experimental data are available, this state can be assumed to be 
reached  when  the  post‐peak  in‐plane  resistance  of  the  infilled  frame  drops  to  60%  of  the  peak.  The 
story‐drift  level  at  which  this  load  degradation  is  reached  can  be  taken  as  the  collapse  criterion  for 
Monte‐Carlo‐type  simulations.  This  critical  drift  limit  can  be  identified  with  a  pushover  analysis  of  a 
planar structure using a detailed finite element model or a simplified model. Experimental results from 
different studies have shown that under in‐plane loads only, this can occur at a story‐drift ratio between 
1.0  and  1.5%.  Quasi‐static  tests  and  finite  element  analyses  have  shown  that  frames  with  weak  infill 
walls or walls with a large opening tend to have lower strengths but higher ductile. Hence, under static 
loads, these frames may not reach 40% strength loss till a relatively large drift level has been reached. 
Nevertheless, walls in these frames will be less stable under dynamic loads as mentioned in a previous 
section. Therefore, to be prudent, the  critical drift limit should not be greater than 1.5%. If a wall has 
been damaged by the out‐of‐plane load, it can be expected that the in‐plane resistance will drop more 
rapidly and a strength loss of 40% can occur at a smaller story‐drift level. Further studies are needed to 
establish a drift limit for such cases. 
 
CONCLUSIONS 
Simplified models using equivalent compression‐only diagonal struts are most efficient for Monte‐Carlo‐
type  collapse  simulations  for  masonry‐infilled  RC  frames.  However,  because  of  the  inability  of  such 
models to represent some main failure mechanisms of an infilled frame, the strut model should be so 
calibrated that the load degradation induced by the unaccounted mechanisms, such as the shear failure 

7.18
@Seismicisolation
@Seismicisolation
Analysis of Seismic Response of Masonry-Infilled RC Frames through Collapse

of the RC columns, can be represented in the load‐displacement response of the entire frame. To this 
end, the axial behavior of a strut can be modeled with a phenomenological law that reflects the damage 
evolution of an entire infilled frame. Such a strut model can be calibrated with a detailed nonlinear finite 
element analysis or the simple procedure described in the paper. The simple procedure presented here 
was  derived  for  structures  that  have  non‐ductile  RC  frames  and  strong  infill  walls.    Further  work  is 
needed to extend the procedure to structures with different frame and wall properties.  
 
In a simplified model, RC beams and columns can be modeled with beam‐column elements with flexural 
hinging  capabilities.  To  simulate  the  damage  and  possible  collapse  of  an  infill  wall  caused  by  out‐of‐
plane  loads,  a  diagonal  strut  can  be  modeled  with  two  bi‐axial  fiber‐section  beam‐column  elements, 
which account for the out‐of‐plane bending as well as the arching mechanism developed in an infill wall. 
Nevertheless, further work is needed to calibrate these models with available experimental data. 
 
For  Monte‐Carlo‐type  collapse  simulations,  a  convenient  criterion  for  collapse  must  be  determined. 
Until more data are available, it can be assumed that an infilled frame will be on the verge of collapse 
when  its  in‐plane  resistance  drops  to  60%  of  the  peak  resistance.  This  is  based  on  test  data  currently 
available.  The  story‐drift  level  at  which  this  load  degradation  will  be  reached  under  monotonically 
increasing  loads  can  be  taken  as  a  collapse  criterion.  This  drift  limit  can  be  identified  with  a  pushover 
analysis of a planar structure using a detailed finite element model or a simplified model, but the critical 
drift ratio for collapse should not be greater than 1.5%. 
 
For  future  numerical  studies  and  the  calibration  of  simplified  models,  finite  element  models  that  can 
simulate collapse in a detailed fashion are desired. To simulate out‐of‐plane failures, 3‐D finite element 
models  are  needed.  Other  simplified  collapse  simulation  strategies  such  as  element  removal  methods 
should  also  be  investigated.  Different  modeling  methods  should  be  compared  and  evaluated  with 
benchmark examples. 
 
ACKNOWLEDGMENTS 
The  modeling  work  presented  in  this  paper  is  based  on  a  prior  research  project  supported  by  the 
National Science Foundation Grant No. 0530709 awarded under the George E. Brown, Jr. Network for 
Earthquake  Engineering  Simulation  Research  (NEESR)  program.  However,  opinions  expressed  in  this 
paper are those of the authors and do not necessarily represent those of the sponsor.  
 
REFERENCES 
Angel, R, Abrams, D, Shapiro, D, Uzarski, J, and Webster, M, 1994, “Behavior of Reinforced Concrete 
Frames  with  Masonry  Infills,”  Report  No.  UILU‐ENG‐94‐2005,  Dept.  of  Civil  Engineering,  University  of 
Illinois, Urbana‐Champaign, IL. 
Bashandy,  T,  Rubiano,  NR,  and  Klingner,  RE,  1995,  “Evaluation  and  Analytical  Verification  of  Infilled 
Frame Test Data,” Report No. 95‐1, Dept. of Civil Engineering, University of Texas, Austin, TX. 
Blackard, B, Willam, K, and Mettupalayam, S, 2009, “Experimental Observations of Masonry Infilled RC 
Frames with Openings,” ACI SP 265‐9, American Concrete Institute, pp. 199‐222. 
Dawe, JL and Seah, CK, 1989, “Out‐of‐Plane Resistance of Concrete Masonry Infilled Panels,” Journal of 
the Canadian Society for Civil Engineering, V. 16, pp. 854‐864. 
Elwood, KJ, 2004, “Modelling Failures in Existing Reinforced Concrete Columns,” Canadian Journal of 
Civil Engineering, V. 31, pp. 846‐859. 
Fiorato,  AE,  Sozen,  MA,  and  Gamble,  WL,  1970,  “An  Investigation  of  the  Interaction  of  Reinforced 
Concrete Frames with Masonry Filler Walls,” Report No. UILU‐ENG 70‐100, Dept. of Civ. Eng., University 
of Illinois at Urbana‐Champaign, Urbana, IL. 

7.19
  @Seismicisolation
@Seismicisolation
P. B. Shing and A. Stavridis

Flanagan, RD and Bennett, RM, 1999, “Bidirectional Behavior of Structural Clay Tile Infilled Frames,” 
Journal of Structural Engineering, V. 125, No. 3, pp. 236‐244. 
Holmes, M, 1961, “Steel Frames with Brickwork and Concrete Filling,” Proceedings of the Institute of 
Civil Engineers, V. 19, No. 6501, pp. 473‐478. 
Kadysiewski, S and Mosalam, KM, 2009, “Modeling of Unreinforced Masonry Infill Walls Considering 
In‐Plane and Out‐of‐Plane Interaction,” PEER Report 2008/102, Pacific Earthquake Engineering Research 
Center, University of California at Berkeley, Berkeley, CA. 
Koutromanos, I, 2011, “Numerical Analysis of Masonry‐Infilled Reinforced Concrete Frames Subjected 
to Seismic Loads and Experimental Evaluation of Retrofit Techniques,” Ph.D. Dissertation, University of 
California at San Diego, La Jolla, CA. 
Koutromanos, I, Stavridis, A, Shing, PB, and Willam, K, 2011, “Numerical Modeling of Masonry‐Infilled 
RC Frames subjected to Seismic Loads,” Computers and Structures, V. 89, pp. 1026‐1037. 
Koutromanos, I and Shing, PB, 2012, “A Cohesive Crack Model to Simulate Cyclic Response of Concrete 
and Masonry Structures,” ACI Structural Journal, V. 109, No. 3, pp. 349‐358. 
Lotfi,  HR  and  Shing,  PB,  1991,”  An  Appraisal  of  Smeared  Crack  Models  for  Masonry  Shear  Wall 
Analysis,” Computers & Structures, V. 41, No. 3, pp. 413–425. 
Mainstone, RJ and Weeks, GA, 1970, “The Influence of Bounding Frame on the Racking Stiffness and 
Strength of Brick Walls,” Proceedings of the 2nd International Conference on Brick Masonry, Stoke‐on‐
Trent, pp. 165‐171. 
Mander,  JB,  Nair,  B,  Wojtkowski,  K  and  Ma,  J,  1993,  “An  Experimental  Study  on  the  Seismic 
Performance  of  Brick‐Infilled  Steel  Frames  with  and  without  Retrofit,”  Report  No.  NCEER‐93‐0001, 
National Center for Earthquake Engineering Research, State University of New York, Buffalo, NY. 
Mehrabi,  AB,  Shing,  PB,  Schuller,  MP,  and  Noland,  JL,  1994,  “Performance  of  Masonry‐Infilled  R/C 
Frames  under  In‐Plane  Lateral  Loads,”  Report  No.  CU/SR‐94‐6,  Dept.  of  Civil,  Environmental,  and 
Architectural Engineering, University of Colorado, Boulder, CO. 
Mehrabi,  AB,  Shing,  PB,  Schuller,  MP,  and  Noland,  JL,  1996,  “Experimental  Evaluation  of  Masonry‐
Infilled RC Frames,” Journal of Structural Engineering, V. 122, No. 3, pp. 228‐237. 
Murcia‐Delso,  J,  2013,  “Bond‐slip  Behavior  and  Development  of  Bridge  Column  Longitudinal 
Reinforcing Bars in Enlarged Pile Shafts,” PhD Dissertation, University of California, San Diego, La Jolla, 
CA. 
Rots, JG, 1991, “Numerical Simulation of Cracking in Structural Masonry,” HERON, Netherlands School 
for Advanced Studies in Construction, The Netherlands, V. 36, No. 2, pp. 49‐63. 
Stafford  Smith,  B,  1967,  “Methods  for  Predicting  the  Lateral  Stiffness  and  Strength  of  Multi‐Story 
Infilled Frames,” Building Science, V. 2, pp. 247‐257. 
Stavridis, A, 2009, “Analytical and Experimental Study of Seismic Performance of Reinforced Concrete 
Frames  Infilled  with  Masonry  Walls,”  Ph.D.  Dissertation,  University  of  California  at  San  Diego,  La  Jolla, 
CA. 
Stavridis, A and Shing, PB, 2010, “Finite Element Modeling of Nonlinear Behavior of Masonry‐Infilled 
RC Frames,” Journal of Structural Engineering, V. 136, No. 3, pp. 285‐296. 
Stavridis,  A,  Koutromanos,  I,  and  Shing,  PB,  2012,  “Shake‐Table  Tests  of  a  Three‐Story  Reinforced 
Concrete Frame with Masonry Infill Walls,” Journal of Earthquake Engineering and Structural Dynamics, 
V. 41, pp. 1089‐1108. 

7.20
@Seismicisolation
@Seismicisolation
SP-297—8

Collapse Assessment of Non-Ductile, Retrofitted and Ductile Reinforced


Concrete Frames
M. Baradaran Shoraka1, K.J. Elwood1, T.Y. Yang1, and A.B. Liel2
1
Department of Civil Engineering, University of British Columbia, 6250 Applied Science Lane, Vancouver BC V6T 1Z4, Canada
2
Department of Civil, Environmental and Architectural Engineering, University of Colorado, Boulder, CO 80309-0428

Synopsis:

Probability of collapse is currently used to set targets for system performance and response measures of new
buildings. This study compares the probability of collapse for new, retrofitted and existing concrete buildings.
Retrofitting plays an important role in reducing seismic risk from older concrete buildings. In order to decide on the
most appropriate and economical retrofit strategy for an existing structure, it is necessary to assess the risk of
collapse of each rehabilitation measure. At present, it is frequently assumed that retrofitting a non-ductile concrete
building will enhance the seismic performance such that it can reach the same performance level as a ductile
building designed based on current seismic codes. However, based on the evaluation of the concrete frames
presented in this paper, typical retrofit schemes (such as: adding an additional lateral force restraint system;
increasing ductility of existing concrete columns; and weakening the existing beams) cannot achieve the same
performance as modern code-conforming structures. The study finds that retrofitting schemes where the columns or
beams are modified such that the frame satisfies the collapse prevention level of ASCE 41-13 have the least
beneficial effect regarding seismic collapse safety; and conversely, adding a shear wall will significantly improve
the seismic performance in terms of the probability of collapse.

Keywords:

Collapse assessment; Nonlinear analysis; Probability of collapse; Retrofitting; Reinforced concrete moment frames;
Seismic safety

8.1
@Seismicisolation
@Seismicisolation
M. Baradaran Shoraka et al.

ACI member Majid Baradaran Shoraka is a post-doctorate researcher at University of British Columbia,
Vancouver, BC, Canada. He received his PhD from the University of British Columbia, Vancouver, BC, Canada.
His research interests include developing performance-based evaluation methodology and code design procedures
for new and existing structures. He is an associate member of the ACI Committee 369.

ACI Fellow, Kenneth J. Elwood is a Professor of Civil Engineering at the University of British Columbia, Canada.
He received his PhD from the University of California, Berkeley. His research interests include the behavior and
performance-based design of reinforced concrete structures under seismic loading. He is Chair of ACI committee
133, Disaster Reconnaissance, past Chair of ACI committee 369, Seismic Repair and Rehabilitation, and member of
Subcommittee H, Seismic Design, of ACI Committee 318.

Tony Yang is an assistant Professor at the University of British Columbia, Vancouver, BC, Canada. He received his
PhD from the University of California, Berkeley, CA. His research focus is on improving the structural performance
through advanced analytical simulation and experimental testing. He is a corresponding member of AISC TC9
which responsible for publishing the seismic design guidelines for steel structures in United States, member of the
Tall Building Initiative project and one of the core developers for the next-generation performance-based earthquake
engineering assessment methodology currently adopted by the ATC-58/FEMA-P58 research project.

Abbie B. Liel is an assistant professor at the University of Colorado Boulder. She received her PhD from Stanford
University. Her research focuses on performance-based assessments of structural safety and collapse risk under
extreme loads, including seismic assessment of non-ductile concrete buildings and alternative retrofit strategies.

INTRODUCTION

With recent earthquakes worldwide, retrofitting plays an important role in mitigating the seismic risk for older RC
structures. Repair and upgrading of existing structures is becoming increasingly important and has significant
economic impacts. Seismic retrofitting of an existing building will either involve adding new components to the
existing structure, increasing the deformation capacity of existing components, or changing the collapse mode of the
structure. In order to decide on the most appropriate and economical retrofitting strategy, it is necessary to assess the
lateral load resistance, deformation capacity, and potential modes of collapse.

Current retrofit standards, including the latest version of ASCE 41 (ASCE, 2013), assist engineers with the seismic
assessment of existing buildings based on component acceptance criteria. In this standard, component demands are
compared with component acceptance criteria for different performance levels, namely immediate occupancy (IO),
life safety (LS), and collapse prevention (CP). The component with the worst performance level will define the state
of the entire structure. The performance of non-ductile reinforced concrete moment frames retrofitted to the LS and
CP performance levels are studied in this paper. Specifically, the collapse vulnerability of different retrofitting
schemes is studied using the system-level assessment procedure introduced in Baradaran Shoraka (2013) and
Baradaran Shoraka et al. (2013).

The collapse vulnerability of existing non-ductile and retrofitted archetype RC structures has previously been
studied by Liel and Deierlein (2013). In that study, 1960s era California type 4- and 8-story non-ductile RC space
and perimeter moment frame structures were chosen. Three retrofit techniques were considered: (1) jacketing of the
RC columns with reinforced concrete, (2) carbon fiber-wrapping of RC columns, and (3) construction of “super
column shear walls” around existing columns. For each type of retrofit, both ‘modest’ and ‘significant’ retrofits were
considered; however, it is not known how these retrofits relate to the criteria specified in ASCE 41-13 as they were
not designed to a specific standard. The collapse performance of these structures was compared to the collapse
performance of the un-retrofitted (original) structure. However, only side-sway collapse mechanisms were
simulated. Non-ductile collapse modes, such as column shear failure and subsequent axial failures, were considered
as non-simulated collapse modes. The study assumes that when the first occurrence of this non-simulated failure
mode occurs, it will lead to collapse of the structure, which produces a conservative assessment of collapse capacity,
ignores the system’s ability to redistribute the load after failure. Figure 1 shows the computed collapse fragilities for
the 4-story space frame structures presented in Liel’s study: including existing (un-retrofitted), retrofitted (using the
three retrofitting techniques) and modern (with 20 ft. and 30 ft. bay widths) buildings. As shown in Figure 1, the

8.2
@Seismicisolation
@Seismicisolation
Collapse Assessment of Non-Ductile, Retrofitted and
Ductile Reinforced Concrete Frames

performance of the retrofitted structures is superior to that of the unretrofitted frames, but does not achieve the low
collapse risk of the modern frames. In addition, Figure 1 shows large variability among the collapse risk of the
retrofitted frames, depending on the unique characteristics of the retrofit. Liel and Deierlein (2013) also conducted a
cost-benefit analysis of the retrofit schemes, the review of which is outside the scope of this paper.

The study presented herein builds on the Liel and Deierlein (2013) findings by explicitly modeling the gravity-load
collapse of the non-ductile frame and by directly considering the nonlinear acceptance criteria for both LS and CP in
ASCE 41-13 in selecting and designing the retrofit for the frames. The first advancement is critical, because it
ensures that the comparison of collapse performance between the unretrofitted and retrofitted frames is more
accurate. The use of ASCE 41-13 ensures that the retrofits considered in this study are consistent with those
designed in practice; ASCE 41-13 is widely used in practice for design and evaluation of seismic retrofit schemes.
The paper will first introduce the case study structure, apply different retrofitting strategies to upgrade the case study
building, and finally compare the seismic performance of each building by means of collapse fragilities and collapse
performance metrics. All the results presented in this paper are original for this manuscript.

Figure 1–Collapse fragility functions for 4-story space frames Liel and Deierlein (2013)

CASE STUDY STRUCTURE

To compare the collapse performance of existing and retrofitted (using the ASCE 41-13 standard) RC moment
frames, a non-ductile perimeter concrete moment resisting frame building designed according to the 1967 UBC
(ICBO 1967) code was selected as the prototype structure. The 8-story perimeter frame is chosen from the Liel
database (Liel, 2008; Liel et al., 2011) and is summarized in Table 1. The prototype building was retrofitted using
three retrofitting techniques, each designed to satisfy two performance objectives (LS and CP) as specified in ASCE
41-13. In addition, the performance of a ductile perimeter moment frame building designed according to the
International Building Code (ICC 2003) was also included in this study.

The prototype building is assumed to be located in Los Angeles, California, site class D, with SS = 1.5g and S1 =
0.6g, where SS is the earthquake spectral response acceleration at short periods and S1 is the earthquake spectral
response acceleration at 1-second periods. The earthquake hazard level defined for the selected performance
objectives has a 2% probability of exceedance in 50 years.

8.3
@Seismicisolation
@Seismicisolation
M. Baradaran Shoraka et al.

Non-ductile perimeter concrete moment resisting frame building

Table 1 provides key characteristics of the prototype building. This structure was designed according to 1967 UBC
for Zone 3 (the highest seismic zoning criteria). Seismic design requirements such as the maximum and minimum
steel reinforcement ratios, maximum stirrup spacing, and requirements on hooks, bar spacing and anchorage were
included. It should be noted that there is no transverse reinforcement in the joints as is typical of buildings of this
era. A concrete strength of 4 ksi (28 MPa) was assumed for all members .

Numerical model– In numerical simulation, there is always the challenge of defining the collapse state during the
analysis and differentiating structural collapse from numerical non-convergence. In collapse analysis used in this
paper, the criterion is based on two types of global failure, gravity-load collapse and side-sway collapse. Side-sway
collapse develops due to P-delta instability at large lateral deformations, with an implicit assumption that the
structure is ductile enough to reach this state. However, because of the limited ductility of older reinforced concrete
structures, an alternate criteria is also defined, entitled here as gravity-load collapse. Gravity-load collapse may be
precipitated by axial-load failure of columns, punching shear failure of slab–column connections, failure of slab-
diaphragm connections, or axial-load failure of beam-column joints. Gravity load collapse is defined in this paper by
the point at which the axial load capacity at a single story drops below the gravity load demand (explained in detail
in Baradaran Shoraka (2013) and Baradaran Shoraka et al. (2013)). Both side-sway and gravity-load collapse have
been considered in the current study.

The numerical analyses were conducted using a two-dimensional frame modeled in OpenSees (PEER, 2009). Only
the perimeter moment resisting frames were included in the numerical model. The analytical model incorporated all
important features required to model collapse of the structure.

 The beam and column elements were modeled using the lumped plastic hinge model developed by Ibarra et
al. (2005) to account for the strength and stiffness degradation under cyclic loads. Strain-softening
behaviour associated with concrete crushing, rebar buckling and fracture, and/or bond failure was
accounted for; including such degradation in the model has an important effect on the collapse response of
these structures. The backbone and cyclic response parameters used in the numerical model were developed
by Haselton et al. (2008).

 The model includes large geometry transformations that take into account P-delta effects resulting from the
mass of the entire structure.

 To ensure the numerical model is capable of modeling the joint failure, a two-dimensional joint model was
used. The numerical model developed by Lowes and Altoontash (2003) was used to define the shear
deformations of the joint and bond-slip behaviour.

 Shear and axial failures in columns were modeled using the limit-state material developed by Elwood
(2004). These models define the shear and axial failures of concrete columns as a function of the
deformation capacity, considering the geometric, material and design parameters.

The first fundamental period of the structure is about 2.4 (s). The ASCE 41 (2013) estimate of the building period is
1.7 (s). The building period is high because this frame was designed to achieve a minimum column size and high
vertical reinforcement percentage.

Ductile perimeter concrete moment resisting frame building

Table 1 also provides key characteristics of the ductile building (Haselton et al., 2011). For comparison, this ductile
building was chosen to be similar to the non-ductile moment frame. This structure was designed according to the
International Building Code (ICC, 2003), ASCE 7 (ASCE, 2002), and ACI 318 (ACI, 2005), and meets all the
governing code requirements and detailing for special moment frames.

8.4
@Seismicisolation
@Seismicisolation
Collapse Assessment of Non-Ductile, Retrofitted and
Ductile Reinforced Concrete Frames

Table 1–Design documentation for 8-story non-ductile (Liel et al., 2011) and ductile (Haselton et al., 2011)
perimeter frame structure (1 in = 0.0254 m)

Column Beam
Column Beam
Design Column Hoop Beam Hoop Floor Bay
Size, Size,
Structure Base Shear Reinforcement Spacing, Reinforcement Spacing, Height Width,
hxb hxb
Coefficient Ratio, tot s Ratio, ’) s (ft) (ft)
(in x in) (in x in)
(in) (in)
8 story-
perimeter
frame 0.054g 30 x 36 0.033 15 26 x 36 0.008 (0.010) 17 15 25
(non-
ductile)
8 story-
perimeter
0.050g 26 x 34 0.018 3.5 26 x 30 0.007 (0.008) 5 15 25
frame
(ductile)

h = height of beam-column tot = ratio of longitudinal reinforcement s = spacing of transverse reinforcement


b = width of beam-column ’ = ratio of longitudinal tension and compression reinforcement

Retrofitted buildings

The non-ductile perimeter concrete moment resisting frame presented in the previous section was evaluated using
the procedures outlined in ASCE 41-13. A pushover analysis, using the first mode load distribution and the target
roof displacement calculated using Eqn. 1, was performed.

Eqn. 1 shows the deformation limit used in the pushover analysis.

(1)

where δ is defined as the roof drift. S represents the elastic spectral displacement of an equivalent single degree
of freedom (SDOF) oscillatory at the earthquake hazard level of interest; chosen Maximum Considered Earthquake,
MCE (equivalent to a probability of occurrence of 2% in 50 years) shaking intensity for this study. Te is the
fundamental period of the existing structure. C0 is characterized as a dimensionless coefficient that relates the
spectral displacement of an equivalent SDOF to the roof displacement. C1 is a dimensionless coefficient that relates
expected maximum inelastic displacements to the displacements calculated using the linear elastic response and
finally C2 is a dimensionless coefficient that adjusts the roof drift ratio to account for the pinched hysteretic shape,
stiffness degradation and strength deterioration effect on the structure.

At such roof drift, the component deformation demands were checked using the acceptance criteria defined in ASCE
41-13. The prototype building was retrofitted to the Life Safety (LS) and Collapse Prevention (CP) performance
levels. For the Collapse Prevention performance level, the structural components can be severely damaged, but the
structure must be able to continue carrying gravity loads without collapse. Because the strength and stiffness
degradation was modeled in all the numerical components, the “Secondary Components” acceptance criteria, as
specified in ASCE 41-13, were used to assess the performance for all components.

There are various retrofitting schemes available. Using the recommendations presented in FEMA 547 (FEMA,
2006), three retrofitting schemes were identified and included in this study:

1) Strengthening the existing columns, beams and joints by steel jacketing, or adding new reinforced concrete,
steel, or fiber-reinforced polymer wrap overlays. This method enhances the strength and deformation
capacity of the existing non-ductile components.

8.5
@Seismicisolation
@Seismicisolation
M. Baradaran Shoraka et al.

2) Weakening a certain portion of the structure by removing or modifying a portion of the existing structure.
This technique, usually applied to weaken the beams, promotes formation of a strong-column weak-beam
mechanism.

3) Adding supplementary lateral force-resisting systems, such as adding ductile reinforced concrete shear walls,
to reduce deformation demands on the existing elements.

Retrofitted building – columns modified (strengthening technique) –The first approach to retrofit non-ductile
RC frames increases the strength and deformation capacity of the critical non-ductile columns. The critical non-
ductile columns were identified based on the acceptance criteria provided in ASCE 41-13. The critical columns were
jacketed with reinforced concrete in order to increase the column deformation capacity with increased transverse
reinforcement. It is assumed that the retrofitted columns will no longer sustain shear/axial failures due to the
additional transverse reinforcement provided in the jacket. The final retrofit design for the selected performance
objective is schematically illustrated in Figure 2a. The highlighted section in this figure shows the retrofitted
columns for each of the performance objectives. The highlighted numbers in the table also show the changes in the
column dimensions and transverse reinforcement for the structures upgraded to satisfy the CP and LS performance
levels.

Retrofitted building – beams modified (weakening technique)–The second approach is to retrofit the non-ductile
RC frames by weakening the existing beams such that the system can form a strong-column, weak-beam
mechanism. The concept of weakening for retrofit of RC frames has been previously proposed in retrofit guidelines
and by several researchers (FEMA-356, 2000; Pampanin, 2005b; Viti et al., 2006, Kam and Pampanin, 2009). This
technique was applied by cutting longitudinal reinforcement in the beams. The longitudinal reinforcement were cut
floor by floor and the demand to the system was recalculated after each iteration, until the component demands for
the critical columns fall within the selected performance level. This retrofitting measure will result in a decrease in
strength and stiffness for the beams. For this technique, it was not possible to weaken the beams to the extent
necessary to reach the LS level. The final design (for the CP level) is schematically illustrated in Figure 2b. The
highlighted section in this figure shows the weakened beams. The highlighted numbers in the table also show the
changes in the beam dimensions and longitudinal reinforcement for the structure upgraded to the CP performance
objectives.
Retrofitted building – wall added (adding technique)–The third approach to retrofit the non-ductile RC frame is
by adding shear walls to decrease the drift demand on the existing moment frame. The shear wall was designed with
the same cross sectional area and reinforcement ratio over the full height of the wall. The strength of the shear wall
was increased by means of expanding the depth and longitudinal steel reinforcement until the component demands
for the critical columns fall into the performance target. Note that in order to assess the performance of a retrofit
satisfying the criteria in ASCE 41-13, only enough wall stiffness was added to get rotations in critical elements
below selected performance target limits, hence final wall dimensions may not resemble a real wall. The final
retrofit design for the selected performance objective is illustrated in Figure 2c. The highlighted section in this figure
shows the wall added to this building. The numbers in the table also show the wall designs for the structures
upgraded to the CP and LS performance levels. The shear wall is modeled in OpenSees using the Nonlinear Beam-
Column element (de Souza, 2000). The shear wall was connected to the frame using elastic truss elements at each
floor (see Figure 2c).

8.6
@Seismicisolation
@Seismicisolation
Collapse Assessment of Non-Ductile, Retrofitted and
Ductile Reinforced Concrete Frames

Columns (CP) Columns (LS)

Existing Columns (CP) Columns (LS)


Floor Columns
sh b (in) h (in) sh b (in) h (in) sh b (in) h (in)
Exterior 0.0036 26 28 0.006 27 32 0.010 30 32
1
Interior 0.005 30 36 0.010 31 40 0.010 34 40
Exterior 0.0031 26 28 0.006 27 32 0.010 30 32
2
Interior 0.004 30 36 0.010 31 40 0.010 34 40
Exterior 0.0031 26 28 0.0031 26 28 0.010 30 32
3
Interior 0.004 30 36 0.004 30 36 0.010 34 40
Exterior 0.0031 26 28 0.0031 26 28 0.010 30 32
4
Interior 0.0036 30 36 0.0036 30 36 0.010 34 40
h = height of beam-column; b = width of beam-columnsh = ratio of transverse reinforcement; 1 in = 0.0254 m
(a) Retrofitted Building – Columns Modified (strengthening technique)

Beams (CP)

Beams Existing Beams (CP)


of b h b h
Floor  '  '
(in) (in) (in) (in)
1 26 36 0.0075 0.01  26 30 0.0075 0.0075
2 26 36 0.0075 0.01  26 30 0.0075 0.0075
3 26 36 0.0075 0.01  26 30 0.0075 0.0075
4 26 36 0.007 0.0093  26 30 0.007 0.007
’ = ratio of longitudinal tension and compression reinforcement; 1 in = 0.0254 m

(b) Retrofitted Building – Beams Modified (weakening technique)

8.7
@Seismicisolation
@Seismicisolation
M. Baradaran Shoraka et al.

Walls (CP) Walls (LS)


Wall (CP)
tw
bf (in) tf (in) L (in) Lw (in)  f  w 
(in)
12 5 50 40 8 0.040 0.0025

Wall (LS)
tw
bf (in) tf (in) L (in) Lw (in)  f  w 
(in)
15 10 75 55 8 0.020 0.0025
fw = ratio of longitudinal reinforcement in the flange and web; bf & tf = width and thickness of the flange;
1 in = 0.0254 m

(c) Retrofitted Building – Walls (adding technique; the shear wall and the truss elements are added)
Figure 2–Design documentation for 8-story retrofitted perimeter frame structure (members highlighted in
this picture are modified for the different retrofitting measures)

PERFORMANCE ASSESSMENT BASED ON ASCE 41-13

The performance assessments for the three types of retrofitted buildings conducted using the ASCE 41-13 document
is presented in the following sections.

Pushover results

ASCE 41-13 Nonlinear Static Procedure uses pushover analysis to assess the deformation demand in the
components. The target roof deformation for all buildings, calculated using Eqn. 1, is presented in Table 2. As seen
in this table, the target roof drift ratio varies from 1.0% to 1.9%. The pushover analysis will represent an idealized
force-deformation curve. The force–deformation was idealized using a bilinear curve to compute the yield and
ultimate drift. Table 3 summarizes the results of the static pushover analyses. Performance levels were assessed
using the acceptance criteria for concrete columns in Table 10-13 of ASCE 41-13.

Figure 3 presents the pushover curve for the non-ductile prototype building. This figure also highlights the sequence
of the critical component failures. In this building, yielding first started in the beams, then the joints and followed by
the columns. Next, the deformation in the non-ductile columns increased and resulted in shear failure in these
critical columns. Axial failure in the columns leads to the collapse of the structure under gravity load, this failure
pre-empts side-sway collapse. It should be noted that this building was not able to reach the target roof drift as
specified by Table 2.

8.8
@Seismicisolation
@Seismicisolation
Collapse Assessment of Non-Ductile, Retrofitted and
Ductile Reinforced Concrete Frames

Table 2–Target drift according to ASCE 41-13

RC type T1 (s) Sa(T1) for 2% in 50 yrs (g) C0 C1 C2  t (in)  t/L (%)


Existing 2.4 0.28 1.3 1 1 19.8 1.56%
Columns (Collapse Prevention) 2.2 0.31 1.3 1 1 19.6 1.54%
Retrofitt

Columns (Life Safety) 2.1 0.34 1.3 1 1 18.6 1.43%


Beams (Collapse Prevention) 3.2 0.18 1.3 1 1 23.7 1.87%
Walls (Collapse Prevention) 2.1 0.33 1.3 1 1 18.4 1.45%
Wall (Life Safety) 1.3 0.65 1.3 1 1 12.9 1.02%
1 in = 0.0254 m

Table 3– Results of static pushover analysis

RC type y/ht Vy (kips) d/ht Vd (kips) T Vd/Vy T1 (s)


Existing 0.42% 563.6 1.00% 714.0 2.40 1.27 2.4
Columns (Collapse Prevention) 0.35% 580.7 0.98% 735.7 2.77 1.27 2.2
Retrofitt

Columns (Life Safety) 0.36% 569.7 1.06% 809.3 2.97 1.42 2.1
Beams (Collapse Prevention) 0.29% 252.1 1.44% 348.7 5.04 1.38 3.2
Walls (Collapse Prevention) 0.40% 633.3 1.36% 834.0 3.38 1.32 2.1
Wall (Life Safety) 0.37% 1163.1 1.02% 1233.4 2.78 1.06 1.3
Ductile 0.16% 360.2 0.76% 455.2 4.66 1.26 1.7
y/ht: yield drift; Vy: shear yield strength; d/ht: drift at maximum base shear; Vd: maximum base shear; T: displacement ductility; 1 kips = 4.448
kN

ASCE 41 Target Roof Drift


First Column Yielding
0.1
First and
First Joint Yielding Second Shear
Base Shear/Seismic Weight [g]

Failure
0.08 in Column

0.06

0.04
First Beam Yielding
Non-Ductile Frame

0.02

0
0.00% 0.20% 0.40% 0.60% 0.80% 1.00% 1.20% 1.40% 1.60% 1.80%
Roof Drift [%]
Figure 3–Force-deformation response of the non-ductile frame
Figure 4a, 4c and 4e present the pushover curves for the retrofitted building using the strengthening, weakening and
adding techniques, respectively. As explained before, the strengthening and adding techniques were selected to
achieve both the CP and LS levels, while using the weakening technique, the structure could only achieve the CP
level. The ASCE 41-13 target drifts in these figures refer to targets for the retrofitted buildings. Figure 4b, 4d and 4f
present the component demands for the non-ductile columns, the critical component, for the strengthening,
weakening and adding technique, respectively. The acceptance criteria specified by ASCE 41-13 for the non-ductile
columns are also presented in these figures.

8.9
@Seismicisolation
@Seismicisolation
M. Baradaran Shoraka et al.
900 70000
ASCE 41 LS
Retrofitted Acceptance Criteria Retrofitted
800 Retrofitted Column
ASCE 41
Frame 60000 Frame CP
LS Level CP Level Acceptance
700 Criteria
Base Shear [kips]

Retrofitted 50000 Retrofitted


600

Moment [kips.in]
Frame Column
CP Level 40000
500 Retrofitted Frame LS
Level
400 30000
Non-Ductile Frame ASCE 41 Non-Ductile
300 Target Roof ASCE 41 CP Frame
Drift 20000 Acceptance
200 ASCE 41
Target Roof CP Level Criteria
UnRetrofitted
100 Drift 10000 Column
LS Level
0 0
0.00% 0.20% 0.40% 0.60% 0.80% 1.00% 1.20% 1.40% 1.60% 1.80% 0 0.02 0.04 0.06 0.08
Roof Drift [%] Plastic Rotation Angle [radian]
(a) Base shear vs. roof drift ratio of the non-ductile frame vs. retrofitted building (b) Force-deformation response of the first story column in the non-ductile
using the strengthening technique. frame vs. retrofitted building using the strengthening technique.
800 70000
ASCE 41 CP
700 ASCE 41 Acceptance
Target Roof 60000 Criteria
UnRetrofitted
Drift
600 Column
Base Shear [kips]

CP Level 50000

Moment [kips.in]
Retrofitted
500 Frame
Non-Ductile Frame
40000 CP Level
400
30000 Non-Ductile
300 Frame

Retrofitted 20000
200
Frame
CP Level
100 10000

0 0
0.00% 0.20% 0.40% 0.60% 0.80% 1.00% 1.20% 1.40% 1.60% 1.80% 2.00% 0 0.02 0.04 0.06 0.08
Roof Drift [%] Plastic Rotation Angle [radian]
(c) Base shear vs. roof drift ratio of the non-ductile frame vs. retrofitted building (d) Force-deformation response of the first story column in the non-ductile
using the weakening technique. frame vs. retrofitted building using the weakening technique.
Figure 4-Pushover results for the retrofitted buildings

8.10
@Seismicisolation
@Seismicisolation
Collapse Assessment of Non-Ductile, Retrofitted and
Ductile Reinforced Concrete Frames
1600 70000
ASCE 41 ASCE 41 LS ASCE 41 CP
Retrofitted Target Roof Acceptance Criteria Acceptance
1400 60000
Frame Drift UnRetrofitted Column Criteria
LS Level LS Level UnRetrofitted
1200 Column
Base Shear [kips]

Retrofitted
Frame 50000
Retrofitted

Moment [kips.in]
1000 CP Level Frame
40000 CP Level
800
30000
600 Non-Ductile
Frame
20000
400
Non-Ductile Frame ASCE 41
Target Roof Retrofitted
200 10000 Frame LS
Drift
CP Level Level
0 0
0.00% 0.20% 0.40% 0.60% 0.80% 1.00% 1.20% 1.40% 1.60% 0 0.02 0.04 0.06 0.08
Roof Drift [%] Plastic Rotation Angle [radian]
(e) Base shear vs. roof drift ratio of the non-ductile frame vs. retrofitted building (f) Force-deformation response of the first story column in the non-ductile
using the adding technique. frame vs. retrofitted building using the adding technique.

Figure 4–Pushover results for the retrofitted buildings (cnt’d)

8.11
@Seismicisolation
@Seismicisolation
M. Baradaran Shoraka et al.

GROUND MOTION SELECTION

To account for the uncertainties of the structural response under different ranges of earthquake shaking intensities, a
suite of earthquake records was used in the nonlinear dynamic analysis. In this study, incremental dynamic analysis
(IDA) proposed by Vamvatsikos and Cornell (2002) was chosen as the method to quantify the maximum structural
response as the earthquake shaking intensity increases.

Typically a site-specific probability seismic hazard analysis (PSHA) is used for the selection and scaling of the
ground motion. However, due to the fact that nonlinear analyses were performed by IDA, ground motions were not
scaled to match the specific target spectrum at each hazard level. Instead, the ground motions were gradually scaled
until global collapse takes place in the structure.

In this study, the ground motions used for the nonlinear dynamic analyses were selected from earthquakes with
moment magnitude between 6.5 to 7.6 and fault rupture distances between 10 to 45 km. A total of 39 pairs of ground
motion records selected by Haselton and Deierlein (2007) were adopted in this study. This set of ground motions
represent an expanded version of the far-field ground motion set which was used in the FEMA P695 document
(FEMA, 2009). Table 4 summarizes the characteristic of the selected ground motions.

DYNAMIC RESULTS, FRAGILITY ANALYSIS AND SYSTEM PERFORMANCE

Incremental dynamic analysis (IDA) was chosen as the method to quantify the structural response as the shaking
intensity increases. The outcome of the IDA was used to identify the earthquake shaking intensities when the
structure collapses. The spectral acceleration at the first mode period, Sa (T1), was recorded when the structure
collapsed and plotted using a log-normal distribution. Figure 5 illustrates the system collapse fragility curve for the
existing non-ductile building. This figure also demonstrates that the relationship between the Sa(T1) and probability
of collapse fits well with the log-normal distribution.

It should be noted that the ground motion records were selected and scaled without considering the distinctive
spectral shape of rare (extreme) ground motions, due to difficulties in selecting and scaling a different set of records
for a large set of buildings having a wide range of first mode periods. To account for the important impact of
spectral shape on collapse assessment, shown by Baker and Cornell (2006), the collapse predictions made using the
general set of ground motions and each of the building models were modified using a method proposed by Haselton
et al. (2011). The expected spectral shape of rare (large) California ground motions is accounted for through a
statistical parameter referred to as epsilon, which is a measure of the difference between the spectral acceleration of
a recorded ground motion and the median value predicted by a ground motion prediction equation. A target value of
=1.5 was used to approximately represent the expected spectral shape of severe ground motions that can lead to
collapse of code-conforming buildings in Los Angeles (Appendix B of FEMA P695, 2009).

As proposed in FEMA P695 (2009), the collapse fragilities were adjusted by a spectral shape factor, SSF, computed
using the following equation:

(2)

where  depends on building inelastic deformation capacity;  depends on the Seismic Design Category (SDC)
and is equal to 1.0 for SDC B/C, 1.5 for SDC D, and 1.2 for SDC E; and is the mean value of the Far-Field
record set listed in Table 4. In all cases, the buildings considered in the current study were designed for a single level
of high seismic ground motions representing Seismic Design Category (SDC) D buildings, therefore, and
are computed using the following equations:

.
. (3a)

. . (3b)

8.12
@Seismicisolation
@Seismicisolation
Collapse Assessment of Non-Ductile, Retrofitted and
Ductile Reinforced Concrete Frames

Table 5 summarizes the SSF values for the different buildings.

Table 4–Far Field Ground Motion Set (adapted from FEMA P695, (2009))

EQ Index Magnitude Year Event Fault Type Station Name Vs30 (m/s) Campell Distance (km)
1 6.7 1994 Northridge Blind thrust Beverly Hills - 14145 Mulhol 356 17.2
2 6.7 1994 Northridge Blind thrust Canyon Country - W Lost Cany 309 12.4
3 6.7 1994 Northridge Blind thrust LA - Saturn St 309 27
4 6.7 1994 Northridge Blind thrust Santa Monica City Hall 336 27
5 6.7 1994 Northridge Blind thrust Beverly Hills - 12520 Mulhol 546 18.4
6 7.1 1999 Duzce, Turkey Strike-slip Bolu 326 12.4
7 7.1 1999 Hector Mine Strike-slip Hector 685 12
8 6.5 1979 Imperial Valley Strike-slip Delta 275 22.5
9 6.5 1979 Imperial Valley Strike-slip El Centro Array 11 196 13.5
10 6.5 1979 Imperial Valley Strike-slip Calexico Fire Station 231 11.6
11 6.5 1979 Imperial Valley Strike-slip SAHOP Casa Flores 339 10.8
12 6.9 1995 Kobe, Japan Strike-slip Nishi-Akashi 609 25.2
13 6.9 1995 Kobe, Japan Strike-slip Shin-Osaka 256 28.5
14 6.9 1995 Kobe, Japan Strike-slip Kakogawa 312 3.2
15 6.9 1995 Kobe, Japan Strike-slip KJMA 312 95.8
16 7.5 1999 Kocaeli, Turkey Strike-slip Duzce 276 15.4
17 7.5 1999 Kocaeli, Turkey Strike-slip Arcelik 523 13.5
18 7.3 1992 Landers Strike-slip Yermo Fire Station 354 23.8
19 7.3 1992 Landers Strike-slip Coolwater 271 20
20 7.3 1992 Landers Strike-slip Joshua Tree 379 11.4
21 6.9 1989 Loma Prieta Strike-slip Capitola 289 35.5
22 6.9 1989 Loma Prieta Strike-slip Gilroy Array 3 350 12.8
23 6.9 1989 Loma Prieta Strike-slip Oakland - Outer Harbor Wharf 249 74.3
24 6.9 1989 Loma Prieta Strike-slip Hollister - South - Pine 371 27.9
25 6.9 1989 Loma Prieta Strike-slip Hollister City Hall 199 27.6
26 6.9 1989 Loma Prieta Strike-slip Hollister Diff. Array 216 24.8
27 7.4 1990 Manjil, Iran Strike-slip Abbar 724 13
28 6.5 1987 Superstition Hills Strike-slip El Centro Imp. Co. Cent 192 18.5
29 6.5 1987 Superstition Hills Strike-slip Poe Road (temp) 208 11.7
30 6.5 1987 Superstition Hills Strike-slip Westmorland Fire Sta 194 13.5
31 7 1992 Cape Mendocino Thrust Rio Dell Overpass - FF 312 14.3
32 7.6 1999 Chi-Chi, Taiwan Thrust CHY101 259 15.5
33 7.6 1999 Chi-Chi, Taiwan Thrust TCU045 705 26.8
34 7.6 1999 Chi-Chi, Taiwan Thrust TCU095 447 45.3
35 7.6 1999 Chi-Chi, Taiwan Thrust TCU070 401 24.4
36 7.6 1999 Chi-Chi, Taiwan Thrust WGK 259 15.4
37 7.6 1999 Chi-Chi, Taiwan Thrust CHY006 438 13.2
38 6.6 1971 San Fernando Thrust LA - Hollywood Stor FF 316 25.9
39 6.5 1976 Friuli, Italy Thrust (part blind) Tolmezzo 425 15.8
1 in = 0.0254 m

8.13
@Seismicisolation
@Seismicisolation
M. Baradaran Shoraka et al.

100%

90%

80%

70%

60%
P [Collapse]

50%

40%

30%

20%

10%

0%
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Sa (T1 = 2.36 s) [g]

Figure 5–Probability of collapse vs. Sa(T1) for non-ductile (unretrofitted) building, before spectral shape
adjustment

Figure 6 shows the fragility curves for all buildings and performance levels considered in this study adjusted for the
spectral shape. The figure clearly illustrates the significant variability in the probability of collapse of the retrofitted
structures, which fill the spectrum between the non-ductile and modern design buildings. It should be noted that the
retrofitted building using shear walls and upgraded to the LS performance level has a collapse performance identical
to the modern designed building (building similar to the 8-story non-ductile perimeter frame but designed using
modern building codes), while the weakening retrofit approach only provides a slight improvement on the
probability collapse over the existing building.

Table 5–Spectral shape factor

RC type b1 0  SSF
Existing 0.16 1.5 0 1.27
Columns (Collapse Prevention) 0.18 1.5 0 1.31
Retrofitt

Columns (Life Safety) 0.19 1.5 0 1.32


Beams (Collapse Prevention) 0.25 1.5 0 1.46
Walls (Collapse Prevention) 0.20 1.5 0 1.35
Wall (Life Safety) 0.18 1.5 0 1.31
Ductile 0.24 1.5 0 1.59

The probability of collapse at the MCE (2% in 50 yr return period) intensity (for each building) is shown with
circles in Figure 6. The circles indicate a variety of collapse ratios at this intensity level for the different retrofitting
techniques and performance objectives. The first mode period of the existing, retrofitted, and modern designed
buildings are all considered as being in the intermediate period range and therefore the dominant period has an
inverse relationship with the spectral acceleration. This variety in collapse performance can be compared by
normalizing the collapse fragility curves with the spectral acceleration at the MCE intensity, as shown in Figure 7.

8.14
@Seismicisolation
@Seismicisolation
Collapse Assessment of Non-Ductile, Retrofitted and
Ductile Reinforced Concrete Frames
100%

90%

80%

70%

60%
P [Collapse]

50%

40%
Existing
Columns (Collapse Prevention)
30%
Columns (Life Safety)
Beams (Collapse Prevention)
20% Walls (Collapse Prevention)
Walls (Life Safety)
10% Ductile
P[Collapse | Sa 2% in 50yrs]

0%
0 0.5 1 1.5 2 2.5
Sa (T1) [g]

Figure 6–Collapse fragility curves for different buildings

Figure 7 clearly indicates when the spectral acceleration is normalized to the MCE intensity, the collapse fragilities
for the different retrofitted buildings have a smaller variability compared to the original collapse fragility curve
(Figure 8). The acceptable probability of collapse at the MCE level for new buildings, defined as 10% by FEMA
P695, is also shown in Figure 7. The non-ductile existing building clearly does not meet this criterion and the
modern code-conforming structure clearly passes this criterion; while most of the retrofitted buildings are very close
to this criterion. The exception is the CP column retrofit which indicates a probability of collapse of 31% at the
MCE level. This poorer performance can be explained by the fact that the CP column retrofit only addresses the
vulnerability of the columns on the first two stories (see Figure 2). Since the 3rd story columns only barely satisfied
the CP acceptance criteria in the assessment of the CP column retrofit, it does not take a significant increase in
ground motion intensity to initiate shear and axial failure of the columns in this story. For 95% of the ground
motions considered, the analysis predicts collapse of the CP column retrofit due to shear and axial failure of the
columns on the 3rd story. In contrast, the CP beam retrofit changes the collapse mode of the structure from gravity-
load collapse due to column failure, to side-sway collapse of a strong-beam weak-column system. Such a system is
less sensitive to small increases in the ground motion intensity. The wall retrofit provides a reduction in drift
demands up the height of the building, and hence, protects all stories from failure.

8.15
@Seismicisolation
@Seismicisolation
M. Baradaran Shoraka et al.

100%

90%

80%

70%

60% Existing
P [Collapse]

Columns (Collapse Prevention)


50% Columns (Life Safety)
Beams (Collapse Prevention)
40% Walls (Collapse Prevention)
Walls (Life Safety)
30% Ductile

20% 10% acceptable Pcol|MCE specified in FEMA P695

10%

0%
0 1 2 3 4 5
Sa (T1)/Sa2%in50yrs

Figure 7–Collapse fragility curves normalized by Sa 2% in 50 yrs

Table 6–Collapse performance metrics


Sa(T1) for P[col] P[col]
RC type MedianSa,col(T1)  LN[Sa,col(T1)] collapse
2% in 50 yrs for Sa(T 1 ) in 50 yrs
Existing 0.28 80.8% 0.19 0.47 1.5E-02 52%
Columns (Collapse Prevention) 0.31 31.0% 0.38 0.47 3.6E-03 16%
Retrofitt

Columns (Life Safety) 0.34 16.5% 0.54 0.50 1.8E-03 8%


Beams (Collapse Prevention) 0.18 10.9% 0.30 0.42 2.3E-03 11%
Walls (Collapse Prevention) 0.33 10.1% 0.52 0.38 1.4E-03 7%
Wall (Life Safety) 0.65 16.6% 0.97 0.41 1.5E-03 7%
Ductile 0.47 2.6% 1 0.40 8.0E-04 4%
MedianSa,col(T): median collapse capacity of structural model, in terms of spectral acceleration at the first-mode period, Sa(T1); LN[Sa,col(T)]: standard
deviation of collapse capacity of structural model, in terms of spectral acceleration at the first-mode period, Sa(T1); P[col] for Sa(T1): probability
of collapse at the first-mode period, Sa(T1); collapse: mean annual frequency of collapse, i.e. collapses per year; P[col] in 50 yrs: probability of
collapse in 50 years (1-e-50)

Different collapse performance metrics are reported in Table 6. The predicted median collapse capacity,
MedianSacol(T1), of the retrofitted structures is approximately 1.5 to 5 times larger than the existing non-ductile
building. The probability of collapse in 50 years shows a higher difference, this performance metric decreases from
52% for the non-ductile building down to a range of 7% - 11% for most of the retrofitted buildings, which is
comparable to the 4% probability of collapse for the modern design structure.

It should be noted that the probability of collapse in 50 years is the preferred collapse metric for comparison to the
probability of collapse at the MCE level. There are two main points for this preference. First of all, the probability of
collapse in 50 years, which is obtained by combining the collapse fragility curve with the hazard curve, covers a
larger range of hazard levels compared to the MCE level. Secondly, the probability of collapse at the MCE level is
influenced by the change in building period as well as change in the fragility curve. As a result, the probability of
collapse in 50 is used for the comparison of the different buildings.

An overall summary of the three sets of buildings is presented in Table 7. This table demonstrates an estimate of the
change in the material required to rehabilitate the existing non-ductile building. Modifying the beams would

8.16
@Seismicisolation
@Seismicisolation
Collapse Assessment of Non-Ductile, Retrofitted and
Ductile Reinforced Concrete Frames

probably cost the least to retrofit (with the assumption that the labor cost to perform all three techniques would
approximately be the same) but has less benefit in terms of the seismic collapse safety. On the other hand, adding
shear walls and retrofitting to the CP and LS performance levels would cost the most but has the lowest probability
of collapse.
Table 7–System performance of the three set of buildings

ASCE 41 (Reinforced
Material List
Concrete Columns) Collapse
Retrofit Performance
Building Type Margin (P[Col]
Technique Level Change (compared to
Acceptance Weight [kips] in 50 yrs)
Demand existing building)
Criteria
Concrete Steel Concrete Steel
Existing (Non-
- - 0.043 0.07 786 48 - - 52%
Ductile)
Enhance Collapse
0.06 0.05 803 49 2% 2% 16%
Columns Prevention
Enhance
Life Safety 0.035 0.03 836 50 6% 4% 8%
Columns
Weaken Collapse
Retrofitt 0.043 0.03 786 48 0% 0% 11%
Beams Prevention
Add Collapse
0.043 0.032 839 50 7% 6% 7%
ShearWall Prevention
Add
Life Safety 0.0161 0.012 868 56 10% 18% 7%
ShearWall
1 kips = 4.448 kN

CONCLUSIONS

Advancements in nonlinear dynamic analysis, seismic hazard analysis, and performance-based earthquake
engineering are enabling more scientific assessment of structural collapse risk. The main objective of this study is to
quantify and compare the seismic safety of non-ductile, retrofitted, and ductile RC frame buildings. At present, the
general assumption when retrofitting a non-ductile building according to ASCE 41-13 is that their seismic
performance is enhanced such that it can reach the performance corresponding to ductile buildings designed based
on current seismic codes. However, as shown in this study, different retrofitting measures do not exhibit the same
collapse performance metrics as modern ductile buildings.

Based on the evaluation of the eight-story RC frame in Los Angeles considered in this study, the following
observations are made:

 The considered retrofit schemes (column strengthening, beam weakening, and adding shear wall) using
ASCE 41-13 acceptance criteria provided an intermediate level of collapse performance, falling between
that achieved by a non-ductile frame and that achieved by a modern code-conforming ductile RC moment
frame.

 The study finds that retrofitting scheme of modifying the columns up to only the CP level has the least
beneficial effect regarding the seismic collapse safety and conversely adding a shear wall has the highest
improvement in collapse performance. The column CP retrofit is vulnerable to collapse due to failure of
unretrofitted columns above the retrofitted stories. Modification of collapse mode due to beam weakening
or reduction of drifts at all stories with an added shear wall resulted in better collapse performance metrics.

 The probability of collapse in 50 years decreases from 52% (for the non-ductile building) down to a range
of 7% - 16% for the retrofitted buildings; higher than the 4% probability of collapse in 50 years for the
modern code-conforming structure.

8.17
@Seismicisolation
@Seismicisolation
M. Baradaran Shoraka et al.

 The predicted median collapse capacities of the retrofitted structures are approximately 1.5 to 5 times larger
than the existing non-ductile building depending on the retrofit technique and the selected performance
level.

The above observations are valid for the building considered in this study. Future research is required to determine
if these observation hold for other structural systems, retrofit techniques and seismic hazard.

ACKNOWLEDGMENTS

This work is supported in part by the Natural Sciences and Engineering Research Council of Canada through the
Canadian Seismic Research Network. Any opinions, findings, and conclusion or recommendations expressed in this
work are those of the authors and do not reflect those of the organizations or individuals noted here.

REFERENCES

American Concrete Institute, 2005. Building Code Requirements for Structural Concrete (ACI 318). American
Concrete Institute, Farmington Hills, MI.
American Society of Civil Engineers, 2013. Seismic Evaluation and Retrofit of Existing Buildings, American
Society of Civil Engineers (ASCE/SEI 41-13), pre-publication edition for public comment and final review.
American Society of Civil Engineers. 2005. ASCE 7-05: Minimum Design Loads for Buildings and Other
Structures, Reston, VA.
Anagnos, T., Comerio, M.C., Goulet, C., Steele, J. and J. P. Stewart, 2010. Development of a concrete building
inventory: Los Angeles case study for the analysis of collapse risk. Proc. 9th US National & 10th Canadian Conf.
on Earthquake Eng., EERI and Canadian Assoc. for Earthquake Eng., July 25-29, 2010, Paper No. 48.
Baker J. W. and Cornell C. A., 2006. Spectral shape, epsilon and record selection, Earthquake Engineering &
Structural Dynamics 35, 1077-1095.
Baradaran Shoraka, M. 2013. Collapse assessment of non-ductile reinforced concrete buildings,PhD thesis,
Civil Engineering Department, University of British Columbia, BC, Canada.
Baradaran Shoraka, M., Yang, T. Y., and Elwood, K. J., 2013. Seismic loss estimation of non-ductile reinforced
concrete buildings, Earthquake Engineering and Structural Dynamics 42, 297–310.
Bazzurro, P., and Cornell, C. A., 1994. Seismic hazard analysis of nonlinear structures. I: Methodology, Journal
of Structural Engineering 120, 3320.
Bazzurro P. and Cornell C. A., 1999. Disaggregation of seismic hazard, Bulletin of the Seismological Society of
America 89, 501-520.
Elwood K. J., Matamoros A. B., Wallace J. W., Lehman D. E., Heintz J. A., Mitchell A. D., Moore M. A.,
Valley M. T., Lowes L. N., Comartin C. D. and Moehle J. P., 2007. Update to ASCE 41 concrete provisions,
Earthquake Spectra 23, 493-523.
Elwood K. J. and Moehle J. P., 2008. Dynamic collapse analysis for a reinforced concrete frame sustaining
shear and axial failures, Earthquake Engineering & Structural Dynamics 37, 991-1012.
Elwood K. J., 2004. Modelling failures in existing reinforced concrete columns, Canadian Journal of Civil
Engineering 31, 846-859.
Elwood, K.J. 2002. Shake table tests and analytical studies on the gravity load collapse of reinforced concrete
frames. Ph.D. Dissertation, Department of Civil and Environmental Engineering, University of California, Berkeley,
CA.
Federal Emergency Management Agency (FEMA P695A), 2009. Recommended Methodology for
Quantification of Building System Performance and Response Parameters, Washington, D.C.

8.18
@Seismicisolation
@Seismicisolation
Collapse Assessment of Non-Ductile, Retrofitted and
Ductile Reinforced Concrete Frames

Federal Emergency Management Agency (FEMA 547), 2006. Techniques for the Seismic Rehabilitation of
Existing Buildings, Washington, D.C., 571 pp.
Federal Emergency Management Agency (FEMA-356), 2000. Pre-Standard and Commentary for the Seismic
Rehabilitation of Buildings. Federal Emergency Management Agency, Washington, D.C.
Haselton, C. B., Liel, A. B., Taylor Lange, S., and Deierlein, G. G., 2008. Beam-Column Element Model
Calibrated for Predicting Flexural Response Leading to Global Collapse of RC Frame Buildings. Pacific Earthquake
Engineering Research Center, Berkeley, CA
Haselton, C. B., Liel, A. B., Deierlein, G. G., Dean, B. S., Chou, J. H., 2011. Seismic Collapse Safety of
Reinforced Concrete Buildings: I. Assessment of Ductile Moment Frames, Journal of Structural Engineering,
137(4), pp. 481-491.
Haselton, C. B., Baker, J. W., Liel, A. B. and Deierlein, G. G, 2011. Accounting for Ground Motion Spectral
Shape Characteristics in Structural Collapse Through an Adjustment for Epsilon, Journal of Structural Engineering,
137(3), pp. 332-344.
Ibarra L. F., Medina R. A. and Krawinkler H., 2005. Hysteretic models that incorporate strength and stiffness
deterioration, Earthquake Engineering & Structural Dynamics 34, 1489-1511.
ICC 2003. International Building Code, International Code Council, Washington, D.C.
Kam, W. Y., and Pampanin, S., 2009. Experimental and numerical validation of selective weakening retrofit for
existing non-ductile RC frames. Proc. Improving the Seismic Performance of Existing Buildings and Other
Structures.
Liel, A.B., 2008. Assessing the collapse risk of California’s existing reinforced concrete frame structures:
Metrics for seismic safety decisions. Ph.D. thesis, Stanford University. United States.
Liel, A. B. and Deierlein, G. G., 2013. Cost-Benefit Evaluation of Seismic Mitigation Alternatives for Older
Reinforced Concrete Frame Buildings, Earthquake Spectra.
Lowes L. N. and Altoontash A., 2003. Modeling reinforced-concrete beam-column joints subjected to cyclic
loading. Journal of Structural Engineering 129, 1686-1697.
Pacific Earthquake Engineering Research Center, 2009. Open system for earthquake engineering simulation
_OpenSees_ framework-Version 2.1.0.
Pampanin, S., 2005. Emerging Solutions for High Seismic Performance of Precast/Prestressed Concrete
Buildings. Journal of Advanced Concrete Technology (ACT), 3(2), 202-223.
Vamvatsikos D. and Cornell A., 2002. Incremental dynamic analysis, Earthquake Engineering and Structural
Dynamics 31, 491-514.
Viti, S., Cimellaro, G. P., and Reinhorn, A. M., 2006. Retrofit of a hospital through strength reduction and
enhanced damping. Smart Structures and Systems, 2(4), 339-355.
Yavari S., Elwood K. J., Lin S., Wu C., Hwang S. and Moehle J. P., 2009a. Experimental study on dynamic
behaviour of multi-story reinforced concrete frames with non-seismic detailing. Proc. Improving the Seismic
Performance of Existing Buildings and Other Structures, pp. 489-499. 489-499.
Yavari S., Elwood K. J. and Wu C., 2009b. Collapse of a nonductile concrete frame: Evaluation of analytical
models, Earthquake Engineering and Structural Dynamics 38, 225-241.

8.19
@Seismicisolation
@Seismicisolation
M. Baradaran Shoraka et al.

8.20
@Seismicisolation
@Seismicisolation
SP-297—9

Seismic Assessment and Retrofit of Existing RC Buildings:


Case Studies from Degenkolb Engineers

Insung Kim, Degenkolb Engineers, Garrett Hagen, Degenkolb Engineers

Synopsis: Case studies on seismic assessment and rehabilitation of reinforced concrete buildings are discussed based
on the projects in which Degenkolb Engineers has been involved in the past 5 years. Design, analysis and challenges
are discussed to present applications of ASCE 31-03, Seismic Evaluation of Existing Buildings and ASCE 41-06,
Seismic Rehabilitation of Existing Buildings.

Keywords: reinforced concrete; seismic evaluation; seismic retrofit, repair, rehabilitation, non-ductile concrete structure

ACI member InSung Kim is a Structural Engineer at Degenkolb Engineers, San Francisco, CA. He received his BS
from Yonsei University, Seoul, South Korea; and his MS and PhD from the University of Texas at Austin, Austin,
TX. He is a member of ACI Committees 369, Seismic Repair and Rehabilitation; 374, Performance-Based Seismic
Design of Concrete Buildings; and 440, Fiber-Reinforced Polymer Reinforcement. His research interests include
seismic design and rehabilitation of reinforced concrete structures.

Garrett Hagen also works with Degenkolb Engineers, in their Los Angeles office. He graduated with his BS and
MS from California Polytechnic University, San Luis Obispo, where his graduate research involved performance-
based analysis of concrete structural walls. He is a member of ACI Committees 369 and 374. He actively
contributes to research regarding the modeling, behavior, and design of reinforced concrete structures.

INTRODUCTION

Degenkolb Engineers is a structural engineering firm specializing in seismic engineering and has five offices
along the West Coast of the USA. The firm has a variety of experience in evaluation and retrofit of building
structures in the high seismic regions of the USA. In this paper, case studies of their seismic evaluation and retrofit
projects of RC (reinforced concrete) buildings in the past 5 years are presented. The projects utilized ASCE 31-03,
Seismic Evaluation of Existing Buildings and ASCE 41-06, Seismic Rehabilitation of Existing Buildings, and its
supplement. Deficiencies of the RC buildings were identified, and retrofit schemes to mitigate the deficiencies were
developed, based on ASCE 31-03 and ASCE 41-06. In this paper, typical deficiencies and retrofit schemes are
discussed along with some of the challenges in the application of the documents. In addition, the areas requiring
further improvement in the concrete chapter of ASCE 41-13, Seismic Evaluation and Retrofit of Existing Buildings,
which is the combined document of the current ASCE 31 and 41 and ACI 369R-11 Guide for Seismic Rehabilitation
of Existing Concrete Frame Buildings and Commentary, are discussed.

9.1
@Seismicisolation
@Seismicisolation
I. Kim and G. Hagen

SEISMIC EVALUATION AND RETROFIT CASE STUDIES

Building Descriptions

Seismic evaluation and retrofit of ten concrete buildings, located in California and built between the 1940’s and
1970’s, are reported in this paper. A summary of the building descriptions and their major deficiencies and retrofit
schemes is given in Table 1. The number of stories of the buildings ranges from 1 story to 10 stories, which can be
considered as low to mid-rise buildings. Most of the buildings had shear walls for their existing lateral force
resisting system (LFRS), and concrete frames were typically used to resist gravity force and/or as supplementary
LFRS. Various types of slabs were used as the diaphragms of the structures. One- or two-way RC slabs were used
throughout the years, and some of the buildings built after the 1960’s had post-tensioned slabs as their diaphragms.
Because the State of California requires mandatory seismic upgrade of hospital buildings, hospital buildings are
evaluated and retrofitted more often than other types of building occupancies. Six out of the ten buildings addressed
in this paper are hospitals.
Upon the request from the building owners, the information which could be used to identify the buildings is not
disclosed.

Major Deficiencies

Some of the classic deficiencies identified in ASCE 31-03 were observed in the building evaluations. Global
issues, such as torsional irregularities, vertical irregularities, potential pounding with adjacent structures, and weak
or soft stories, were observed in the various buildings. Specific issues, such as non-ductile concrete frames with
shear critical captive columns, large openings in shear walls, and unconfined boundary elements, were found as well.
Severe vertical irregularities were typically observed for the buildings with podiums (Building 1, 2, 4 and 8).
An example of a building with a podium is shown in Fig. 1. At the top of podium shown in Fig. 1 where the setback
occurred, the strength and stiffness of the building changed abruptly, requiring large forces to be transferred through
podium diaphragms and sometimes creating significant stress concentrations. For some of the buildings, changes in
LFRS occurred in the transition from above the podium level to below.

Floor plan above podium

Floor plan at podium level

Figure 1 — Vertical discontinuity, setback at top of podium

Captive columns were typically observed in the perimeter frames with spandrel beams (Building 2 and 8, Fig.
2). The spandrels framed into the columns eccentrically, and the eccentricity combined with high shear force in the
captive columns created the potential for damage concentration in these frames (Fig. 2). Eccentric beam column
connections were common not only in the spandrel frames but also in moment frames located in the building
perimeters.

9.2
@Seismicisolation
@Seismicisolation
Seismic Assessment and Retrofit of Existing RC Buildings:
Case Studies from Degenkolb Engineers

Figure 2 — “Captive” columns (elevation view) and eccentrically connected spandrel and column (plan view)

Diaphragm capacity and diaphragm connections to the LFRS and/or gravity force resisting system were
problematic in most of the buildings evaluated. Reinforcement in the slab of the existing buildings was typically
designed for gravity force only, so the amount of existing reinforcement which could be utilized for seismic load
transfer was typically not sufficient. The connection from diaphragm to the vertical lateral or gravity force resisting
system had deficiencies as well. For example, Fig. 3 represents a case in which the dowels connecting the slab to
shear wall were not sufficient and were not adequately anchored to develop yield. Large diaphragm openings
observed in Buildings 9 and 10 created a weak link for the lateral force transfer. PT slabs in Building 4 and 6 did not
contain conventional reinforcement so the diaphragm did not have enough capacity to transfer the seismic load, and
the connections to the columns were not adequate. The failure of the diaphragm and its connections may be one of
the main causes for collapse of a structure as observed in the CTV Building and Pyne Gould Building after the 2011
Christchurch Earthquake (C. Hyland and A, Smith 2012, Canterbury Earthquake Royal Commission, 2012).
Rehabilitation schemes for these issues generally consisted of installing additional collectors and/or adding new
frames or shear walls to reduce demand in the diaphragm.
High tensile or compressive force in the vertical member located close to the end of the LFRS, due to global
building overturning, was one of the common deficiencies observed in the mid-rise buildings. The continuous deep
(stiff) spandrel beams in the perimeter of the building created high uplift and compressive demands in the walls and
columns at the building corners. For building analysis, precautions were taken to accurately capture stiffness at
building transitions, as it is possible for axial forces to be overestimated when the building base is numerically
modeled with restraints to vertical transition (e.g. fixed or pinned supports).
Many of the deficiencies observed in the buildings are well summarized in Program Plan for the Development
of Collapse Assessment and Mitigation Strategies for Existing Reinforced Concrete Buildings (NEHRP, 2010).

9.3
@Seismicisolation
@Seismicisolation
I. Kim and G. Hagen

Figure 3 — Diaphragm to wall connection

Retrofit Schemes

Addition of New Vertical LFRS


Addition of new vertical LFRS was required for all but one of the ten buildings to retrofit the structures. Shear
walls and conventional braced frames or buckling restrained braced frames (BRB) were the most common vertical
LFRS added to the buildings. Depending on the accessibility to the building during construction, the new system
was added to either the interior or exterior of the building. The installation of new vertical LFRS required new
foundations, which often controlled feasibility of the new system. The connections between the new vertical LFRS
system and existing horizontal LFRS (diaphragm and collectors) were critical to utilize the new system. Addition of
new vertical LFRS often required new collectors as well. An example of installation of new shear walls and their
foundation is shown in Fig. 4.

Strengthening of Existing Vertical LFRS


Shear walls were strengthened with shotcrete overlay or Fiber Reinforced Polymer (FRP) overlay (Building 2,
5 and 10). Installation of FRP on an existing shear wall is shown in Fig. 5. Non-ductile columns in moment frames
were strengthened with FRP jacketing to improve ductility and shear strength (Building 2, Fig. 5). In Building 5, the
openings in the shear walls were infilled with concrete to create solid shear walls.

Addition and Strengthening of Horizontal LFRS


Retrofit of all ten buildings included strengthening of horizontal LFRS’s (e.g. diaphragm and collectors).
Connections between diaphragms and vertical LFRS’s were often the weak links for seismic load transfer. In order
to improve the load transfer between vertical and horizontal LFRS’s, connections needed to be strengthened and/or
collectors needed to be added or strengthened. Collector strengthening was achieved by addition of a concrete or
steel member (Fig. 6).
Because it is not easy to improve concrete diaphragm shear capacity, if shear demand on the diaphragm is
excessive, addition of new vertical LFRS is generally considered as a mitigation method. The shear capacity of bare
metal deck or plywood sheathing diaphragm can be strengthened by improving connections between the deck or
sheathing and the roof or floor framing (Building 7).

9.4
@Seismicisolation
@Seismicisolation
Seismic Assessment and Retrofit of Existing RC Buildings:
Case Studies from Degenkolb Engineers

Strengthening of Existing Gravity Force Resisting System


To achieve drift compatibility between the LFRS and gravity force resisting system, the deformation capacity
of the gravity system for some buildings needed to be improved. A common technique to improve deformation
capacity of existing gravity systems was FRP jacketing of the gravity RC columns (Building 1 and 8, Fig. 5). Where
the slab-column connection did not have adequate deformation capacity, the existing connections were strengthened
(Building 8) or supplementary connections were added to support the slab after the existing connections failed
during an earthquake (Building 2).

Figure 4 — Installation of new shear wall and foundation.

Figure 5 — FRP strengthened column and wall.

9.5
@Seismicisolation
@Seismicisolation
I. Kim and G. Hagen

Figure 6 — Collector strengthening (plan view).

Removal of Existing Elements


In certain cases, removal of existing elements is the most efficient way of improving seismic performance of
the building. In Building 2, the existing stair towers and infilled concrete walls were removed to assure the intended
seismic load path of the building. In addition, vertical slots were added at the ends of some of the perimeter
spandrels to reduce lateral force attracted by them.

Connection Between Two Buildings


To prevent pounding of adjacent buildings, existing seismic joints were made larger (Buildings 1 and 2) or the
two buildings were connected so that for the buildings would share LFRS’s without pounding.

Analysis Techniques
After Tier 1 evaluation using the checklists in ASCE 31-03, 3D numerical models are typically created for
further evaluation or retrofit design. The initial analysis generally begins with linear static and dynamic analysis, but
nonlinear analysis may be adopted depending on the characteristics of the project. Depending on the analysis
techniques selected, different structural analysis programs may be used. 3D numerical models of a building
developed using different analysis programs are shown in Fig. 7. Typically, after an engineer reports that the
building does not comply with Tier 1 criteria to the owner, the engineer provides the evaluation results with
recommendations for detailed evaluation of the Tier 1 deficiencies, to determine whether the deficiencies in fact
require mitigation through retrofit.
Based on review of the existing drawings and completion of the checklists during the Tier 1 evaluation and
linear analysis, the engineer usually can identify the critical deficiencies which jeopardize building performance,
The engineer can then provide several options of retrofit schemes to the building owner. Some retrofit schemes may
require advanced nonlinear analysis to justify the design. In those cases, the added cost for advanced analysis has the
potential for making the most cost-effective retrofit scheme for the project possible. It is important that engineers
understand they should not use nonlinear analysis with the intent of justifying performance of existing non-ductile
components which are critical elements for resisting gravity or seismic load and may experience significant
deformation during earthquakes, even if they show acceptable performance in the nonlinear analysis. More complex
numerical models do not always result in more reliable or accurate results and nonlinear internal load distribution
calculated in a model can be very different from the load distribution in a real building. Engineers should be more
cautious about uncertainty of seismic loads and approximate natures of analysis results when their decisions are

9.6
@Seismicisolation
@Seismicisolation
Seismic Assessment and Retrofit of Existing RC Buildings:
Case Studies from Degenkolb Engineers

based on the insights obtained from nonlinear analysis. Of course, nonlinear analysis can be utilized to resolve
unrealistic concentrations of internal forces shown in linear analysis by allowing redistribution when the global
structural system of a building has adequate strength, deformation capacity and redundancy. However, the best use
of nonlinear analysis may be the verification of the retrofit design, employing the capacity-based design concept and
ductile details for load redistribution.
A good example for application of nonlinear analysis is the use of foundation springs. The springs allow the
reduction of unrealistic high axial forces due to global building overturning (Building 2, 5 and 8). The analysis
should be performed with upper and lower bound spring stiffness as described in ASCE 41 in order to capture
uncertainty of the foundation stiffness.
Nonlinear analysis is also effective for buildings in which deformation compatibility is of primary concern, in
which case localized deformations are critical.

Figure 7 — 3D numerical models developed using ETABS (left) and PERFORM 3D (right)

9.7
@Seismicisolation
@Seismicisolation
I. Kim and G. Hagen

Table 1 — Summary of the case study buildings


Building Description
Year of construction
Analysis
Number of stories Major Deficiencies Major Retrofit Schemes
Techniques
Building Lateral force resisting system
Building use
Location
New exterior buttress walls with
1971
Torsional irregularity footing
10 stories + 2 story basement
Vertical irregularity New BRB Linear static and
 8 story moment frame tower
Setback at podium Collector strengthening with steel dynamic analysis
Bldg. 1 and 2 story shear wall podium
Eccentric beam column joints plate Nonlinear static and
and 2 story shear wall basement
Column deformation capacity FRP strengthening of gravity dynamic analysis
Hospital
columns
Southern California
Seismic joint modification
Shotcrete overlay over existing wall
Weak story New external concrete walls with
new foundation
Soft story
1963 Shotcrete overlay over spandrel
Deflection incompatibility Linear static and
9 stories beam for collector
Short "captive" columns dynamic analysis
Bldg. 2 Shear walls and moment frames FRP strengthening of columns and
Strong columns/weak beam s Nonlinear static and
Hospital beams and shear walls
Non ductile concrete frames dynamic analysis
Southern California Removal of existing infill
Columns bar splices
Removal and relocation of stair
High drift towers
Seismic joint modification

1968
Linear static and
3 stories New RC walls and columns with
Bldg. 3 Vertical discontinuity dynamic analysis
Shear walls and moment frames new foundation
Diaphragm strength Nonlinear static
School New collectors
analysis
Northern California
Vertical discontinuity
Torsional irregularity
1968 New RC walls and boundary
Diaphragm strength Linear static and
3 stories elements
Shear wall strength dynamic analysis
Bldg. 4 Shear walls and moment frames Steel plate collector
Deformation capacity of post Nonlinear static
School Steel channel support at slab
tensioned slab with column (no analysis
Northern California column joints
continuous reinforcement
through column)
New shear wall
New structural slab-on-grade at the
first floor level to connect new wall
Weak story to basement wall
1949 (4 floors were added in Vertical discontinuities RC infill of wall openings
1965) Weak shear walls and coupling Strengthening of existing concrete
Linear static and
3 stories + basement original beam wall and spandrel
dynamic analysis
7 stories + basement after 1965 Torsional irregularity New below-grade concrete shear
Bldg. 5 Nonlinear static
Shear walls Non ductile concrete frames walls to strengthen load path at sub-
analysis
Hospital High drift basement foundation level
Northern California New structural connection between
basement floor slab and perimeter
concrete walls
Connecting two buildings
Chimney (stack) brace
1969
3 stories Non-ductile columns
New exterior BRB with foundation Linear static
Bldg. 6 Moment frames Two-way PT slab without
Concrete collectors analysis
Office conventional reinforcement
Northern California

9.8
@Seismicisolation
@Seismicisolation
Seismic Assessment and Retrofit of Existing RC Buildings:
Case Studies from Degenkolb Engineers

Table 1, continued — Summary of the case study buildings


Building Description
Year of construction
Building number of stories Major Deficiencies Major Retrofit Schemes Analysis Method
Lateral force resisting system
Building use
Location

Steel HSS strongbacks


Inter-panel connection using
1968 anchors or through bolts with
Out of plane bending of wall
steel plates
1 story with mezzanine Connections between precast
Above and below roof steel HSS Linear static and
Tiltup concrete shear walls and wall panels
Bldg. 7 and channel collectors dynamic analysis
a wood roof diaphragm Footings on liquefiable soil
Re-nailing of roof diaphragm
Warehouse Weak collector and diaphragm
Additional braced frames and
Northern California
shear walls to cut down roof
diaphragm span

1948
Lack of diaphragm strength in New shear wall with footing Linear static and
4 stories
Transverse direction Slab-column connection dynamic analysis
Bldg. 8 Pier-spandrel shear walls
Weak story strengthening Nonlinear static
Hospital
Non-ductile concrete frames FRP column jacketing analysis
Northern California
1968
8 stories Vertical discontinuity New shear walls with axial and
Linear dynamic
Bldg. 9 Shear walls Inadequate load path near lateral force-resisting piles
analysis
Hospital diaphragm openings Steel plate collectors
Southern California
Vertical discontinuity
1966 Shear wall strength
FRP strengthening of walls
6 stories Diaphragm strength
FRP column jacketing Linear dynamic
Bldg 10 Shear walls Inadequate load path near
Addition of and strengthening of analysis
Hospital diaphragm openings
collector elements
Southern California Column strength under
discontinuous shear walls

9.9
@Seismicisolation
@Seismicisolation
I. Kim and G. Hagen

AREAS IN ASCE 41 CONCRETE CHAPTER FOR FURTHER IMPROVEMENT (IMPROVEMENT IN


GUIDELINES AND RESEARCH NEEDS)

The ASCE 31-03 and 41-06 standards have been combined into a new addition of ASCE 41-13, Seismic
Evaluation and Retrofit of Existing Buildings, and most of the updates in the concrete chapter of ASCE 41-13 were
based on ACI 369R-11, Guide for Seismic Rehabilitation of Existing Concrete Frame Buildings and Commentary.
The following section discusses areas requiring further improvement in ASCE 41 and ACI 369R based on the
experience in seismic evaluation and retrofit of the ten buildings in this study.

Modeling Parameters and Acceptance Criteria of Strengthened Members


The concrete chapters in ASCE 41-06 and ACI 369R-11 provide modeling parameters (MP’s) and acceptance
criteria (AC) of deformation controlled components. The MP’s and AC are determined based on detailing conditions
and loading conditions of conventional concrete members. However, when the existing concrete members are
strengthened, they are not always strengthened with conventional details of concrete members. The current ASCE
41 chapter does not provide clear guidelines for methods of selecting MPs and AC of the strengthened members.
The following are some of the example cases for which more guidelines on strengthened members are required
based on the case studies of the buildings.

 When a RC column, shear wall, or diaphragm element is strengthened with fiber reinforced polymer (FRP),
the guides for selection of MPs and AC for the strengthened members are required (Fig. 5).
 If a RC column or collector is strengthened with steel plates, engineers need not only MP’s and AC
information, but also guides for calculating stiffness of the strengthened member (Fig. 6).
 If a shear wall has a concrete overlay on one side or an existing column is jacketed by additional concrete
and transverse reinforcement, guides for selecting the MP’s and AC are required.

Further guides for modeling and selection of AC and detailing requirements of strengthened members to
achieve specific performance (i.e. deformation capacity) are necessary for more reliable retrofit. Guides for
strengthened members with techniques commonly used in practice would be helpful for engineers to understand and
adopt these techniques.

Diaphragm and Collectors


As described earlier, retrofit of all ten buildings included the strengthening of diaphragms or collectors. The
horizontal LFRS is as important as the vertical LFRS for transferring and resisting seismic force to achieve intended
performance. However, Chapter 6.10 and 6.11 of ASCE 41-06 (Cast-in-Place and Precast Diaphragm) reference the
shear wall and frame sections for the diaphragm and collector, respectably. Because in many cases the retrofit design
is governed by diaphragm capacity, more specific guidelines for considering the behavior of the horizontal LFRS are
desirable. In addition, Chapter 6.10.2.4 requires that diaphragm connections must be considered force-controlled,
but definitions for diaphragm connections might need more clarification (e.g. diaphragm to collector connection;
diaphragm or collector to vertical LRFS connection) so engineers can include evaluation of such connections in their
scope of work per the standard, and the analysis criteria remain consistent for different engineers For example, the
current requirement may therefore be interpreted that the shear friction connection between the diaphragm and the
vertical system needs to be force-controlled, which likely becomes a weak link in most concrete buildings with
concrete diaphragms because the previous building codes did not provide detailing requirements for such
connections. In this case, the connection shown in Fig. 3 should be repaired in most of cases to develop ductility of
horizontal and vertical LFRS’s by strengthening the connection or providing collector elements so diaphragm loads
can be transferred to the vertical LFRS through tension or compression in the collectors. This type of strengthening
typically has major cost impact on a retrofit project, so more specific requirements in the standard may help
engineers to determine the scope of their project.

Criteria for Gravity System


Current ASCE 41 concrete chapters and ACI 369R-11 do not provide modeling and acceptance criteria that are
specifically applicable to systems considered to resist gravity force only. For examples, specific guidelines for
evaluating the acceptable deformation of gravity frames with beams containing discontinuous longitudinal

9.10
@Seismicisolation
@Seismicisolation
Seismic Assessment and Retrofit of Existing RC Buildings:
Case Studies from Degenkolb Engineers

reinforcement in the beam-column joint, or deformation capacities of perimeter spandrel frames with captive
columns, may be helpful.
 
Nonlinear Response History Analysis
The current provisions for nonlinear analysis in the Concrete Chapter of ASCE 41 are mainly applicable to the
nonlinear static (push-over) analysis. Because nonlinear dynamic analysis is commonly used in practice and is
sometimes required by the jurisdiction, more guidelines for nonlinear dynamic analysis are needed. The hysteresis
information of a member (i.e., energy dissipation during cyclic behavior) and stiffness modification for the nonlinear
dynamic procedure might be critical information required by engineers to apply provisions for the nonlinear
dynamic analysis. Recommendations for magnitude of damping (both within and outside of the elastic range) are
needed to ensure engineers are not over-accounting for damping when explicitly modeling hysteretic behavior.
Additionally, in ASCE 41, there are currently not sufficient guidelines for cyclic degradation.

CONCLUDING REMARKS

Based on experiences of Degenkolb Engineers in seismic evaluation and retrofit of concrete buildings in the
past 5 years, major structural deficiencies, strengthening schemes, and analysis techniques were discussed in this
paper. The major structural deficiencies observed in the buildings corresponded to those recognized throughout past
earthquakes. To mitigate those deficiencies, new vertical and/or horizontal LFRS were added or the existing systems
were strengthened. New lateral systems or materials for the existing members were added to the buildings and some
of these consisted of materials other than concrete. Knowledge regarding the interaction between new and existing
elements was critical to achieve the intended performance of the building after rehabilitation. It would be helpful to
engineers if more guides regarding element interaction were provided in the future editions of ASCE 41 and ACI
369R.
Horizontal LFRS, diaphragm, and collectors were strengthened in all the buildings in this study and are as
important as the vertical LRFS for seismic force resistance and transfer. More guidelines for evaluation and retrofit
of the horizontal LFRS are desirable in ASCE 41 and ACI 369R to enhance engineers’ understanding of these
system and to achieve performance consistent with the performance of the vertical LFRS.
Advanced analysis like the nonlinear dynamic procedure is often used in the evaluation of existing and
retrofitted structures. It is desirable that more information regarding nonlinear dynamic analysis, such as cyclic
hysteresis and variation of stiffness of conventional concrete members or members strengthened with materials other
than concrete, should be provided in ASCE 41 and ACI 369R so engineers could have more consistent procedures
for determining load redistribution and energy dissipation in concrete members. It should be emphasized that
engineers should not use nonlinear analysis with the intent of justifying performance of existing non-ductile
components which are critical elements for resisting gravity or seismic load and may experience significant
deformation during earthquakes. Engineers should be more cautious about uncertainty of seismic loads and
approximate natures of analysis results when their decisions are based on the insights obtained from nonlinear
analysis. The best use of nonlinear analysis may be the verification of the retrofit design, employing the capacity-
based design concept and ductile details for load redistribution

ACKNOWLEDGEMENTS

The authors would like to extend their gratitude to their colleague engineers within Degenkolb Engineers who
provided the information regarding the projects for these case studies.

REFERENCES

ACI Committee 369 (2011), Guide for Seismic Rehabilitation of Existing Concrete Frame Buildings and
Commentary (ACI 369R-11), American Concrete Institute, Farmington Hills, Mich., USA

American Society of Civil Engineers (ASCE), (2003), Seismic Evaluation of Existing Buildings, ASCE 31-03,
Reston, VA., USA

9.11
@Seismicisolation
@Seismicisolation
I. Kim and G. Hagen

American Society of Civil Engineers (ASCE), (2007), Seismic Rehabilitation of Existing Buildings, ASCE 41-06,
Reston, VA., USA

American Society of Civil Engineers (ASCE), (2013, expected), Seismic Evaluation and Retrofit of Existing
Buildings, ASCE 41-13, Reston, VA.,

C. Hyland, A, Smith (2012), CTV Building Collapse Investigation, Department of Building and Housing, New
Zealand

Canterbury Earthquake Royal Commission (2012), Final Report Volume 2, The Performance of Christchurch CBD
Buildings, Christchurch, New Zealand

NEHRP Consultants Joint Venture, (2010), Program Plan for the Development of Collapse Assessment and
Mitigation Strategies for Existing Reinforced Concrete Buildings, US Department of Commerce, National
Institute of Standards and Technology, USA

9.12
@Seismicisolation
@Seismicisolation
SP-297—10

Towards an Accurate Determination of Collapse Vulnerable


Reinforced Concrete Buildings
By Khalid M. Mosalam and Selim Günay

Synopsis: There are many vulnerable reinforced concrete (RC) buildings located in earthquake-
prone areas around the world. These buildings are characterized by the lack of seismic details
and corresponding non-ductile behavior and significant potential of partial and global collapse.
One of the current challenges of the earthquake engineering profession and research
communities is the identification of such buildings and determination of effective and
economical retrofit methods for response enhancement. Identification of these buildings is not a
trivial task due to the various sources of non-ductile behavior and the large number of involved
sources of uncertainty. Furthermore, accurate determination of collapse-prone buildings is
important from an economical perspective. Unfortunately, there are not enough economical
resources to retrofit all the non-ductile buildings that have the symptoms for collapse potential.
In order to use the available monetary resources in an effective manner, these buildings should
be accurately and reliably ranked to identify those that are most vulnerable to collapse. This
paper intends to provide a contribution to the accurate determination of the most collapse
vulnerable non-ductile RC buildings by discussing the methods from existing literature and
exploring the research needs related to (a) gravity load failure modeling and (b) consideration of
sources of uncertainty in an efficient manner.

Keywords: Collapse simulations; element removal; failure detection; gravity load


failure; non-ductile reinforced concrete buildings; post-failure modeling; tornado
diagram analysis.

10.1
@Seismicisolation
@Seismicisolation
K. M. Mosalam and S. Günay

Khalid M. Mosalam is a professor at the Civil and Environmental Engineering Department of


the University of California, Berkeley. Mosalam teaches structural engineering, finite element
methods, and behavior and design of concrete structures. He conducts research on the
performance and health monitoring of structural systems of concrete, masonry, and wood
subjected to extreme loads. He is also active in the areas of energy-efficient buildings and
assessment and rehabilitation of essential facilities such as bridges and electrical substations.

Selim Günay is a research scientist at the NEES laboratory at the University of California,
Berkeley. He executes the experiments of various NEES projects and conducts research on
various aspects of earthquake engineering including progressive collapse, hybrid simulations,
and performance-based earthquake engineering of different structural systems.

INTRODUCTION

It is a well-known fact that there are many vulnerable reinforced concrete (RC) buildings located in earthquake-
prone areas around the world. These buildings are characterized by the lack of seismic details (such as lack of
confinement at the beam and column ends and the beam-column joints, strong beam-weak column proportions, and
presence of shear-critical columns) and the corresponding non-ductile behavior and significant potential of partial
and global collapse, posing threats to human life. One of the current challenges of the earthquake research and
profession is the identification of such buildings and the determination of effective and economical retrofit methods
for enhancing their seismic response. Identification of these buildings is not an easy task due to the various sources
of non-ductile behavior and the large uncertainties involved in material characteristics, amount of reinforcement,
geometry, etc. Furthermore, accurate determination of collapse-prone buildings is important from an economical
perspective. Unfortunately, there are insufficient economical resources to retrofit all the non-ductile buildings that
have the symptoms for collapse potential. In order to effectively use these limited monetary resources, these
buildings should be accurately and reliably ranked to identify those that are most vulnerable to collapse. This paper
intends to provide a contribution to the accurate determination of the most collapse vulnerable non-ductile buildings
by discussing relevant methods from existing literature and exploring the research needs related to (a) gravity load
failure modeling and (b) consideration of uncertainty in an efficient manner.
Current state of knowledge and practice in nonlinear static and dynamic analyses has the ability to determine
side-sway collapse that occurs due to lateral dynamic instability at excessive lateral displacements, when the lateral
strength of the structure degrades significantly. However, non-ductile RC buildings mostly collapse by losing
gravity load carrying capacity, much before reaching these excessive displacements, as demonstrated in Figure 1. At
point A of this figure, the lateral load resistance starts to degrade because of various events such as in-plane or out-
of-plane failure of infill walls or shear failure of columns. This is followed by the loss of the gravity load carrying
capacity of the lateral and gravity load resisting components, starting with points B and C, respectively. The first
part of this paper explores this relatively neglected, but significantly important, issue of gravity load carrying
capacity by discussing the methods of gravity load failure modeling in collapse simulations and pointing out further
research needs. It is to be noted that the term “gravity load failure” is used in the rest of the paper to refer to the loss
of gravity load carrying capacity.
There are two alternative options that can be considered for gravity load failure modeling of the elements of a
structure: (a) explicit modeling, (b) implicit modeling, i.e. non-simulated failure. Explicit modeling of gravity load
failure consists of two stages: (1) detection of gravity load failure and (2) post-failure modeling. In the next section
of the paper, the advantages and disadvantages of the implicit modeling are presented. Moreover, three different
approaches are discussed for the purpose of employing them in the second stage of explicit modeling. These
approaches are (i) element removal, (ii) assignment of low stiffness to a failed element, and (iii) representation of
the post-failure response of failed elements with degradation. A subsequent section of the paper explores the models
that can be used for failure detection of various structural members in explicit modeling. The final section related to
gravity failure modeling is comprised of the modeling of the structural elements primarily designed to resist the
gravity loads with insignificant contribution to the lateral load resistance. While the gravity load resisting system can

10.2
@Seismicisolation
@Seismicisolation
Towards an Accurate Determination of Collapse
Vulnerable Reinforced Concrete Buildings

be approximately considered for the case of side-sway collapse, its explicit modeling significantly complements the
explicit gravity load failure modeling of the primary lateral resisting system, as shown in Figure 1.
Lateral
Capacity

Vertical Ductile RC
Capacity Building

Lateral System Range Gravity System


Lateral Range
A
Capacity
B

Non-ductile RC
Vertical Building
Capacity

Lateral System Gravity System


Range Range

Figure 1: Lateral and vertical responses of RC buildings (Modified from Holmes, 2000)

As mentioned above, non-ductile RC buildings generally involve significant amount of uncertainties. Therefore,
a substantial number of collapse simulations may need to be conducted for accurate identification of buildings which
are most vulnerable to collapse. The final part of the paper presents a deterministic sensitivity analysis method from
the literature, the so-called tornado diagram, as a method to identify the sources of uncertainties which are most
influential on the seismic response. Accordingly, this practical method can handle the effect of uncertainty more
efficiently in collapse simulations of non-ductile RC buildings by only incorporating these influential sources of
uncertainty as probabilistic, while those that are less influential can be incorporated as deterministic.

GRAVITY LOAD FAILURE MODELING OF LATERAL LOAD RESISTING COMPONENTS

In order to model the gravity load failure of the lateral load resisting components of non-ductile RC buildings, there
are two alternative options that can be considered. First option is the explicit modeling of gravity load failure, while
the second one is the determination of gravity failure of the components implicitly, through post-processing without
modeling of gravity failure (non-simulated failure). In the second approach, which has been the commonly utilized
approach, engineering demand parameters (EDPs), such as drifts or accelerations, obtained as a result of the
analyses can be used to determine collapse by comparing these EDPs with the limits provided by available gravity
loss models (e.g. Elwood and Moehle, 2005).
Implicit modeling of gravity load failure (non-simulated failure) can only be feasible in some cases where the
first element failure, for example a column axial failure, is sufficient to define global collapse. Such a case may
occur when all the columns at a story have similar properties and failure of all columns is likely to take place almost
simultaneously. In this case, there is no need to explicitly consider the consequences of an axially-failed column and
first column axial failure can be sufficient to define global collapse. In all other cases, explicit modeling of gravity
load failure is essential for accurate determination of collapse. This distinction can be further supported by
considering one of the collapse indicators discussed in the NIST report for collapse assessment and mitigation
strategies for existing RC buildings (2010). The considered collapse indicator is the “maximum fraction of columns
at a story experiencing axial failures,” which requires the identification of the number of columns experiencing
gravity load failures. Such identification would potentially be inaccurate if executed by post-processing of results
without an explicit consideration of gravity load failure, because the gravity load failure of a column (or a beam-
column joint) is likely to affect the response of the other columns and the overall system.

10.3
@Seismicisolation
@Seismicisolation
K. M. Mosalam and S. Günay

Explicit modeling of gravity load failure consists of two stages. The first stage is the detection of gravity load
failure. Available models in literature for columns and beam-column joints that can be used for this purpose are
presented in the next section. The second stage is the post-failure modeling, possible options of which are: (1)
Element removal, (2) Assigning low stiffness to a collapsed element, and (3) Representing the post-failure response
with degradation. The first approach consists of the direct removal of the element from the structural model upon its
failure (e.g. Talaat and Mosalam, 2007, 2009). The second approach consists of reducing the stiffness of the
collapsed element using a small multiplier (e.g. Grierson et al., 2005) in order to eliminate its contribution to the
global structural stiffness matrix. In the third approach, the post-failure response is represented with a degraded
force-displacement relationship (e.g. Elwood and Moehle, 2005). Advantages and disadvantages of these three
approaches are summarized in Table 1 along with the implicit gravity load failure approach mentioned above.

Table 1: Advantages and disadvantages of different gravity load failure modeling methods
Component
Advantages Disadvantages
Failure Method
Explicit Modeling 1. Numerical problems associated with ill- 1. Requirement of additional book-keeping
Option 1: conditioned stiffness matrices are operations to update the nodal masses and to
eliminated. check nodal forces, constraints, restraints,
Element removal
2. Enforcing dynamic equilibrium dangling nodes, floating elements, etc.
enables:
a) Computation of the resulting 2. An additional computational burden
increase in nodal accelerations introduced by the redefinition of degrees of
b) Inclusion of the system’s complete freedom of a structural model and the
kinematic state at time of element corresponding connectivity upon removal of
collapse to determine if it can an element or several elements.
survive to a new equilibrium state.
3. Motion of the collapsed element can be 3. Convergence problems, not on the element or
tracked relative to the damaged system material levels, but on the numerical
to estimate the time and kinetics of a integration level, as a result of the sudden
subsequent collision with the intact updating of mass, stiffness and damping
structural part. matrices, as well as the local vibrations
4. Elimination of the numerical triggered as a consequence of the resulting
convergence problems related to the transient effect (refer to discussions on the
iterative formulation of some element methods to overcome these convergence
and material types by removing them problems).
(refer to item 1 in the disadvantages of
degraded post-failure response).
Explicit Modeling Additional tasks (Items 1 and 2 in 1. Numerical problems associated with ill-
Option 2: disadvantages of element removal) related conditioned stiffness matrices.
to the element removal process are 2. Not possible to explicitly consider the
Assigning low
avoided. consequences of component failure (Items 2
stiffness to a
and 3 in advantages of element removal).
failed element

Explicit Modeling Additional tasks (Items 1 and 2 in 1. Numerical convergence problems related to
Option 3: disadvantages of element removal) related the iterative formulation of some types of
to the element removal process are elements, e.g. force-based beam-column, and
Degraded post-
avoided. materials, e.g. Bouc-Wen, since the failure
failure response
state generally occurs at a negatively sloped
portion of the constitutive relationship.
2. Not possible to explicitly consider the
consequences of component failure (Items 2
and 3 in advantages of element removal).
Implicit Modeling Suitable for fast and simplified analyses Inaccurate results due to the lack of realistic
(Non-simulated and in some special cases, e.g. having representation of the post-failure response.
failure) similar columns in one story.

10.4
@Seismicisolation
@Seismicisolation
Towards an Accurate Determination of Collapse
Vulnerable Reinforced Concrete Buildings

The element removal approach of Talaat and Mosalam (2007) is based on dynamic equilibrium and the
resulting transient change in system kinematics. It constitutes the basis of a corresponding progressive collapse
algorithm. This algorithm is implemented in OpenSees for automatic removal of collapsed elements during an
ongoing simulation, Figure 2. The implementation is carried out as a new OpenSees module, designed to be called
by the main analysis module after each converged integration time step to check each element for possible violation
of its respective removal criteria, where the relevant models presented in the next section can be used for defining
the removal criteria. A violation of a pre-defined removal criterion triggers the activation of the algorithm on the
violating element before returning to the main analysis module. Activation of the element removal algorithm
includes updating nodal masses, checking if the removal of the collapsed element results in leaving behind dangling
nodes or floating elements, which must be removed as well as removing all associated element and nodal forces,
imposed displacements, and constraints. It is noted that the gravity loads at the node of a column, which is common
with the other elements, is not removed. Accordingly, the gravity loads on the structure are not reduced upon
removal of a column, allowing for the analysis model to capture the redistribution of the gravity loads to the other
intact columns.

Update
Start from main code  Remove dangling nodes structural
End,
 Remove floating elements model, back to
Check for dangling nodes,  Delete element/node loads time step
 Remove element and main
floating elements, and
 Update nodal masses solution code
element loads and masses
parameters

Figure 2: Element removal algorithm (Talaat and Mosalam, 2007)

Since the structural elements lose their ability to support gravity loads after gravity load failure, the removal of
a failed element is the most representative approach to model gravity load failure. Hence, the discussion in the
following two paragraphs are based on the comparison of the element removal approach (Option 1) with Options 2
and 3. It should be noted that Option 1 approach assumes that the gravity load support is lost instantaneously. As
mentioned previously, the first stage of gravity load failure in explicit modeling is the detection of this failure. Such
detection is based on equations derived from tests where the loss of the gravity load support of the test specimen was
defined by a single point and there is no data obtained from the tests beyond this point. Accordingly, the assumption
of instantaneous gravity load failure is dictated by the gravity failure detection models. As an alternative, the
detection equations can be constructed with a probability distribution, e.g. in the form of a set of equations for the
median and median plus/minus a dispersion. However, such equations require further experimental research.
The removal of a collapsed element requires several book-keeping operations to update the nodal masses and to
check nodal forces, constraints, restraints, dangling nodes, floating elements, etc. Also, there is an additional
computational burden introduced by the redefinition of degrees of freedom of a structural model and the
corresponding connectivity upon removal of one or more elements. Such additional tasks are avoided in Option 2,
which consists of assigning low stiffness to failed elements. However, there are three important advantages of
Option 1 compared to Option 2. First, it avoids numerical problems due to ill-conditioned stiffness matrices. Second,
enforcing the dynamic equilibrium enables: (1) the computation of the resulting increase in nodal accelerations and
(2) the inclusion of the system’s complete kinematic state at time of element collapse to determine if the structure
can successfully redistribute the forces from the removed element and survive to a new equilibrium state. Third, the
motion of the collapsed element can be tracked relative to the damaged system to estimate the time and kinetics of a
subsequent collision with the intact structural part.
Although representing the post-failure response with a degraded force-displacement relationship in Option 3 is
realistic for most of the failed components, it may introduce numerical problems. The failure state generally
corresponds to a negatively sloped portion of the constitutive relationship, where the iterative formulation of some
types of elements, e.g. force-based beam-column, and materials, e.g. Bouc-Wen, are likely to experience
convergence problems. The removal of such elements automatically eliminates the associated numerical problems.
Analyses conducted to estimate the responses obtained from shaking table tests of a non-ductile RC frame showed
that the analyses considering and not considering the element removal (ER) were both successful in predicting the
observed collapse of the non-ductile members (Mosalam et al., 2009). On the other hand, the response after collapse
was rather jagged and close to being unstable for the case without element removal whereas the analysis with

10.5
@Seismicisolation
@Seismicisolation
K. M. Mosalam and S. Günay

element removal provided a more reasonable response, Figure 3. It should be noted that the analyses without the
element removal used the degraded post-failure approach (Option 3).

Ductile Columns Non-ductile Columns

Slip spring
C4 C3 C2 C1
Shear and
axial springs

50 50
40 Column C1 40 Column C2

Column Shear (kN)


Column Shear (kN)

30 30
20 20
10 10
0 0
-10 -10
-20 -20 Test
-30 -30 Analysis (ER: Option1)
-40 -40
Analysis ( w/o ER: Option 3)
-50 -50
-0.1 -0.05 0 0.05 0.1 -0.1 -0.05 0 0.05 0.1
Drift Ratio Drift Ratio

Figure 3: Response of a system with non-ductile columns experiencing axial failure from shaking table tests
and analyses with and without element removal (Mosalam et al., 2009)

Element removal may introduce convergence problems, not on the element or material levels, but on the
numerical integration level, as a result of the sudden updating of mass, stiffness and damping matrices, and the
triggered local vibrations as a consequence of the resulting transient effect due to the sudden changes in the
matrices. It is reminded that the external forces due to gravity loading and the ground motion excitation remain
unchanged when the matrices are suddenly updated. A possible solution to such convergence problems is adaptive
switching of solver type and convergence criteria and reduction of the integration time step (Talaat and Mosalam,
2007). It is to be noted that this strategy has been used for the analyses of the non-ductile RC frame mentioned in the
above paragraph. Another effective solution is the use of transient integrators which do not require iterations, e.g.
operator-splitting methods (Hughes et al., 1979). Analyses conducted on bridge systems (Mosalam et al., 2013)
showed that the operator splitting method results in exactly the same solution as the commonly used implicit
Newmark integration even for cases with highly nonlinear response.
It is noted that the objective of this paper is not the recommendation of an explicit modeling option; rather it is
to present the available options of collapse determination to the interested reader. Furthermore, the most suitable
option may change from one building to another. Hence, the presented advantages and disadvantages are expected to
provide guidance to the readers in choosing the best starting point and possibly switching between different options.

DETECTION OF GRAVITY LOAD FAILURE

As mentioned previously, the first stage of the explicit modeling of gravity load failure is the detection of gravity
load failure. Models that can be used for the detection of gravity load failure of columns, beam-column joints, slab-
column joints and infill walls are described in the following sub-sections. Collapse of buildings with shear-walls,
due to loss of gravity load carrying capacity, has rarely been observed in the last 50 years (Wallace et al., 2008).
Therefore, detection of gravity load failure of shear-walls is not covered in this paper.

10.6
@Seismicisolation
@Seismicisolation
Towards an Accurate Determination of Collapse
Vulnerable Reinforced Concrete Buildings

Columns

One of the models that can be used to detect the gravity load failure of columns is proposed by Elwood and Moehle
(2005), where the drift at axial failure of a shear-damaged column is represented as follows:

1  tan  
2
 4 (1)
  
 L  axial 100 tan   P s Ast f yt d c tan  

where (Δ/L)axial is the drift ratio at axial failure, P is the column axial force, Ast, fyt, and s are respectively area, yield
strength, and spacing of the transverse reinforcement, dc is the column core depth (center to center of tie) and θ is the
critical crack angle from the horizontal (assumed 65°).
Elwood and Moehle (2005) stated that this axial failure model is based on data from 12 columns, where all
columns were constructed from normal strength concrete, had the same height-to-width ratio, were designed to yield
the longitudinal reinforcement prior to shear failure, and were tested in uniaxial bending. Despite these limitations,
Fardipour et al. (2011) mentioned that the drifts estimated with a modified version of Equation (1) were in
reasonable agreement with the test results of four cantilever columns with axial load ratios of 20% to 40%, a
nominal transverse reinforcement ratio of 0. 07% and vertical reinforcement ratio of 0.5% to 1%. The modifications
of Fardipour et al. (2011) to Equation (1) consisted of: (1) change of the crack angle to 55+35P/Po for P/Po < 0.25
and 59 for P/Po > 0.25 where Po is the axial force capacity of the undamaged column and (2) addition of the yield
drift to the right-hand side of Equation (1) to calculate the drift corresponding to axial failure. However, it is to be
noted that the above change of the crack angle is inappropriate due to the way Equation (1) was developed.
This axial capacity model is implemented in OpenSees as a limit state material model and used as a spring
connected to a column end. Removal of the corresponding spring is also implemented in an earlier version of
OpenSees. It is to be noted that the authors can provide this version to an interested reader, until the code is
rearranged to comply with the standard version of OpenSees. When the drift during a simulation reaches the drift
corresponding to axial failure, the corresponding spring and the column, to which the spring is connected, are
removed using the element removal algorithm.

Beam-Column Joints

Beam column joints of old non-ductile RC buildings, e.g. designed in the sixties, are generally unreinforced without
any transverse steel bars. The beams connected to such joints rotate relative to the columns, i.e. right angle between
the beam and column is not maintained, due to joint shear failure and corresponding deformation. Joint panel
flexibility can be modeled by using a rotational spring located between the beam and column end nodes. It is noted
that rigid end offsets are used at the beam and column ends to consider the joint physical dimensions (Figure 4d).
The rotational spring is defined by a nonlinear constitutive relationship, which is characterized by a backbone curve
and a set of hysteresis rules (Park and Mosalam, 2013a). These characteristics are recently developed empirically
based on the measured joint responses and visual observations from tests of four corner beam-column-slab joint
specimens (Park and Mosalam, 2013b) and verified by comparison with other exterior and interior beam-column
joint tests. A strength model is developed to determine the peak force of the backbone curve, which also corresponds
to the joint shear strength (Park and Mosalam, 2009, 2012a). The practical strength model in Figure 4a, which
accounts for the effects of two main parameters, namely (1) the joint aspect ratio, defined as the ratio of beam to
column cross-sectional heights and (2) the beam reinforcement ratio, is verified through its accurate predictions of
various beam-column joint test results available in the literature (Park and Mosalam, 2012b, 2012c).
The rotational spring and the constitutive relationship mentioned above can be used to represent the axial failure
and the corresponding removal of a beam-column joint using the proposed extension by Hassan (2011). Hassan
proposed an axial capacity model for beam-column joints where the drift, i.e. beam tip displacement normalized by
the beam length, refer to the test setup in (Park and Mosalam, 2012), at axial collapse is represented as a function of
the axial force and the beam bottom reinforcement strength (Figure 5). Equation (2) is suitable to be used to remove
a beam-column joint as a part of the progressive collapse algorithm when the drift in this equation is replaced by the
joint rotation of the considered analytical model, Figure 4c. Note that the difference between the joint rotation and
the above-mentioned drift in the tests of Hassan (2011) is the sum of the flexural deformation of the beam and the
beam displacement due to column rotation (Figure 6). However, it should be noted that this expression is based on a
rather small database of joint axial failures. Therefore, more joint axial failure tests are needed to further verify this

10.7
@Seismicisolation
@Seismicisolation
K. M. Mosalam and S. Günay

relationship. Hassan (2011) also mentioned that the case of high axial load on a joint where the beam flexural
capacity is much smaller than the direct joint failure capacity is excluded from the application of this model.

Asb f yb
 axial  0.057 (2)
P tan 

where θaxial is the joint rotation at the joint axial failure, P is the axial force, Asb and fyb are the respective area and
yield strength of the beam bottom reinforcement, and θ is the crack angle.

Figure 4: Overview of unreinforced beam-column joint model (Park and Mosalam, 2012c)

The removal of a beam-column joint is not implemented in OpenSees yet. However, the idea is similar to the case of
column failure, i.e. when the rotation of the spring (representing the joint) during a simulation reaches the rotation
corresponding to axial failure defined by Equation (2), the joint (rotational spring and rigid end offsets) is removed
using the element removal algorithm. Because the generic removal algorithm is already implemented, including the
removal of beam-column joints using the above mentioned criteria is rather straightforward.

Slab-Column Joints

Gravity support loss of slab-column joints can be defined with the punching shear failure. In order to detect the
punching shear failure, available limit models in literature such as Hueste and Wight (1999) or Elwood et al. (2007)
can be used (Figure 7). In these models, drift or plastic rotation values corresponding to the punching shear failure
are determined as functions of the gravity shear ratio defined as the value of the vertical gravity shear divided by the
punching shear strength of the joint. Kang et al. (2009) used these limit models for the detection of punching shear
failure, while conducting analytical simulations to predict the results of shaking table tests. In that regard, they used
option 3, representing the post-failure response with degradation, for post-failure modeling.

10.8
@Seismicisolation
@Seismicisolation
Towards an Accurate Determination of Collapse
Vulnerable Reinforced Concrete Buildings

Figure 5: Axial capacity model in Equation (2) for beam-column joints (Hassan, 2011)


  : Drift at axial collapse
 L  axial
θc  j ,axial : Joint rotation at axial collapse
θj
 c ,axial : Column rotation at axial collapse

  bf 
  : Beam flexural drift at axial collapse
bf  L  axial

Figure 6: Contribution of different components to beam tip displacement in a beam-column joint test

(a) Hueste and Wight (1999) (b) Elwood et al. (2007)


Figure 7: Detection of gravity load failure for slab-column joints

10.9
@Seismicisolation
@Seismicisolation
K. M. Mosalam and S. Günay

Unreinforced Masonry Infill Walls

When the seismic vulnerabilities present in the RC system are combined with the complexity of the interaction
between the infill walls and the surrounding frame and the brittleness of the unreinforced masonry (URM) materials,
the URM infill walls can increase the vulnerability and collapse potential of non-ductile RC buildings. Earthquakes
in the last two decades, e.g., 1999 Kocaeli, 2008 Wenchuan and 2009 L’Aquila earthquakes, led to several
observations related to URM infill walls (Mosalam and Günay, 2012), which are listed as follows:
1. URM infill walls contribute to the stiffness and strength of the frames as evidenced from the weak/soft
story damage of the open ground story buildings and the torsional response created by the non-uniform
distribution of infill walls around the building perimeter.
2. URM infill wall failure is a combination of in-plane (IP) and out-of-plane (OOP) effects as evidenced from
some of the URM infill wall failures taking place at the upper stories instead of the lower stories where the
story shear forces are the highest.
3. Failure of infill walls at a story lead to the formation of weak/soft stories during the earthquake, which may
result in the failure of a story as evidenced by intermediate story collapses.
4. Infill walls interact with the frame members as evidenced by shear cracks and failures of columns and
beam-column joints in infilled bays.
Accordingly, URM infill walls should be modeled to consider these observations. The first two observations can be
reflected by employing a practical model that considers IP-OOP interaction of infill walls, Figure 8, (Kadysiewski
and Mosalam, 2008). The fourth observation can be taken into account by modeling nonlinear shear springs at the
column ends to consider the effect of additional horizontal forces transferred from the infill walls to the columns.
The third observation can be considered by removal of failed infill walls where detection of failure is based on a
combination of IP and OOP displacements (OpenSees Wiki, 2011, Mosalam and Günay, 2013).
Modeling of infill walls considering the element removal due to IP/OOP interaction has been recently used for
the investigation of the earthquake response of buildings designed according to modern seismic codes without
considering the infill walls in the design process (Mosalam et al., 2013). It is noted that, different from the other
discussed elements, failure of infill walls is not directly considered as gravity load failure. It is rather considered to
negatively affect the lateral response. However, detection and post-failure modeling of infill walls are still important
modeling aspects for the objective of the accurate determination of collapse, because the consequences are likely to
affect the gravity load failure of other discussed elements.
IP displacement

Displacement history

Failure Curve (symmetric


OOP displacement about x and y axes)

Figure 8: Failure detection for URM infill walls considering IP-OOP interaction

EXPLICIT MODELING OF GRAVITY SYSTEMS

Modeling of the gravity system as a single column that accounts for P-Δ effects (Figure 9) is a commonly utilized
approach (Liel et al., 2009, Lai and Mahin, 2013). Such modeling does not generally account for the strength and
stiffness of the gravity system, either because collapse is not of interest or the investigated collapse mechanism is
side-sway. For the collapse type investigated herein, namely the gravity load failure, explicit modeling of the gravity
system is essential. As shown in Figure 1, gravity support may not be completely lost after the lateral and axial

10.10
@Seismicisolation
@Seismicisolation
Towards an Accurate Determination of Collapse
Vulnerable Reinforced Concrete Buildings

failures of the primary system. Therefore, explicit modeling of the gravity system generally leads to a more accurate
and realistic determination of the global collapse.

Leaning
column

Figure 9: Modeling of the gravity system as a leaning column

TORNADO DIAGRAM ANALYSIS

Non-ductile RC buildings generally involve significant number of uncertainty sources. Some of these sources
are concrete strength, longitudinal and transverse reinforcement yield strength, concrete and masonry modulus of
elasticity, masonry compressive and shear strengths, damping ratio, story mass, and ground motion record-to-record
variability. Considering all these uncertainty sources in the process of accurate determination of the most collapse-
vulnerable buildings is likely to result in an extensive number of collapse simulations. However, uncertainties of
some of these parameters may have insignificant effects on the variability of the structural response. Tornado
diagram analysis is a practical method used to identify and rank the effect of parameter uncertainties on the response
variability (Lee and Mosalam, 2006). Considering that time is of essence to rapidly determine the most collapse
vulnerable buildings, so that they can be retrofitted or demolished (if needed) before the next big earthquake,
Tornado diagram analysis comes forward as a suitable method to eliminate the burden of unnecessary simulations
by treating the parameters with insignificant effect as deterministic.
The tornado diagram, commonly used in decision analysis, has been used in sensitivity analysis in earthquake
engineering (Porter et al., 2002). The diagram consists of a set of horizontal bars, referred to as swings, one for each
random variable, i.e. considered parameter. The length of each swing represents the variation in the output, i.e. EDP,
due to the variation in the respective random variable. Thus, a variable with larger effect on the EDP has larger
swing than those with lesser effect. In a tornado diagram, swings are displayed in a descending order from top to
bottom. This wide-to-narrow arrangement of swings resembles a tornado. In order to determine the swing due to a
considered parameter, two extreme values, e.g. 10th and 90th percentiles, corresponding to pre-defined upper and
lower bounds of the assumed probability distribution for the parameter are selected. Considered EDP is determined
as a result of nonlinear response history analysis using the upper and lower bound values of the considered
parameter while the other input random variables are set to their best estimates such as the medians. This process
yields two bounding values of the EDP variation for each input parameter. The absolute difference of these two
values is the swing of the EDP corresponding to the selected input parameter. This process is repeated for all the
input parameters to compute the swings of the EDP, Figure 10. Finally, one builds the tornado diagram by arranging
the obtained swings in a descending order as mentioned above. The resulting Tornado diagram generally provides an
indicative picture for the selection of the necessary random variables to be used in the collapse simulations to
develop the fragility curves (Lee and Mosalam, 2005).

10.11
@Seismicisolation
@Seismicisolation
K. M. Mosalam and S. Günay

Figure 10: Construction of the Tornado diagram (Lee and Mosalam, 2006)

CONCLUDING REMARKS

Non-ductile RC buildings are one of the main seismic safety concerns worldwide. In order to use the limited
monetary resources in an effective manner, these buildings should accurately and reliably be ranked to identify those
that are most vulnerable to collapse. This paper attempted to provide a contribution to the accurate determination of
the most collapse vulnerable non-ductile buildings by discussing the methods from existing literature and exploring
the research needs related to gravity load failure modeling and consideration of uncertainty in an efficient manner.
These topics are presented to facilitate the accurate and efficient application of refined seismic assessment.
Concluding remarks of the above discussions are as follows:
 Non-ductile RC buildings mostly collapse by losing gravity load carrying capacity, much before reaching
the lateral displacements that would be experienced in a side-sway collapse mechanism.
 Implicit (non-simulated) modeling of gravity load failure is only adequate in the case where determining
the first element failure is sufficient to define global collapse. In all other cases, explicit modeling should
be utilized for accurate prediction of global collapse.
 There are various advantages and disadvantages of the methods that can be used for post-failure modeling.
Furthermore, the most suitable option may change from one building to another. Advantages and
disadvantages presented in this paper are expected to provide guidance to the readers for choosing the best
starting point and switching between different options.
 There are various methods in literature that can be used for failure detection of the RC elements of a
structure. However, there is still a need for enhancement of the failure detection of some of the elements,
for example the non-ductile RC beam-column joints.
 There is an existing element removal algorithm implemented in OpenSees. This algorithm is used in the
standard version of OpenSees for the removal of URM infill walls. Removal of columns and beam-column
joints are planned to be included in the standard version of OpenSees in near future.
 Gravity support may not be completely lost after the lateral and axial failures of the primary lateral load
resisting system. Therefore, explicit modeling of the gravity system generally leads to a more accurate and
realistic determination of the global collapse.
 Non-ductile RC buildings generally involve significant amount of uncertainties. The presented Tornado
Diagram Analysis can be used as a method to handle uncertainty in an efficient and practical manner.

REFERENCES

Elwood K.J., Matamoros A., Wallace J.W., Lehman D.E., Heintz J.A., Mitchell A., Moore M.A., Valley M.T.,
Lowes L.N., Comartin C., Moehle J.P., 2007, “Update of ASCE/SEI 41 Concrete Provisions,” Earthquake
Spectra, 23(3):493-523.
Elwood K.J., Moehle J.P., 2005, “Axial Capacity Model for Shear-Damaged Columns,” ACI Structural Journal,
102: 578-587.

10.12
@Seismicisolation
@Seismicisolation
Towards an Accurate Determination of Collapse
Vulnerable Reinforced Concrete Buildings

Fardipour M., Lam N., Gad E., Wilson J., 2011, “Collapse Limit State Assessment of Lightly Reinforced Concrete
Columns,” Australian Earthquake Engineering Society 2011 Conference, 18-20 November, Barossa Valley,
South Australia.
Grierson D.E., Xu L., Liu Y., 2005, “Progressive-Failure Analysis of Buildings Subjected to Abnormal Loading,”
Computer-Aided Civil and Infrastructure Engineering 20(3):155-171.
Hassan W.M., 2011, “Analytical and Experimental Assessment of Seismic Vulnerability of Beam-Column Joints
without Transverse Reinforcement in Concrete Buildings,” PhD Dissertation, University of California,
Berkeley, CA.
Holmes W.T., 2000, “Risk Assessment and Retrofit of Existing Buildings”, 12th World Conference on Earthquake
Engineering, Auckland, New Zealand, Paper No: 2826
Hueste, M.B.D., Wight J.K., 1999, “Nonlinear Punching Shear Failure Model for Interior Slab-Column
Connections,” Journal of Structural Engineering, 125(9):997-1007.
Hughes T.J.R., Pister K.S., Taylor R.L., 1979, “Implicit-Explicit Finite Elements in Nonlinear Transient Analysis,”
Computer Methods in Applied Mechanics and Engineering 17/18:159-182.
Kadysiewski S., Mosalam K.M., 2008, “Modeling of Unreinforced Masonry Infill Walls Considering In-Plane and
Out-of-Plane Interaction,” PEER Report 2008/102, Berkeley, CA, 144 pp.
Kang TH.-K., Wallace J.W., Elwood K.J., 2009, “Nonlinear Modeling of Flat Plate Systems,” ASCE Journal of
Structural Engineering, 135(2): 147–158.
Lai J.W., Mahin S.A., 2013, “Experimental and Analytical Studies on the Seismic Behavior of Conventional and
Hybrid Braced Frames,” PEER Report 2013/20, Berkeley, CA, 633 pp.
Lee, T.-H. and K.M. Mosalam, 2005, "Seismic Demand Sensitivity of Reinforced Concrete Shear-Wall Building
Using FOSM Method," Earthquake Engineering and Structural Dynamics, 34(14): 1719-1736.
Lee T.-H., Mosalam K.M., 2006, “Probabilistic Seismic Evaluation of Reinforced Concrete Structural Components
and Systems,” PEER Report 2006/04, Berkeley, CA, 181 pp.
Liel A., Haselton C., Deierlein G., Baker J., 2009, “Incorporating Modeling Uncertainties in the Assessment of
Seismic Collapse Risk of Buildings,” Structural Safety, 31(2):197–211.
Mosalam K.M., Park S., Günay M.S., 2009, “Evaluation of an Element Removal Algorithm For Shear Critical
Reinforced Concrete Frames,” COMPDYN, Thematic Conf. on Computational Methods in Structural Dynamics
and Earthquake Engineering, June 22-24, Greece.
Mosalam, K.M. and Günay M.S., 2012, “Behavior and Modeling of Reinforced Concrete Frames with Unreinforced
Masonry Infill Walls”, in Structural Engineering and Geomechanics, [Ed. S.K. Kunnath], in Encyclopedia of
Life Support Systems (EOLSS), Developed under the Auspices of the UNESCO, Eolss Publishers, Oxford ,UK,
[http://www.eolss.net] [Retrieved February 19, 2013]
Mosalam K.M., Günay M.S., Sweat H.D., 2013, “Simulation of Reinforced Concrete Frames with Unreinforced
Masonry Infill Walls with Emphasis on Critical Modelling Aspects,” 12th Canadian Masonry Symposium, June
2-5, Vancouver.
Mosalam, K.M. and Günay S., 2013, “Progressive Collapse Analysis of RC Frames with URM Infill Walls
Considering In-Plane/Out-of-Plane Interaction,” Earthquake Spectra, In press.
Mosalam, K.M., Liang X., Günay S., and Schellenberg A., 2013, “Alternative Integrators and Parallel Computing
for Efficient Nonlinear Response History Analyses,” COMPDYN 2013, 4th ECCOMAS Thematic Conference
on Computational Methods in Structural Dynamics and Earthquake Engineering, M. Papadrakakis, V.
Papadopoulos, V. Plevris (eds.), June 12-14, Kos Island, Greece.
National Institute of Standards and Technology (NIST), 2010, “Program Plan for the Development of Collapse
Assessment and Mitigation Strategies for Existing Reinforced Concrete Buildings,” Report No. NIST GCR 10-
917-7, NEHRP.
OpenSees Wiki, 2011, http://opensees.berkeley.edu/wiki/index.php/Infill_Wall_Model_and_Element_Removal
OpenSees Wiki, 2012, http://opensees.berkeley.edu/wiki/index.php/Rigid_Diaphragm_Consequences
Park, S. and K.M. Mosalam, 2009, “Shear Strength Models of Exterior Beam-Column Joints without Transverse
Reinforcement,” PEER Report 2009/106, Berkeley, CA, 101 pp.

10.13
@Seismicisolation
@Seismicisolation
K. M. Mosalam and S. Günay

Park, S. and K.M. Mosalam, 2012a, “Parameters for Shear Strength Prediction of Exterior Beam-Column Joints
without Transverse Reinforcement,” Engineering Structures, 36(3):198-209.
Park, S. and K.M. Mosalam, 2012b, “Analytical Model for Predicting the Shear Strength of Unreinforced Exterior
Beam-Column Joints,” ACI Structural Journal, 2012, 102(2): 149-160.
Park S., Mosalam K.M., 2012c, “Experimental and Analytical Studies on Reinforced Concrete Buildings with
Seismically Vulnerable Beam-Column Joints”, PEER Report 2012/03, Berkeley, CA, 224 pp.
Park, S. and K.M. Mosalam, 2013a, “Simulation of Reinforced Concrete Frames with Non-Ductile Beam-Column
Joints,” Earthquake Spectra, 29(1): 233-257.
Park, S. and K.M. Mosalam, 2013b, “Experimental Investigation of Non-Ductile Reinforced Concrete Corner Beam-
Column Joints with Floor Slabs,” ASCE, Journal of Structural Engineering, 139(1): 1-14.
Porter, K.A., Beck, J.L., and Shaikhutdinov, R.V., 2002, “Sensitivity of Building Loss Estimates to Major Uncertain
Variables,” Earthquake Spectra, 18(4): 719-743.
Talaat M., Mosalam K.M., 2007, “Computational Modeling of Progressive Collapse in Reinforced Concrete Frame
Structures,” PEER Report 2007/10, Berkeley, CA, 310 pp.
Talaat, M. and K.M. Mosalam, 2009, “Modeling Progressive Collapse in Reinforced Concrete Buildings Using
Direct Element Removal,” Earthquake Engineering and Structural Dynamics, 38(5): 609-634.
Wallace, J.W., Elwood, K.J., and Massone, L.M., 2008, “An Axial Load Capacity Model for Shear Critical RC Wall
Piers,” Journal of Structural Engineering, 134(9): 1548-1557.

10.14
@Seismicisolation
@Seismicisolation
SP-297—11

 
IDENTIFYING BUILDINGS WITH HIGH SEISMIC RISK UNDER URBAN
RENEWAL LAW IN TURKEY

Baris Binici, Ahmet Yakut, Sadun Taniser, Guney Ozcebe

SYNOPSIS:

A new law known as the "Urban Renewal Law" for risk mitigation was passed in May 2012 with the
objective of reducing seismic risk associated with the existing building stock in Turkey. As stated in the
law, new provisions are set forth to assess and to identify seismically vulnerable residential buildings as
quickly as possible. The buildings that are classified as high risk are either demolished or strengthened.
New buildings are constructed through the financing options provided by the government. In this study,
first, the technical provisions of seismic risk assessment, based on linear elastic analysis, are briefly
described with special emphasis on the deformation limits. Because of the inability of the linear elastic
analysis to allow for redistribution, some flexibility is provided on how many vertical load bearing
elements are allowed to exceed their performance limits. Afterwards, three case study buildings are
analyzed by using the new provisions and ASCE/SEI 41-06 linear elastic procedure. Level and sources of
conservatism in the two approaches are critically evaluated.

Keywords: reinforced concrete buildings, seismic assessment, linear elastic procedure

11.1
@Seismicisolation
@Seismicisolation
B. Binici et al.

Biography:

Baris Binici is a Professor at Middle East Technical University, Department of Civil Engineering, Turkey. He
obtained his B.Sc. degree at Middle East Technical University and M.S.E. and Ph.D. degrees at The University
of Texas at Austin. His areas of research are reinforced concrete mechanics, seismic assessment and retrofit of
concrete structures.

Ahmet Yakut is a Professor at Middle East Technical University, Department of Civil Engineering, Turkey. He
completed his B.Sc. and M.Sc. degrees at Middle East Technical University and Ph.D. degree at The University
of Texas at Austin. His research interests are earthquake engineering, seismic vulnerability and risk assessment.

Sadun Taniser obtained his B.Sc. and M.Sc. degrees in civil engineering at Middle East Technical University.
He is currently a Ph.D candidate at Middle East Technical University. His research field is earthquake
engineering.

Guney Ozcebe is the Dean of Engineering and Architecture Faculty at TED University. He completed his B.Sc.
and M.Sc. degrees at Middle East Technical University and his Ph.D. degree at University of Toronto. He is a
member of ACI Committees 314 and 318-WA. His research interests include repair and upgrade of concrete
structures, high strength concrete, performance-based design, evaluation and assessment of structures.

11.2
@Seismicisolation
@Seismicisolation
Identifying Buildings with High Seismic Risk Under Urban Renewal Law in Turkey

INTRODUCTION

Seismic performance of the Turkish building stock during recent earthquakes revealed that majority of
existing buildings is extremely vulnerable to earthquakes (Kocaeli and Duzce 1999, Bingol 2003, Van 2011).
The poor performance prompted officials and government organizations to take some major actions in the last
decade including the assessment and strengthening of school buildings, hospitals, industrial facilities and some
of the government buildings. Although, the existing seismic code was updated in 2007 to include a part on the
assessment and rehabilitation of existing buildings, most of the vulnerable residential buildings received no
attention and they are at high seismic risk. A new campaign was started with a new urban renewal law passed on
May 16 2012 (MEU 2012) to mainly address the vulnerable residential building stock. According to the law,
local municipality authorities, Ministry of Environment and Urbanization (MEU) or any of the apartment owners
may request the seismic assessment of a building. If a building is found to be seismically vulnerable, occupants
are given 60 days before demolishing the building or strengthening it with a methodology approved by the local
municipality authorities. Within ten days of the technical report claiming that building is not expected to perform
satisfactorily in a future earthquake, seismically vulnerable buildings are banned from rent or sale. The
applicants are provided either 18 months of rent support as a grant or offered reduced interest rates for mortgage
by the government for the process of refinancing of the building as an encouragement.

According to the law a building is classified as high risk or critical if the building is expected to experience
collapse or very heavy damage under the earthquake level. The number of buildings in Turkey to be examined
under this law in the next 10 years is estimated to be in the order of several millions. The time and budget
required for the assessment of these buildings using existing code procedures is overwhelming. For these
reasons, The Ministry of Environment and Urbanization set up a committee to draft a relatively fast and
acceptable procedure for assessment of residential buildings (named as Specifications for Classification of High
Risk Buildings-SCHRB). The procedure is based on linear elastic analysis of the building model that may be
generated from the information collected for the ground floor. This article first summarizes the provisions with
emphasis on the member performance limit states. Then, applications of the new provisions and existing Turkish
code on a set of representative new and existing buildings are presented. The linear assessment procedure of
ASCE/SEI 41-06 was also applied on three of these buildings in order to investigate the differences in
assessment results among these procedures.

SUMMARY OF THE NEW PROVISIONS

The new assessment provisions are intended to be practical, efficient and as accurate as possible. The speed
of the assessment procedure was established by allowing the building data (concrete strength, member
dimensions, detailing etc.) to be collected from the ground floor only unless vertical element discontinuity is
found. A number of methods are available for seismic assessment based on ground floor member dimensions
(Hassan and Sozen 1997, Gulkan and Sozen 1999, Yakut et al. 2006). These practical procedures, although
provide valuable information on the seismic vulnerability of a building stock, are not tailored to estimate the
expected seismic performance of a single building. On the other hand, nonlinear static procedures (FEMA 356,
ASCE/SEI 41-06, ACI 562-13) are not deemed essential by the committee to solely identify the buildings with
many critical deficiencies leading to unacceptable seismic performance. Hence, the committee employed the
linear elastic (equivalent lateral force or mode superposition) procedure with some major differences as
compared to that of the Turkish Earthquake Code (TEC 2007) and ASCE/SEI 41-06. The new provisions are
applicable to buildings having 8 stories or less with total building height less than 25 meters. It is noteworthy to
mention that the new provisions are not intended to be used for seismic retrofit as more detailed provisions are
given in TEC (2007). A brief summary of the seismic assessment procedure detailed in the new provisions is
given below:

1. A three dimensional model of the building is generated based on a detailed survey performed for the
critical floor (generally the ground floor) only. If preferred, a complete survey can be carried out.
Since the assessment is done for only columns and shear walls at the critical floor, material
properties and reinforcement detailing are determined for these members. At least five concrete core
samples shall be taken from the columns and walls to determine the concrete strength. Removing
the cover concrete for several columns and walls is required to determine the reinforcement details.
The complete building model may be obtained by the replication of the ground floor layout over the
building height for regular buildings. However, one must consider variations in height and

11.3
@Seismicisolation
@Seismicisolation
B. Binici et al.

irregularities in plan and elevation as defined in the TEC (2007) including the torsional irregularity,
plan irregularity, and possible discontinuity of vertical load bearing elements. If discontinuity of
vertical load bearing elements exists in a building the engineer must resort to obtaining the building
data from all floor levels.

2. Based on the results of linear elastic analysis (equivalent lateral load or response spectrum) under
the design response spectrum (given in TEC 2007) and using no response modification factor, the
bending moment demand capacity ratios (DCR) at member ends and interstory drift deformations
are determined. Moment capacities for columns are computed based on axial loads obtained from
G+nQ±E/6 load combinations (G: dead load, Q: live load, n: live load reduction factor, E:
earthquake load). In this way, over conservatism regarding expected earthquake induced axial loads
is avoided. As a further simplification in the new provisions, flexural stiffnesses are taken as 30
percent of EIg for beams and walls, and 50 percent of EIg for columns. Given the lack of accurate
modulus of elasticity information for each column in the buildings, dependence of flexural stiffness
on axial load ratio was ignored. Furthermore, sensitivity of results on case study buildings showed
almost no influence in the final decision upon using the proposed simplified rigidities.

3. For each column and wall at the critical floor, DCR and interstory drift deformations are compared
with the corresponding limit values. If the interstory drift ratios in any floor are higher than the ones
obtained in the ground floor then assessment of the members for the interstory drift ratio at that floor
is also carried out. If the column/wall does not satisfy either one of the limits, it is classified as
unacceptable. Depending on the number of unacceptable members, buildings are classified as
“critical” or “not critical.” A critical building represents a building that is expected to suffer heavy
damage or collapse under the design earthquake effect. Because of the inability of the linear elastic
analysis to allow for redistribution, some flexibility is provided on how many columns are allowed
to exceed their performance limits. When the average axial load ratio resulting from gravity loads in
the considered floor exceeds 0.65, none of the members are allowed to exceed their performance
limit to classify the building as "not critical, (NC).” When the average axial load ratio is less than or
equal to 0.1, columns/walls that carry up to 35 percent of the story shear are allowed to exceed their
performance limits in order to classify the building as NC. Linear interpolation is used to determine
the acceptable story shear ratio for intermediate average axial load ratios.

It is worth mentioning that the proposed provisions are applicable to low to mid rise residential buildings, in
which significant member size changes from floor to floor are not expected. For these buildings, possible minor
changes in the member dimensions do not lead to significant differences in the assessment results as evidenced
by the analysis of several typical residential buildings in Turkey during the process of calibration of the
provisions. Similar to ASCE/SEI 41-06 provisions based on linear elastic procedure, DCRs are used to compare
the demand with the performance limits for columns and walls. Additionally, interstory drift limit is used as
deformation based assessment parameter. When these limits are exceeded, the component is very likely to
experience significant strength degradation. Translating the expected damage into quantitative engineering
parameters is a quite challenging task, especially within the context of urban renewal as the outcome may mean
demolishing buildings under the current law in Turkey. The approach of the committee was to set the
performance limits between the life safety definition of ASCE/SEI 41-06 (i.e. 75% of the deformation level at
which 20% loss of lateral maximum strength occurs for a member) and axial collapse. Accurate estimation of
actual collapse of a column or wall member is very challenging, for which a number of recent studies were
conducted (for example Elwood 2003). The limit values proposed in the new provisions are given for various
groups of columns and walls similar to that of ASCE/SEI 41-06. In the new provisions, columns are classified
into three, walls are classified into two groups based on their expected failure modes. The performance limit
values were determined for each group based on the mean of experimental data and analytical simulations when
the expected failure modes are in the presence of low axial load and low shear force. For high axial load and/or
shear cases, limits were selected from the lower bound values. Some adjustments were made for certain
components in order to ensure that buildings that fully comply with the provisions of TEC (2007) are found as
NC.

The following methodology was employed while establishing DCR and interstory drift limits: First, ductility
and ultimate displacements of columns and walls were obtained from experimental results, empirical/mechanical
models or relevant guidelines (Solmaz 2010, Erguner 2009, Bae and Bayrak 2009, ASCE/SEI 41-06 Supplement

11.4
@Seismicisolation
@Seismicisolation
Identifying Buildings with High Seismic Risk Under Urban Renewal Law in Turkey

1 and Eurocode 8 were employed for this purpose). Ultimate deformations were taken as the displacement values
corresponding to 20 percent capacity drop beyond the peak in experimental results or collapse prevention
displacement limit when guidelines are considered. Afterwards displacement ductility was divided by a factor of
1.25 in order to obtain DCR limits, and ultimate inelastic displacement was divided by a factor of 1.4 in order to
obtain corresponding interstory drift ratio limits. In this way, inelastic deformation limits were converted to
elastic deformation limits to be used along with linear elastic analysis results as neither the TEC nor SCHRB
employ displacement coefficient while estimating the target elastic displacement demand.

PERFORMANCE LIMITS FOR COLUMNS

The columns are classified in SCHRB (similar to ASCE/SEI 41-06) based on the shear demand capacity
ratio and transverse reinforcement detailing of columns. Tables 1 and 2 show the variables used in the
classification of columns in SCHRB and ASCE/SEI 41-06. Three different possible failure modes are considered
in the classification of columns. Class A and type i columns are expected to fail in flexure, class B and type ii
columns are expected to fail in shear-flexure, and class C and type iii columns are expected to fail in shear. The
calculation of shear strength Vr and Vn are similar in SCHRB and ASCE/SEI 41-06 if one uses the ACI 318-11
shear strength equation. Although the approach in general is similar, the definition of Ve (Vp in ASCE/SEI 41-06)
is slightly different. In SCHRB, Ve is calculated as the smaller of shear force computed based on plastic hinges
forming at column ends and based on plastic hinges forming at beam ends. On the other hand Vp is determined as
the shear demand at flexural yielding of plastic hinges in ASCE/SEI 41-06.

Table 1. Classification of Columns in SCHRB

Ve/Vr s≤100 mm*, 135o hooks, Ash≥0.06×s×bk×(fcm/fywm)* All other cases


Ve/Vr ≤ 0.7 A B
0.7 < Ve/Vr ≤ 1.1 B B
1.1 < Ve/Vr B C
* within the plastic hinge zone

Table 2. Classification of Columns in ASCE/SEI 41-06

Transverse Reinforcement Details


ACI conforming details with
Closed hoops with 90° hooks Others
135° hooks
Vp/(Vn/k) ≤ 0.6 i* ii ii
0.6 < Vp/(Vn/k) ≤ 1.0 ii ii iii
1.0 < Vp/(Vn/k) iii iii iii
* To qualify for condition i, a column must have Av/bws ≥0.002 and s/d ≤ 0.5 within flexural plastic hinge region. Otherwise, the column shall
be assigned to condition ii.

Performance limits for class A columns according to SCHRB are given in Table 3. Comparisons of SCHRB
and ASCE/SEI 41-06 DCR limits are presented in Figure 1. SCHRB limits were determined based on studies by
Bayrak and Bae (2009) and Solmaz (2010) and further comparisons of performance limits of typical column
sections using ASCE/SEI 41-06 and Eurocode 8 (2005). It can be observed that for class A columns there is a
general agreement between the recommended values of SCHRB and ASCE/SEI 41-06. Columns with low Ve/Vr
ratios and/or "insufficient" confining steel reinforcement experience failure due to a combined effect of flexure
and shear with limited ductility. A database of 41 columns compiled by Erguner (2009), which is extracted from
a wider database reported by Berry et al. (2004), was employed to study the columns classified as class B in
Table 1. The demand-capacity-ratios and drift deformation limits of the columns from the database are plotted
along with the proposed limits in Figure 2. Recommendations were made as a function of axial load ratio and
transverse steel reinforcement amounts. The shear stress ratio (Ve/bdfctm) was not considered as a separate
parameter due to the wide scatter of the data that shows no clear dependency as shown in Figure 2. For class B

11.5
@Seismicisolation
@Seismicisolation
B. Binici et al.

columns, recommended values of DCR for low and high axial load ratios are based on mean and lower bound
values, respectively. Interstory drift ratio limits are based on lower bound values. It can be observed that the
number of experiments involving high shear force together with high axial force is very limited.

In SCHRB and ASCE/SEI 41-06 approaches, m-factors are used to represent DCR (moment demand
calculated from G+nQ±E loading divided by moment capacity) and to assess performance of the member. IN
SCHRB, drift ratios of the critical floor are also checked in addition to the DCR limits. The major differences
between the performance limits of SCHRB and ASCE/SEI 41-06 can be summarized as follows:

- Class “A” and Type “i" columns:


o In SCHRB: m-factors depend only on axial load level (NK / (fcmAc)) assuming that Class A
columns will exhibit sufficiently ductile response.
o In ASCE/SEI 41-06: m-factors depend on both (NK / (fcmAc)) and the transverse reinforcement
ratio within the confining zone of columns (ρ = Av / bws)

- Class “B” and Type “ii” columns:


o In SCHRB: m-factors depend on (NK / (fcmAc)) and (ρ = Av / bws).
o In ASCE/SEI 41-06: m-factors depend on (NK / (fcmAc)), (ρ = Av / bws), and (V / (bwd√f’c)).

- Class “C” and Type “iii” columns:


o In SCHRB: m-factors are directly taken as 1.0.
o In ASCE/SEI 41-06: m-factors depend on (NK / (fcmAc)), (ρ = Av / bws).

NK (P in ASCE/SEI 41-06) parameter is determined using different load combinations in SCHRB and
ASCE/SEI 41-06. NK is calculated using (G + Q + E / 6) load combination in SCHRB while ASCE/SEI 41-06
leads to consider the axial load as a force-controlled parameter and use (Eq. 3-19) in the calculation of NK which
is (G + Q + E / C1C2J). Use of different NK values leads to consideration of different Vr values (Vn in ASCE/SEI
41-06) in column classification, and use of different section moment capacities in the evaluation of each column.

Table 3. Performance Limits for Columns

a) Class A b) Class B

Nk /(fcm×Ag) DCRlimit (δ/h)limit Nk /(fcm×Ag) Ash /(s×bk) DCRlimit (δ/h)limit


≤ 0.1 5.0 0.03 ≤ 0.0005 2.0 0.01
≤ 0.1
≥ 0.6 2.5 0.0125 ≥ 0.006 5.0 0.03
≤ 0.0005 1.0 0.005
≥ 0.6
≥ 0.006 2.5 0.0075

11.6
@Seismicisolation
@Seismicisolation
Identifying Buildings with High Seismic Risk Under Urban Renewal Law in Turkey

Figure 1. Comparisons of SCHRB and ASCE/SEI 41-06 DCR and Interstory Drift Ratio Limits for Class A and
Type i Columns

Figure 2. Comparisons of DCR and Interstory Drift Ratio Limits with Experimental Results for Class B and
Type ii Columns

PERFORMANCE LIMITS FOR WALLS

The reinforced concrete walls are classified into two groups in SCHRB, whereas they are classified into
three groups in ASCE/SEI 41-06. Kazaz et al.(2012) compiled a database for shear walls. The compiled
experimental work indicated that, unlike columns, the number of experiments on walls failing in flexure-shear
and shear mode is rather limited. Hence, it was not deemed necessary to differentiate between these two failure
modes for walls. The classification for walls is done in terms of (Ve/Vr) and (Hw/lw) ratios in SCHRB, whereas
only (Hw/lw) ratio is decisive in classification in ASCE/SEI 41-06. Tables 4 and 5 show the classification of RC
walls. According to SCHRB, type A walls are controlled by flexural behavior; type B walls are controlled by
either flexure-shear or shear behavior.

11.7
@Seismicisolation
@Seismicisolation
B. Binici et al.

Table 4. Classification of Walls in RBTY Table 5. Classification of Walls in ASCE/SEI 41-06

Hw/lw Ve/Vr < 1.0 1.0 ≤ Ve/Vr Hw/lw Controlled by


2.0 ≤ Hw/lw A B Hw/lw < 1.5 Flexure
Hw/lw < 2.0 B B 1.5 ≤ Hw/lw ≤ 3.0 Flexure and Shear
Hw/lw > 3.0 Shear

Similar to the columns, m-factors and interstory drift ratios are used in the determination of acceptance
criteria of RC walls. If the lateral load carrying capacity of shear walls are above 50 percent of the total base
shear demand and the maximum interstory drift ratio demand is less than 0.75, such buildings can be evaluated
solely based on interstory drift ratio limits. The parameters used for each wall class can be summarized as
follows:

- Walls controlled by flexure: The same parameters control the m-factors in both guidelines. These
parameters are axial load, shear, and confinement in boundary of walls. However, the effect of axial load in m-
factors is slightly different in ASCE/SEI 41-06. Whereas the effect of axial load is represented by NK in SCHRB,
it is formulated by ((As–A’s)fy + P) in ASCE/SEI 41-06. In other words, the effect of longitudinal reinforcement
contribution to the ductility on axial load level of wall is taken into account in ASCE/SEI 41-06.

- Walls controlled by shear: m-factor is a function of normalized shear (Ve / bwdfctm) in SCHRB whereas it is
a function of axial load (((As–A’s)fy + P)/Acf’c) in ASCE/SEI 41-06.

The interstory drift limits were determined for the same parameters as DCR limits. Table 6 presents the
performance limits in SCHRB for columns and walls. The numerical simulation results of Kazaz et al. (2012)
were employed along with the compiled test results for class A walls as shown in Figure 3.a. The selected
performance limits for low axial load levels correspond to 50 percent probability of being exceeded based on the
experimental data points. For high axial loads, DCR limits were selected as lower bounds to the simulation
results as there were no experimental data. However, such an approach was not considered for the interstory drift
ratio limits owing to the superior performance of buildings with significant structural walls under earthquakes.
Instead, they were determined considering the assessment results of the case study buildings. For class B walls,
performance limits were selected for a 30 percent probability of being exceeded and their dependency on axial
force was neglected (Figure 3.b). This can be justified on the ground of having low axial force levels on walls for
practical purposes and lack of experimental data for such structural elements. Additionally, numerical results of
Kazaz et al. (2012) indicated that dependence of the performance limits on axial load ratio for type B walls is not
significant within the practical limits of axial load ratios.

Table 6. Performance Limits for Walls

TEC
(2007)
Nk /(fcm×Ag) Ve /(b×d× fc) Compliant DCRlimit (δ/h)limit Ve /(b×d× fcm) DCRlimit (δ/h)limit
Boundary
Region
Yes 6.0 0.030 ≤ 0.9 4.0 0.020
≤ 0.9
No 4.0 0.015 ≥ 1.3 2.0 0.010
< 0.1
Yes 3.5 0.015
≥ 1.3
No 2.0 0.0075
Yes 3.5 0.020
≤ 0.9
No 2.0 0.010
> 0.25
Yes 2.0 0.010
≥ 1.3
No 1.5 0.005

11.8
@Seismicisolation
@Seismicisolation
Identifying Buildings with High Seismic Risk Under Urban Renewal Law in Turkey

a) Class A Walls

b) Class B Walls

Figure 3. DCR and Interstory Drift Ratio Limits for Walls

CASE STUDY BUILDINGS

Analysis results from three case study buildings are presented in this section. Building 1 is a 6-story
residential RC moment resisting frame building designed in accordance with the provisions of TEC (2007). The
floor plan is designed to represent typical buildings in Turkey. The height of first story is 3.5 m and the height of
other stories is 3 m. The story plan and three dimensional model of the building are presented in Figure 4.
Building 2 is also a 6-story residential building having RC moment frame-wall system designed according to the
code. The building has a commonly used configuration of shear walls in both principal directions. The story
height is 3 m for all story levels. The story plan and three dimensional model of the building are presented in
Figure 5. The third building is building 10 of case study buildings that is a 5-story RC moment frame-wall
system. This building experienced 1999 Düzce earthquake and suffered moderate damage. The story height is
2.8 m for all story levels. The story plan and three dimensional model of the building are presented in Figure 6.
Material and site properties of the buildings are summarized in Table 7.

11.9
@Seismicisolation
@Seismicisolation
B. Binici et al.

Figure 4. Building 1 Story Plan and Three Dimensional Model

Figure 5. Building 2 Story Plan and Three Dimensional Model

Figure 6. Building 3 Story Plan and Three Dimensional Model

11.10
@Seismicisolation
@Seismicisolation
Identifying Buildings with High Seismic Risk Under Urban Renewal Law in Turkey

Table 7. Material Properties of the Buildings

Material Properties Building 1 Building 2 Building 3


Concrete compressive cylinder strength, fc (MPa) 25 25 12
Tensile strength of reinforcing steel, fy (MPa) 420 420 220
Modulus of elasticity of concrete, Ec (MPa) 30250 30250 25260
Modulus of elasticity of reinforcing steel, Es (MPa) 200000 200000 200000

Linear dynamic procedure (modal analysis) is used for all case studies by employing mode superposition
method with sufficient modes that contribute 90 percent of the participating mass of the building in each
orthogonal horizontal direction. The design response spectrum for the highest seismic zone in Turkey, which is
defined in TEC (2007), is used for Building 1 and Building 2. The assessment of Building 3 is carried out under
the response spectrum of 1999 Düzce earthquake, as this building was moderately damaged under Düzce
earthquake.

Table 8. Number of Members Classified according to m/mlimit Results

Building 1 Building 2 Building 3


Provision Condition
Column Shear Wall Column Shear Wall Column Shear Wall
Acceptable 16 - 28 1 13 0
ASCE/SEI 41-06
Unacceptable 16 - 0 5 22 2
Acceptable 32 - 28 5 24 0
SCHRB
Unacceptable 0 - 0 1 11 2

Figures 7 to 9 present m/mlimit results of the ground floor from the +X direction analysis of Buildings 1, 2,
and 3, respectively. Table 8 summarizes the number of acceptable and unacceptable column and shear walls of
buildings in terms of m/mlimit results. Analyses for other directions were also conducted, but they are not
presented here for brevity as the conclusions are similar. For Building 1, ASCE/SEI 41-06 produces more
conservative results as compared to SCHRB for all columns. A similar conclusion can be reached upon
examining the results of Building 2. All columns were found as acceptable according to both provisions for this
building. However ASCE/SEI 41-06 estimations were very conservative for the seismic performance of shear
walls. SCHRB provisions classified Building 3 as NC according to the assessment based on interstory drift ratio.
ASCE/SEI 41-06 results, on the hand, estimated insufficient seismic performance for about half of the columns
and all of the shear walls. This result is again on the over safe side from a point of seismic assessment.

Two main reasons of the ASCE/SEI 41-06 in providing overly conservative results for the seismic
assessment are attributed to the following factors:

1- The m-factors provided in SCHRB are more liberal than those given ASCE/SEI 41-06.

2- When calculating the component capacities the associated axial loads are obtained under a load
combination with reduced earthquake force with a reduction factor of 6 in SCHRB. However in the same
combination, ASCE/SEI 41-06 leads to use of larger earthquake force with a reduction factor of (C1C2J)
which is usually found between 2 and 2.5. This situation leads to use of smaller component capacities under
higher axial loads.

3- Owing to the superior performance of building with shear walls, SCHRB provisions allow drift based
assessment when the drift demands are sufficiently small (i.e. less than 0.75%). Such a favorable
consideration is not present in ASCE/SEI 41-06 and one must always employ the m-factors for linear elastic
assessment.

11.11
@Seismicisolation
@Seismicisolation
B. Binici et al.

2.00
ASCE/SEI 41‐06  S: Column 
SCHRB 
m/mlimit
1.00

0.00 S1
S2
S3
S4
S5
S6
S7
S8
S9
S10
S11
S12
S13
S14
S15
S16
S17
S18
S19
S20
S21
S22
S23
S24
S25
S26
S27
S28
S29
S30
S31
S32
Member Names

Figure 7. Performance Comparison of Members according to ASCE/SEI 41-06 and SCHRB for Building 1

CONCLUSIONS

In this article, the technical provisions and the relevant background for the seismic assessment
under Urban Renewal Law in Turkey were presented. Special emphasis was given to member
ductility and deformation limits for use along with linear elastic assessment procedures. Afterwards
comparisons of assessment results on three case study buildings from the new provisions and
guidelines of ASCE/SEI 41-06 were discussed.
SCHRB provisions estimated that well designed buildings would exhibit better performance
levels than collapse prevention level. Similarly, for a building that experienced moderate damage
due to the presence of significant shear walls, the building was classified as NC meaning that it is
not expected to e collapse or suffer heavy damage. ASCE/SEI 41-06 estimated that for the well
designed buildings, many of the vertical load bearing elements would be in the collapse
performance state. A similar result was found from the examination of the third building. These
analyses results showed that ASCE/SEI 41-06 guidelines are on the over safe side as compared to
SCHRB provisions and observed building performance. When such overly safe building
performance estimations are made, retrofit costs tend to increase so high and discourage people
from employing seismic retrofit solutions. Apart from being conservative that is on the safe side,
another reason which may lead to overly conservative estimations may be the fact that these
guidelines were calibrated based on member deformation limits (from member tests) without
placing sufficient emphasis on structural system behavior and performance. The authors believe
that ASCE-SEI 41-06 approach can further be improved in the upcoming revision cycles from the
output of the research that reveals the uncovered link between the member and system
performance.

11.12
@Seismicisolation
@Seismicisolation
Identifying Buildings with High Seismic Risk Under Urban Renewal Law in Turkey

1.00
ASCE 41‐06
ASCE/SEI 41‐06  S: Column
m/mlimit
RBTY 2012

0.00
S1
S2
S3
S4
S5
S6
S7
S8
S9
S10
S11
S12
S13
S14
S15
S16
S17
S18
S19
S20
S21
S22
S23
S24
S25
S26
S27
S28
Member Names
a) Columns
2.00
P: Shear Wall
m/mlimit

1.00

0.00
P1

P2

P3

P4

P5

P6
Member Names
b) Walls
Figure 8. Performance Comparison of Members according to ASCE/SEI 41-06 and SCHRB for Building 2
2.00
S: Column ASCE 41‐06
ASCE/SEI 41‐06
P: Shear Wall RBTY 2012
m/mlimit

1.00

0.00
S1
S2
S3
S4
S5
S6
S7
S8
S9
S10
S11
S12
S13
S14
S15
S16
S17
S18
S19
S20
S21
S22
S23
S24
S25
S26
S27
S28
S29
S30
S31
S32
S33
S34
S35 Member Names
 
a) DCR 
4.00
Interstory Drift Ratio Demand S: Column 
3.50
3.00 Interstory Drift Ratio Demand acc. to SCHRB
P: Shear Wall 
2.50
IDR (%)

2.00
1.50
1.00
0.50
0.00
S1
S2
S3
S4
S5
S6
S7
S8
S9
S10
S11
S12
S13
S14
S15
S16
S17
S18
S19
S20
S21
S22
S23
S24
S25
S26
S27
S28
S29
S30
S31
S32
S33
S34
S35
P1
P2
P3
P4
P5
P6

Member Names
 
b) Interstory Drift Ratio

Figure 9. Performance Comparison of Members according to ASCE/SEI 41-06 and SCHRB for Building 3

11.13
@Seismicisolation
@Seismicisolation
B. Binici et al.

APPENDIX: NOTATION

Ash    Area of transverse reinforcement  
A g    Area of cross section 
bk    Width of core concrete  
b    Width of cross section perpendicular to shear force 
m, DCR   Demand capacity ratio 
mlimit, DCRlimit     Demand capacity ratio limit 
d    depth of cross section 
d b    Diameter of the tension reinforcement 
DRy    Yield interstory drift ratio 
DRu    Maximum interstory drift ratio 
EIg    Flexural rigidity of gross cross section 
EIeff    Effective flexural rigidity  
fcm, fc    Concrete compressive strength in MPa  
fywm    Yield strength of transverse reinforcement in MPa 
fy    Yield strength of longitudinal reinforcement in MPa 
fctm    Tensile strength of concrete in MPa 
h    Column depth  
Mu    Moment capacity 
Nk, N    Axial load 
N/N0    Axial load ratio 
R    Strength reduction factor 
s    Spacing of transverse reinforcement  
Ve, V    Flexural shear demand 
Vr     Shear strength 
    Effective stiffness factor 
(/h)sınır    Interstory drift ratio limit 
w    Wall length 
µ, µ    Displacement ductility  
ρs = Ash/bks   Transverse reinforcement ratio 

REFERENCES

ACI Committee 562, 2013, Code Requirements for Evaluation, Repair, and Rehabilitation of Concrete Buildings
(ACI 562-13) and Commentary, American Concrete Institute, Farmington Hills, MI 48331, U.S.A.

ACI Committee 318, 2011. Building Code Requirements for Structural Concrete (ACI 318-11), American
Concrete Institute, Farmington Hills, MI 48331, U.S.A.

American Society of Civil Engineers, 2000. Prestandard and Commentary for the Seismic Rehabilitation of
Buildings, Report No. FEMA 356. Reston, Virginia.

American Society of Civil Engineers, 2007. Seismic Rehabilitation of Existing Buildings, Report No. ASCE/SEI
41-06, Reston, Virginia, 428.

American Society of Civil Engineers, 2007. Seismic Rehabilitation of Existing Buildings, Report No. ASCE/SEI
41-06-Supplement 1, Reston, Virginia, 428.

Bae S, Bayrak O, 2009. Drift Capacity of Reinforced Concrete Columns, ACI Structural Journal, Vol. 106(4),
405-415.

11.14
@Seismicisolation
@Seismicisolation
Identifying Buildings with High Seismic Risk Under Urban Renewal Law in Turkey

Berry, M. P., M. Parrish, and M. O. Eberhard. 2004. PEER structural performance database user’s manual.
Pacific Earthquake Engineering Research Center. University of California, Berkeley, California.

CEN. Eurocode 8, 2005., Design Provisions for Earthquake Resistance of Structures, European Prestandard
ENV- Part 3: Assessment and Retrofitting of Existing Buildings, Comité Européen de Normalisation,
Brussels.

Elwood, J.K., 2003. Shake table tests and analytical studies on the gravity collapse of reinforced concrete
frames. Ph.D. dissertation, University of California at Berkeley, Berkeley, California.

Ergüner K., 2009. Analytical examination of performance limits for shear critical reinforced concrete columns,
M.S. Thesis, Middle East Technical University, Department of Civil Engineering, Ankara, Turkey.

Gulkan, P., Sozen, A.M., 1999. Procedure for Determining Seismic Vulnerability of Building Structures, ACI
Structural Journal 96(3), 336-342.

Hassan, A.F., Sozen, M.A., 1997. Seismic Vulnerability Assessment of Low-rise Buildings in Regions with
Infrequent Earthquakes, ACI Structural Journal, 94(1), 31-39.

Kazaz, I., Gülkan, P. and Yakut, A. 2012. Deformation limits for structural walls with confined boundaries,
Earthquake Spectra, 28 (3):1019–1046.

Ministry of Public Works and Settlement, 2007. Specification for Buildings to be Built in Seismic Zones (TEC
2007), Government of Republic of Turkey.

Middle East Technical University-EERC., 2011. 23 October 2011 Mw=7.2 Van-Ercis Earthquake
Reconnaissance Report, METU-EERC Report No:2011-04, METU, Ankara.

Ministry of Environment and Urbanization (MEU), 2012. The Urban Renewal Law for Regions under Disaster
Risk, Law No: 6306, Official Gazette: Date, 28(309), Volume 52, Turkey.

Pujol, S., Sözen, M., Ramírez, J., 2000. Transverse Reinforcement for Columns of RC Frames to Resist
Earthquakes, Journal of Structural Engineering, 126(4), 461-466.

Sezen, H., 2002. Seismic behavior and modeling of the reinforced concrete building columns, Ph.D. dissertation,
University of California at Berkeley, Berkeley, California.

Solmaz T., 2010. Evaluation of performance based displacement limits for reinforced concrete columns under
flexure, M.S. Thesis, Middle East Technical University, Department of Civil Engineering, Ankara, Turkey.

Yakut, A., 2004, Preliminary seismic performance assessment procedure for existing RC buildings, Engineering
Structures, 26(10), 1447-1461.

Yakut A., Ozcebe, G., Yucemen M.S., 2006. Seismic Vulnerability Assessment using Regional Empirical Data,
Earthquake Engineering & Structural Dynamics, 35(10), 1187–1202.

Yakut, A., Binici, B., Ozcebe, G., Ay, B.O., Akansel, V.H., 2012. Calibration and Validation of the Provisions
for the Seismic Risk Evaluation of Existing RC Buildings in Turkey under Urban Renewal Law, Report
Submitted to Ministry of Environment and Urbanization.

11.15
@Seismicisolation
@Seismicisolation
B. Binici et al.

11.16
@Seismicisolation
@Seismicisolation

You might also like