You are on page 1of 16

www.springer.

com/journal/13296
International Journal of Steel Structures
March 2014, Vol 14, No 1, 43-58
DOI 10.1007/s13296-014-1006-4
Steel Concentrically Braced Frames using
Tubular Structural Sections as Bracing Members:
Design, Full-Scale Testing and Numerical Simulation
Jiun-Wei Lai
1,
*
and Stephen A. Mahin
2
1
Postdoctoral Researcher, Department of Civil & Environmental Engineering, University of California, Berkeley, USA
2
Byron L. and Elvira E. Nishkian Professor of Structural Engineering,
Department of Civil & Environmental Engineering, University of California, Berkeley, USA
Director, Pacific Earthquake Engineering Research Center, USA
Abstract
This paper presents the results of experimental and analytical studies carried out on two full-scale, one-bay, two-story steel
concentrically braced frames. Square hollow and round hollow structural sections were used for the bracing components. The
specimens were designed and detailed according to the 2005 AISC Seismic Provisions, and tested cyclically under displacement
control but with a fixed lateral load distribution over height. Numerical computational models including the brace components,
gusset plate details and frame members were implemented in OpenSees. Numerical simulations were then performed to
investigate the cyclic behavior of brace components, brace failure mechanisms and overall system response. Satisfactory
agreement was obtained in comparisons of experimental and numerical results. Premature failures observed suggest that beam-
to-gusset plate connections could be pinned to accommodate large rotational demands at this location without the need to form
plastic hinges. Test results also showed that for the braced frames having the same configuration, designed for similar base shear
capacities, and subjected to the same roof level displacement history, the braced frame specimen using round tubular sections
as diagonal braces was able to sustain larger story drifts without brace fracture than the specimen employing square tubular
sections. Fracture of the column base in the second specimen, although inconclusive from a single test, suggests more study
is needed of design requirements for column to base plate connections where large variations of axial, bending and shear load
are expected.
Keywords: concentrically braced frames, hysteresis loops, low-cycle fatigue, hollow structural sections, experimental study
1. Introduction
Steel braced frames are considered one of the most
efficient and economical lateral load resisting systems
available to control deformations in civil structures under
wind or earthquake loading. The behavior of braced frames
in the elastic range, as expected during wind loading or
minor seismic excitations, has been quite good; however,
a review of the structural performance of steel braced
frames after several major earthquakes has identified
some anticipated and unanticipated damage (AIJ, 1995;
Bonneville and Bartoletti, 1996; Kelly et al., 2000). This
damage has prompted many engineers and researchers to
consider ways of improving the post-elastic behavior of
braced frame systems. One approach has been to increase
the inelastic deformability of the brace by using manufactured
elements, such as buckling restrained braces (Watanabe et
al., 1988) and self-centering braces (Christopoulos et al.,
2008). Others have carried out experimental studies (Lee
and Goel, 1987; Tremblay et al., 2002; Roeder and Lehman,
2008) or analytical (Uriz and Mahin, 2008, Chen, 2010)
investigations to assess the ability of braces having different
slenderness and compactness to increase the drift capacity
of braced frames. Another approach has been to reduce
deformation demands by strengthening the system as a
whole (Chen, 2010) or by mitigating local concentrations
of overall system deformation at one or a few stories
(Khatib et al., 1988; Tremblay and Merzouq, 2004).
Although many experimental studies of conventional
buckling brace components and several braced frame
specimens have been investigated in the past thirty years
(Black et al., 1980; Ballio and Perotti, 1987; Lee and
Note.-Discussion open until August 1, 2014. This manuscript for this
paper was submitted for review and possible publication on April 29,
2012; approved on December 7, 2013.
KSSC and Springer 2014
*Corresponding author
Tel: +1-510-965-7738
E-mail: adrian.jwlai@berkeley.edu
44 Jiun-Wei Lai and Stephen A. Mahin / International Journal of Steel Structures, 14(1), 43-58, 2014
Goel, 1987; Bertero et al., 1989; Tremblay, 2002; Roeder
et al., 2004; Yang and Mahin, 2005; Clark et al., 2008;
Roeder and Lehman, 2008; Uriz and Mahin, 2008), the
number of large-scale tests of complete concentric braced
frames needed to assess ultimate system behavior is still
limited.
Since the overall behavior of braced frames is sensitive
to relative proportions, strengths and stiffnesses of members
and local details, system tests are needed to assess fully
the adequacy of current code provisions and suggested
improvements. As such, integrated experimental and
analytical studies are carried out to examine likely seismic
behavior of representative concentrically braced steel frame
systems, and the ability of current analysis methods to
simulate this behavior. In this study, two full-scale, one-
bay, two-story steel concentrically braced frames were
constructed and tested under a series of cyclic lateral
displacement excursions increasing in amplitude up to a
maximum roof drift ratio about 4%. In a companion study
of code-compliant special concentric braced frames (Lai
et al., 2010), nonlinear dynamic analyses showed that
under the most severe hazard level considered (i.e. 2%
probability of exceedence in 50 years), the median
expected maximum story drift ratio is about 3.3%.
2. Design and Analysis of Frame Specimens
2.1. Selection of brace configuration and preliminary
analysis
As part of a George E. Brown, Jr., Network for
Earthquake Engineering Simulation (NEES) Small Group
Project entitled International Hybrid Simulation of
Tomorrows Braced Frame Systems, a coordinated set of
large-scale tests were carried out on various configurations
of concentrically braced frames representative of systems
used in western North America. These tests included two-
story and three-story double story X-braced frames tested
at the National Center for Research in Earthquake
Engineering (NCREE) in Taiwan (Clark et al., 2008;
Lumpkin et al., 2010), a three-dimensional frame tested
at the University of Minnesota (Palmer et al., 2010), and
two-story frames with diamond-shaped brace configurations
(with a V configuration in the lower story and an
inverted-V shape in the upper story) tested at University
of California, Berkeley. This paper focuses on two of the
Berkeley test specimens. The diamond brace configuration
used for these specimens was selected to complement
other tests carried out in the overall program and to focus
attention on floors where gusset plates are used to
connect braces to columns. Due to space and test rig
limitations, typical story height and beam span were
selected as 2,743 mm and 6,096 mm, respectively, as
shown in Fig. 1. In one specimen, square HSS sections
were used for the braces, while circular HSS sections
were used in the other specimen.
2.2. Design of test rig
Several configurations were considered and carefully
evaluated during the design of the test setup (Lai, 2009).
An overview of the final test setup is shown in Fig. 2.
Thirty reconfigurable reaction blocks (ten blocks per
stack) were grouted together and post-tensioned horizontally
and vertically over a strong floor to create an integrated
reaction wall. The maximum base shear capacity of the
test setup is 4,003 kN (with 2,669 kN applied at the upper
level and 1,334 kN at the lower level), assuming the load
applied by the lower actuator is half that acting in the
upper one. An Atlas 6,672 kN actuator with 304.8 mm
stoke was attached at each floor level. This stroke capacity
corresponds to about 5% of the total height of the test
specimen. A heavy built-up beam (Fig. 2) was provided
between the specimen and the laboratory strong floor to
spread out concentrated reaction forces at the base of the
specimen. Four additional stiff load transfer beams were
provided below the top flange of the cellular strong floor
to spread out the concentrated uplift forces. Out-of-plane
support of the test specimen was provided along the top
of both beams and at the beam-to-column connections by
a stiff and strong transverse support frame shown in Fig.
Figure 2. Overview of the final test setup.
Figure 1. Dimension of test specimen.
Steel Concentrically Braced Frames using Tubular Structural Sections as Bracing Members: Design, Full-Scale Testing ... 45
2. The lateral supports were designed to move longitudinally
with the test specimen. Details of the test setup design are
summarized in a report by Lai (Lai, 2009).
2.3. Design of braces
For the ideal condition where the top floor is subjected
to twice the lateral force as the lower floor, the braces
were selected to protect the test setup such that story
shears would not exceed 4,003 kN and 2,669 kN in the
lower and upper stories, respectively. By assuming the
braces resist about 80% of the maximum permitted story
shear, using a system over-strength factor of 2 as
stipulated in the AISC Seismic Provisions (AISC, 2005b)
and considering the geometry of the bracing system, the
maximum permitted brace forces were estimated. Detailing
requirements from the AISC Specification (AISC, 2005a)
and AISC Seismic Provisions (AISC, 2005b), such as
limitations on brace slenderness ratio and width-thickness
ratio, were imposed to finalize the brace size and the detailing
of the specimen. For instance, net section reinforcing
plates designed in accordance with AISC Seismic Provisions
(AISC, 2005b) were welded on all tubular brace connection
regions, as shown in Fig. 3, to prevent possible premature
fractures (Yang and Mahin, 2005).
2.4. Design of columns
A tributary gravity load (computed based typical design
dead and live loads for office occupancies and a tributary
floor area of 18.6 m
2
) was included in the column design.
The column axial force due to overturning was estimated
by dividing the maximum overturning moment permitted
by the test set up at the base level by the moment arm
between the two columns. Accordingly, the estimated
axial force was 214 kN from the tributary gravity loading
and 3,003 kN from overturning. The demands of bending
moment in the columns were calculated from structural
analysis. Since the specimen was only two-stories tall, a
single piece column extended over both stories. Because
tributary gravity loads contributed only a minor portion of
the total column axial load, these loads were not imposed
during the tests.
2.5. Design of beams
The design approach used for the upper and lower
beams differed slightly. The roof beam was designed
assuming a maximum axial force of 2,669 kN (corresponding
to the maximum force permitted in the upper actuator).
The bending moment demands for the top beam were
extracted from structural analysis results, considering a
vertical unbalanced force produced by the two braces
intersecting at the beams midspan, as required by the
AISC Seismic Provisions (2005b). For the lower beam,
the maximum axial force was calculated to be 1,415 kN
under the loading conditions shown in Fig. 4 (for
Specimen TCBF-B-1). Bending moments in the lower
beam were calculated from structural analysis. No vertical
unbalanced load was required for the lower beam level
due to the brace configuration.
2.6. Design of gusset plates
All gusset plates were designed for out-of-plane buckling
considering the uniform force method suggested in AISC
Specification (AISC, 2005a). Typical details used (Fig. 5)
incorporated 30-degree tapers and a plastic hinge fold gap
equal to twice the thickness of the gusset plate. To explore
the possibility of simplifying field erection, a single piece
gusset plate was attached to each column at the lower
level. This plate was intended to be welded to the column
in the shop, and to the beam and braces in the field. Two
finger plates were shop welded to each of these gusset
plates to simulate the flanges of a beam continuing to the
column face. The fingerplates extend slightly beyond the
ends of the gusset plate tapers (Fig. 6) to facilitate welding
of the beam to the gusset plate.
2.7. Design of connections
All connections were designed considering the maximum
probable force in the interface using the capacity design
approach outlined in the AISC Seismic Provisions (2005b).
2.8. Design of base plates
Base plates and end plates were designed according to
the AISC steel design guide (Fisher and Kloiber, 2006).
Table 1 as well as Figs. 7 and 8 summarize the final
Figure 4. The load condition for lower beam.
Figure 3. The net section reinforcing plate.
46 Jiun-Wei Lai and Stephen A. Mahin / International Journal of Steel Structures, 14(1), 43-58, 2014
member sizes and the material types used. The members
satisfy 2005 AISC requirements for compactness, as can
be noted in Table 2, and very nearly satisfy those in the
draft 2010 edition for highly ductile members. In interpreting
results, it is useful to note that the round HSS brace
sections selected satisfy the compactness requirement by
Figure 7. Member sizes of test specimen TCBF-B-1.
Figure 8. Member sizes of test specimen TCBF-B-2.
Figure 5. Typical gusset plate detail.
Figure 6. The one-piece gusset plate.
Table 1. Name, member size, and material type of the specimen components
Name
Column & Beams Braces
Section and Material b/t (h/t
w
) Section and Material b/t (D/t) kL/r
TCBF-B-1
W1296 (Column)
W24117 (Roof Beam)
W2468 (Lower Beam)
(ASTM A992)
6.76 (17.7)
7.53 (39.2)
7.66 (52.0)
HSS 555/16
HSS 663/8
(ASTM A500B)
14.2
14.2
51
47
TCBF-B-2
HSS 51/2
HSS 61/2
(ASTM A500B)
10.8
12.9
60
55
Table 2. AISC Seismic Provision Limitations
Seismic Provision
Limitations
Wide Flange HSS
Flange (b/t) Web (h/t
w
)

Square (b/t) Round (D/t)
2005
Seismic
Compactness

ps
=7.22

ps
=52.6 (Column, C
a
=0.38)

ps
=52.3 (Roof Beam, C
a
=0.39)

ps
=53.4 (Lower Beam, C
a
=0.35)

ps
=16.1
ps
=27.7
2010
Highly Ductile

hd
=7.22

md
=9.15

hd
=47.3 (Column, C
a
=0.38)

md
=52.6 (Column, C
a
=0.38)

hd
=47.1 (Roof Beam, C
a
=0.39)

md
=52.3 (Roof Beam, C
a
=0.39)

hd
=47.8 (Lower Beam, C
a
=0.35)

md
=53.4 (Lower Beam, C
a
=0.35)

hd
=13.8

md
=16.1

hd
=24.0

md
=27.7
Moderately
Ductile
Steel Concentrically Braced Frames using Tubular Structural Sections as Bracing Members: Design, Full-Scale Testing ... 47
a far greater margin than do the square sections. All wide
flange beams, columns and braces in the specimens were
ASTM A992 steel sections. All braces were made of
ASTM A500 Grade B steel. The 19 mm thick gusset
plates, 51 mm thick base plates, 51 mm stub beam end
plates, 13 mm shear tabs, 16 mm finger plates, 16 mm
continuity plates, 10 mm washer plates for all-thread
anchor rods and brace reinforcing cover plates were made
of ASTM A572 Grade 50 steel plate. Beam web stiffener
plates, lifting lugs, shim plates and miscellaneous parts
were made of ASTM A36 steel plates. High strength
structural fasteners that satisfy the ASTM A490 standard
were used at beam-column connections and the one-piece
gusset plate-to-beam web splices. High strength, all-thread
anchor rods (ASTM A193 Grade B7) were used at
column bases and gusset-to-floor beam base plates.
Figures 9(a) and 10(a) show the photo of specimen
TCBF-B-1 and TCBF-B-2 before testing. To reduce
overall cost of specimen fabrication, the top beam and
two columns in Specimen TCBF-B-1 were reused in
Specimen TCBF-B-2.
3. Loading History
The displacement of the roof beam was monitored and
used to control the overall motion of the specimens. The
lower level actuator was force controlled to have half of
the load applied at the upper level. The resulting lateral
force pattern is a typical inverted triangular distribution.
The test protocol was adapted from the Appendix T of
the AISC Seismic Provisions (AISC, 2005b). An additional
six cycles corresponding to one half of the elastic design
drift (0.5D
be
) and two cycles at the elastic design drift
(D
be
) were added to the beginning of the test protocol.
Figure 11 shows a plot of the cyclic test protocol in terms
of roof displacement and roof drift ratio. Specimen
movement towards the east side of the laboratory (Fig. 1)
is defined as a positive displacement.
During the test process, loading was paused to document
major events, such as brace fracture, weld cracking, or
significant yielding or local buckling of members. Testing
was terminated following any cycle in which both braces
at a single story were completely fractured.
Figure 9. Specimen TCBF-B-1.
Figure 11. Loading protocol for Specimens TCBF-B-1 and TCBF-B-2.
Figure 10. Specimen TCBF-B-2.
48 Jiun-Wei Lai and Stephen A. Mahin / International Journal of Steel Structures, 14(1), 43-58, 2014
4. Instrumentation
More than two hundred instruments, including linear
strain gages, strain gage rosettes, linear variable differential
transformers (LVDTs), wire pots and tilt-meters were
installed and monitored throughout the entire test. The
specimen was whitewashed to help in visual detection of
yielding and damage.
Multiple time lapse digital images were taken using
high-resolution digital single-lens reflex (DSLR) cameras
and stored using desktop computers. Four Canon EOS 5D
Mark-II DSLR cameras were connected to the data
acquisition system and triggered to take still photos every
5 mm of roof displacement. Three Canon EOS D1 DSLR
cameras were connected to desktop computers and shot
still images every 10 seconds continuously throughout the
tests. Additional high definition (HD) videos captured
global and certain local behavior of individual braces. A
three-dimensional Leica HD laser scanner was also used
to capture the specimens deformed shape throughout the
cyclic testing.
5. Test Results
In many respects, Specimens TCBF-B-1 and TCBF-B-
2 behaved as expected. The braces all buckled out-of-
plane (see Figs. 9(b) and 10(b) for details) as intended.
The net section reinforcing plates at the brace-to-gusset
plate connections achieved their purpose, with no weld or
other yielding or fractures noted in these areas. In the
gusset plates, the expected yield pattern was developed
within the 2t fold line width provided (Figs. 12(a) and
12(b)). However, as discussed later, significant damage
occurred at the ends of the lower level beam and at the
base of the columns that limited the ultimate displacement
capacity of the specimens. A more detailed description of
the testing process and key observations is provided by
Lai (Lai, 2009).
5.1. Overall behavior
The peak actuator forces (Figs. 13 and 14) throughout
the test in both experiments were all less than the maximum
forces permitted for each actuator, which were 2669 kN
for upper level actuator and 1334 kN for lower level
actuator. Thus, no excessive overturning moment was
developed during the test that would overload the test
setup or reaction floor. These figures illustrate that the
degradation of the peak story shear forces occurs rapidly
once the crack in the brace initiated (Figs. 13 and 14) and
also points out that the peak story shear forces degrade
more slowly with cycling for Specimen TCBF-B-2 than
for Specimen TCBF-B-1.
Specimen base shear versus roof displacement hysteresis
loops are presented in Figs. 15 and 16. Loops for both
specimens are fairly stable and repeatable during the early
cycles. However, there is a reduction of load capacity
during repeated cycles to the same displacement. This
cyclic strength deterioration increases with increasing
displacement amplitude.
A more detailed assessment of the hysteretic characteristics
of the specimens can be made by examining the relation
between story shear and story deformations (or story
drifts). Figures 17 and 18 indicate that lateral drifts were
larger in the lower story, especially for Specimen TCBF-
B-1. Interestingly, the individual story deformation excursions
are not symmetric even though the roof displacements are
symmetrically applied. This is mainly associated with the
unequal distribution of story drifts due to the weak story
mechanism that occurs once the braces buckle. The
distribution of drift is also influenced to a smaller extent
by the configuration of the braces and the way that
actuators are attached to the specimen.
5.2. Behavior of braces
Brace axial force versus axial deformation hysteresis
loops are plotted in Figs. 19 and 20. All four braces in
both specimens yielded in tension and buckled in
compression. Since load cells were not installed in the
braces, brace axial forces were estimated in several ways
using strain gauge measurements from portion of the
braces and adjacent elements that remained essentially
elastic in conjunction with elementary mechanics and
equilibrium considerations. The different methods considered
produced similar results and best estimates are plotted
(see Lai and Mahin, 2013 for more details). Minor
deformation hardening during inelastic elongation of the
tubular braces is noted in all four braces in both
specimens. The peak compression strength decreased
Figure 12. The yield patterns in the first story gusset plates after testing for both specimens.
Steel Concentrically Braced Frames using Tubular Structural Sections as Bracing Members: Design, Full-Scale Testing ... 49
significantly from cycle to cycle. The braces buckled
laterally, and formed plastic hinges in the gusset plates
(Fig. 12) and at midspan. The midspan plastic hinge
locally buckled on the compression-most (inside) face of
the section. With increased cycling and deformation,
rupture initiated in the locally buckled region and
propagated during further cycling until complete fracture
of the section occurred. This behavior is representative of
that observed in past component tests (Black et al., 1980;
Lee and Goel, 1987; Tremblay, 2002; Yang and Mahin,
2005). At the end of tests, both braces in the lower story
of both specimens had completely fractured, while the
braces in the upper story had not. A partial fracture was
observed in the upper story east brace in Specimen
TCBF-B-1 (upper part of Fig. 19). No cracks initiated in
the upper story braces in Specimen TCBF-B-2.
The out-of-plane displacements at the mid-span of each
brace are shown in Figs. 21 and 22. The maximum brace
out-of-plane displacements for Specimen TCBF-B-1 were
closed to 380 mm in the lower story and 240 mm in the
upper story. For Specimen TCBF-B-2, the maximum
brace out-of-plane displacements were close to 500 mm
in the lower story and 340 mm in the upper story. These
brace out-of-plane displacements were as large as about
eight times the axial deformations recorded in the braces
(Figs. 19, 20, 21 and 22). This potential out-of-plane
deformation should be considered when braces are placed
near safety related nonstructural components, such as
stairways, or cladding elements that may pose a falling
hazard.
Figure 21 and the lower part of Fig. 22 show that during
the last cycles of response out-of-plane displacements
were developed at the peak tension loading. In earlier
Figure 15. Base shear vs. roof displacement relationship
for Specimen TCBF-B-1.
Figure 16. Base shear vs. roof displacement relationship
for Specimen TCBR-B-1.
Figure 13. Actuator force and story displacement time
history for Specimen TCBF-B-1.
Figure 14. Actuator force and story displacement time
history for Specimen TCBF-B-2.
50 Jiun-Wei Lai and Stephen A. Mahin / International Journal of Steel Structures, 14(1), 43-58, 2014
cycles, the out-of-plane displacements generally returned
to zero as the brace was loaded in tension. However, once
ruptures initiated at the midspan of a brace, the center of
the remaining material at this section shifted from the
mid-depth of the section. As a result of this eccentricity,
the brace displaced laterally in the direction opposite in
sign from that occurring during the compression phase of
the cycle. This local bending at the ruptured section
during the tension phase of a cycle is believed to accelerate
the complete fracture of the section. This phenomenon
can be observed in both experiments where the brace
fractured or cracked in the specimen (Fig. 21 and the
lower part of Fig. 22).
5.3. Behavior of columns, beams and connections
During cycles with small lateral displacement, the axial
forces in the east- and west-side columns were nearly of
equal magnitude but opposite in sign. This can be seen in
Figs. 23 and 24. Once braces buckled and began to loose
compression capacity, the internal forces in the frames
redistributed so that the tension braces still developed
their full tensile capacity. The unbalance between the
peak compression and tensile forces developed in the
braces resulted in the peak compression forces in the
columns becoming far greater than the peak tensile forces.
The column axial compression force drop gradually with
the deterioration of the compression capacity of the
contiguous brace, and rapidly as the tension brace begins
to rupture. Even with the complete fracture of both braces,
the remaining beams and columns act as a moment-
resisting frame, and some variation of column axial loads
occurred during continued lateral displacement.
Note that in TCBF-B-2 specimen, the CJP weld
connecting on the outermost flange of the west column to
the thick base plate fractured suddenly in the early
portion of the planned test protocol; at a roof drift ratio of
about 0.9%. The fracture initiates in the vicinity of the
heat-affected zone in the base plate and is shown in Fig.
25. The column base then underwent emergency repair
using a large cover plate on the fractured flange (which
Figure 18. Story shear vs. story drift relationship for
Specimen TCBF-B-2.
Figure 17. Story shear vs. story drift relationship for
Specimen TCBF-B-1.
Steel Concentrically Braced Frames using Tubular Structural Sections as Bracing Members: Design, Full-Scale Testing ... 51
also suffered a small amount of local buckling) and
stiffeners around the column base as shown in Fig. 26.
The modification of the column base for the west-side
column changed that columns stiffness and moved the
potential plastic hinge location upwards.
The strong axis bending moment in the columns at top
and bottom ends in each story were estimated from
readings of strain gages located away from potential plastic
hinge regions. These readings reveal that the columns at
both stories in Specimen TCBF-B-1 remained essentially
elastic throughout the entire experiment, except at the
column bases. Minor flaking of whitewash was also
noted at the column bases. During the TCBF-B-2 test,
significant yielding (identified through Lders lines seen
in Figs. 25 and 27) occurred in the lower story columns
and also at the bottom end of the west side column in the
upper story (Fig. 28). The peak base shear and column
axial loads in this specimen were slightly higher than
those in Specimen TCBF-B-2. Fracture may have been
also associated with the cumulative effects of prior
inelastic actions that occurred at the column base during
testing of TCBF-B-1. However, as seen in Fig. 25, the
pattern of yield lines, and observed distortion of the
specimen during the tests, suggests that the column base
was subjected to torsional and out-of-plane bending as a
result of the eccentric transfer of forces from the buckled
braces to the column.
The total column shear forces estimated at each story
Figure 19. Estimated brace axial forces vs. brace axial deformations for Specimen TCBF-B-1.
Figure 20. Estimated brace axial forces vs. brace axial deformations for Specimen TCBF-B-2.
52 Jiun-Wei Lai and Stephen A. Mahin / International Journal of Steel Structures, 14(1), 43-58, 2014
for both specimens are shown in Figs. 29 and 30. Almost
symmetric shear force response was found in Specimen
TCBF-B-1, while un-symmetric response was found in
Specimen TCBF-B-2 after the installation of the repair. It
is believed that this asymmetry is due to local strains in
the gages near the repair exceeding yield (Fig. 27(b)).
From the relationships plotted between total column shear
force and story shear force in Figs. 31 and 32, the
columns took at the beginning of the tests about 17% of
total story shear at both stories. After the braces buckled,
the columns took a greater and greater portion of the total
story shear. A slope of unity in these plots suggests that
the columns took the entire story shear, as is observed for
the later cycles for the lower story. It is also interesting to
note that the lower story column webs yielded for both
specimens, as identified from significant flaking of white
wash and from shear strain values derived from rosette
gages installed on the columns web. The webs in the
upper level columns remained elastic. The distribution of
whitewash flaking can be seen in Figs. 27(b) and 28(a)
for the west-side column.
The top-level beam behaved as expected. Vertical
deflections developed at the mid-span of the beam. For
both specimens, the peak deflection was about 5 mm (less
than 1/1000 beam span) and the top beam remained elastic
(confirmed by strain gage readings). The maximum
unbalanced force applied to the upper beam was estimated
from strain gage readings on the adjacent beams to be
Figure 21. Estimated brace axial forces vs. brace out-of-plane displacements for Specimen TCBF-B-1.
Figure 22. Estimated brace axial forces vs. brace out-of-plane displacements for Specimen TCBF-B-2.
Steel Concentrically Braced Frames using Tubular Structural Sections as Bracing Members: Design, Full-Scale Testing ... 53
690 kN and 835 kN for Specimens TCBF-B-1 and TCBF-
B-2, respectively. These are smaller than design values
based on AISC Seismic Provisions (AISC, 2005b) of 830
kN and 960 kN. Note that the actual R
y
value computed
for the square and round HSS braces used were 1.2 and
1.3, respectively, (which are less than 1.4 specified in
AISC Seismic Provisions for ASTM A500B steel). In
addition, the observed residual post-buckling compressive
strengths were higher than 30% of the braces initial
compressive strengths. These factors reduced the unbalanced
loads resisted by to the top beam. Since no braces were
connected to the midspan of the lower beam, this beams
vertical deflection was quite small for either specimen.
However, plastic hinges began to form at both ends of
the lower beam in both specimens at a drift of about
1.25%, and with further cycling, significant local buckling
rapidly developed in both specimens at the top and
bottom beam flanges (and to a significant, but lessor
extent, to the beam web) adjacent to the gusset plates
(Figs. 33 and 34). The estimated axial forces in the lower
beam derived from strain gage readings at different
locations indicate that a maximum axial force of about
1334 kN developed during the tests in both compression
and tension. This compares to values of 1415 kN and
1350 kN in compression used in design of Specimens
TCBF-B-1 and TCBF-B-2, respectively (Fig. 4). Based
on simple equilibrium, one would expect tensile loads in
this beam to be quite small. However, beam plastic hinging
and local buckling resulted in a net overall shortening of
the beam when it was subjected to compression loads. To
maintain compatibility with the rest of the specimen,
significant transient and residual tension forces were
developed in the beam. Nonetheless, the 2005 AISC
Seismic Provisions do not provide any direct guidance on
the need to design such beams for axial load. Due to the
severe local buckling of the lower beams plastic hinge
regions, one of the flanges fractured on each end of the
beam. This occurred in both specimens at a drift ratio of
about 2%.
The one-piece gusset plate details used here may have
aggravated the initiation and severity of the flange local
bucking due to the weld access holes provided between
the beam flange and web at the end of the finger plate
stiffeners. Once one of the flanges fractured, further
rotation at this location was adequately accommodated by
yielding and local buckling of the web and bending of the
Figure 23. Roof displacement vs. column axial forces for
Specimen TCBF-B-1.
Figure 24. Roof displacement vs. column axial forces for
Specimen TCBF-B-2.
Figure 25. West column base weld fractured during the
first trial for Specimen TCBF-B-2.
Figure 26. West column base after repair for Specimen
TCBF-B-2.
54 Jiun-Wei Lai and Stephen A. Mahin / International Journal of Steel Structures, 14(1), 43-58, 2014
remaining flange. This suggests that a simple pin connection
at this location might be acceptable.
5.4. Panel zone and gusset plates
No web doubler plates were provided in the panel
zones. In both specimens, the lower floor level panel
zones were attached to large gusset plates and did not
yield, while the panel zones in the roof beam-column
connections yielded in both tests (detected by whitewash
flaking and readings from strain rosettes). This inelastic
deformation in the upper level panel zone was compatible
with the plastic hinging noted at the ends of the lower
level beam. No significant adverse effects on system
behavior of either specimen resulted from this yielding.
The one-piece-gusset plates formed a yield line,
perpendicular to the axis of the brace yielded as expected,
near the 2t yield line region in the first and second story
gusset plates (Figs. 12, 33 and 34). In the TCBF-B-2 test,
eleven linear strain gages were attached on one face of
Figure 28. Second story column yield patterns for
Specimen TCBF-B-2.
Figure 27. Column base yield patterns for Specimen
TCBF-B-2.
Figure 29. Time history of the sum of column shear
forces in both stories for Specimen TCBF-B-1.
Figure 30. Time history of the sum of column shear
forces in both stories for Specimen TCBF-B-2.
Figure 31. Story shear component from two columns vs.
total story shear forces for Specimen TCBF-B-1.
Figure 32. Story shear component from two columns vs.
total story shear forces for SpecimenTCBF-B-2.
Steel Concentrically Braced Frames using Tubular Structural Sections as Bracing Members: Design, Full-Scale Testing ... 55
the one-piece-gusset plate located on the east side of the
first story. These were used to monitor the axial stain
distribution in the tapered region of the gusset plate.
Figure 35 shows plots of the difference in strains on
corresponding locations on opposite sides of the gusset
plate. The increase in these strains with distance from the
centerline of the brace suggests that the braces contribute
to in-plane frame action.
6. Numerical Simulation of Test Results
In the computational phase of this study, a low-cycle
fatigue sensitive material model (Uriz and Mahin, 2008)
implemented in the computer analysis framework OpenSees
(McKenna, 1997) was used. All wide flange beams,
columns and tubular braces were modeled using force-
based nonlinear beam-column elements with initial
transverse imperfections (about 1/1000 of the member
length). Corotational geometric transformations were used
to capture member geometric nonlinearities under large
deformation (i.e., to simulate member lateral buckling).
Fiber sections were used to model the geometric cross
sectional shape of members. The total number of
elements used to represent each brace was 24, and 64
fibers were used to represent each square HSS and round
HSS, as shown in Fig. 36(a). The uniaxial Giuffre-
Menegotto-Pinto steel material with initial elastic tangent
of 200 GPa and a strain-hardening ratio of 0.003 was
used. Yield strength of the steel material was based on the
mill certificate report. Rigid end zones were assumed in
the model connections and gusset plates. Figure 36(b)
schematically illustrates the two-story braced frame
specimen model used with OpenSees.
A series of material calibration trials were done to
identify low-cycle fatigue and other properties for the
braces using existing experimental data on tubular braces
(i.e., Yang and Mahin, 2005). Two typical calibration
results under different loading histories are shown in Figs.
37 and 38. From the plots we can clearly see that the
overall cyclic behaviors in the component tests are
Figure 35. The bending strain time history in the tapered gusset plate at eastern side of Specimen TCBF-B-2.
Figure 33. Plastic hinges formed in the lower beam of
Specimen TCBF-B-1.
Figure 34. Plastic hinges formed in the lower beam of
Specimen TCBF-B-2.
56 Jiun-Wei Lai and Stephen A. Mahin / International Journal of Steel Structures, 14(1), 43-58, 2014
captured well by the OpenSees low cycle fatigue model.
Using the input parameters obtained from the calibration
results, analyses of the two-story specimens were performed
using the recorded top-level displacement history as
Figure 36. Illustration of OpenSees Two-Story Braced Frame Model.
Figure 37. Square HSS brace axial force-axial displacement
relationships (test vs. OpenSees).
Figure 38. Round HSS brace axial force-axial displacement
relationships (test vs. OpenSees).
Figure 39. Test results vs. OpenSees cyclic pushover
results for Specimen TCBF-B-1.
Figure 40. Test results vs. OpenSees cyclic pushover
results for Specimen TCBF-B-2.
Steel Concentrically Braced Frames using Tubular Structural Sections as Bracing Members: Design, Full-Scale Testing ... 57
input. The numerical results obtained are compared with
the actual test results for both specimens in Figs. 39 and
40. It is clear that the overall behaviors of two braced
frame specimens are well predicted by the simulation
models, including the deterioration and member failure.
7. Conclusions
Based on the experimental and analytical results
presented, several conclusions can be drawn. It should be
recognized, however, that these are based on two
specimens, each of which includes unique details. As
such, these conclusions may not apply to other brace
configurations, other braces having different slenderness
and compactness ratios, or material properties.
When designed for similar base shear capacity and
subjected to the same loading protocol, the specimen
(TCBF-B-2) with braces fabricated from round tubular
sections exhibits better displacement capacity than the
more traditional specimen (TCBF-B-1) that used square
tubular sections as braces. The peak base shear developed
in inelastic cycles was also observed to degrade slower
and local buckling of the braces occurred later in
Specimen TCBF-B-2 specimen under the same test
sequence. The particular round HSS braces used satisfied
AISC requirements for compactness by a far greater
margin than did the square HSS braces. This selection of
a size for the hollow round brace was a consequence of
using a small diameter circular section with a strength
and slenderness similar to that used for the square HSS
brace.
For the inverted triangular pattern of lateral load used
in design and imposed during the tests, story drifts
tended, once brace buckling initiated, to be larger in the
lower level than in the upper story. This concentration of
damage is associated with braces at each level buckling at
different times and degrading at different rates during the
test. This difference is a consequence of the finite number
of sizes from which to pick bracing members satisfying
design requirements, and small differences between the
boundary and initial conditions for the as-built specimens
and the analytical models used in design. Because the
story with the greatest lateral displacement tends have its
strength deteriorate more, damage tends to concentrate
eventually in a single story. In both specimens, selection
of brace sizes as close to the expected demands resulted
in initial lateral buckling of braces in both stories.
However, because of the slower deterioration of strength
exhibited by the compact round HSS braces, the distribution
of lateral displacement for Specimen TCBF-B-2 was
more uniform than for Specimen TCBF-B-1.
The gusset plate details performed as intended. However,
at large lateral displacements, frame action resulted in
plastic hinges at the faces of the gusset plate used to
attach the beam to the column. The single-piece gusset
used worked well, but the complexities of the connection
of the beam to the gusset plate may have accelerated local
buckling and fracture of the beam at this location. A
pinned beam-to-gusset detail might be advisable to avoid
such local damage. Test results indicate large axial forces,
including large tensile forces, develop in the beams as a
result of the unbalance in horizontal force components of
the braces joining at the ends of the beam and shortening
of the beam plastic hinge regions. More study is needed
to better understand the required axial strength and
stiffness of beams in concentric braced frames.
The columns from Specimen TCBF-B-1 were reused
for Specimen TCBF-B-2. The failure at one of the
column base welds may be an indication of inadequate
workmanship, or possibly low-cycle fatigue. The behavior
of this particular connection is complicated due to the
absence of a gusset plate at the base of the column to
assist in transferring axial load to the base plate, and the
presence of torsional and out-of-plane bending in the
column associated with the out-of-plane buckling of the
braces at the level above. More study is warranted.
The numerical model implemented in OpenSees was
able to match the experimental results with considerable
accuracy. The key to this modeling was the use of fiber-
based models that captured the brace hysteretic loops
well, and the use of an empirically calibrated low-cycle
fatigue model to capture the progressive rupture the
bracing elements. Such fiber models do not directly
account for the effects of local buckling, and their use
requires careful validation using relevant test results.
Acknowledgment
This research is supported by National Science
Foundation (NSF) under grant number CMMI-0619161.
Financial Support from NSF is greatly appreciated. The
conclusions and opinions expressed in this paper are only
those of the authors and do not necessarily represent the
views of the sponsor. Tests would not have been possible
without the assistance of the Herrick Corporation; their
help on fabricating and erecting the test specimens are
greatly appreciated.
References
AIJ (1995). Reconnaissance report on damage to steel
building structures observed from the 1995 Hyogoken-
Nanbu (Hanshin/Awaji) earthquake. Steel Committee of
Kinki Branch, Architectural Institute of Japan, Tokyo,
Japan.
AISC (2005a). Specification for structural steel buildings.
American Institute of Steel Construction, Chicago,
Illinois.
AISC (2005b). Seismic provisions for structural steel
buildings. American Institute of Steel Construction,
Chicago, Illinois.
Ballio, G. and Perotti, F. (1987). Cyclic behavior of axially
loaded members: Numerical simulation and experimental
58 Jiun-Wei Lai and Stephen A. Mahin / International Journal of Steel Structures, 14(1), 43-58, 2014
verification. Journal of Constructional Steel Research, 7,
pp. 3-41.
Bertero, V. V., Uang, C.-M., Llopiz, C., and Igarashi, K.
(1989). Earthquake simulator testing of concentric
braced dual system. Journal of Structural Engineering,
ASCE, 115(8), pp. 1877-1894.
Black, G. R., Wenger, B. A., and Popov, E. P. (1980).
Inelastic buckling of steel struts under cyclic load
reversals. UCB/EERC-80/40, Earthquake Engineering
Research Center, University of California, Berkeley, CA.
Bonneville, D. and Bartoletti, S. (1996). Case study 2.3:
Concentric braced frame, Lankershim Boulevard, North
Hollywood. 1994 Northridge earthquake; building case
Studies project; proposition 122: product 3.2, SSC 94-06,
Seismic Safety Commission State of California, pp. 305-
324.
Chen, C.-H. (2010). Performance-based seismic demand
assessment of concentrically braced steel frame
buildings. Ph.D. Dissertation, University of California,
Berkeley, CA.
Christopoulos, C., Tremblay, R., Kim, H. J., and Lacerte, M.
(2008). Self-centering energy dissipative bracing system
for the seismic resistance of structures: Development and
validation. Journal of Structural Engineering, ASCE,
134(1), pp. 96-107.
Clark, K., Powell, P., Lehman, D., Roeder, C., and Tsai, K.C.
(2008). Experimental performance of multi-story X-
braced frame systems. Proc. SEAOC 77
th
Annual
Convention, Big Island, Hawaii.
Fisher, J. M. and Kloiber, L. A. (2006). Base plate and
anchor rod design. 2
nd
Edition, American Institute of
Steel Construction, Chicago, Illinois.
Kelly, D. J., Bonneville, D., and Bartoletti, S. (2000). 1994
Northridge earthquake: damage to a four-story steel
braced frame building and its subsequent upgrade. Proc.
12
th
World Conference on Earthquake Engineering, New
Zealand Society for Earthquake Engineering, Upper Hutt,
New Zealand.
Khatib, I. F., Mahin, S. A., and Pister, K. S. (1988). Seismic
behavior of concentrically braced steel frames. UCB/
EERC-88/01, Earthquake Eng. Research Center,
University of California, Berkeley, CA.
Lai, J. W. (2009). Test results of two special concentric
braced frame specimens. CE299 Report, Department of
Civil and Environmental Engineering, University of
California, Berkeley, CA.
Lai, J. W., Chen, C. H., and Mahin, S. A. (2010).
Experimental and analytical performance of concentrically
braced steel frames. Proc. NASCC the Steel Conference
and the Structures Congress, Orlando, FL.
Lai, J. W. and Mahin, S. A. (2013). Experimental and
analytical studies on the seismic behavior of conventional
and hybrid braced frames. PEER 2013/20, Pacific
Earthquake Engineering Research Center, University of
California, Berkeley, CA.
Lee, S. and Goel, S. C. (1987). Seismic behavior of hollow
and concrete filled square tubular bracing members.
UMCE87-11, University of Michigan, Ann Arbor, MI.
Lumpkin, E., Hsiao, P., Roeder, C. W., Lehman D. E., Tsai,
K. C., Lin, C. H., Wu, A. C., and Wei, C. Y. (2010).
Cyclic response of three-story, full-scale concentrically
braced frames. Proc. 9
th
U.S. National and 10
th
Canadian Conference on Earthquake Engineering,
Toronto, Canada.
McKenna, F. (1997). Object oriented finite element
programming: Frameworks for analysis, algorithms and
parallel computing. Ph.D. Dissertation, University of
California, Berkeley, CA.
Palmer, K., Okazaki, T., Roeder, C., and Lehman, D. (2010).
Three-dimensional tests of two-story, one-bay by one-
bay, steel braced frames: Specimen design. Proc. 9
th
U.S.
National and 10
th
Canadian Conference on Earthquake
Engineering, Toronto, Canada.
Roeder, C. W., Lehman, D. E., and Yoo, J. H. (2004).
Performance-based seismic design of braced-frame
gusset plate connections. Proc. Connections in Steel
Structures V, Amsterdam, Netherlands.
Roeder, C. W. and Lehman, D. E. (2008). Seismic design
and behavior of concentrically braced steel frames.
Structure Magazine, pp. 37-39.
Tremblay, R. (2002). Inelastic seismic response of steel
bracing members. Journal of Constructional Steel
Research, 58, pp. 665-701.
Tremblay, R. and Merzouq, S. (2004). Dual buckling
restrained braced steel frames for enhanced seismic
response. Proc. Passive Control Symposium 2004,
Yokohama, Japan.
Uriz, P. and Mahin, S. A. (2008). Toward earthquake-
resistant design of concentrically braced steel-frame
structures. PEER 2008/08, Pacific Earthquake Engineering
Research Center, University of California, Berkeley, CA.
Watanabe, A., Hitomi, Y., Saeki, E., Wada, A., and Fujimoto,
M. (1988). Properties of brace encased in buckling-
restraining concrete and steel tube. Proc. Ninth World
Conference on Earthquake Engineering, Tokyo-Kyoto,
Japan, Paper No. 6-7-4, pp. 719-724.
Yang, F. and Mahin, S. A. (2005). Limiting net section
failure in slotted HSS braces. Structural Steel Education
Council, Moraga, CA.

You might also like