You are on page 1of 13

Performance-Based Seismic Design of Shape Memory

Alloy–Reinforced Concrete Bridge Piers. I: Development


of Performance-Based Damage States
A. H. M. Muntasir Billah, A.M.ASCE 1; and M. Shahria Alam, M.ASCE 2
Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Performance-based seismic design aims to dictate the structural performance in a predetermined fashion given the possible
seismic hazard scenarios the structure is likely to experience. Identifying and assessing the probable performance is an integral part of
performance-based design. Before implementation, accurate and practical definitions of different performance levels and corresponding limit
states must be determined. This study aims to develop performance-based damage states for shape memory alloy (SMA)–reinforced concrete
(RC) bridge piers considering different types of SMAs and seismic hazard scenarios. Using incremental dynamic analysis (IDA), this study
develops quantitative damage states corresponding to different performance levels (cracking, yielding, and strength degradation) and specific
probabilistic distributions for RC bridge piers reinforced with different types of SMAs. Based on an extensive numerical study, this study
also proposes residual drift–based damage states for SMA-RC piers. Finally, an analytical expression is proposed to estimate the residual
drift of SMA-reinforced concrete elements as a function of the expected maximum drift and superelastic strain of SMA. Comparison with
experimental results revealed that the proposed equation could very well predict the residual drift obtained from the experimental results.
DOI: 10.1061/(ASCE)ST.1943-541X.0001458. © 2016 American Society of Civil Engineers.
Author keywords: Performance-based design; Shape memory alloy; Damage states; Incremental dynamic analysis; Residual drift; Special
design issues.

Introduction region), which are prone to experience more damage during an


earthquake.
In recent years, seismic design guidelines have focused on Numerous experimental and numerical studies proved the effi-
performance-based design in order to predict and better manage ciency of SMA reinforced structures in seismic regions. However,
the postearthquake functionality and condition of structures. there exists no proper design guideline for utilizing SMA in high-
Recent developments in performance-based seismic design and as- way bridges. Moreover, most of the design guidelines are moving
sessment approaches have emphasized the importance of properly towards performance-based design. AASHTO has already devel-
assessing and limiting the residual (permanent) deformations that oped performance-based design guidelines for bridges referred
are typically sustained by a structure after a seismic event (Pettinga as AASHTO SGS (AASHTO 2011). Moreover, the recent edition
et al. 2006). If reinforced concrete (RC) structures can be designed of the Canadian highway bridge design code (National Research
in such a way that they are not only capable of undergoing large Council of Canada 2014) has also adapted performance-based
displacements with adequate energy dissipation capacity during a design and defined some performance levels and performance
seismic event, but can also eliminate the problem of permanent de- criteria for different types of bridges. To successfully apply the
formation, this will substantially scale down the repair and main- performance-based design concept to SMA-reinforced concrete
tenance cost of structures and make them safer against earthquakes. (SMA-RC) bridge piers, the performance objectives and their as-
Superelastic shape memory alloy (SMA) possesses the distinct abil- sociated limit state criteria must be clearly defined first. Most of the
ity to experience large deformations and retrieve its original shape current researches on SMA-RC bridge piers are focused on seismic
upon load removal along with possessing a high resistance to cor- performance assessment and comparison with regular RC bridge
rosion (Alam et al. 2009). This is a distinct property that makes piers (Billah and Alam 2014; Cruz and Saiidi 2012; Saiidi et al.
SMA a smart material and a strong contender for reinforcement 2009). Although a good number of studies exist on performance-
in RC structures particularly at critical locations (e.g., plastic hinge based damage states for steel RC bridge piers (Lehman et al. 2004;
Hose et al. 2000), no study so far has focused on the performance-
based damage states for SMA-RC piers. This is mainly due to lim-
1
Former Graduate Student, School of Engineering, Univ. of British ited number of experimental studies performed on SMA-RC piers
Columbia, Kelowna, BC, Canada V1V1V7; presently, Bridge Engineer, where the high cost of SMA was the main restraining factor. Since
Parsons, Burnaby, BC, Canada V5H4M2. E-mail: muntasir.billah@ the behavior of SMA-RC piers are significantly different from their
alumni.ubc.ca steel counterpart, using those damage states for SMA-RC piers
2
Associate Professor, School of Engineering, Univ. of British might lead to faulty design. Moreover, the mechanical properties
Columbia, Kelowna, BC, Canada V1V1V7 (corresponding author).
of SMAs (superelastic strain, elastic modulus, and yield strength)
E-mail: shahria.alam@ubc.ca
Note. This manuscript was submitted on December 16, 2014; approved vary widely where several compositions of SMAs have been
on October 20, 2015; published online on July 22, 2016. Discussion period developed and used by different researchers in civil engineering
open until December 22, 2016; separate discussions must be submitted for applications (Alam et al. 2007). In a recent study, Billah and
individual papers. This paper is part of the Journal of Structural Engineer- Alam (2014) assessed the comparative seismic vulnerability of
ing, © ASCE, ISSN 0733-9445. SMA-RC and steel-RC bridge piers in terms of drift limits based

© ASCE 04016140-1 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


Table 1. Properties of Different Types of SMA
εs εr E fy f p1 f T1 fT2
SMA type Alloy (%) (%) (GPa) (MPa) (MPa) (MPa) (MPa) fy=E Reference
SMA-1 NiTi45 6 0.5 62.5 401.0 510 370 130 0.0065 Alam et al. (2008)
SMA-2 NiTi45 8 0.5 68 435.0 535.0 335 170 0.0063 Ghassemieh et al. (2012)
SMA-3 FeNCATB 13.5 1.5 46.9 750 1,200 300 200 0.0159 Tanaka et al. (2010)
SMA-4 CuAlMn 9 0.4 28 210.0 275.0 200 150 0.0075 Shrestha et al. (2013)
SMA-5 FeMnAlNi 6.13 0.7 98.4 320.00 442.5 210.8 122 0.0033 Omori et al. (2011)
Note: f y = austenite to martensite starting stress; f P1 = austenite to martensite finishing stress; fT1 = martensite to austenite starting stress; f T2 = martensite to
austenite finishing stress, εs = superelastic plateau strain length; εr = recovery strain; and E = modulus of elasticity.
Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

on performance criteria. The results showed that the performance Fe-Mn-Si, Fe-Ni-Co-Ti, Fe-Ni-Nb, Cu-AL-Mn, etc. Fe-based
limits of SMA-reinforced piers vary considerably from their steel SMAs show good workability, machinability, weldability, and wide
counterparts. Hence, the current study aims at developing perfor- transformation hysteresis as compared to Ni-Ti SMAs. Although
mance-based damage states for SMA-RC bridge piers considering several compositions of Fe-based SMAs have been developed,
five different SMAs with three different earthquake hazard levels. large-scale applications are still limited due to the poor shape
The ultimate goal of this study is to provide a technical basis for the recovery limit and associated costly training treatment. Recently
development of performance-based seismic design, and other Tanaka et al. (2010) developed a ferrous polycrystalline SMA
evaluation methodologies, for SMA-RC bridge piers. (Fe-Ni-Co-Al-Ta-B) which has a very high superelastic strain range
Using an incremental dynamic analysis (IDA)-based analytical of over 13% at room temperature. This SMA has approximately 20
approach (Vamvatsikos and Cornell 2002), here performance-based times higher SE than Fe-Ni-Co-Ti alloy and almost double that of
damage states (based on drift limits) have been developed for five conventional Ni-Ti alloy. This Fe-based SMA has extremely high
different SMA-RC bridge piers and validated against experimental ductility, greater strength, and also energy dissipation capacity sev-
data. Application of such technique may palliate the burden of eral times higher than that of commercially available Ni-Ti SMA.
gathering a large amount of test data and cost of experiments, More recently, Omori et al. (2011) developed another Fe-based
which were used in the past to develop different damage sates SMA (Fe-Mn-Al-Ni) which has superelasticity similar to the con-
for RC bridge piers. Past studies have demonstrated that it is nec- ventional Ni-Ti SMA but much lower austenite finish temperature,
essary to consider residual drifts to fully characterize the perfor- which allows this SMA to operate in superelastic range even at a
mance of a structural system after a seismic excitation and the very low temperature. In order to improve the machinability and
potential damage that the system can experience (Christopoulos reduce the cost, a Cu-based SMA (Cu-Mn-Al) has been developed
et al. 2003; Erochko et al. 2011). Since SMA has the ability to re- (Araki et al. 2010) that has comparable superelasticity to that of
duce the residual drift significantly after unloading, the residual NiTi SMAs. Moreover, these Cu-Al-Mn SMAs have comparatively
drift of different SMA-RC bridge piers under varying intensity lower strain rate effects than Ni-Ti SMAs (Araki et al. 2012) and
earthquakes needs to be investigated. This study also developed also been reported to provide recentering and crack recovery capa-
residual drift–based damage states for SMA-RC bridge piers and bilities (Shrestha et al. 2013).
proposed an analytical expression that can be used for predicting Many different types of SMAs have been developed that have
the residual drift in SMA-reinforced concrete elements. huge potential for smart structural applications, such as in bridge
piers. In this study, two Ni-Ti, one Cu-based, and two Fe-based
shape memory alloys have been selected for use in bridge piers.
Shape Memory Alloys The selected SMAs along with their mechanical properties such
as the elastic modulus (E), austenite to martensite starting stress
Smart materials like SMAs have demonstrated a wide range of en- (fy ); austenite to martensite finishing stress (fP1 ); martensite to
gineering applications namely, biomedical, robotics, aerospace, austenite starting stress (fT1 ); martensite to austenite finishing
civil, and mechanical engineering. Two distinct properties such stress (fT2 ); superelastic strain (εs ); residual strain (εr ) are listed
as the shape memory effect (SME, ability to recover plastic strain in Table 1.
upon heating) and superelasticity (SE, ability to recover plastic
strain upon load removal) make SMA a strong contender against
conventional metals and alloys for the application in various sec- Design and Geometry of Bridge Piers
tors. Several compositions of SMAs have been developed to date
such as Ni-Ti, Cu-Zn, Cu-Zn-Al, Cu-Al-Ni, Fe-Mn, Mn-Cu, Fe-Pd, This section briefly describes the design and configurations of dif-
and Ti-Ni-Cu, etc. Numerous applications of SMAs in the civil en- ferent SMA-RC bridge piers used in this study. Since SMA is a
gineering field have been documented (Ocel et al. 2004; Saiidi and costly material, it is only used in the bottom plastic hinge region
Wang 2006; Lindt and Potts 2008; Alam et al. 2009; Araki et al. of the bridge piers. Five different SMAs are used in this study to
2010; Billah and Alam 2012; Dezfuli and Alam 2013). Most ap- develop the performance-based damage states for SMA-RC bridge
plications have focused on the use of Ni-Ti alloy while very few piers. The bridge pier is assumed to be located in Vancouver,
focused on the application of Cu-based SMAs (Araki et al. 2010; British Columbia, Canada, and was seismically designed following
Shrestha et al. 2013), or Fe-based SMAs (Dezfuli and Alam 2013; the Canadian highway bridge design code (CSA 2010). The con-
Czaderski et al. 2014). Although Ni-Ti SMAs show large recover- sidered bridge is a lifeline bridge as per CHBDC (CSA 2010). In an
able strains, good superelasticity, and exceptional resistance to cor- event of the design earthquake (return period of 475 years), a life-
rosion, the high cost of Ni-Ti SMAs and their machinability restrict line bridge needs to remain open for immediate use to all traffic.
their large-scale applications. Therefore, the piers need to be designed to achieve this targeted
In an attempt to reduce the cost of SMAs, researchers have come performance. According to CHBDC (CSA 2010), the importance
up with various Fe-based and Cu-based low cost SMAs such as factor of I ¼ 3 and the response modification factor of R ¼ 3 were

© ASCE 04016140-2 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Cross section and elevation of SMA reinforced concrete bridge pier

considered for this lifeline bridge. Fig. 1 shows the cross section is being used to refer to the initiation of phase transformation of
and elevation of the bridge pier. The diameter of all the columns SMA and the yield strain was calculated by defining the austenite
was 1.83 m; the piers were reinforced with 48 longitudinal reinforc- to martensite starting stress (fy ) by the elastic modulus (E).
ing bars of varying diameter (reinforcement ratio of 1.2–1.75%) Five different SMA rebars as presented in Table 1 were used to
and 16-mm-diameter steel spirals at a 76 mm pitch (transverse design the different bridge piers. The bridge piers are designated
reinforcement ratio of 0.7%). The height of the pier is 9.14 m with as SMA-RC-1 (reinforced with SMA-1), SMA-RC-2 (reinforced
an aspect ratio of 5, which ensured the flexure dominated behavior. with SMA-2), and so on. SMA-RC-1 and SMA-RC-2 are
A constant mass of 85t was applied at the top which represents the reinforced with 48-28M SMA-1 and SMA-2 bars, SMA-RC-3 is
weight of the superstructure. The design moment and axial force reinforced with 48-20M SMA-3 bars, SMA-RC-4 is reinforced
were 11,265 kN · m and 9,500 kN, respectively. The shear capacity with 48-35M SMA-4 bars, and SMA-RC-5 is reinforced with
of the pier was checked using modified compression field theory 48-32M SMA-5 bars. Rebar sizes were selected in such a way
(Vecchio and Collins 1986), which predicts the experimentally de- that the axial forces developed in the rebar are almost similar.
termined shear failure within 1% error (Bentz et al. 2006). The The bridge piers were designed in such a way that they have com-
shear resistance of the pier was found to be 3,912 kN, which parable moment capacities. Fig. 2(a) shows the moment-curvature
was much higher than the shear force demand of 1,230 kN. The response of different SMA-RC sections. From this figure, it is evi-
moment-shear interaction and the axial load-moment interaction dent that all the sections have similar initial stiffness and compa-
of the bridge piers were investigated. However, for brevity, the re- rable moment capacities. Since SMA-5 has higher elastic modulus,
sults are not presented here. It was found that the maximum mo- SMA-RC-5 showed higher initial stiffness which is 1.78, 1.72,
ment and shear force as well as the applied maximum axial load and 2.21, and 3.87 times higher than that of SMA-RC-1, SMA-RC-2,
moment are within the safe boundary. Wang et al. (2008) recom- SMA-RC-3, and SMA-RC-4, respectively. Moment-curvature re-
mended that the shear capacity of the pier should be greater than 1.6 sponse of all the sections revealed that this design process led
times the base shear corresponding to the design moment which to comparable moment capacities for the five different SMA-
was also satisfied. reinforced bridge piers. The elastic periods of the SMA-RC-1,
Different diameter bars were used for different SMAs since dif- SMA-RC-2, SMA-RC-3, SMA-RC-4, and SMA-RC-5 were
ferent SMAs have different elastic modulus and yield strength. calculated as 0.513 s, 0.513 s, 0.514 s, 0.515 s, and 0.511 s, respec-
Although SMA does not have a yielding process, the term yield tively which were close and expected to attract similar earthquake

18000 2000
16000
1600
Base Shear (kN)

14000
Moment (kN-m)

12000
1200
10000
SMA-1
8000 800
SMA-2
6000 SMA-3 SMA-1 SMA-2
4000 SMA-4 400 SMA-3 SMA-4
SMA-5 SMA-5
2000
0 0
0 0.02 0.04 0.06 0 0.2 0.4 0.6 0.8 1 1.2
(a) Curvature (1/m) (b) Displacement (m)

Fig. 2. (a) Moment-curvature relationship of RC sections with different types of SMA; (b) static pushover curves for bridge piers reinforced with
different types of SMA

© ASCE 04016140-3 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


Table 2. Material Properties for SMA-RC Bridge Pier the Paulay and Priestley (1992) equation provided different plastic
Material Property Value hinge lengths for different SMA-reinforced bridge piers. Therefore,
the Paulay and Priestley (1992) expression for plastic hinge length
Concrete Compressive strength (MPa) 42.4
calculation in SMA/RC elements can be considered with reasonable
Corresponding strain 0.0029
Tensile strength (MPa) 3.5
accuracy.
Elastic modulus (GPa) 23.1
Steel Elastic modulus (GPa) 200
Yield stress (MPa) 475 Analytical Modeling of Bridge Piers
Ultimate stress (MPa) 692
Ultimate strain 0.14 In this study, a fiber element–based nonlinear analysis pro-
Plateau strain 0.016 gram SeismoStruct version 6 has been employed to develop
Strain hardening parameter 0.0125 performance-based damage states for SMA-RC bridge piers. Incre-
Transition curve initial shape parameter 19.5 mental dynamic analyses (IDA) have been performed to determine
Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

Transition curve shape calibrating coefficient (A1) 18.5 the various damage states of the bridge piers. The program has
Transition curve shape calibrating coefficient (A2) 0.15 the ability to determine the large displacement behavior and the
collapse load of framed structures accurately under either static
or dynamic loading, while taking into account both geometric non-
forces. Fig. 2(b) shows the pushover response curves for these five linearities and material inelasticity (Pinho et al. 2007). The bridge
different SMA-RC bridge piers. From this figure, it can be observed piers were modeled with three-dimensional (3D) inelastic beam–
that all the bridge piers had similar stiffness and load-carrying column elements (force-based elements), with circular sections
capacities. for the piers; the constitutive laws of the reinforcing steel and con-
The material properties of the concrete and steel rebar used crete were as per the Menegotto and Pinto (1973) and Mander
in the bridge piers are summarized in Table 2. In the SMA-RC et al. (1988) models, respectively. This steel model takes into ac-
bridge piers, SMA was used as longitudinal reinforcement only count the Bauschinger effect, which is relevant for the representa-
at the plastic hinge region. In the remaining part, steel rebars were tion of the columns’ stiffness degradation under cyclic loading. The
used as reinforcement. The plastic hinge length, Lp was calculated superelastic SMA model developed by Auricchio and Sacco (1997)
according to the Paulay and Priestley (1992) equation has been employed for modeling SMA using the parameters
provided in Table 1. The material models were calibrated (values
Lp ¼ 0.08L þ 0.022db f y ð1Þ provided in Table 1) to capture the strength and stiffness degrada-
tion under cyclic loading.
where L = length of the member in mm; db = bar diameter in mm; and The accuracy of the program in predicting the seismic response of
f y = yield strength of the rebar in MPa. Previously, Alam et al. (2008) bridge structures has been demonstrated by several researchers
and O’Brien et al. (2007) showed that the Paulay and Priestley through comparisons with experimental results (Casarotti and Pinho
(1992) equation can reasonably estimate the plastic hinge length 2006; Alam et al. 2008, 2009; Billah and Alam 2013). Fig. 3 shows
of SMA-reinforced concrete element. Moreover, Saiidi and Wang the comparison of experimental and analytical results from two
(2006), Saiidi et al. (2009), and Cruz and Saiidi (2012) also used different studies using two different SMAs. Fig. 3(a) shows the
the Paulay and Priestley (1992) equation to calculate the plastic comparison of shake table test results and analytical results of a
hinge length for their experimental studies where SMA rebars were SMA-steel RC bridge pier where SMA was particularly used in
placed in the bottom plastic hinge region of bridge piers. AASHTO the plastic hinge region. The numerical results obtained from
SGS (AASHTO 2011) also suggests using the same equation for SeismoStruct could predict the experimental result of Saiidi and
calculating the plastic hinge length of concrete bridge piers. Wang (2006) accurately where the variations were only 5.6%,
However, CHBDC (2010) recommends using the greatest of the 6.1%, and 9.4% for base shear, tip displacement, and amount of en-
maximum cross sectional dimension of column, one-sixth of clear ergy dissipation, respectively. Fig. 3(b) shows the load-rotation re-
height of the column or 450 mm. However, according to CHBDC sponse of concrete beams reinforced with Cu-Al-Mn SMA (SMA-4)
(2010), the plastic hinge length does not depend on strength and in the midspan under four-point reverse cyclic loading (Shrestha
diameter of longitudinal rebar. Since this study utilized different et al. 2013). From this figure, it is evident that the adopted analytical
types of SMA rebars of varying diameter and yield strength, using model was capable of predicting the experimental response very well

Fig. 3. Comparison of experimental and numerical results: (a) SMA-RC (SMA-1) bridge pier; (b) SMA-RC (SMA-4) beam

© ASCE 04016140-4 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


Extensive experimental investigations on bridge piers performed in
Selection of Shape Memory Alloy
the past were utilized to develop the damage states for reinforced
concrete bridge piers (Berry and Eberhard 2003; Hose et al. 2000).
Due to the fact that very limited experimental results are available
Design of bridge piers with different SMA for SMA-reinforced bridge piers, an IDA-based approach, as illus-
trated in Fig. 4, was developed in this study to generate the neces-
sary data for developing performance-based damage states for
Moment-curvature analysis of different bridge piers reinforced with different types of SMAs. First, five
SMA-RC section different types of SMAs were selected and five different SMA-RC
bridge piers were designed. In the next step, moment-curvature
analyses of the designed piers were carried out to compare their
Selection of suitable ground motions and
hazard spectra moment capacities. After that, a suite of suitable ground motions
Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

was selected and IDA was carried out. Based on the strain limits,
different performance criteria were captured from IDA. Using the
IDA of different SMA-RC bridge pier IDA results, dynamic pushover curves were generated and the per-
formance limits were identified. In the final step, suitable probabi-
listic distributions corresponding to different limit states were
Capture the different performance limits
identified and drift limits corresponding to different limit states
for each EQ at different intensity were defined.
IDA (Vamvatsikos and Cornell 2002) was employed to deter-
mine the performance limit states of different bridge piers using
Develop the dynamic pushover curves for an ensemble of 10 selected ground motions. IDA is a useful method
each EQ and compute the median, for more-detailed seismic performance predictions of structures
5 percentile and 95 percentile curves
subjected to different seismic excitation levels. In IDA, the finite-
element model is subjected to numerous inelastic time history
Obtain the drift limit at different analyses using one or a set of ground motion record(s), each scaled
performance levels and determine the (up and/or down) to different seismic intensity levels while tracking
suitable distribution the response of the structure (e.g., displacements, accelerations,
etc.). This procedure of scaling and time history analysis is repeated
until dynamic instability in the form of large drifts occurs, indicat-
Define drift limits at different performance
levels for different SMAs ing structural collapse.

Fig. 4. Steps for the development of performance-based damage states Selection of Ground Motions
for SMA-RC bridge pier
The incremental dynamic analyses were carried out using the 10
selected ground motions as shown in Table 3. These ground motion
records were obtained from the Pacific Earthquake Engineering
along with strength and stiffness degradation where the variations Research (PEER) Center ground motion database (PEER 2011).
were only 3.4% and 5.9% for maximum force and beam rotation, These accelerograms were chosen such that they represent the seis-
respectively. mic characteristics of the site of the structure. The ratio between the
peak ground acceleration (PGA) and peak ground velocity (PGV)
is an indicator of the frequency content of seismic motion. The
IDA-Based Approach for Developing characteristic seismic motions for the western part of Canada have
Performance-Based Damage States a PGA=PGV ratio of around 1.0 (Naumoski et al. 1988). The se-
lected ensemble of earthquake records is presented in Table 3 where
For successful implementation of the performance-based design the PGA=PGV ratio varies between 0.8 and 1.3. These 10 ground
concept in SMA-RC bridge pier, the performance objectives and motion records were obtained from the PEER strong motion data-
their corresponding damage state criteria need to be clearly defined. base. The recent edition of Canadian highway bridge design code

Table 3. Selected Earthquake Ground Motion Records (Data from PEER 2011)
Number Event Year Record station M a R b (km) PGA (g) PGA/PGV
1 Imperial Valley 1979 El Centro array#11 6.5 21.9 0.36 0.8
2 Imperial Valley 1979 Chihuahua 6.5 28.7 0.254 0.84
3 Kobe 1995 Takatori 6.9 4.3 0.56 0.9
4 Kobe 1995 JMA 6.9 3.4 0.77 1.02
5 Loma Prieta 1989 Holister South and Pine 6.9 28.8 0.371 0.97
6 Loma Prieta 1989 16 LGPC 6.9 16.9 0.605 1.19
7 Nothridge 1994 Rinaldi 6.7 7.5 0.87 0.93
8 Nothridge 1979 Olive View 6.7 6.4 0.721 0.95
9 Superstition Hill 1987 Wildlife liquefaction array 6.7 24.4 0.134 1.0
10 Superstition Hill 1987 Wildlife liquefaction array 6.7 24.4 0.132 1.03
a
Moment magnitudes.
b
Closest distances to fault rupture.

© ASCE 04016140-5 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


1
Table 4. Proposed Damage State Framework
0.9 2%/50 Year (Target)
Damage Damage Functional
Spectral Acceleration (g)
0.8 5%/50 Year (Target)
10%/50Year (Target) parameter state level Description
0.7
2%/50 Year (Mean) Cracking DS-1 Immediate Onset of hairline cracks
0.6
5%/50 Year (Mean) Yielding DS-2 Limited Theoretical first yield
0.5
10%/50 Year (Mean) of longitudinal rebar
0.4
Spalling DS-3 Service disruption Onset of concrete spalling
0.3 Core crushing DS-4 Life safety Crushing of core concrete
0.2
0.1
0
0 1 2 3 4 states. The yielding of SMA rebar was monitored by defining
Time (sec) the yield strain of SMA bar and tracking the occurrence of first
Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

yield in SMA rebar. The spalling strain was assumed to be 0.004


Fig. 5. Design and mean response spectrum of 10 records used for IDA
as suggested by Priestley et al. (1996). Paulay and Priestley (1992)
analysis matching the three different CHBDC spectrum (2, 5, and 10 in
found that the crushing strain of confined concrete ranges between
50 years)
0.015 and 0.05. In this study, the crushing strain of confined con-
crete for different SMA-RC bridge piers was calculated using the
Paulay and Priestley (1992) equation
(CHBDC) (National Research Council of Canada 2014) requires
that highway bridges should meet target performance levels εcu ¼ 0.004 þ 1.4ρs f yh εsm =f c0 ð2Þ
under seismic ground motions with different return periods. In this
study, three different levels of seismic ground motions were where εcu = ultimate compression strain (0.019); εsm = steel strain
considered according to CHBDC 2014 [CSA S6-14 (National at maximum tensile stress (0.14); f c0 = concrete compressive
Research Council of Canada 2014)]. These records correspond strength in MPa (42.4 MPa); f yh = yield strength of transverse steel
to three different hazard levels with a 2, 5, and 10% probability in MPa (450 MPa); and ρs = volumetric ratio of confining
of exceedance in 50 years. The respective return periods are steel (0.7%).
2,475, 975, and 475 years. For each hazard level, 10 ground mo- Most of the damage states available in literature are discrete in
tions given in Table 3 were used. The selected ground motions nature and quantify the damage deterministically (Marsh and
were scaled to specific hazard levels using SeismoMatch. This soft- Stringer 2013). Practically, the drift level corresponding to certain
ware is able to adjust any ground motion accelerogram to match a damage is not a discrete deterministic quantity and each damage
specific design response spectrum using wavelet algorithm pro- level is associated with a distribution of values. The drift limits de-
posed by Abrahamson (1992) and Hancock et al. (2006). Matching fined at different damage states should clearly indicate whether it
was done within the period range of interest, which was 0.05–4 s represents the lower bound, median, or some intermediate value for
as suggested by Baker et al. (2011). The mean spectra and the the onset of a particular damage type. In order to develop compre-
target spectra corresponding to different hazard levels are shown hensive performance-based damage states, in this study, the prob-
in Fig. 5. abilistic distribution of each damage state was also identified and
the median of the distribution was defined as the drift limit corre-
sponding to each damage state.
Performance-Based Damage States Criterion
In order to determine the limit state drift values for different per-
Performance-based seismic design largely relies on the identifica- formance levels, the drift limits corresponding to the strain values
tion and selection of proper limit/damage states. Often damage were determined using IDA results for different hazard levels for
states are defined in terms of drift or displacement. Damages are the five different SMA-RC bridge piers. The drift limits at various
usually defined as discrete observable damage states (e.g., rebar performance levels were identified using the dynamic pushover
yielding, concrete spalling, longitudinal bar buckling, and bar frac- curves obtained from IDA. Dynamic pushover curves represent
ture) (Marsh and Stringer 2013). For example, the recent edition of the relation between the maximum drift and its corresponding base
the Canadian highway bridge design code (National Research shear obtained from IDA while being subjected to an earthquake
Council of Canada 2014) defined four different performance levels, record (Elnashai and Luigi 2008). These curves represent the struc-
such as, immediate service, limited service, service disruption, and tural capacity under specific earthquake loading. Dynamic push-
life safety. Based on the material strain limit (concrete and steel), over curves, obtained from IDA, take into account progressive
four different damage states corresponding to the four performance structural stiffness degradation, change of modal characteristics,
levels are defined. For example, limited service indicates a repair- and period elongation of the structure for increasing values of ex-
able damage, which corresponds to a limiting reinforcing steel ternal action, which is not achievable through static pushover
strain of 0.015 for concrete structures, and no buckling of primary analysis. Inelastic characteristics such as strength degradation
members are allowed in steel structures. and energy dissipation largely affect the seismic performance of
In this study, four quantitative performance limit states were de- structures, which are also required for developing performance-
fined for the SMA-RC bridge piers based on the performance levels based damage states for performance-based design. The drift levels
and damage states proposed by Hose et al. (2000). Table 4 shows for different performance levels obtained from IDA were used to
the four performance limit states and their associated functional find a suitable distribution for each damage state that describes
level definition adopted in this study. The performance limit states the statistical distribution of the developed damage states. Statisti-
considered here are, the drift (%) at the onset of hairline cracks, cal analyses were carried out to find the most suitable probability
longitudinal rebar yielding, cover concrete spalling, and crushing density function (PDF) to represent the data related to each damage
of core concrete. In this study, a strain-based damage detection ap- state. Using statistical tools and analyses, a suitable distribution
proach was used for defining the drift levels at different damage for each damage state was determined using goodness-of-fit tests.

© ASCE 04016140-6 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


The following section discusses the development of performance- particular damage state is represented by the vertical solid line.
based damage states for five different SMA-reinforced bridge piers. From Figs. 6–10 it can be observed that the uncertainty of each
damage state is unique, as indicated by the dispersion or width
of the distribution. The median drift level of each damage state
Probabilistic Distribution of Drift-Based Damage States is defined as the limiting drift value for each performance level.
Using the results obtained from IDA, the probabilistic distribution The drift levels corresponding to different damage states for
of each damage state corresponding to different hazard levels was different hazard levels are shown in Table 5. Table 5 also shows
determined to represent the statistical variability of damage states the probabilistic distribution of each damage state. From IDA,
at different hazard levels. The probabilistic distribution of each measurements of drift levels corresponding to each damage state
damage state, i.e., yielding, spalling, and crushing, was superim- were obtained and statistically processed to find out the most
posed on the dynamic pushover curves obtained from IDA and suitable distribution. The suitability of the selected distributions
are shown in Figs. 6–10. The expected (median) drift level at a for representing each damage state was evaluated using a
Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. Dynamic pushover response and different damage states with distribution for SMA-RC-1 for (a) 2% in 50 years; (b) 5% in 50 years; (c) 10% in
50 years probability of exceedance

Fig. 7. Dynamic pushover response and different damage states with distribution for SMA-RC-2 for (a) 2% in 50 years; (b) 5% in 50 years; (c) 10% in
50 years probability of exceedance

Fig. 8. Dynamic pushover response and different damage states with distribution for SMA-RC-3 for (a) 2% in 50 years; (b) 5% in 50 years; (c) 10% in
50 years probability of exceedance

© ASCE 04016140-7 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 9. Dynamic pushover response and different damage states with distribution for SMA-RC-4 for (a) 2% in 50 years; (b) 5% in 50 years; (c) 10% in
50 years probability of exceedance

Fig. 10. Dynamic pushover response and different damage states with distribution for SMA-RC-5 for (a) 2% in 50 years; (b) 5% in 50 years; (c) 10%
in 50 years probability of exceedance

Table 5. Damage States of Different SMA-RC Bridge Pier and Their Associated Distribution
Drift (%)
Probability of exceedance
SMA-1 SMA-2 SMA-3 SMA-4 SMA-5
Damage Damage 2% 5% 10% 2% 5% 10% 2% 5% 10% 2% 5% 10% 2% 5% 10%
parameter state in 50 in 50 in 50 in 50 in 50 in 50 in 50 in 50 in 50 in 50 in 50 in 50 in 50 in 50 in 50 Distribution
Cracking DS-1 0.28 0.28 0.28 0.30 0.28 0.28 0.28 0.28 0.28 0.28 0.28 0.28 0.28 0.28 0.28 Uniform
Yielding DS-2 1.68 1.76 1.86 1.66 1.72 1.80 2.28 2.42 2.58 1.74 1.83 1.95 1.10 1.16 1.21 Lognormal
Spalling DS-3 2.66 2.79 2.88 2.69 2.77 2.87 1.64 1.72 1.80 2.52 2.61 2.68 1.97 2.02 2.10 Normal
Crushing DS-4 5.05 5.68 5.94 5.51 5.91 6.05 7.65 7.81 7.94 5.56 5.63 5.72 4.73 4.79 4.84 Gamma

Kolmogorov-Smirnov (K-S) goodness-of-fit test. The following drift. This drift value of 0.28% can be identified as Damage
conclusions are derived from the distribution of different damage State-1(DS-1);
states: • From statistical analyses, it was found that the lognormal
• Irrespective of the type of SMA and earthquake hazard level, distribution better represents the uncertainty in drift limits for
cracking occurs at a drift of 0.28%, and it can be represented DS-2 (yielding). Usually the variation in metal strength, such
better with a uniform distribution. Since the cracking strain as yield strength of steel, is better represented by a lognormal
of concrete depends only on the tensile strength of concrete, distribution (Ellingwood 1977; Ghobarah et al. 1998). Similar
small variations in concrete cracking drift were observed. Uni- distributions for yield drift limits for SMA-RC bridge piers were
form distribution is a preferable choice when all of the outcomes obtained, which is largely dependent on the yield strength
have an equal probability of occurring. Since the cracking drift of SMA;
of all the SMA-RC bridge piers ranged between 0.28 and 0.30% • A normal distribution was found to be the best fit for repre-
and have equal probability of occurrence, the cracking drift is senting the variability in drift limits corresponding to DS-3
assumed to follow a uniform probability distribution. Results of (spalling) based on the K-S goodness-of-fit test. A normal dis-
the K-S goodness-of-fit test also confirmed the suitability of a tribution is better suited for representing the spalling drift since
uniform distribution for representing the distribution of cracking all the SMA-RC bridge piers showed a strong tendency towards

© ASCE 04016140-8 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


the central value of spalling drift as well as the positive and ne- for SMA-RC-3 at DS-2 were 52% higher than that of SMA-
gative deviations from this central value are equally likely. The RC-5, irrespective of the hazard level;
selected distribution seems reasonable since concrete strength • From Table 5, it can be observed that except for SMA-RC-3,
can be better represented by a normal distribution (Ellingwood yielding occurred in all the bridge piers before the initiation
1977; Mirza et al. 1979); and of cover spalling. The delayed rebar yielding of SMA-RC-3
• A K-S goodness-of-fit test was performed to identify the most can be attributed to its higher yield strength. A similar observa-
suitable distribution for defining the variation of DS-4 (crush- tion of the SMA-RC column has been reported by Saiidi and
ing). The K-S goodness-of-fit test indicated that the gamma dis- Wang (2006) where spalling of cover concrete took place before
tribution, which usually indicates an extreme event, provides the initiation of SMA yielding;
best fit to the data and was the most suitable for representing • DS-3 is considered at the onset of cover concrete spalling. All
the crushing drift. the piers experienced yielding before spalling where the only
exception was SMA-RC-3. For SMA-RC-1, DS-3 occurred
Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

at a drift level of 2.66%, 2.79%, and 2.88% for 2%, 5%, and
Maximum Drift–Based Damage States 10% in a 50-years hazard level, respectively. This is expected
Figs. 6–10 show the dynamic pushover curves for SMA-RC-1 since a hazard level with lower probability indicates a more da-
through SMA-RC-5 under different levels of earthquakes, respec- maging earthquake. A similar trend is also observed for the
tively. The dynamic pushover curves derived from 10 earthquakes other SMA-RC bridge piers where the limiting drift value in-
(for each bridge pier) were statistically processed to obtain the creased with decreased return period. In terms of drift limit,
median, 5th percentile, and 95th percentile capacity curves. SMA-RC-1 and SMA-RC-2 performed better than the other
Comparisons of Figs. 6–10 reveal that: three SMA-RC piers as they could sustain more drift before
• SMA-RC-3 (Fig. 8) has higher deformation and strength capa- entering into DS-3;
city as compared to the other SMA-RC bridge piers; • At DS-4 (crushing of concrete), all the SMA-RC bridge piers
• For seismic hazard level of 2% in 50 years, the median capacity sustained more than 5% drift under various hazard levels
of SMA-RC-3 was 2,743 kN which was 16%, 15%, 20%, and whereas the SMA-RC-3 exceeded 7.5%. For a hazard level with
17% higher than that of SMA-RC-1, SMA-RC-2, SMA-RC-4, a 2% of probability of exceedance in 50 years, SMA-RC-3
and SMA-RC-5, respectively; and sustained a drift of 7.65% before crushing, which was 34%,
• Maximum base shear demand is also significantly influenced by 28%, 27%, and 38% higher than that of SMA-RC-1, SMA-
the earthquake hazard level. For example, the median maximum RC-2, SMA-RC-4, and SMA-RC-5, respectively; and
base shear of SMA-RC-1, for 2% in 50 years is 2,305 kN which • For a particular SMA-RC bridge pier, crushing drift also varied
is 5% and 7% higher than that of 5% and 10% in 50-years re- significantly at different hazard levels. In the case of SMA-RC-
cords, respectively. 1, the crushing drift corresponding to 2% in 50-years hazard
Evaluation of the results presented in Table 5 provides a valu- level is 11.5% and 15% lower than the crushing drift at 5%
able insight on the damage states developed for different SMA-RC and 10% in 50-years hazard level, respectively. However, in
bridge piers. The damage states are defined for different hazard the case of SMA-RC-5, the crushing drift at 2% in 50-years ha-
levels. From Table 5 it can be observed that zard level was 2.3% and 1.25% lower than the crushing drift at
• Damage State-2 or yielding occurs at a drift level below 2% 5% and 10% in 50-years hazard level, respectively.
except for SMA-RC-3. At DS-2, there is significant variation The drift limits presented in Table 5 can be used for perfor-
in drift limits for different SMA-RC bridge piers. For SMA- mance-based design of SMA-RC bridge pier. Based on the design
RC-1 and SMA-RC-2, the drift limit is quite similar irrespective earthquake scenario, the designer can define the target performance
of the earthquake hazard levels which ranges between 1.68– level and associated drift limits. Since, performance-based damage
1.86% and 1.66–1.80%, respectively. Since SMA-RC-1 and states are proposed for different types of SMA, the designer can
SMA-RC-2 are reinforced with Ni-Ti SMAs (different compo- select any particular SMA and design the bridge pier according
sitions) with similar mechanical properties, they tend to have to the owners expected performance level.
similar drift limits at DS-2;
• Before yielding, SMA-RC-3 sustained higher drift compared to
the other SMA-RC piers. At DS-2, under 2% in 50 years hazard Residual Drift–Based Damage States for SMA-RC
level, the drift limit for SMA-RC-3 was 2.28%, which was sig- Bridge Piers
nificantly higher than the drift limits obtained for the other
SMA-RC bridge piers. This is expected since SMA-3 has higher Residual drift has been considered as one of the significant perfor-
yield strength and post yield stiffness as compared to the other mance indicators in judging a structure’s postearthquake safety and
SMAs considered in this study; the economic feasibility for repairing (Ramirez and Miranda 2012).
• Although SMA-4 has a low elastic modulus, its yield strength Although residual drift dictates the postearthquake functionality of
is very high, which eventually increased the yield strain and highway bridges, no other design guidelines except the Japanese
resulted in higher drift values. At DS-2, the drift limits for code for highway bridge design (JRA 2006) provide any residual
SMA-RC-4 were 3.4%, 3.8%, and 4.6% higher than that of drift limit of bridge piers. In a recent study, Saiidi and Ardakani
SMA-RC-1 for 2%, 5%, and 10% in 50-years probability of ex- (2012) found that bridge piers meeting current seismic require-
ceedance, respectively; ments can withstand larger traffic loads even when the residual
• Exceptional performance was observed for SMA-RC-5, which drift is 1.2% or more. Lee and Billington (2011) considered a
yielded at a very low drift level (1.1–1.2%) as compared to the 1% residual drift to be large enough for a bridge replacement.
other SMA-RC piers. This is due to SMA-5’s very low yield In order to develop the damage states (DS) for SMA-RC bridge
strength to elastic modulus ratio (0.0033), which reduced the piers, a probabilistic approach has been adopted in this study.
drift capacity of SMA-RC-5; Based on the existing literature (O’Brien et al. 2007; Billah and
• Although both SMA-3 and SMA-5 are Fe-based, due to the var- Alam 2014), four different damage states have been identified
iation in their yield strength and elastic modulus, the drift limits and a range of limiting residual drifts were considered. It was

© ASCE 04016140-9 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

Fig. 11. Fragility curves in terms of residual drift at (a) 10% in 50 years; (b) 5% in 50 years; (c) 2% in 50 years probability of exceedance

assumed that a residual drift below 0.25% would meet the service- the limiting value for the corresponding damage state. For example,
ability requirement (DS-1) while a residual drift larger than 1% in Fig. 11(a) (10% in 50 years), the 50% probability of occurrence
would be characterized as a collapse damage state (DS-4). The in- of DS-2 corresponds to a RD of 0.48% while the limiting RD values
termediate Damage States DS-2 and DS-3 are assumed to take for DS-2 for hazard levels of 5% and 2% in 50 years correspond to
place at a residual drift larger than 0.5% and 0.75%, respectively. 0.55% and 0.62%, respectively. It can be observed that the limiting
DS-1 requires that no structural realignment is necessary and the RD value for DS-2 was assumed to be 0.5% and the values obtained
bridge is fully operational. DS-2 consists of minor structural repair- from the median probability of exceedance are quite close. Simi-
ing and requires the bridge to be operational without requiring larly, the limiting RD values with a 50% probability of occurrence
bridge closure. A pier experiencing DS-3 will require major repair at different damage states and hazard levels were developed as out-
and may require bridge closure but should be usable for restricted lined in Table 6. From Table 6, it can be observed that as the ground
emergency traffic after inspection. DS-4 corresponds to the case motion return period decreases (probability of occurrence in-
when the residual drift is sufficiently large that the structure is creases), the limiting residual drift corresponding to different DS
in danger of collapse from earthquake aftershocks. decreases. For example, at DS-4, the limiting drift value for an
Once the damage states were identified, fragility curves for earthquake with a 2,475 years return period is 1.22%, which is
residual drifts were developed using the IDA results for three differ- 6.5% and 13.1% higher than an earthquake with 975 years and
ent seismic hazard levels. In this study, fragility functions were 475 years return periods, respectively. Observations from Table 6
developed using Eq. (3), which takes the form of lognormal cumu- indicate that, as the damage level increases (DS-1 to DS-4) the dif-
lative distribution functions having a median value of θ and loga- ference in limiting RD values at different hazard levels decreases.
rithmic standard deviation or dispersion of β For instance, at DS-2, the limiting RD value corresponding to a
  2,475 years return period is 11% and 22.5% higher than that of
lnðRD=θÞ 975 years and 475 years return period, respectively. However, this
FðRDÞ ¼ ϕ ð3Þ
β difference goes down to 6.5 and 13.1% for DS-4.

where FðRDÞ = conditional probability that the bridge pier will


be damaged to a given DS as a function of the residual drift Prediction of Residual Drift
(RD); F = standard normal cumulative distribution function; and
θ and β = median value of the probability distribution and the log- For performance-based design, prediction of residual drift as
arithmic standard deviation corresponding to the DS, respectively. a function of the target or maximum drift would be very useful.
Fig. 11 shows the fragility curves for SMA-RC bridge piers for Previous research has shown that residual drift predictions using
different damage states at three different hazard levels. Here, the nonlinear analysis are highly variable and subjected to different
fragility curves are plotted irrespective of the SMA types to general- modeling features (Applied Technology Council, ATC-58) (ATC
ize the associated damage states. Using these fragility curves, the 2012). Recently ATC-58 (ATC 2012) recommended some general
residual drift–based damage states for SMA-RC bridge piers were equations for predicting residual drift using peak transient drift
developed. From the fragility curves corresponding to each damage and yield drift. ATC-58 (ATC 2012) suggested that prediction of
state, the RD value with a 50% probability of occurrence indicates residual drift requires advanced nonlinear simulation with careful

Table 6. Residual Drift Damage States of SMA-RC Bridge Pier


Residual drift, RΔ (%)
Probability of exceedance
Damage state Functional level Description 10% in 50 5% in 50 2% in 50
Slight (DS ¼ 1) Fully operational No structural realignment is necessary 0.24 0.28 0.33
Moderate (DS ¼ 2) Operational Minor structural repairing is necessary 0.48 0.55 0.62
Extensive (DS ¼ 3) Life safety Major structural realignment is required to restore safety margin 0.73 0.82 0.87
for lateral stability
Collapse (DS ¼ 4) Collapse Residual drift is sufficiently large that the structure is in danger of 1.04 1.16 1.22
collapse from earthquake aftershocks

© ASCE 04016140-10 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


3 0.5

2.5 0.4

Predicted RD (%)
Predicted RD (%)
2
0.3
1.5
0.2
1

0.5 0.1

0 0
0 0.5 1 1.5 2 2.5 3 0 0.1 0.2 0.3 0.4 0.5
(a) Experimental RD (%) (b) Experimental RD (%)

Fig. 12. Comparison of residual drift prediction with experimental results: (a) O’Brien et al. (2007); (b) Youssef et al. (2008)
Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

attention to cyclic hysteretic response of the models and numerical bridge piers may not be applicable for SMA-RC bridge piers.
accuracy of the solution. In this study, the residual drift responses In this study, a set of performance-based damage states for
were obtained using IDA, which is one of the most advanced non- bridge piers reinforced with five different types of SMAs has been
linear analysis techniques and the models were validated against developed in terms of both maximum and residual drift. Using an
experimental results. From the residual drift responses of the IDA-based approach, this study proposed performance-based drift
SMA-RC bridge piers, it was found that the residual drift in levels for different damage states considering three different hazard
SMA-RC bridge piers is a function of maximum drift and supere- levels. To predict the residual deformation of SMA-RC bridge
lastic strain of the SMA. Using the residual drift responses obtained piers, a relationship between maximum drift, residual drift, and
from different SMA-RC bridge piers under a wide range of ground superelastic strain of SMA has been developed. The prediction
motions, a nonlinear regression analysis was conducted to inves- equation was validated against experimental observations from
tigate the effect of maximum drift and superelastic strain on the SMA reinforced concrete members. Based on the results of this
residual drift response. Using a nonlinear regression analysis, study, the following major conclusions can be drawn:
the following equation was developed for predicting the residual • The progression of damage was similar for all the RC bridge
drift of SMA-RC bridge piers: piers reinforced with different SMAs (except for SMA-RC-3):
      concrete cracking, longitudinal reinforcement yielding, cover
εs εs 1 spalling, and core crushing. For all SMA-RC bridge piers, the
RD ¼ 0.5 MD2 − MD þ ð4Þ
100 100 εs cracking occurred at the same level of drift (due to same cross
section) while the drift at other performance levels varied based
where RD = residual drift (%); εs = superelastic strain; and MD = on the mechanical properties of SMA used;
maximum drift (%). • Different performance-based drift limits, i.e., cracking, spal-
In order to investigate the accuracy of the proposed residual drift ling, yielding, and crushing of SMA-RC bridge piers, strongly
prediction equation, a comparison was carried out with experimen- follow uniform, normal, lognormal, and gamma distributions,
tal results. Fig. 12 shows the comparison of residual drifts obtained respectively. These distributions can be used for selecting the
from experimental investigations and the prediction equation. target drift levels for performance-based design of SMA-RC
Fig. 12(a) shows the comparison with experimental results of piers;
O’Brien et al. (2007), where a SMA-RC bridge pier was tested • Except for DS-1(cracking), other three damage states are signif-
under reverse cyclic loading. The bridge pier was constructed using icantly influenced by the type of SMA used. For DS-2 (yield-
Ni-Ti SMA in the plastic hinge region and the superelastic strain of ing), the limiting maximum drift varies from 1.18% (SMA-5) to
SMA rebar was 6%. Maximum drift values and the corresponding 2.28% (SMA-3) and for DS-3 (spalling), the limiting maximum
residual drifts were obtained from experimental results. Using the drift varies from 1.64% (SMA-3) to 2.69% (SMA-2) for hazard
maximum drift value and the superelastic strain of the SMA rebar, level corresponding to a 2,475 years return period;
the residual drifts were predicted. Fig. 12(a) shows that the pro- • In terms of maximum drift, consideration of different hazard
posed equation predicted the residual drift very well with an aver- levels does not have any significant impact on DS-1 and DS-2.
age absolute error (AAE) of 4.65% and average standard deviation On the other hand, different hazard levels have substantial im-
of 0.03. Fig. 12(b) shows the comparison of residual drift predic- pacts on DS-3 and DS-4;
tion with the experimental results of Youssef et al. (2008) where • The proposed residual drift limit states tend to decrease with
Ni-Ti SMA with superelastic strain of 6% was used as reinforce- increased probability of occurrence (decreases return period).
ment in the beam-column joint. From Fig. 12(b) it is evident that The damage states developed in terms of residual drifts correlate
the proposed equation is capable of predicting the residual drift well with damages observed from experimental studies;
with reasonable accuracy with an AAE of 2.06% and average stan- • Residual drift can be expressed as a function of maximum drift
dard deviation of 0.015. and the superelastic strain of SMA rebar;
• Based on the residual drift responses of all the SMA-RC piers
under different levels of ground motions, a prediction equation
Conclusions was developed to predict the residual drift response of an SMA-
RC bridge pier. The proposed equation can correlate very well
In order to develop a performance-based design guideline for with experimental observations; and
SMA-RC bridge piers, structural performance objectives and their • The proposed equation can be used for predicting the residual
corresponding limit state criteria must be clearly defined. Due to drift of SMA-RC bridge pier when designing the pier for a target
significant differences in the behavior of SMA-reinforced bridge residual drift. Based on the maximum drift and residual drift, the
piers as compared to conventional piers, damage states for typical designer would be able to select an SMA with the required

© ASCE 04016140-11 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


superelastic strain, thereby ensuring the safety of bridges under Casarotti, C., and Pinho, R. (2006). “Seismic response of continuous span
extreme earthquake events. bridges through fibre-based finite element analysis.” Earthquake Eng.
The current study considers one particular type of bridge pier Eng. Vibr., 5(1), 119–131.
model without considering uncertainty in geometry and material Christopoulos, C., Pampanin, S., and Priestley, M. J. N. (2003).
properties such as concrete and steel properties. A further study “Performance-based seismic response of frame structures including
residual deformations. I: Single-degree-of-freedom systems.” J. Earth-
using bridge piers with various aspect ratios, axial load levels,
quake Eng., 7(1), 97–118.
and with different sets of geometry/material properties should be
Cruz, N. C., and Saiidi, M. S. (2012). “Shake-table studies of a four-span
conducted for better understanding the contributions of other bridge model with advanced materials.” J. Struct. Eng., 10.1061/
parameters to the performance-based damage states for SMA- (ASCE)ST.1943-541X.0000457, 183–192.
reinforced concrete bridge piers. However, the results of this study CSA (Canadian Standards Association). (2010). “CSA S6-10, CHBDC
will work as a benchmark for evaluating future studies, developing 2010, Canadian highway bridge design code.” CAN/CSA S6-1,
more-refined damage states, and improving the capacity limit states Canadian Standards Association, Rexdale, ON.
Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

for SMA-RC bridge piers. Czaderski, C., Shahverdi, M., Brönnimann, R., Leinenbach, C., and
Motavalli, M. (2014). “Feasibility of iron-based shape memory alloy
strips for prestressed strengthening of concrete structures.” Constr.
Build. Mater., 56, 94–105.
Acknowledgments
Dezfuli, F. H., and Alam, M. S. (2013). “Shape memory alloy wire-
The financial contributions of Natural Sciences and Engineering based smart natural rubber bearing.” Smart Mater. Struct., 22(4),
45013–45029.
Research Council of Canada (NSERC) through Discovery Grant
Ellingwood, B. (1977). “Statistical analysis of RC Beam column interac-
and Industrial Postgraduate Scholarship Program were critical to
tion.” J. Struct. Eng., 103, 1377–1388.
conduct this study and are gratefully acknowledged.
Elnashai, A. S., and Luigi, D. S. (2008). Fundamentals of earthquake
engineering, Wiley, New York.
Erochko, J., Christopoulos, C., Tremblay, R., and Choi, H. (2011).
References “Residual drift response of SMRFs and BRB frames in steel buildings
designed according to ASCE 7-05.” J. Struct. Eng., 10.1061/(ASCE)ST
AASHTO. (2011). AASHTO guide specifications for LRFD seismic bridge .1943-541X.0000296, 589–599.
design, LRFDSEIS-2-M, 2nd ed., Washington, DC. Ghassemieh, M., Mostafazadeh, M., and Sadeh, M. S. (2012). “Seismic
Abrahamson, N. A. (1992). “Non-stationary spectral matching.” Seismol. control of concrete shear wall using shape memory alloys.” J. Intell.
Res. Lett., 63(1), 30. Mater. Syst. Struct., 23(5), 535–543.
Alam, M. S., Youssef, M. A., and Nehdi, M. (2007). “Utilizing shape Ghobarah, A., Aly, N. M., and El-Attar, M. (1998). “Seismic reliability
memory alloys to enhance the performance and safety of civil infra- assessment of existing reinforced concrete buildings.” J. Earthquake
structure: A review.” Can. J. Civ. Eng., 34(9), 1075–1086. Eng., 2(4), 569–592.
Alam, M. S., Youssef, M. A., and Nehdi, M. (2008). “Analytical prediction Hancock, J., et al. (2006). “An improved method of matching response
of the seismic behaviour of superelastic shape memory alloy reinforced spectra of recorded earthquake ground motion using wavelets.”
concrete elements.” Engr. Struct., 30(12), 3399–3411. J. Earthquake Eng., 10(1), 67–89.
Alam, M. S., Youssef, M. A., and Nehdi, M. (2009). “Seismic performance
Hose, Y., Silva, P., and Seible, F. (2000). “Development of a performance
of concrete frame structures reinforced with superelastic shape memory
evaluation database for concrete bridge components and systems under
alloys.” Smart Struct. Syst., 5(5), 565–585.
simulated seismic loads.” Earthquake Spectra, 16(2), 413–442.
Applied Technology Council. (2012). “Seismic performance assessment of
JRA (Japan Road Association). (2006). “Specifications for highway
buildings.” ATC-58, Redwood City, CA.
bridges.” Tokyo.
Araki, Y., et al. (2010). “Potential of superelastic Cu-Al–Mn alloy bars for
seismic applications.” Earthquake Eng. Struct. Dyn., 40(1), 107–115. Lee, W. K., and Billington, S. K. (2011). “Performance-based earthquake
Araki, Y., Maekawa, N., Omori, T., Sutou, Y., Kainuma, R., and Ishida, K. engineering assessment of a self-centering, post-tensioned concrete
(2012). “Rate-dependent response of superelastic Cu–Al–Mn alloy rods bridge system.” Earthquake Engr. Struct. Dyn., 40(8), 887–902.
to tensile cyclic loads.” Smart Mater. Struct., 21, 032002. Lehman, D., Moehle, J., Mahin, S., Calderone, A., and Henry, L. (2004).
Auricchio, F. and Sacco, E. (1997). “Superelastic shape-memory-alloy “Experimental evaluation of the seismic performance of reinforced con-
beam model.” J. Intell. Mater. Syst. Struct., 8(6), 489–501. crete bridge columns.” J. Struct. Eng., 10.1061/(ASCE)0733-9445
Baker, J. W., Lin, T., Shahi, S. K., and Jayaram, N. (2011). “New ground (2004)130:6(869), 869–879.
motion selection procedures and selected motions for the PEER trans- Lindt, J. W., and Potts, A. (2008). “Shake table testing of a superelastic
portation research program.” PEER Technical Rep. 2011/03, PEER shape memory alloy response modification device in a wood shear
Center, Univ. of California, Berkeley, CA. wall.” J. Struct. Eng., 10.1061/(ASCE)0733-9445(2008)134:8(1343),
Bentz, E. C., Vecchio, F. J., and Collins, M. P. (2006). “Simplified modified 1343–1352.
compression field theory for calculating shear strength of reinforced Mander, J. B., Priestley, M. J. N., and Park, R. (1988). “Theoretical stress-
concrete elements.” ACI Struct. J., 103(4), 614–624. strain model for confined concrete.” J. Struct. Eng., 10.1061/(ASCE)
Berry, M., and Eberhard, M. (2003). “Performance models for flexural 0733-9445(1988)114:8(1804), 1804–1826.
damage in reinforced concrete columns.” PEER Technical Rep. Marsh, L. K., and Stringer, S. J. (2013). “Performance-based seismic bridge
2003/18, PEER Center, Univ. of California, Berkeley, CA. design, a synthesis of highway practice.” NCHRP Synthesis-440, TRB,
Billah, A. H. M. M., and Alam, M. S. (2012). “Seismic performance of Washington, DC.
concrete columns reinforced with hybrid shape memory alloy Menegotto, M., and Pinto, P. E. (1973). “Method of analysis for cyclically
(SMA) and fiber reinforced polymer (FRP) bars.” Constr. Build. Mater., loaded R.C. plane frames including changes in geometry and non-
28(1), 730–742. elastic behaviour of elements under combined normal force and
Billah, A. H. M. M., and Alam, M. S. (2013). “Statistical distribution of bending.” Symp. on the Resistance and Ultimate Deformability of
seismic performance criteria of retrofitted multi-column bridge bents Structures Acted on By Well-Defined Repeated Loads, International
using incremental dynamic analysis: A case study.” Bull. Earthquake Association for Bridge and Structural Engineering, Zurich, Switzerland,
Eng., 11(6), 2333–2362. 15–22.
Billah, A. H. M. M., and Alam, M. S. (2014). “Seismic fragility assessment Mirza, S. A., Hatzinikolas, M., and MacGregor, J. G. (1979). “Statistical
of concrete bridge piers reinforced with superelastic shape memory descriptions of strength of concrete.” J. Struct. Eng., 105(ST6),
alloy.” Earthquake Spectra, 31(3), 1515–1541. 1021–1036.

© ASCE 04016140-12 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140


National Research Council of Canada. (2014). “Canadian highway bridge Saiidi, M. S., and Ardakani, S. M. S. (2012). “An analytical study of
design code.” CAN/CSA-S6-14, Ottawa, ON. residual displacements in RC bridge columns subjected to near-fault
Naumoski, N., Tso, W. K., and Heidebrecht, A. C. (1988). “A selection of earthquakes.” Bridge Struct., 8(1), 35–45.
representative strong motion earthquake records having different A/V Saiidi, M. S., O’Brien, M., and Zadeh, M. S. (2009). “Cyclic response
ratios.” EERG Rep. 88-01, Dept. of Civil Engineering, McMaster Univ., of concrete bridge columns using superelastic nitinol and bendable
Hamilton, ON, Canada. concrete.” ACI Struct J., 106(1), 69–77.
O’Brien, M., Saiidi, M. S., and Zadeh, M. S. (2007). “A study of concrete Saiidi, M. S., and Wang, H. (2006). “Exploratory study of seismic response
bridge columns using innovative materials subjected to cyclic loading.”
of concrete columns with shape memory alloys reinforcement.” ACI
Rep. No. CCEER-07-01, CCEER, Dept. of Civil Engineering, Univ. of
Struct. J., 103(3), 435–442.
Nevada, Reno, NV.
Ocel, J., et al. (2004). “Steel beam-column connections using shape SeismoMatch version 2.1 [Computer software]. SeismoMatch Earthquake
memory alloys.” J. Struct. Eng., 10.1061/(ASCE)0733-9445(2004) Engineering Software Solutions, Pavia, Italy.
130:5(732), 732–740 . Seismostruct version 6.0 [Computer software]. Pavia, Italy.
Omori, T., et al. (2011). “Superelastic effect in polycrystalline ferrous Shrestha, K. C., et al. (2013). “Feasibility of Cu–Al–Mn superelastic alloy
Downloaded from ascelibrary.org by University Of British Columbia on 04/12/19. Copyright ASCE. For personal use only; all rights reserved.

alloys.” Science, 333(6038), 68–71. bars as reinforcement elements in concrete beams.” Smart Mater.
Paulay, T., and Priestley, M. N. J. (1992). Seismic design of reinforced Struct., 22(2), 025025.
concrete and masonry buildings, Willey, New York. Tanaka, Y., Himuro, Y., Kainuma, R., Sutou, Y., Omori, T., and Ishida, K.
PEER (Pacific Earthquake Engineering Research). (2011). “New ground (2010). “Ferrous polycrystalline shape-memory alloy showing huge
motion selection procedures and selected motions for the PEER trans- superelasticity.” Science, 327(5972), 1488–1490.
portation research program.” PEER Rep. 2011/03, Univ. of California, Vamvatsikos, D., and Cornell, C. A. (2002). “Incremental dynamic
Berkeley, CA. analysis.” Earthquake Engr. Struct. Dyn., 31(3), 491–514.
Pettinga, D., Pampanin, S., Christopoulos, C., Priestley, N. (2006). Vecchio, F. J., and Collins, M. P. (1986). “The modified compression-filed
“Accounting for residual deformations and simple approaches to their
theory for reinforced concrete elements subjected to shear.” ACI Struct.
mitigation.” 1st European Conf. on Earthquake Engineering and
J., 83(2), 219–231.
Seismology, Geneva, Switzerland.
Pinho, R., Casarotti, C., and Antoniou, S. (2007). “A Comparison of single- Wang, D. S., Ai, Q. H., Li, H. N., Si, B. J., and Sun, Z. G. (2008).
run pushover analysis techniques for seismic assessment of bridges.” “Displacement based seismic design of RC bridge piers: Method
Earthquake Eng. Struct. Dyn., 36(10), 1347–1362. and experimental evaluation.” 14th World Conf. on Earthquake Engi-
Priestley, M. J. N., Seible, F., and Calvi, G. M. (1996). Seismic design and neering, Beijing, China.
retrofit of bridges, Wiley, New York. Youssef, M. A., Alam, M. S., and Nehdi, M. (2008). “Experimental
Ramirez, C. M., and Miranda, E. (2012). “Significance of residual drifts investigation on the seismic behavior of beam-column joints reinforced
in building earthquake loss estimation.” Earthquake Eng Struct Dyn., with superelastic shape memory alloys.” J. Earthquake Eng., 12(7),
41(11), 1477–1493. 1205–1222.

© ASCE 04016140-13 J. Struct. Eng.

J. Struct. Eng., 2016, 142(12): 04016140

You might also like