Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Mechanics Based Design of Structures and Machines

An International Journal

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/lmbd20

An experimental study on a self-centering damper


based on shape-memory alloy wires

Afsaneh Falahian, Payam Asadi, Hossein Tajmir Riahi & Mahmoud


Kadkhodaei

To cite this article: Afsaneh Falahian, Payam Asadi, Hossein Tajmir Riahi & Mahmoud
Kadkhodaei (2021): An experimental study on a self-centering damper based on shape-
memory alloy wires, Mechanics Based Design of Structures and Machines, DOI:
10.1080/15397734.2021.1939048

To link to this article: https://doi.org/10.1080/15397734.2021.1939048

Published online: 21 Jun 2021.

Submit your article to this journal

Article views: 65

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=lmbd20
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES
https://doi.org/10.1080/15397734.2021.1939048

An experimental study on a self-centering damper based on


shape-memory alloy wires
Afsaneh Falahiana, Payam Asadia, Hossein Tajmir Riahib , and Mahmoud
Kadkhodaeic
a
Department of Civil Engineering, Isfahan University of Technology, Isfahan, Iran; bDepartment of Civil
Engineering, University of Isfahan, Isfahan, Iran; cDepartment of Mechanical Engineering, Isfahan University of
Technology, Isfahan, Iran

ABSTRACT ARTICLE HISTORY


This experimental study investigates the behavior of a proposed damper Received 31 March 2021
based on austenite Nitinol wires. The damper’s design is so that the wires Accepted 1 June 2021
are always in tension under both tensile and compressive forces. Also, the
KEYWORDS
damper can employ steel wires as a hybrid to enhance the performance of
Passive self-centering
the damper. Steel wires can be employed in parallel with the SMA wires damper; shape-memory
and in the same manner. Therefore, SMA and steel wires are stretched sim- alloy wires; residual strain;
ultaneously in the hybrid damper state. This experimental study includes hybrid damper; heat study
the process of training the wires and the main cyclic tests with different
strain rates which leads to the extraction of the hysteresis curves of the
proposed damper. A thermal study is also conducted on the employed
shape-memory alloy (SMA) wires during the cyclic loadings. The results
indicate that the increase of loading frequency causes the SMA wires to
heat; thus, the inner area of the hysteresis curve decreases, resulting in
diminished damping capacity. Also, the damper stiffness and the residual
strain are increased by using the steel wires. The proposed SMA damper is
employed for seismic control of a three-story steel frame to evaluate its
efficiency. The results reveal that the residual drift ratios of the controlled
structure are significantly decreased.

1. Introduction
Reduction of damages, repairs, restoration costs, and residual deformations following probable
earthquakes during the lifetime of buildings, are essential for designing a control system. In this
regard, passive control devices are useful in structural engineering because of the ease of installa-
tion and low maintenance costs (Soong and Spencer 2002; Beheshti and Asadi 2020). Passive
dampers reduce structural damages caused by severe earthquake events. Still, after these events,
there are often large residual deformations in the structure, making reuse of the structure and
dampers impossible. Thus, the use of smart materials such as shape-memory alloys, capable of
undergoing deformation up to 8% strain without leaving significant residual deformations, has
been considered in this study. Resistance to fatigue and corrosion, long shelf-life with moderate
deterioration, the infrequent need for maintenance, good durability, and elimination of residual
strains due to heat and magnetic field are the main features making this alloy usable in dampers
(Casciati, Faravelli, and Veca 2018; Casciati, Torra, and Vece 2018; Chang and Araki 2016; Dolce

CONTACT Payam Asadi asadi@iut.ac.ir Department of Civil Engineering, Isfahan University of Technology, Isfahan
8415683111, Iran.
Communicated by Corina Sandu.
ß 2021 Taylor & Francis Group, LLC
2 A. FALAHIAN ET AL.

Figure 1. The changes in the lattice of SMA under the heating and loading effects.

and Cardone 2001a; Raniecki, Lexcellent, and Tanaka 1992). NickelTitanium Naval Ordnance
Laboratory (Nitinol) alloy has a shape-memory property. Two austenite and martensite phases of
this alloy are dependent on temperature variations. The crystal lattice of the austenite phase is
symmetric body-centered cubic, while the crystal lattice of the martensite phase has less lattice
symmetry than that of the austenite phase (Czechowicz and Langbein 2015). Figure 1 displays the
critical phase-transformation temperatures. In Fig. 1, Af, As, Ms, and Mf indicate the temperature
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 3

of the finish of austenitic phase, the beginning of the austenitic phase, the beginning of the mar-
tensite phase, and the finish of the martensite phase, respectively. Figure 1 also exhibits the load-
ing and heating transfer of the lattice of the shape-memory alloy between twinned martensite,
detwinned martensite, and austenite. The inelastic strain developed in the memory alloy disap-
pears after heating, where the material returns to its original shape. Also, the super-elastic alloy
can return to its original shape after high-strain loading (Raniecki, Lexcellent, and Tanaka 1992;
Smith 2005; Czechowicz and Langbein 2015; Chang and Araki 2016). Lim and McDowell (1995)
subjected Nitinol alloy wires to cyclic loading and observed a super-elasticity behavior, where the
residual stresses were significantly diminished. The evaluation of temperature variations and
stress-strain curves in successive cycles indicated that the temperature and stress-strain profiles
became stable after several cycles (Casciati, Faravelli, and Veca 2018). The SMA’s martensite
phase requires external heat to return to the original shape and exhibits a greater damping cap-
acity and energy dissipation than that of the austenite phase. On the other hand, super-elastic
memory alloy in the austenite phase has an enormous recentering force making it capable of
returning the structure to the original shape and diminishing residual strains (Song, Ma, and
Li 2006).
These smart properties of SMA materials lead to extensive applications of them in different
aspects of the industry (Moradi, Alam, and Asgarian 2014; Abid et al. 2021; Issa and Alam 2020;
Kamarian et al. 2020). Moradi, Alam, and Asgarian (2014) demonstrated that the NiTi SMA bra-
ces can significantly reduce steel frames’ permanent deformations up to 96%. Issa and Alam
(2020) employed NiTi SMA bars to reinforce the piston-based self-centering bracing for seismic
control of four and six story steel buildings. The results indicated a significant reduction in the
residual drifts of the buildings subjected to the natural accelerograms.
So far, Nitinol alloy has been used in the manufacturing of many dampers (Salichs, Hou, and
Noori 2001; Masuda and Noori 2002; Han et al. 2003; Ma and Cho 2008; Ozbulut et al. 2010;
Parulekar et al. 2012; Torra et al. 2014; Gao et al. 2016; Morais et al. 2017; Li, Liu, and Fu 2018;
Zhou et al. 2018). Salichs, Hou, and Noori (2001), as well as Torra et al. (2014), indicated that
dampers’ behavior depends on the excitation frequency. They demonstrated that the SMA-based
dampers in low frequency and amplitude of ground motion had some advantages over vis-
cous dampers.
Since steel wires slack immediately after yielding, hybrid steel-SMA-based damper may better
decrease the residual and peak drifts of the structure (Fang, Wang, and Feng 2019; Wang et al.
2019; Fang et al. 2019a; Wang et al. 2020). The purpose of hybrid dampers is to enhance the
damper’s capacity (stiffness or energy-dissipation capacity) and the self-centering property (Pan
and Cho 2007; Zhu and Zhang 2008; Tang and Lui 2014; Ozbulut, Silwal, and Michael 2015;
Asgarian, Salari, and Saadati 2016; Alipour, Kadkhodaei, and Safaei 2017). Zhu and Zhang (2008)
proposed a friction damper containing SMAs to create the self-centering property. Nonlinear
time-history analysis of the frames equipped with this damper indicated that the peak floor dis-
placements were limited, while there was a low residual drift. This conclusion is similar to
another study finding obtained by Tang and Lui (2014). They presented a hybrid damper using
viscous fluid and SMA wires employed for energy dissipation and self-centering property, respect-
ively. They proposed a device that could be connected to the apex of the chevron brace and
included a set of SMA wires in parallel with a fluid viscous damper. Pan and Cho (2007) eval-
uated a micro-hybrid damper with three rows of SMA wires and springs. Asgarian, Salari, and
Saadati (2016) investigated the cyclic behavior of a self-centering hybrid damper consisting of a
steel tube as a vertical energy dissipation member and two pairs of SMA wires as the self-center-
ing members. Alipour, Kadkhodaei, and Safaei (2017) proposed a damper consisting of a steel
spring for energy dissipation and SMA wires to restore the inelastic strain. Experiments revealed
that the equivalent damping coefficient of the proposed damper was about 11%. They also dem-
onstrated that the effect of changes in the diameter and length of the wires on the cyclic
4 A. FALAHIAN ET AL.

Figure 2. The general shape of the proposed damper.

hysteresis’ general shape was ignorable, whereas the energy dissipation and the system input force
rose at a constant rate by increasing the diameter of SMA wires. However, Casciati, Torra, and
Vece (2018) indicated that thin wires (0.5 mm of diameter or less) are weaker than thick wires,
for dissipating energy by hysteresis loops, in cold-winter areas (i.e. down to 20  C). The SMA-
based dampers are often used in structures with temperatures higher than the mentioned tem-
perature (20  C).
The above-mentioned studies suggested that dampers based on the SMA wires diminished
residual deformations. Also, hybrid dampers with SMA wires were produced to augment the
damper’s performance. Increasing the hybrid damper’s stiffness reduces the peak inter-story drifts
of the structure subjected to the design base earthquake. The purpose of this study is to evaluate
the behavior of the proposed damper based on SMA wires while using steel wires. This damper
has some similarities with the proposed damper by Qiu and Zhu (2016, 2017a, 2017b). However,
some details like the general shape of the damper, the number of used SMA wires, and the
anchor details are different. Furthermore, the experimental and numerical assessments of the pro-
posed damper in the current study are different than those proposed by Qiu and Zhu. Qiu and
Zhu (2016) and Qiu et al. (2020). They focused on the experimental and numerical evaluation of
the frames equipped with SMA-based dampers. However, the main focus of the current study is
to experimentally explore the performance of the SMA-based dampers in both cases of using the
SMA wires and hybrid SMA with steel wires. Another goal of the study is to investigate the
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 5

Figure 3. The upper and bottom force transmitter steel plates.

effects of the temperature of the wires and the rate of cyclic loading on the behavior of the
damper. Subsequently, a numerical investigation was also conducted on a three-story steel frame
controlled with the proposed SMA damper.

2. The general properties of the proposed damper


The proposed damper is somewhat similar to the proposed dampers in some studies (e.g.,
Ozbulut and Hurlebaus 2011; Morais et al. 2017) while offering some advantages and novelties.
The damping capacity and stiffness of this damper can increase by employing additional SMA or
steel wires (unlike the one proposed by Haque, Issa, and Shahria Alam (2019)). Also, the anchor
detail of wires has been improved. It was designed such that the fracture would occur in the mid-
dle region of the wires and away from the support areas where an unexpected fracture may occur.
It is worth mentioning that the proposed damper by Ozbulut and Hurlebaus (2011) was not man-
ufactured and experimentally studied in real practice. Also, in the study of Morais et al. (2017),
few experimental studies were conducted on the damper’s behavior; whereas, this study investi-
gates the effects of load frequency and temperature variation. It also explores the damper’s behav-
ior in the hybrid form by using different ratios of SMA wires to steel wires. As indicated in Figs.
2 and 3, the damper design is in a manner that causes half of the wires to be in tension under
either tensile or compressive force. Therefore, the damper has the same behavior while exposed
to both tension and compression forces. The increase in the damper’s stiffness (to reduce the
inter-story drifts of the structure subjected to the design base earthquake) and residual deform-
ation were estimated in the hybrid damper state. Shape-memory alloy wires, steel plates, and steel
bars constitute the main components of the proposed damper. Figure 2 displays the pro-
posed damper.
The shear and moment capacity of the rods were more than the expected maximum applied
force, so they did not yield during the tests, nor did they undergo significant deformation. The
length and diameter of the rods were 300 mm and 24 mm, respectively. The grades of rods and
steel plates were DIN.10.9 and ST-37, in that order. The width, length, and thickness of the upper
and bottom plates were 270, 340, and 10 mm. The baseplate and endplates were intended to be
used under different external loads from the brace’s other components when the damper is used
as a part of a bracing system.
Cyclic loadings with different frequencies were applied to the endplates in the form of tensile
or compressive forces. These force actions were transferred to the steel rods holding the wires,
and eventually, the force was transferred to the wires. Figure 3 shows the upper and bottom steel
plates. The long-slot and standard holes arrangement cause half of the wires to be stretched each
time (under either compressive or tensile load).
The NiTi shape-memory alloy wires have a low weight compared to their strength, high corro-
sion and wear resistance, and high reversible strain (around 8% without any permanent
6 A. FALAHIAN ET AL.

Table 1. The mechanical properties of the NiTi wires.


Symbol Description Unit Value
E Elastic modulus (Young’s modulus) MPa 346.22
Ep Plastic modulus MPa 37.3

Mf Martensitic finish temperature at zero stress C 2.28

Ms Martensitic start temperature at zero stress C 16.59

As Austenitic start temperature at zero stress C 0.13

Af Austenitic finish temperature at zero stress C 16.89
eL Maximum recoverable strain (see Fig. 14) % 4.5
ep Maximum plastic strain % 19
rcrs Critical stress for start of martensite detwinning MPa 377
rcrf Critical stress for finish of martensite detwinning MPa 507
D Diameter mm 0.5

deformation for low fatigue cycles) (Dolce and Cardone 2001b). In this study, the wires were in
the austenite phase. The diameter of the SMA wires was 0.5 mm. Table 1 presents the employed
SMA wires’ mechanical properties, which are calculated from the tension test. The maximum
recoverable strain (eL ) and maximum plastic strain (eP ) are obtained by the STM-50 servo trac-
tion machine. The employed SMA wires were in their super-elastic region, so their formation and
anchoring were difficult.
Due to the SMA wires’ small diameter, holding the wires is challenging (Zareie and Zabihollah
2020). Therefore, various methods have been employed to restrain the SMA wires. The anchoring
details, in this work, include the integrated use of SMA wires in touch with a cone (Ma and Cho
2008), the use of special clamps (Ozbulut et al. 2010; Morais et al. 2017), and the application of
the locking screw (Parulekar et al. 2012). Zareie and Zabihollah (2020) also employed an in-house
holding mechanism composed of plate, nuts, and bolts to keep the SMA wires in place. In this
research, using a simple trick, even if some of the wires were torn, the damper could operate
without having lost a significant amount of its capacity. Also, as the number of wires increased,
no-slip condition occurred among the wires, while the damper’s properties did not change. This
trick involved bending the two ends of the wire in a circular shape and pressing them into three
steel nuts (M1.6 based on DIN 934) to employ materials’ frictional properties (Fig. 4). Note that
pressing did not cause a considerable change in the SMA wires’ thermodynamic properties, nor
did it create any unexpected stress concentration. It did, however, directed the wire’s yielding
point to the middle parts (will be shown in Section 6.3). These rings were mounted on the non-
threaded portion of the rods. The number of SMA wires was chosen such that their ultimate cap-
acity would be far less than the rods’ shear yield strength. A comparison of the hysteresis curves
of the wires (will be shown in Fig. 10) and dampers (will be shown in Fig. 11) shows that slip
reduction has not happened. Consequently, this means that the contact surface of the wires and
rods is low and is not effective. The tension test revealed that the effective length of the wires
is 63 mm.

3. Experiments
Cyclic tests with low-frequency and high-frequency rates were conducted on the SMA wires and
dampers. The experiments were carried out in the mechanics of materials laboratory of the
department of mechanical engineering of Isfahan University of Technology. The STM-50 servo
traction machine (Fig. 5) was employed for low-frequency static tests. The minimum and max-
imum load cell capacities were 500 N and 50 kN, respectively. Also, the minimum and maximum
operable loading frequencies were 0.001 and 0.1 Hz, in that order. Additionally, the SAF-50 axial
fatigue hydraulic machine was employed for the high-frequency dynamic tests (Fig. 6). This
device had two mobile and fixed jaws capable of imposing cyclic loads with frequencies of
0.01–5 Hz and a maximum magnitude of 10 kN.
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 7

Figure 4. The anchor detail of the SMA rings.

Displacement-controlled cyclic tests were applied to the damper according to the defined load-
ing protocol by using these machines. During all tests, the laboratory temperature lay within the
range of 20–22  C. Figure 7 illustrates how the damper works under pressure and tension. As can
be observed, half of the wires were in the tensile state under both tensile and compressive exter-
nal loads.

4. Loading protocol
Although several loading protocols as a quasi-static testing method have already been proposed
for cyclic testing of devices and structural members. Most of them have focused on the perform-
ance of conventional structural systems. There are a few protocols for self-centering devices. The
study on self-centering devices is distinct in terms of deformation demand and the number of
damaging cycles (Fang et al. 2019b; Fang, Ping, and Chen 2020). The parameters that determine
the protocol include the total number of cycles, primary sub-cycles, main cycles, and duplicate
cycles. ATC-24 (1992), Clark et al. (1997), and DesRoches, McCormick, and Delemont (2004)
loading protocols are among the most common protocols employed in experiments. In the
DesRoches, McCormick, and Delemont (2004) protocol (SMA protocol), the displacement contin-
ues until a complete transformation, followed by several cycles of displacement. In this study, as
indicated in Fig. 8, the employed incremental protocol was similar to the SMA protocol. The
negative parts of the loading were not performed in the cyclic tests on the wires.

5. Results of the experiments on the wires


5.1. Results of the tests on the SMA wires
In this study, various experiments were carried out to investigate the SMA wires’ mechanical and
thermo-mechanical properties. These experiments included the differential scanning calorimetry
(DSC) test for calculating the transformation temperatures, the cyclic test for evaluating the
mechanical properties, and extra tests for calculating the effective wire length and the required
numbers of SMA training wires. The training cycles are performed to minimize residual strains
and degradation of energy dissipation convergence. Accordingly, the number of training cycles
was obtained. In the DSC test, the test sample was heated from temperatures well below the aver-
age ambient temperature (60  C) to the maximum temperature (60  C), then cooled to low tem-
peratures. Transition energy changes by temperature were plotted in two heating and cooling
8 A. FALAHIAN ET AL.

Figure 5. The experimental setup for the low-frequency cyclic loading tests by STM50 machine.

cycles. The SMA wire’s transformation temperatures are estimated based on the tangent slopes of
the curves within the range of variations (Shayanfard, Kadkhodaei, and Jalalpour 2019). The DSC
test results indicated that the temperature of the Mf, Ms, As, and Af are 2.28  C, 16.59  C,
0.13  C, and 16.89  C, respectively. Accordingly, to use the SMA wires’ super-elastic property in
the austenite phase, the ambient temperature during testing should be above 16.89  C.
Since wire rings (a ring on each side of the wires) have been used in the damper, the wires’
effective tensile length with two rings at each side should be obtained. Initially, a sample wire was
inserted from the nuts between the jaws for the cyclic test at a low strain rate (0.00016 Hz loading
frequency in three cycles). Then, a similar test was conducted on three sample wires with two
rings on each side. Afterward, the effective length was obtained equal to 63 mm by comparing the
wires’ strain value with two rings on each side with the value of the bare SMA wires.
The SMA wires were trained by cyclic loading to stabilize the response and the nonelastic
strain (Casciati, Faravelli, and Vece 2018). As the number of cycles grows, the plastic strains
increase. When the stabilized response was achieved, the macroscopic strain was stored within
the wire. Figure 9 indicates that after 50 cycles of loading (with 0.001 Hz loading frequency and
up to 8% strain), the equivalent viscous damping and the residual extension were stabilized.
Afterward, the equivalent viscous damping was calculated as follows (Jennings 1968):

ED
neq ¼ (1)
2pKs d2
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 9

Figure 6. The setup of the high-frequency cyclic loading tests by SAF-50 machine.

Figure 7. The state of the damper under (a) compressive, (b) tensile loads.

Fmax  Fmin
Ks ¼ (2)
dmax  dmin

where neq represents the equivalent viscous damping ratio, ED is the dissipation energy in a
force-deformation cycle, d shows the maximum deformation in a cycle, and Ks indicates the
10 A. FALAHIAN ET AL.

Figure 8. The protocol utilized for the cyclic tests.

Figure 9. The stabilization process of the equivalent viscous damping and the residual extension under the cyclic test.

equivalent stiffness. Fmax and Fmin denote the maximum and minimum forces in a cycle, respect-
ively. Also, dmax and dmin are the maximum and minimum deformations related to maximum
and minimum forces, in that order.
After training the SMA wires, cyclic tests were carried out on them. Figure 10 indicates the stress-
strain curves of the SMA wires under cyclic tests (with 0.00016 Hz loading frequency) to the point of
failure. The maximum strains of samples are 6–7%, which reached this value in three loading cycles.
This figure illustrates the self-centering behaviors of the utilized SMA wires. Also, as expected from
previous studies (Dolce and Cardone 2001b; Casciati, Faravelli, and Vece 2017), fatigue does not
occur in the hysteresis curves of wires under low loading cycles (compared to earthquake excitation).
Casciati, Faravelli, and Vece (2017) showed that the hysteresis curves of thin wires under loads of less
than 120 cycles are almost constant and failure occurs in cycles more than 3000.
Testo camera was used for the thermal analysis of SMA wires. The ambient temperature was
21  C during the cyclic test, which is higher than 16.89  C. Therefore, the wires were in the aus-
tenite phase which then transformed into a detwinned martensite phase during the cyclic loading
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 11

Figure 10. The stress-strain curve of the SMA wires.

Figure 11. The hysteresis curves of the damper.

and returned to austenite after unloading. The transformation of austenite to martensite phase
increased the temperature of the wire (Casciati, Torra, and Vece 2018). It was indicated that the
temperature of the selected point was 23  C before loading, which increased to 33.1  C after load-
ing at the time of transformation.
12 A. FALAHIAN ET AL.

Figure 12. The equivalent viscous damping ratio of the damper in terms of damper extension.

5.2. Results of tests on the steel wires


In this part of the study, three samples of steel wires of 0.6 mm diameter were stretched at the
rate of 1 mm s1 (strain rate: 0.00016 s1). The average yield stress and yield strain of the wires
were 240 MPa and 0.12%, respectively. Also, the average poststiffness of the steel wires was 0.6%
of the initial elastic stiffness.

6. Experiments on the damper


This section presents the results of the cyclic tests on the proposed damper. Therefore, the effect
of loading frequency and temperature on the hysteresis curve of the damper are investigated.
Additionally, the results for the hybrid damper using steel wires are presented.

6.1. Investigation of the performance of the damper


The displacement-controlled cyclic test was conducted with a maximum displacement of 12 mm
(19% strain) at a strain rate of 0.0003 s1. The total cross-sectional area of the SMA wires, which
were trained before the test, was 11.78 mm2. Figure 11 displays the hysteresis curves of the test.
As mentioned, the effective length of the wires was 63 mm. This figure illustrates that the damper
has self-centering behavior and can withstand significant strains with trivial residual displacement.
Also, the damper’s hysteresis curves, especially in the case of elastic stiffness, are very similar to
the hysteresis curves of the SMA wires (Fig. 10). Thus, it is revealed that the contribution of
other components, such as friction between the steel rods, to dissipating the energy is ignorable.
Figure 12 indicates the equivalent viscous damping ratio in terms of the damper deformation.
It can be observed that the equivalent viscous damping ratio was increased by increasing the
applied strain up to 4.5% and then remained approximately constant by increasing the applied
strain up to 16%. Initially, the equivalent viscous damping ratio grew almost linearly; however, it
remained constant in the later cycles. Figure 10 indicates that the SMA wires’ stress has remained
almost constant during the martensite detwinning phase. Therefore, as shown in Fig. 12, the
equivalent viscous damping has also remained almost constant in this phase.
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 13

Figure 13. The force-extension curves for the different strain rates; (a) pseudo-static test, (b) dynamic test.

Figure 14. The effect of loading frequency on the equivalent damping ratio and residual extensions of the damper

6.2. The effect of loading frequency


Generally, the increase in the loading frequency reduces the energy dissipation and thus the
damping ratio (Pieczyska et al. 2005). Since most civil engineering structures’ natural frequency is
less than 5 Hz, the damper performance was investigated up to this frequency. In classifying the
type of cyclic tests on NiTi wires, the experiments with the strain rate up to 0.0033 s1 are
defined as pseudo-static, while those with higher strain rates are considered dynamic (Dayananda
and Rao 2008; Soul et al. 2010). Figure 13 displays the force-extension curve of the damper for
different strain rates (up to 8% strain with six cycles).
An increase in the strain rate increases the temperature of the shape-memory alloy (Torra et
al. 2014); thus, the high and low transformation stresses approach each other (from 400 to
200 MPa), while the damping energy decreases. This phenomenon is called the effect of the adia-
batic heating and cooling process due to the high frequency of loading and unloading. Figure 14
presents the effect of loading frequency on the equivalent viscous damping ratio and residual
deformations (for cycles with a maximum strain of 8%). The residual drifts are created when the
14 A. FALAHIAN ET AL.

Figure 15. Thermal diagram of SMA wires under the cyclic loading; (a) unloading, (b) peak deformation: 1 mm, (c) peak deform-
ation: 2 mm, (d) peak deformation: 3 mm, (e) peak deformation: 4 mm, (f) peak deformation: 5 mm (the left side is the selected
window where the point with the peak temperature is displayed, and the right side is the temperature diagram for the
selected window).

SMA wires experience plastic strains (Alam, Youssef, and Nehdi 2007; Zareie et al. 2020a, 2020b).
However, it is not easy to distinguish them from Fig. 11. Thus, Fig. 14 reports the great residual
drifts for the high equivalent damping ratios. Also, as the strain rate increases, the residual drift,
and the equivalent damping ratio decrease. The rate of change is relatively linear. Almost all tem-
perature changes are dissipated and have ignorable effects during the transformation in low-fre-
quency loading (up to 0.01 Hz). On the other hand, the dissipated heat is very small and
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 15

Figure 15. Continued.

negligible in high-frequency loading (more than 1 Hz). Therefore, the maximum contribution to
the hysteresis curve happens in intermediate-frequency loading (Soul et al. 2010).

6.3. The effect of cyclic loading on the temperature of the SMA wire
The effect of amplitude and strain rate of cyclic loading was investigated on the SMA wires’ tempera-
ture. Austenite is a stable phase at high temperatures and low stress; whereas, martensite is stable at
16 A. FALAHIAN ET AL.

Figure 16. Thermal diagram of SMA wires under the strain rates of (a) 0.0013 s1, (b) 0.015 s1.

Table 2. The results of cyclic tests on the hybrid dampers.


Area of steel Area of SMA Axial Stiffness The ultimate Residual
Test No. a wires (mm2) wires (mm2) (N mm1) force (N) strain (%)
1 0 0 11.78 6.975 5718.74 0.35
2 0.11 1.13 10.6 9.863 5519.29 0.57
3 1.44 6.78 4.71 24.312 5168.24 0.76

lower temperatures and higher stress. The transformation of the austenite to martensite phase begins
at a point with higher stress (Morais et al. 2017). Initially, the strain rate was 0.015 s1, while the
amplitude of the cyclic loadings rose incrementally. Figure 15 exhibits the location of nucleation, for-
ward movement of temperature, and temperature distribution by using a thermal camera (Testo cam-
era). The increase in the temperature of a point of the wire can indicate the nucleation of the new
martensite phase since initially the temperature of all areas is uniform. Nucleation is the beginning of
the transformation in the potential points and starts from the points that were pressed by the nuts,
due to the stress concentration in them. As the phase transformation develops, the high-temperature
front moves along the wires, until all parts of the wires transform from austenite to martensite phase.
During the cyclic loadings, the SMA wires’ temperature increases from 21.6 to 33.1  C.
The thermal camera recorded the SMA wires’ temperature at two strain rates of 0.0013 and
0.015 s1 (as shown in Fig. 16). The results revealed that the elevation of the strain rate increases
the number of nucleation points and the wires’ temperature. The wires’ maximum temperatures
at the deformation of 5 mm under the strain rates of 0.0013 and 0.015 s1 were 26.3  C and
33.1  C, respectively. This is due to wires’ isothermal behavior at low strain rates (less than
0.001 s1) and adiabatic state at higher strain rates (greater than 0.01 s1).
As mentioned in the previous section, the increase in the temperatures augments the alloy’s
stress level and narrows the hysteresis curves, diminishing energy dissipation capacity.

6.4. The experimental results of the hybrid damper


Steel wires have more modules of elasticity than SMA wires. At the same time, SMA wires pos-
sess self-centering property. The optimum hybrid damper has an adequate stiffness along with
the self-centering property to limit the structure’s peak inter-story drift to a permitted value,
while subjected to the design base earthquake hazard level. According to Eq. (3), steel wires with
different ratios (a) were added to hybridize the dampers:

ASteel
a¼ (3)
ASMA
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 17

Figure 17. The hysteresis curves of the hybrid damper with the different ratios of steel to SMA wires, (a) a ¼ 0, (b) a ¼ 0.11, (a)
a¼ 1.44, (a) comparison of final curves with each other.

where ASMA and ASteel represent the areas of the SMA and steel wires, respectively. When the
damper includes only the SMA wires, a is equal to zero. Table 2 reports the cyclic test results of
the hybrid dampers with different ratios of a, exposed to a 0.0003 s1 strain rate, where axial stiff-
ness denotes AE/L (L is the effective length of the wires which is 63 mm).
Figure 17 reveals the hysteresis curves of the hybrid dampers. Once the steel wires yield,
residual displacements occur, while for the SMA wires, the residual strain is always negligible.
The steel wires’ stiffness is greater than that of the SMA wires, but the SMA wires’ yield and
ultimate strengths are greater than those of the steel wires. Therefore, a minor decrease occurs in
the ultimate capacity of hybrid dampers.
18 A. FALAHIAN ET AL.

Figure 17. Continued.

As expected, although further use of steel wires increases the hybrid dampers’ residual defor-
mations, the damper’s stiffness is significantly augmented. Considering 0.5% for the permissible
residual strain ratio (0.63 mm deformation) and according to the second damage state of FEMA
P-58 (2012), using linear extrapolation, a is approximately 0.075. Also and the damper’s stiffness
using linear interpolation is about 8.944 N mm1. This means that the stiffness of the hybrid
damper is elevated by about 28% compared to the pure SMA damper, while the ultimate capacity
force change of the damper is negligible. Increasing the hybrid damper’s stiffness reduces the
peak inter-story drifts of the structure subjected to the design base earthquake.
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 19

Figure 18. The general shape of the three-story frame and the cyclic curves of its dampers.

Table 3. The characteristics of the designed braced frames.


Spectral response
Short Period of Wires Elastic stiffness of Postactivation stiffness Elastic base-shear of Fundamental
period (Ss) 1 s (S1) area (cm2) damper (kN m1) of damper (kN m1) the structure (kN) Period (s)
1.0 g 0.44 g 8.55 9330.53 1011.89 322.57 0.90

7. Application in the seismic control of the structures


This section investigates the proposed damper’s application in the seismic control of an inter-
mediate steel frame of three stories and three bays. Although the experimentally investigated
damper is small-scale, it is modeled based on the obtained results in previous sections. The
authors recommend using the SMA cables in a large-scale damper as well. The results of a previ-
ous study showed that SMA cables have similar properties and characteristics to SMA wires
(Fang et al. 2019b). The seismic force-resistant system consists of a dual system of intermediate
steel moment frames and special concentrically braced frames (permitted earthquake resistant sys-
tem in accordance with ASCE0707 (2016). The frame is loaded and designed according to
ASCE0707 (2016) and AISC360360 (2010). Figure 18 indicates the general shape of the structure.
Table 3 lists the characteristics of the designed braced frames.
The design of the damper is so that the axial stiffness of the dampers, for this structure, is
equal to the braces’ axial stiffness (Ozbulut et al. 2010). The demand ductility of the uncontrolled
structure (with steel braces) was about 3.0, so it is estimated that the demand ductility of the con-
trolled structure falls in the same range as well. Also, since the inter-story drift ratio of the struc-
ture was limited to 4%, the target inter-story of the SMA wires is limited to the same value. Note
that the P-Delta effect is considered by the first-order treatment of geometric nonlinearity, while
the degradation of connection strength is not considered. The structural modeling is performed
in OpenSees (McKenna 2011) which is a powerful software for modeling structures and SMA-
based elements (Zareie et al. 2020a, 2020b). Steel02 and self-centering elements are assigned to
the structural steel elements and wires, respectively. Beta (the forward to reverse activation stress/
force ratio (McKenna, 2011) and “sigAct” (the forward activation stress/force (McKenna 2011)
are obtained from the hysteresis curve of Fig. 11 to be 0.9 and 376.949, respectively. Due to the
few numbers of cycles of the ground motions and the low period of the three-story structure, a
low-frequency hysteresis curve is used. Figure 19 illustrates the accuracy of the hysteresis curves
modeled in OpenSees. Table 4 compares the energy dissipation capacity, initial or postelastic stiff-
ness, and residual drift of the experimental and numerical results. The structure is studied by
20 A. FALAHIAN ET AL.

Figure 19. The comparison of hysteresis curves of the numerical and experimental studies.

Table 4. The comparison of the energy dissipation capacity, initial or postelastic stiffness, and residual drift of the experimen-
tal and numerical results.
Experimental Numerical Error
Energy dissipation capacity (J) (for the cycle with peak strain) 245.2 216.7 28.5
Initial stiffness (N mm1) 350.3 407.5 57.2
Postelastic stiffness (N mm1) 62.3 43.7 18.6
Residual drift (for the cycle with peak strain) 0.5% 0% 0.5%

Table 5. The selected accelerograms from FEMA 440 (2005).


No. Earthquake names Magnitude (Ms) Station number Component ( ) PGA (cm s2)
1 Imperial Valley 6.8 5051 315 200.2
2 San Fernando 6.5 8005 90 107.9
3 San Fernando 6.5 269 21 133.4
4 Landers 7.5 12,149 0 167.8
5 Loma Prieta 7.1 58,065 0 494.5
6 Loma Prieta 7.1 47,006 67 349.1
7 Loma Prieta 7.1 58,135 360 433.1
8 Loma Prieta 7.1 1652 270 239.4
9 Morgan Hill 6.1 57,383 90 280.4
10 Palmsprings 6.0 5069 45 129.0
11 Northridge 6.8 24,278 360 504.2

performing nonlinear dynamic analyses. Thus, 11 far-fault and nonpulse-like accelerograms on-
site class C were selected from FEMA 440 (2005) (listed in Table 5). The accelerograms are scaled
according to ASCE0707 (2016).
Figure 20 displays the inter-story drift ratios of the structure’s third story (roof) exposed to
“Northridge” ground motion (a selected ground motion from Table 5). The results reveal that the
SMA damper significantly decreases the residual drift of the structure.
Table 6 indicates the maximum and average peak inter-story as well as residual drift ratios of
the structure exposed to all ground motions. It is observed that the SMA damper considerably
diminishes the residual drift ratios, while the changes of the inter-story drift ratios are ignorable.
The SMA-based damper diminished the residual inter-story drift ratio of the uncontrolled struc-
ture by about 80%. Consequently, the controlled structure satisfies the second damage state of
FEMA P-58 (2012).
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 21

Figure 20. The inter-story drift ratio of the roof story subjected to the Northridge ground motion.

Table 6. The inter-story and residual drift ratios of the structure subjected to 11 ground motions.
Inter-story drift ratio (%) Residual drift ratio (%)
Maximum Average Maximum Average
Uncontrolled 2.95 2.28 1.14 0.62
Controlled 3.03 2.24 0.14 0.07
Decrease percentage 2.71 1.75 87.72 88.71

8. Summary and conclusions


A proposed damper based on the SMA wires was experimentally studied. The damper’s behavior
was investigated while being subjected to the variations of loading (in case of amplitude and
strain rate) and in the hybrid state. Temperature studies were conducted on the SMA wires as
well. The damper was designed so that half of the wires were in tension under both tensile and
compressive forces. The friction property was employed to restrain the wires, which did not alter
the material’s thermodynamic properties. After the damper’s initial training using cyclic tests, it
provided a stable cyclic behavior with considerable ductility. Therefore, due to the self-centering
behavior of the damper, it is possible to use it repeatedly. A numerical study was then conducted
on a three-story steel frame to evaluate the proposed damper’s capabilities in structural seismic
control. The findings can be summarized as follows:

 The minimum deformation applied to transform the damper’s wires was about 1.5 mm (about
2.4% strain). By increasing the cyclic loading amplitude, the equivalent viscous damping ratio
was enhanced by 11%.
 The damper has self-centering behavior and can withstand up to 12 mm extension (19%
strain) with trivial residual displacement (up to 0.65%).
 When the damper was exposed to the cyclic loading of 0.0013 s1 strain rate, the maximum
wire temperature at the deformation of 5 mm was 26.3  C. Meanwhile, the damper’s tempera-
ture increased to 33.1  C for the greater frequency of the cyclic loading of 0.015 s1. This
increase in the loading frequency augmented the SMA’s internal heat, thus reducing the
damping capacity to 70%.
22 A. FALAHIAN ET AL.

 The damper’s stiffness increased by combining SMA and steel wires, though residual strain
occurred in this hybrid damper. The stiffness of a hybrid damper with the permissible residual
strain (0.5%) was about 28% greater than that of the pure SMA damper.
 The proposed damper reduced the residual drifts of the three-story steel frame up to 87%,
while the effects on the peak inter-story drift ratios were negligible.

Disclosure statement
No potential conflict of interest was reported by the authors.

ORCID
Hossein Tajmir Riahi http://orcid.org/0000-0001-8891-1455
Mahmoud Kadkhodaei http://orcid.org/0000-0002-9353-4583

References
Abid, F., Hami, A. E., Merzouki, T., Walha, L., and Haddar, M. 2021. An approach for the reliability-based design
optimization of shape memory alloy structure. Mechanics Based Design of Structures and Machines 2, 49:155–71.
doi:10.1080/15397734.2019.1665541.
AISC360. 2010. Specification for structural steel buildings. AISC 360-10. Chicago, IL: American Institute of Steel
Construction.
Alam, M. S., Youssef, M. A., and Nehdi, M. 2007. Utilizing shape memory alloys to enhance the performance and
safety of civil infrastructure: review. Canadian Journal of Civil Engineering 34(9):1075–86. doi:10.1139/l07-038.
Alipour, A., Kadkhodaei, M., and Safaei, M. 2017. Design, analysis, and manufacture of a tension–compression
self-centering damper based on energy dissipation of pre-stretched superelastic shape memory alloy wires.
Journal of intelligent material systems and structures 28(15):2129–39. doi:10.1177/1045389X16682839.
ASCE07. 2016. Minimum design loads and associated criteria for buildings and other structures. Reston, VA:
American Society of Civil Engineers.
Asgarian, B., Salari, N., and Saadati, B. 2016. Application of intelligent passive devices based on shape memory
alloys in seismic control of structures. Structures 5:161–169. doi:10.1016/j.istruc.2015.10.013.
ATC-24. 1992. Guidelines for cyclic seismic testing of component of steel structures. Redwood City, CA: Applied
Technology Council.
Beheshti, M., and Asadi, P. 2020. Optimal seismic retrofit of fractional viscoelastic dampers for minimum life-cycle
cost of retrofitted steel frames. Structural and Multidisciplinary Optimization 61(5):2021–35. doi:10.1007/s00158-
019-02454-w.
Casciati, S., Faravelli, L., and Vece, M. 2017. Investigation on the fatigue performance of Ni-Ti thin wires.
Structural Control and Health Monitoring 24(1):e1855. doi:10.1002/stc.1855.
Casciati, S., Faravelli, L., and Vece, M. 2018. Long-time storage effects on shape memory alloy wires. Acta
Mechanica 229:697–705. doi:10.1007/s00707-017-1993-2.
Casciati, S., Torra, V., and Vece, M. 2018. Local effects induced by dynamic load self-heating in NiTi wires of
shape memory alloys. Structural Control and Health Monitoring 25(4):e2134. doi:10.1002/stc.2134.
Chang, W.-S., and Araki Y. 2016. Use of shape memory alloy in construction: A critical review. Proceedings of the
ICE – Civil Engineering 169:87–89. doi:10.1680/jcien.15.00010.
Clark, P., Frank, K., Krawinkler, H., and Shaw, R. 1997. Protocol for fabrication, inspection, testing, and documenta-
tion of beam-column connection tests and other experimental specimens. SAC Steel Project Background
Document. October, Report No. SAC/BD-97/02.
Czechowicz, A., and Langbein, S. 2015. Shape memory alloy valves: Basics, potentials, design. Switzerland: Springer.
Dayananda, G., and Rao, M. S. 2008. Effect of strain rate on properties of superelastic NiTi thin wires. Materials
Science and Engineering: A 486(1–2):96–103. doi:10.1016/j.msea.2007.09.006.
DesRoches, R., McCormick, J., and Delemont, M. 2004. Cyclic properties of superelastic shape memory alloy wires
and bars. Journal of Structural Engineering 130(1):38–46. doi:10.1061/(ASCE)0733-9445(2004)130:1(38).
Dolce, M., and Cardone, D. 2001a. Mechanical behaviour of shape memory alloys for seismic applications 1.
Martensite and austenite NiTi bars subjected to torsion. International Journal of Mechanical Sciences 43(11):
2631–56. doi:10.1016/S0020-7403(01)00049-2.
MECHANICS BASED DESIGN OF STRUCTURES AND MACHINES 23

Dolce, M., and Cardone, D. 2001b. Mechanical behaviour of shape memory alloys for seismic applications 2.
Austenite NiTi wires subjected to tension. International Journal of Mechanical Sciences 43(11):2657–77. doi:10.
1016/S0020-7403(01)00050-9.
Fang, C., Ping, Y., & Chen, Y. 2020. Loading protocols for experimental seismic qualification of members in con-
ventional and emerging steel frames. Earthquake Engineering & Structural Dynamics 49(2):155–74. doi:10.1002/
eqe.3231.
Fang, C., Wang, W., & Feng, W. 2019. Experimental and numerical studies on self-centring beam-to-column con-
nections free from frame expansion. Engineering Structures 198:109526. doi:10.1016/j.engstruct.2019.109526.
Fang, C., Wang, W., Zhang, A., Sause, R., Ricles, J., and Chen, Y. 2019a. Behavior and design of self-centering
energy dissipative devices equipped with superelastic SMA ring springs. Journal of Structural Engineering
145(10):04019109. doi:10.1061/(ASCE)ST.1943-541X.0002414.
Fang, C., Zheng, Y., Chen, J., Yam, M. C., and Wang, W. 2019b. Superelastic NiTi SMA cables: Thermal-mechan-
ical behavior, hysteretic modelling and seismic application. Engineering Structures 183:533–49. doi:10.1016/j.eng-
struct.2019.01.049.
FEMA 440. 2005. Improvement of nonlinear static seismic analysis procedures. FEMA Region II. Redwood City, CA:
Applied Technology Council.
FEMA P-58. 2012. Seismic performance assessment of buildings volume 1 – Methodology. Tech. Rep. FEMA-P58.
Washington, DC: Federal Emergency Management Agency.
Gao, N., Jeon, J.-S., Hodgson, D. E., and DesRoches, R. 2016. An innovative seismic bracing system based on a
superelastic shape memory alloy ring. Smart materials and structures 25(5):055030. doi:10.1088/0964-1726/25/5/
055030.
Han, Y. L., Li, Q., Li, A. Q., Leung, A., and Lin, P. H. 2003. Structural vibration control by shape memory alloy
damper, Earthquake Engineering & Structural Dynamics 32(3):483–94. doi:10.1002/eqe.243.
Haque, A. B. M. R., Issa, A., and Shahria Alam, M. 2019. Superelastic shape memory alloy flag-shaped hysteresis
model with sliding response from residual deformation: Experimental and numerical study. Journal of Intelligent
Material Systems and Structures 30(12):1823–49. doi:10.1177/1045389X19844328.
Issa, A., and Alam, M. S. 2020. Comparative seismic fragility assessment of buckling restrained and self-centering
(friction spring and SMA) braced frames. Smart Materials and Structures 29(5):055029. doi:10.1088/1361-665X/
ab7858.
Jennings, P. C. 1968. Equivalent viscous damping for yielding structures. Journal of the Engineering Mechanics
Division 94(1):103–16. doi:10.1061/JMCEA3.0000929.
Kamarian, S., Bodaghi, M., Isfahani, R. B., and Song, J. I. 2020. A comparison between the effects of shape memory
alloys and carbon nanotubes on the thermal buckling of laminated composite beams. Mechanics Based Design of
Structures and Machines 1–24. doi:10.1080/15397734.2020.1776131.
Li, H.-N., Liu, M.-M., and Fu, X. 2018. An innovative re-centering SMA-lead damper and its application to steel
frame structures, Smart Materials and Structures 27(7):075029. doi:10.1088/1361-665X/aac28f.
Lim, T. J., and McDowell, D. L. 1995. Path dependence of shape memory alloys during cyclic loading. Journal of
Intelligent Material Systems and Structures 6(6):817–30. doi:10.1177/1045389X9500600610.
Ma, H., and Cho, C. 2008. Feasibility study on a superelastic SMA damper with re-centring capability. Materials
Science and Engineering: A 473(1–2):290–6. doi:10.1016/j.msea.2007.04.073.
Masuda, A., and Noori, M. 2002. Optimization of hysteretic characteristics of damping devices based on pseudoe-
lastic shape memory alloys. International Journal of Non-Linear Mechanics 37(8):1375–86. doi:10.1016/S0020-
7462(02)00024-0.
McKenna, F. 2011. OpenSees: A framework for earthquake engineering simulation. Computing in Science &
Engineering 13(4):58–66. doi:10.1109/MCSE.2011.66.
Moradi, S., Alam, M. S., and Asgarian, B. 2014. Incremental dynamic analysis of steel frames equipped with NiTi
shape memory alloy braces. The Structural Design of Tall and Special Buildings 23 (18):1406–25. doi:10.1002/tal.
1149.
Morais, J., de Morais, P. G., Santos, C., Costa, A. C., and Candeias, P. 2017. Shape memory alloy based dampers
for earthquake response mitigation. Procedia Structural Integrity 5:705–12. doi:10.1016/j.prostr.2017.07.048.
Ozbulut O. E., and Hurlebaus S. 2011. Re-centering variable friction device for vibration control of structures sub-
jected to near-field earthquakes. Mechanical Systems and Signal Processing 25:2849–62. doi:10.1016/j.ymssp.2011.
04.017.
Ozbulut, O., Roschke, P., Lin, P., and Loh, C. 2010. GA-based optimum design of a shape memory alloy device for
seismic response mitigation, Smart Materials and Structures 19(6):065004. doi:10.1088/0964-1726/19/6/065004.
Ozbulut, O. E., Silwal, B. and Michael, R. 2015. Design and component testing of an SMA-based seismic control
device. In Structures Congress 2015, 1237–52. Reston, VA: American Society of Civil Engineers. doi:10.1061/
9780784479117.106.
Pan, Q., and Cho, C. 2007. The investigation of a shape memory alloy micro-damper for MEMS applications.
Sensors (Basel) 7(9):1887–900. doi:10.3390/s7091887.
24 A. FALAHIAN ET AL.

Parulekar, Y., Reddy, G., Vaze, K., Guha, S., Gupta, C., Muthumani, K., and Sreekala, R. 2012. Seismic response
attenuation of structures using shape memory alloy dampers. Structural control and health monitoring 19(1):
102–19. doi:10.1002/stc.428.
Pieczyska, E., Gadaj, S., Nowacki, W. K., Hoshio, K., Makino, Y., and Tobushi, H. 2005. Characteristics of energy
storage and dissipation in TiNi shape memory alloy. Science and Technology of Advanced Materials 6(8):889–94.
doi:10.1016/j.stam.2005.07.008.
Qiu, C., Qi, J., and Chen, C. 2020. Energy-based seismic design methodology of SMABFs using hysteretic energy
spectrum. Journal of Structural Engineering 146(2):04019207.
Qiu, C. X., and Zhu, S. 2016. High-mode effects on seismic performance of multi-story self-centering braced steel
frames. Journal of Constructional Steel Research 119:133–43. doi:10.1016/j.jcsr.2015.12.008.
Qiu, C. X., and Zhu, S. 2017b. Performance-based seismic design of self-centering steel frames with SMA-based
braces. Engineering Structures 130:67–82. doi:10.1016/j.engstruct.2016.09.051.
Qiu, C., and Zhu, S. 2017a. Shake table test and numerical study of self-centering steel frame with SMA braces.
Earthquake Engineering & Structural Dynamics 46(1):117–37. doi:10.1002/eqe.2777.
Raniecki, B., Lexcellent, C., and Tanaka, K. 1992. Thermodynamic models of pseudoelastic behaviour of shape
memory alloys. Archive of Mechanics, Archiwum Mechaniki Stosowanej, 44:261–84.
Salichs, J., Hou, Z., and Noori, M. 2001. Vibration suppression of structures using passive shape memory alloy
energy dissipation devices, Journal of Intelligent Material Systems and Structures 12(10):671–80. doi:10.1177/
104538901320560319.
Shayanfard, P., Kadkhodaei, M., and Jalalpour, A. 2019. Numerical and experimental investigation on electro-
thermo-mechanical behavior of NiTi shape memory alloy wires. Iranian Journal of Science and Technology,
Transactions of Mechanical Engineering 43(1):621–29. doi:10.1007/s40997-018-0183-8.
Smith, R. C. 2005. Smart material systems: Model development, SIAM. doi:10.1137/1.9780898717471.
Song, G., Ma, N., and Li, H.-N. 2006. Applications of shape memory alloys in civil structures. Engineering
Structures 28(9):1266–74. doi:10.1016/j.engstruct.2005.12.010.
Soong T., and Spencer Jr., B. 2002. Supplemental energy dissipation: State-of-the-art and state-of-the-practice.
Engineering Structures 24(3):243–59. doi:10.1016/S0141-0296(01)00092-X.
Soul, H., Isalgue, A., Yawny, A., Torra, V., and Lovey, F. 2010. Pseudoelastic fatigue of NiTi wires: Frequency and
size effects on damping capacity, Smart Materials and Structures 19(8):085006. doi:10.1088/0964-1726/19/8/
085006.
Tang, W., and Lui, E. M. 2014. Hybrid recentering energy dissipative device for seismic protection. Journal of
Structures. doi:10.1155/2014/262409.
Torra, V., Carreras, G., Casciati, S., and Terriault, P. 2014. On the NiTi wires in dampers for stayed cables. Smart
Structures and Systems 13(3):353–74. doi:10.12989/sss.2014.13.3.353.
Wang, W., Fang, C., Feng, W., Ricles, J., Sause, R., and Chen, Y. 2020. SMA-based low-damage solution for self-
centering steel and composite beam-to-column connections. Journal of Structural Engineering 146(6):04020092.
doi:10.1061/(ASCE)ST.1943-541X.0002649.
Wang, W., Fang, C., Zhang, A., and Liu, X. 2019. Manufacturing and performance of a novel self-centring damper
with shape memory alloy ring springs for seismic resilience. Structural Control and Health Monitoring 26(5):
e2337. doi:10.1002/stc.2337.
Zareie, S., Alam, M. S., Seethaler, R. J., and Zabihollah, A. 2020a. Stability control of a novel frame integrated with
an SMA-MRF control system for marine structural applications based on the frequency analysis. Applied Ocean
Research 97:102091. doi:10.1016/j.apor.2020.102091.
Zareie, S., Issa, A. S., Seethaler, R. J., and Zabihollah, A. 2020b. Recent advances in the applications of shape mem-
ory alloys in civil infrastructures: A review. Structures 27:1535–50. doi:10.1016/j.istruc.2020.05.058.
Zareie, S., and Zabihollah, A. 2020. A study of pre-straining shape memory alloy (SMA)-based control elements
subject to large-amplitude cyclic loads. Ships and Offshore Structures 16:306–13. doi:10.1080/17445302.2020.
1726647.
Zhou, P., Liu, M., Li, H., and Song, G. 2018. Experimental investigations on seismic control of cable-stayed bridges
using shape memory alloy self-centering dampers. Structural control and health monitoring 25(7):e2180. doi:10.
1002/stc.2180.
Zhu, S., and Zhang, Y. 2008. Performance based seismic design of steel braced frame system with self-centering
friction damping brace. In Proceedings of the Structures Congress 2008: 18th Analysis and Computation Specialty
Conference, 1–13. Vancouver, Canada. doi:10.1061/41000(315)32.

You might also like