Fatigue of Nitinol

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

Available online at www.sciencedirect.com

www.elsevier.com/locate/jmbbm

Review Article

Fatigue of Nitinol: The state-of-the-art


and ongoing challenges

M.J. Mahtabia,b, Nima Shamsaeia,b,n, M.R. Mitchellc


a
Department of Mechanical Engineering, Mississippi State University, Box 9552, Mississippi State, MS 39762, USA
b
Center for Advanced Vehicular Systems (CAVS), Mississippi State University, Box 5405, Mississippi State, MS 39762,
USA
c
Mechanics & Materials Consulting, LLC, 4447 Acrete Lane, Flagstaff, AZ 86004, USA

ar t ic l e in f o abs tra ct

Article history: Nitinol, a nearly equiatomic alloy of nickel and titanium, has been considered for a wide
Received 31 December 2014 range of applications including medical and dental devices and implants as well as
Received in revised form aerospace and automotive components and structures. The realistic loading condition in
4 June 2015 many of these applications is cyclic; therefore, fatigue is often the main failure mode for
Accepted 7 June 2015 such components and structures. The fatigue behavior of Nitinol involves many more
Available online 16 June 2015 complexities compared with traditional metal alloys arising from its uniqueness in

Keywords: material properties such as superelasticity and shape memory effects. In this paper, a

Fatigue review of the present state-of-the-art on the fatigue behavior of superelastic Nitinol is

Shape memory alloys presented. Various aspects of fatigue of Nitinol are discussed and microstructural effects

Nitinol are explained. Effects of material preparation and testing conditions are also reviewed.

Superelastic Finally, several conclusions are made and recommendations for future works are offered.

Cyclic behavior & 2015 Elsevier Ltd. All rights reserved.

Literature review

Contents

1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
2. Fatigue concepts and definitions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
2.1. Fatigue specimens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
2.2. Fatigue testing techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
3. Fatigue of Nitinol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
3.1. Structural fatigue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
3.1.1. Uniaxial fatigue behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
3.1.2. Torsional fatigue. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237

n
Corresponding author at: Department of Mechanical Engineering, Mississippi State University, Box 9552, Mississippi State, Starkville,
MS 39762, USA. Tel.: þ1 662 325 2364.
E-mail address: shamsaei@me.msstate.edu (N. Shamsaei).

http://dx.doi.org/10.1016/j.jmbbm.2015.06.010
1751-6161/& 2015 Elsevier Ltd. All rights reserved.
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 229

3.1.3. Multiaxial fatigue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238


3.1.4. Mean stress/strain effects on fatigue behavior. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
3.1.5. Phase contribution to fatigue behavior. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
3.2. Thermo-mechanical fatigue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
3.3. Functional fatigue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
4. Failure mechanism and microstructural effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
5. Treatment and processing effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
6. Other parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
7. Future works and challenges. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
8. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251

1. Introduction components (Bigeon and Morin, 1996). Expansion and con-


traction in stented blood vessels and the combined torsional
Nitinol, a nearly equiatomic alloy of nickel and titanium, and bending deformations in endodontic files are other
exhibits unique material properties such as superelasticity examples of cyclic deformation in components made of
(also known as pseudo-elasticity) and shape memory effects, Nitinol alloys. Several researchers have studied the fatigue
resistance to corrosion, and an unusual combination of behavior of Nitinol and investigated different aspects of
strength and ductility. The unique properties of this material property deterioration in this material (Melton and Mercier,
together with its biocompatibility and resistance to corrosion 1979a; Pelton et al., 2008, 2003; Plotino et al., 2009; Robertson
have made it a preferred material for biomedical applications et al., 2012). However, due to the complexity in mechanical
(Duerig et al., 1999; El Feninat et al., 2002). Many new response of Nitinol, its fatigue behavior is not yet completely
applications of the Nitinol alloys have been extended to understood.
several industries due to the discovery of their unique Compared to other metal alloys, Nitinol can withstand
properties over the years. In biomedical engineering, for greater strains prior to failure. For instance, superelastic
example, the popularity of Nitinol is increasing because of Nitinol may endure cyclic strain amplitudes ranging from
its unique stress/strain response, similar to biomaterials such 4% to 12% for the number of cycles at which most other alloys
as bone (Duerig et al., 1999). Therefore, Nitinol has been used would fail under cyclic strain amplitudes of 1% or less (Wilkes
extensively in various biomedical applications such as endo- and Liaw, 2000). The complexity in fatigue response of Nitinol
vascular stents, vena cava filters, and endodontic files (Duerig arises from the pronounced effects of shape memory proper-
et al., 1999; Mohd Jani et al., 2014; Pelton et al., 2000; Plotino ties and the phase transformation phenomenon on its
et al., 2009). mechanical response. Any changes in the shape memory
In addition to bioengineering, Nitinol has been used in properties, resulted from material compositions (i.e. amount
aerospace, automotive, civil and structural engineering. of nickel or titanium in a Nitinol alloy), treatment process, or
Actuators (Hartl and Lagoudas, 2007), structural connectors, test temperature, may affect significantly the fatigue beha-
seals, vibration dampers, release or deployment mechanisms vior of these alloys. One of the main challenges in evaluating
(Carpenter and Lyons, 2001; Huett and Willey, 2000), inflata- the fatigue resistance of Nitinol is the formation of stress-
ble structures (Peng et al., 2005), manipulators (Prahlad and induced martensite under mechanical deformation, and
Chopra, 2001), and the Mars Pathfinder (Godard et al., 2003) therefore, classical fatigue theories may not be applicable
are only some applications of this material in the aerospace directly to superelastic Nitinol alloys (Maletta et al., 2012).
industry. Nitinol is also used as an energy dissipating tool for Furthermore, Nitinol alloys exhibit a noticeable asymme-
shock and seismic application (Mohd Jani et al., 2014; Song try in the stress–strain behavior under tension and compres-
et al., 2006). sion (Gall et al., 1998; Gall and Sehitoglu, 1999). This
Nitinol components and structures are often subjected to asymmetry in stress response for strain-controlled tests
deformations or stresses that result in some kinds of may result in a significant compressive mean stress in low-
mechanical failures. As mentioned previously, Nitinol alloys cycle regime, where stress-induced martensite exists. Simi-
are used widely in various industries such as automotive, larly for force-controlled tests, there may be a large amount
aerospace, civil and bioengineering. The realistic loading of tensile mean strain as a result of asymmetry in strain
conditions in many of these applications are cyclic. There- response. Presence of mean strains/stresses may significantly
fore, fatigue is often the main failure mode for such compo- influence the fatigue behavior of Nitinol; compressive mean
nents and structures. Fatigue is generally defined as the stresses generally increase the fatigue life and tensile mean
gradual and progressive deterioration of materials’ strength strains are known to reduce the fatigue life. Considering
under cyclic loading. Fatigue failure is unexpected and may these challenges, not many fatigue life predictive models
cause serious damage to mechanical and biomedical systems have been proposed for these alloys in the open literature.
without any warning. It has been reported that 50 to 90% of Miniature components made from Nitinol, require mon-
all mechanical failures are caused by fatigue (Stephens et al., itoring the fatigue initiation rather than the fatigue crack
2000). growth since the critical crack size, in a fracture mechanics
Some early studies have considered the fatigue of Nitinol sense, is quite small. The interest in the current state-of-
actuators and discovered low cycle fatigue risks for these practice for Nitinol alloys is mainly to control crack initiation
230 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

Nomenclature constant amplitude strain-controlled cyclic load-


ing: (εmax  εmin)/2
Af austenite finish temperature—the temperature εeq equivalent (von Mises) strain—that is calculated
above which Nitinol is fully austenitic based on the von Mises
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi equivalent strain equa-
As austenite start temperature—the temperature at tion: εeq ¼ ε2t þ 43 ε2s
which, upon heating, martensitic Nitinol begins εs shear strain—the angular change in a square
transforming to austenite element under torsional stress
EA austenite modulus—modulus of elasticity for εt axial strain—the change in length divided by the
superelastic Nitinol corresponding to the initial initial length
linear section in the stress–strain curve ε0f fatigue ductility coefficient in the Coffin–Manson
EM stress-induced martensite modulus—modulus of equation is the intercept at one reversal of the
elasticity for superelastic Nitinol corresponding plastic strain-life line
to the linear section beyond stress plateau in εm mean strain—average of the maximum and mini-
stress–strain curve mum strains in a constant amplitude strain-
b fatigue strength exponent in Basquin or Coffin– controlled cyclic loading: εm ¼(εmaxþεmin)/2
Manson equations, the slope of the stress/elastic εmax maximum strain—maximum strain in a constant
strain-life line in log–log coordinates amplitude strain-controlled cyclic loading
c fatigue ductility exponent in Coffin–Manson equa- εmin minimum strain—minimum strain in a constant
tion, the slope of the plastic strain-life line in log– amplitude strain-controlled cyclic loading
log coordinates εp plastic strain—the permanent, non-recoverable
d wire diameter component of the total strain
Mf martensite finish temperature—the temperature at εpr proportional strain—the strain value in a super-
which, upon cooling, Nitinol is fully martensitic elastic Nitinol, for which, the stress–strain rela-
Ms martensite start temperature—the temperature at tionship is essentially linear. This value
which, upon cooling, austenitic Nitinol begins corresponds to σpr
transforming to martensite εAM
F martensite finish strain—the strain related to the
Nf number of cycles to failure—number of cycles that a ending point of the stress plateau in a super-
fatigue sample experiences before failure occurs elastic Nitinol. Subscript F stands for finish and
under an applied cyclic stress or strain. NOTE: 2Nf superscript AM indicates the Austenite-Marten-
is defined as the number of reversals to failure. site transformation
NiXTiY Nitinol of nickel (X atomic percent) and titanium εAM
S martensite start strain—the strain related to the
(Y atomic percent) starting point of the stress plateau in a super-
Rσ stress ratio—ratio of the minimum to maximum elastic Nitinol. Subscript S stands for start and
stresses (σmin/σmax) in a constant amplitude force- superscript AM indicates the Austenite- Marten-
controlled cyclic loading site transformation. In some articles this has
Rε strain ratio—ratio of the minimum to maximum been referred to as the “elastic limit” for super-
strains (εmin/εmax) in a constant amplitude strain- elastic Nitinol.
controlled cyclic loading εtr transformation strain—the range of the strain asso-
T test temperature ciated with the transformation between the aus-
α cyclic strain coefficient in the modified Coffin–Man- tenite and martensite phases in a superelastic
son equation Nitinol, calculated as the difference between the
β cyclic strain exponent in the modified Coffin–Man- total strain and the strain related to the starting
son equation point of stress-induced martensite, εAM
S :
ΔT relative temperature—the difference between the εtr ¼ ε εS
AM

test temperature, T, and austenite finish tempera- ν Poisson’s ratio—the negative of the ratio of trans-
ture, Af: ΔT ¼T  Af verse to axial strain
Δεs/2 shear strain amplitude—the half of the shear strain ρ radius of curvature in a rotating-bending test setup
range in a torsional fatigue test for fatigue testing of wires
Δε strain range—difference between the maximum σa stress amplitude—half of the difference between
and minimum strains in a constant amplitude the maximum stress and the minimum stress in
cyclic loading: Δε ¼εmax εmin a constant amplitude force-controlled cyclic load-
Δσ stress range—difference between the maximum ing; σa ¼(σmax σmin)/2
and minimum stresses in a constant amplitude σeq equivalent (von Mises) stress—that is calculated
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
force-controlled cyclic loading: Δσ¼ σmax σmin based on the following equation: σ eq ¼ σ 2 þ 3 τ2
εa strain amplitude—half of the difference between σ 0f fatigue strength coefficient in Basquin or Coffin–
the maximum strain and minimum strain in a Manson equations as the intercept at one reversal
of the stress/elastic strain-life line
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 231

σm mean stress—average of the maximum and mini- σ AM


F martensite finish stress—stress related to the end-
mum stresses in a constant amplitude force- ing point of the stress plateau in a superelastic
controlled cyclic loading: σm ¼ (σmaxþσmin)/2 Nitinol. Subscript F stands for finish and super-
σmax maximum stress—maximum stress in a constant script AM indicates the Austenite-Martensite
amplitude force-controlled cyclic loading transformation
σmin minimum stress—minimum stress in a constant σ AM
S martensite start stress—stress related to the start-
amplitude force-controlled cyclic loading ing point of the stress plateau in a superelastic
σpr proportional stress—the stress value in a super- Nitinol. Subscript S stands for start and super-
elastic Nitinol for which the stress–strain rela- script AM indicates the Austenite-Martensite
tionship is essentially linear. This value transformation
corresponds to εpr

because the failure in these materials takes place shortly equation is often presented as in Eq. (1):
after the crack nucleation (Robertson et al., 2012). On the  b
σ a ¼ σ 0f 2Nf ð1Þ
other hand, difficulties concerned with experimental work
and measurements have been an obstacle in such studies. In this equation, σa is the applied alternating stress, 2Nf is
In this paper, some basic definitions and concepts of the number of reversals before the material’s failure under σa.
0
fatigue behavior and analysis for metallic materials are first The parameters σ f and b are the fatigue strength coefficient
introduced. Various aspects of fatigue of Nitinol alloys are and fatigue strength exponent, respectively.
presented subsequently and results from previous studies are In terms of strain, the total strain amplitude, Δε/2, can be
discussed. Effects of material preparation and testing techni- separated into elastic strain amplitude, Δεe/2, and plastic
ques are reviewed and recommendations to improve the strain amplitude, Δεp/2, as presented in Eq. (2):
fatigue resistance of Nitinol are investigated. An attempt is Δε Δεe Δεp
made here to describe the complexity in the fatigue behavior ¼ εa ¼ þ ð2Þ
2 2 2
of Nitinol based on its mechanical and microstructural
For elastic strain, this relationship can be formulated
features. The contradictions obtained in the literature are
using Basquin’s relation as follows:
analyzed critically and some challenging topics for future
Δεe σa σ 0f  b
research are recommended. Finally, the overall conclusions ¼ ¼ 2Nf ð3Þ
on the fatigue behavior of Nitinol alloys are presented. The 2 E E
main focus here is on the macroscopic aspects of the fatigue where E is the elastic modulus of the material. The plastic
of Nitinol and minimal discussions on the microstructural strain amplitude, Δεp/2, can be related to fatigue life using the
aspects are presented. It should be noted that most of the Coffin–Manson relationship (Coffin and Tavernelli, 1959;
discussions are related to the superelastic Nitinol that is used Manson and Dolan, 1966) as follows:
in the vast majority of present applications. However, some Δεp  c
¼ ε0f 2Nf ð4Þ
discussions on Nitinol in the martensitic phase are also 2
made. Although we have tried to collect information from Finally, the total strain-life can be presented using the
many published documents in the open literature, this article following equation:
is by no means comprehensive. 0
σf  b 0  c
εa ¼ 2Nf þ εf 2Nf ð5Þ
E
In Eqs. (3)–(5), ε0f , σ 0f , b and c are material constants that are
0 0
2. Fatigue concepts and definitions determined experimentally; εf and σ f are the material fatigue
ductility coefficient and fatigue strength coefficient, respec-
In this section, some terminologies and basic concepts of tively. Constant b is the strength exponent and constant c is
fatigue of metallic materials are reviewed. Two major called the fatigue ductility exponent.
approaches to represent the fatigue data for metallic materi- Another important factor that can influence the fatigue
als are the stress-life and strain-life methods. In these life of a material is the presence of a constant force, in
methods, the fatigue data is presented as the stress or strain addition to the cyclic force. Since the constant force can be
amplitude against the fatigue life in terms of the number of treated as the mean of the alternating force, several attempts
cycles, Nf, or reversals, 2Nf, to failure. Different analytical (Dowling, 2009; Ince and Glinka, 2011) have been made to
models have been used to represent mathematically the consider the effects of mean stress/strain on the fatigue
fatigue strength of the materials. These methods are briefly behavior of materials. The tensile mean stress usually
described in this section. reduces the fatigue resistance, i.e. life, while the compressive
Basquin (1910)’s relation has been used widely to relate mean stress may increase the fatigue life. Two of the more
the stress amplitude to the life of the material at failure. This common methods for evaluating the mean stress effects on
relationship is simply a power-law function that expresses fatigue life are Goodman and Morrow models (Stephens et al.,
the stress amplitude as a function of number of cycles (or 2000). Both models illustrate a reducing tolerable alternating
reversals) to failure. The general form of the Basquin’s stress with increasing mean stress amplitude.
232 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

a b c d

Solid circular Thin-walled tubular Diamond-shape Flat dogbone


dogbone specimen specimen components specimen

Fig. 1 – Typical test specimens used for fatigue characterization of Nitinol alloys.

2.1. Fatigue specimens


Rotating
motor
ρ Guide
Different types of specimens are required to facilitate various Nitinol wire
Nitinol wire
testing techniques for fatigue evaluation of Nitinol. Wires are
the simplest and the most commonly used shape for Nitinol
Bushing
fatigue characterizations. Fatigue testing of wires is per- Rotating Mandrel
motor ρ
formed mainly with a rotating-bending testing technique
according to ASTM E2948-2014-Rotating Bending Fatigue
Fig. 2 – Configurations for the rotating-bending fatigue test of
Tests of Solid Round Fine Wire (ASTM E2948-14, 2014).
Nitinol wires.
However, wires may not be the most appropriate type of
specimens for fatigue characterization under all types of
deformation such as axial compression or torsion. Dogbone- wire is bent to 1801 and approximately 901 as shown in Fig. 2
shaped specimens, shown in Fig. 1(a) and (b), are more (a) and (b), respectively.
suitable for fatigue testing specifically under compression or Rotating-bending test method is an appropriate technique
torsional deformations (e.g. Jensen, 2005; Subramaniam, for collecting strain-based fatigue data for wire specimens
1998). Dogbone specimens are used widely in the form of under zero mean strain conditions (i.e. fully reversed loading,
solid or hollow. Hollow specimens are more suitable for Rε ¼ 1). Testing apparatus is generally comprised of a
torsional fatigue analysis because the shear stress is consid- variable-speed motor and a sensor for counting the number
ered to be approximately uniform on the cross section of of rotations (i.e. cycles or reversals). In some cases, a silicon
thin-walled tubular specimens. In some studies, stent-like oil or de-ionized water bath and an electric heater with a
diamond-shape components, shown in Fig. 1(c), have been temperature control unit may be used to facilitate a constant-
employed in displacement-controlled fatigue testing (Pelton temperature test environment. In this method, wire is bent to
et al., 2008; Tolomeo et al., 2000). Finite element analysis is a specific curvature with known radius of curvature (ρ) in
required to calculate the deformation-induced stress and order to apply an alternating strain amplitude. The strain
strain terms at the critical elements of these components. amplitude is then approximated by the ratio of wire radius
In some studies, flat dogbone Nitinol specimens (Fig. 1(d)) (d/2) to the radius of the curvature as follows:
have also been used under uniaxial cyclic tests (Nayan et al.,
d
2008). εa ¼ ð6Þ

2.2. Fatigue testing techniques Various strain amplitudes can be attained by altering the
wire diameter and the radius of the curvature. Alternating
A simple and commonly used fatigue test method for Nitinol strains are then applied by rotating the curved wire about its
wires is the rotating-bending test, which is suitable for longitudinal axis as presented in Fig. 2. For the rotating-
testing Nitinol wires used in several medical devices sub- bending test setup shown in Fig. 2(a), the wire is bent to an
jected to a bending (or axial) loading. Several experimental ellipse-like shape to achieve its minimum radius of curvature
studies have used rotating-bending tests to evaluate the and maximum strain at the mid-length. The length of the
fatigue behavior of Nitinol alloys (e.g. Bertacchini et al., wire specimen, excluding the length embedded in chucks, is
2009; Miyazaki et al., 1999; Pelton et al., 2013). In certain equal to 5.25ρ (ASTM E2948-14, 2014). This testing can also be
cases, the results from rotating-bending tests may be suitable conducted in guided condition to reduce the vibration and
for fatigue evaluation purposes (Duerig et al., 1999; ASTM maintain the wire specimen in plane. The support guides are
E2948-14, 2014). A rotating-bending testing machine is com- located out of the area of maximum curvature, where the
prised mainly of a motor to rotate the wire and a constraint to fatigue failure typically occurs.
maintain the strain amplitude. Rotating-bending fatigue Pelton et al. (2013) used the same concept by employing
testing has become a common test setup for the structural mandrels of different radii to bend the wires into smooth
fatigue evaluation of Nitinol wires (Tobushi et al., 1998). Two curves as shown in Fig. 2(b). This method is a guided test
common configurations of the rotating-bending test setup are setup where mandrels act as support guides. The mandrels
shown in Fig. 2. Both types of test setups are used to conduct should be made of low friction material, such as polyoxy-
fully reversed, rotating-bending fatigue. In these setups, the methylene or polytetrafluoroethylene (ASTM E2948-14, 2014)
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 233

and machined carefully to reduce the effects of friction under thermal cyclic loads (Pelton, 2011). From the micro-
between the wire and the mandrel that may result in scuffing scopic point of view, the failure process in Nitinol is generally
the Nitinol wires. A comparison of the various rotating- comprised of two main stages including crack initiation and
bending test setups for fatigue testing of Nitinol is reported crack propagation. The crack initiation stage is typically very
by Norwich (2014) and more details on fatigue testing of thin slow, compared with the crack propagation stage. The rapid
wires can be found in the recently published standard ASTM rupture of the cracked surface is reported in (Condorelli et al.,
E2948-14 (2014). In addition to the rotating-bending test, 2010), which suggests the critical crack size of Nitinol alloys
uniaxial and multiaxial testing using closed-loop fatigue to be comparatively small.
testing machines are also common for Nitinol. Standard test
methods for the evaluation of the fatigue behavior of metallic 3.1. Structural fatigue
materials can be found in ASTM Committee E08 Fatigue and
Fracture series of the test standards. ASTM E466 Force- Similar to any other metallic material, Nitinol alloys experi-
Controlled (ASTM E466-07, 2007) and ASTM E606 Strain- ence a reduction in strength under repeated forces and
Controlled (ASTM E606/E606M-12, 2012) provide the regula- consequently fail at a stress level smaller than their nominal
tions for such fatigue testing. tensile strength. As previously mentioned, a majority of
A major difference between the rotating-bending fatigue fatigue analysis of Nitinol alloys has been focused on the
machines and the more common closed-loop uniaxial fatigue crack initiation approaches. However, there are a number of
machines is the distribution of stress/strain over the speci- studies using the fatigue crack-growth approach (McKelvey
men’s cross section (i.e. volume). In the rotating-bending test, and Ritchie, 2001; Robertson and Ritchie, 2007; Vaidyanathan
the outer layer of the specimen (i.e. the surface of the wire) et al., 2000). Therefore, the primary focus of this paper is on
experiences the maximum amount of stress/strain during the the fatigue initiation stage and the fatigue crack-growth
test, while the stress/strain values are proportionately less in studies are discussed minimally.
inner layers. Nevertheless, the entire specimen’s cross sec-
tion (i.e. volume) experiences the same amount of stress/ 3.1.1. Uniaxial fatigue behavior
strain in the uniaxial fatigue machines. Due to the difference Different deformation approaches can be employed to study
in stress/strain distribution across the section, the effects of the fatigue behavior of engineering materials. Each approach
surface finish quality on fatigue behavior may be more illustrates various aspects of the fatigue response of the
pronounced in the rotary-bending tests. However, impurities material under study. Stress-life and strain-life methods are
and inclusions in the material may govern the fatigue the two more common approaches in fatigue life analysis of
behavior of the specimen in the uniaxial fatigue test to a materials. Selection of the most appropriate method is
greater degree as compared to the one in the rotating- usually based on several parameters such as type of the
bending test, particularly in the gigacycle range. deformation or loading of the component in real-world
application, mechanical behavior of the material, and the
feasibility of the testing method. In this section, results from
different studies on the fatigue behavior of the Nitinol alloys
3. Fatigue of Nitinol are reviewed.
The mechanical behavior of Nitinol depends significantly
The fatigue of Nitinol alloys is classified into three different
on its “characteristic temperatures”. Typically, there are four
categories in the literature (Van Humbeeck, 1991); (1) struc-
characteristic temperature values for Nitinol: Mf, Ms, As, Af.
tural fatigue that is a reduction in strength under cyclic loads,
The martensite finish temperature, Mf, is the temperature
(2) functional fatigue, which is a loss in functional properties
below which the material is fully martensitic. The martensite
such as superelasticity and shape memory effects (Eggeler
start temperature, Ms, is the temperature that, upon cooling,
et al., 2004), and (3) thermal fatigue that is a change in
austenitic Nitinol alloy starts transforming to martensitic
material properties such as transformation temperatures
phase. The austenite start temperature, As, is the tempera-
ture at which, upon heating, martensitic Nitinol tends to
transform to the austenitic phase. Finally, the austenite finish
temperature, Af, is the temperature above which Nitinol
maintains the austenitic phase under unstrained condition.
The relations of these temperatures with respect to the test
temperature dictate the overall mechanical response of the
material. As illustrated in Fig. 3, the stress–strain relation-
ships of Nitinol alloys are dependent on the working tem-
perature. The material may also show superelastic or shape
memory behavior depending on the test temperature with
respect to the transformation temperatures. Therefore, it is
important to determine the characteristic temperatures for
any Nitinol alloy prior to further analyses.
Fig. 3 – Schematic stress–strain curves of Nitinol in different
phases. T represents the test temperature (Pelton et al., 3.1.1.1. Stress-life fatigue behavior. The force-controlled tests
2000). and stress-life method are very common for fatigue analysis
234 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

of materials. Nevertheless, the majority of fatigue studies on


Nitinol have followed the strain-controlled test setups and
strain-life approach. It should be noted that Nitinol alloys,
specifically superelastic alloys, are also often used for force
carrying applications such as reinforcing bars in concrete
bridges (Saiidi and Wang, 2006) and endodontic files. There-
fore, it seems necessary to study the fatigue behavior of
shape memory alloys employing the stress-life approach as
well. Employing different techniques for cyclic fatigue testing
of Nitinol, force- (stress-) controlled or displacement- (strain-)
controlled, can also affect the fatigue resistance of Nitinol by
Fig. 4 – Stress-life fatigue curves for different phases of
influencing the kinematic irreversibility, resulted from dis-
Nitinol alloys (Melton and Mercier, 1979a)—Austenite finish
location movements, leading to the accumulation of the
temperature is approximated here as: Af ¼Msþ40 1C
permanent deformation (Nayan et al., 2008). On the other
(Robertson et al., 2012).
hand, representing the fatigue data in different forms may
help to better understand the fatigue behavior of the material
in different applications. In other words, using stress-life data
for the applications in which the loading is force-based can
be more appropriate than the strain-life representation. For
instance, a rotary dental instrument made of Nitinol is a
force-based application as it is designed to carry certain
forces and the stress-life approach more appropriately repre-
sents the physics of the situation.
Two of the earliest comprehensive studies on the fatigue of
Nitinol alloys were performed by Melton and Mercier (1979a,
1979b). Their works included both stress-life and strain-life
approaches at room temperature (  22 1C) using specimens
with different martensite start temperatures, Ms. Although
they reported the martensite start temperature, Ms, of mate- Fig. 5 – Fully reversed stress-life data for superelastic Nitinol
rials, the austenite finish temperature, Af, is also used in alloys (Miyazaki et al., 1988). T is the test temperature and Af
latter literature as an indicator of the Nitinol shape memory is the material’s austenite finish temperature.
character. Robertson et al. (2012) suggested that Af is approxi-
mately 40 1C greater than Ms (i.e. Af ¼ Msþ40 1C). This approx-
imation is used in this document.
Force-controlled tests were performed in a fully reversed
condition (Rσ ¼  1) in Melton and Mercier (1979a)’s study.
They presented stress-life diagrams for two different compo-
sitions of Nitinol; Ni49.9Ti50.1 and Ni49.7Ti50.3 (throughout this
paper NixTiy indicates a Nitinol alloy with x amount of nickel
and y amount of titanium, both in atomic percentage, except
as stated otherwise). Their results revealed the strong influ-
ence of the Nitinol metallurgical phases on fatigue behavior,
as shown in Fig. 4. Superelastic Nitinol exhibited superior
stress-life fatigue resistance compared to the mixed-mode
Fig. 6 – Stress-life fatigue data for Nitinol wires at different
and martensitic Nitinol.
temperatures (Kim and Miyazaki, 1997). T is the test
Miyazaki et al. (1988) provided a comprehensive study on the
temperature.
fatigue of Nitinol. Their results include stress-life fatigue for
unique specimens (Ni50.8Ti49.2) with Af ¼ 27 1C at different test
temperatures as shown in Fig. 5. The stress-life curves from with As   1 1C and Af ¼ 23 1C at room temperature (i.e. 22 1C).
their study are linear at high test temperatures and bi-linear These tests were conducted under non-zero mean stress
at lower test temperatures. The lower test temperatures are condition with Rσ ¼ 0.1. Stress-life data from Nayan et al.
very close to the austenite finish temperature, Af. These data (2008)’s study also indicate two linear sections that have a
clearly indicate that for the stress-life fatigue approach and at common point at a stress level very close to the critical stress
test temperatures above Af, where Nitinol exhibits super- required for inducing martensite.
elastic behavior, the fatigue life increases as the test tem- Stress-life fatigue data from Kim and Miyazaki (1997)’s
perature increases. Moumni et al. (2009, 2005) also reported study) was later analyzed by Pelton et al. (2000). These data
that the stress-life fatigue results for superelastic Ni48.7Ti51.3 are presented in Fig. 6 for Ni50.9Ti49.1 with Af ¼ 60 1C and
can be presented well in a linear form in a semi-log plot. As ¼ 37 1C. As seen from these plots, the stress-life data for
Nayan et al. (2008) also investigated the uniaxial tension– superelastic Nitinol wires can be expressed linearly by a
tension fatigue behavior of flat dogbone Ni50.7Ti49.1 specimens Wöhler curve in the semi-log plot that are similar to the
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 235

results from Moumni et al. (2005) and Miyazaki et al. (1988). It


can be concluded from these studies that in the stress-life
approach the fatigue lives of Nitinol alloys increase with
increasing test temperature. Pelton et al. (2000) also reported
a temperature dependent fatigue limit for the stress-life data
in which the fatigue limit increases with increasing test
temperature toward Af.
Analysis of the fatigue behavior of Nitinol alloys using
stress-life results illustrates a drastic reduction in fatigue life
with increasing stress amplitude, which is similar to other
metal alloys. It also reveals the sensitivity of the results to
test temperature. However, the fatigue resistance of Nitinol
does not reduce with an increase in the test temperature,
which is in contrast to other metallic materials.
Although stress-life representation of the fatigue data is
very common in metal fatigue analysis, it is not used widely
for fatigue analysis of Nitinol alloys; hence, there are only
limited studies in the literature employing this approach for
Nitinol. Currently there is no reliable formulation in the
present literature that relates the stress amplitude to the
failure life for Nitinol alloys. However, it seems that the
stress-life fatigue data for Nitinol may be expressed by a
straight line on the semi-log plot, as illustrated in Fig. 6.

3.1.1.2. Strain-life fatigue behavior. Superelastic Nitinol alloys


may tolerate a large pseudo-elastic (recoverable) strain at
almost a constant stress level when the stress reaches a Fig. 7 – Effects of test temperature on the rotating-bending
certain level, so-called the plateau stress as seen in Fig. 3. fatigue life for a Nitinol: (a) (Miyazaki et al., 1999) (b) (Kim,
Consequently, the stress–strain “modulus” relationship is 2002). T is the test temperature and Af is the austenite finish
linear only for a minimal strain range, compared to the total temperature.
recoverable strain. A reasonable linear correlation between
the number of cycles to failure and strain amplitude was
reported for Nitinol alloys in the log–log plot (McNichols et al.,
1981; Melton and Mercier, 1979a). Moreover, in many biome- worth mentioning that a few studies have suggested a fatigue
dical applications of Nitinol (e.g. stents), the component is limit in the long life regime for strain-life fatigue curves of
under cyclic deformations, requiring the material to resist Nitinol alloys (Eggeler et al., 2004; Pelton et al., 2013).
fatigue failure. Therefore, strain-controlled testing proce- Kim (2002) studied the fatigue behavior of three binary
dures are more appropriate for the fatigue analysis of Nitinol combinations of nickel and titanium with Ni contents of
alloys. In addition, strain-based testing at Rε ¼ 1 does not 50%, 50.5%, and 50.85%. The corresponding austenite finish
require very specialized equipment since these tests can be temperatures of the materials were 87.7, 85, and 60.4 1C. The
conducted using the simple rotating-bending test setup for specimens were annealed, water quenched, and electro-
wires, as was described previously in Fig. 2. For these reasons, polished before subjected to fatigue tests at four different
majority of fatigue studies on shape memory alloys and temperatures; 20, 50, 80, and 110 1C. Kim (2002) also observed
specifically for Nitinol are based on the strain-life (ε–N) an increase in the fatigue life with decreasing test tempera-
approach. ture, as illustrated in Fig. 7(b) for 50.5% Ni content, which
Miyazaki et al. (1999) observed different trends for strain-life verified the observations made by Miyazaki et al. (1999).
fatigue data of martensitic and austenitic Ni50Ti50 at different Tobushi et al. (1997) performed rotating-bending tests on
test temperatures as shown in Fig. 7. The specimens were Ni55.3Ti44.7 wires at different temperatures; 30, 60, and 80 1C.
water quenched after annealing at 400 1C. A silicon oil bath The austenite finish temperature for this Nitinol alloy was
was used and the temperature was controlled by an electric Af ¼50 1C; therefore, the specimens were martensitic at 30 1C
heater and a fan in order to maintain a constant test and superelastic at 60 and 80 1C. Tests were conducted in air
temperature. In strain-life representation of the fatigue data, and also in the water bath to maintain a constant tempera-
there is a reverse relationship between the test temperature ture condition. Strain-life data from this study, shown in
and the fatigue life of Nitinol for test temperatures above Af Fig. 8, indicates that the martensitic Nitinol alloy has greater
and at εa40.7% (Miyazaki et al., 1999) as presented in Fig. 7(a). fatigue life as compared to the superelastic Nitinol. Tobushi
Therefore, as the test temperature decreases, the fatigue life et al.’s data is in agreement with the findings of other
of the superelastic Nitinol specimens increases. Furthermore, researchers such as Miyazaki et al. (1999) and Kim (2002),
a convergence of strain-life curves at 105 cycles can be seen where they found greater fatigue lives for mixed-phase
in Fig. 7(b), which implies the insensitivity of the strain-life Nitinol than those for superelastic Nitinol under strain-life
curve to temperature in the long life regime (Kim, 2002). It is fatigue as shown in Fig. 7. Tobushi et al. (1997) observed a
236 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

strain-controlled fatigue tests on flat dogbone-shaped


Ni50.8Ti49.2 specimens at room temperature (25 1C). Austenite
finish temperature of this alloy was Af ¼ 14 1C, thus, the alloy
was superelastic. Cyclic strain amplitudes were chosen such
that the stress-induced martensite was formed in the speci-
mens. The value of β calculated from their study was 0.914 for
the residual amplitude and 0.171 for the elastic strain ampli-
tude. The elastic strain amplitude in their study was con-
sidered as strain amplitudes less than the strain
corresponding to the proportional limit, εpr (i.e. the initial
linear “modulus” portion of stress–strain curve in Fig. 3).
However, they used only the Coffin–Manson relation for the
Fig. 8 – Strain-life fatigue data for Ni55.3Ti44.7 wt% at constant stress-induced martensitic Nitinol, i.e. the stress plateau in
test temperatures, T (Tobushi et al., 1997). The constant test Fig. 3; therefore, the equations may not be applicable to the
temperatures were maintained using water bath. fully austenitic or thermally martensitic Nitinol.
In another study, Kollerov et al. (2013) considered the Coffin–
Manson relationship for low-cycle fatigue of superelastic and
shape memory Nitinol alloys. Materials were binary Nitinol
with three different nickel contents; 55.7, 55.8, and 54.7 in
weight percent, and except for one material, all the speci-
mens were superelastic at the test temperature of 21 1C.
Specimens were tested using rotating-bending test technique
with strain amplitudes between 0.75% and 4.8%. Kollerov
et al. (2013) proposed a modified Coffin–Manson equation for
strain levels above 2% as a function of the maximum
recoverable strain, which can be completely recovered after
unloading for superelastic Nitinol, or by heating up to a
temperature above Af, for thermally martensitic Nitinol.
Kollerov et al. also found a good correlation between their
Fig. 9 – Schematic strain-life fatigue regimes of superelastic modified Coffin–Manson equation in low-cycle fatigue (i.e. 2%
Nitinol in log–log scale. T is the test temperature and Af is oεao5%) for both superelastic and shape memory Nitinol
the material austenite finish temperature. alloys. In the proposed modified Coffin–Manson equation, the
strain exponent was a function of stress plateau and the
modulus of elasticity, whereas the strain coefficient was a
linear relationship between strain and failure life in log–log function of maximum recoverable strain, stress plateau, and
plots for εa40.8%, as shown in Fig. 8. The strain-life data from the modulus of elasticity. They, however, did not find a
the tests in this high-strain region (corresponding to Nfo104 correlation between the high-cycle fatigue strength of Nitinol
cycles) were expressed by a power law function as: alloys and the maximum recoverable strain. Furthermore,
they suggested that the high-cycle fatigue resistance of
εa ¼ αNf β ð7Þ
Nitinol alloys could be improved by microstructural enhance-
where α is the cyclic strain coefficient and β is the strain ment such as decreasing impurities or by providing nano-
exponent. The β ¼0.235 was obtained from the regression size, Ni-rich particle precipitation and increasing dislocation
analysis for experiments conducted in air in Tobushi et al. density.
(1997)’s study). They mentioned that the β value was less Pelton et al. (2013) recently divided the strain-life fatigue
than the typical strain exponent for other metals. However, data of superelastic Nitinol (i.e. the test temperature above Af)
the β value obtained from the experiments that were con- to four distinct regimes in log–log plots; very low-cycle
ducted in the water bath, was calculated to be 0.47, which (Nr 103 cycles), low-cycle (N103 cycles), mid-cycle
was comparable to that for other metallic materials (Tobushi (103 r Nr105 cycles), and high-cycle (NZ105 cycles). This
et al., 1997). The comparison made by Tobushi et al. on the classification is comprehensively discussed by Pelton et al.
magnitude of the exponent β may not be appropriate since (2013).
there is a basic difference in the exponents of the original A schematic, explaining the general strain-life fatigue beha-
Coffin–Manson relation and the equation proposed by vior of the superelastic Nitinol in log–log scale is presented in
Tobushi et al.; the c (fatigue ductility exponent) in the Fig. 9. This figure is mainly generated in this review article
Coffin–Manson relationship is an indicator of the change in based on the observations by Miyazaki et al. (1999) and Pelton
the plastic strain, while there is no plastic strain for super- et al. (2013). As can be seen in this figure, the very low-cycle
elastic Nitinol alloys unless it is fully martensitic. regime or first regime in the strain-life plot is related to very
Maletta et al. (2012) also attempted to use the Coffin–Manson high strain amplitudes, εa ZεAM F , region D in Fig. 10. The
relationship for both elastic and plastic (i.e. residual strain) material in this region is fully martensitic and exhibits both
strain amplitudes and the failure life of the superelastic elastic and plastic behaviors. The fatigue life decreases
Nitinol alloy in low-cycle fatigue regime. They conducted almost linearly in logarithmic form with increasing strain
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 237

Fig. 10 – Multiple phases for superelastic Nitinol; (A)


austenitic linear elastic region, (B) transition from linear
elastic to stress plateau region, (C) stress plateau region, and
(D) fully detwinned (deformed) martensite region.

amplitude. The low-cycle or second regime contains the data


for high strain amplitudes (εAM S rεa r εF ), which corresponds
AM

to the mixed austenite/stress-induced martensite phase (i.e.


the region C of Fig. 10). In this region, the strain-life fatigue
behavior has a very steep slope and seems to be almost
independent of the strain amplitude. The fatigue for this
range of strains is nearly constant around 1000 cycles. The
seemingly independence of the fatigue life in this regime may
indicate the dominating role of stress, rather than strain, in Fig. 11 – Torsional fatigue behavior of Nitinol alloys using (a)
the fatigue failure of the superelastic Nitinol in the stress shear stress-life approach for martensitic Nitinol (Jensen,
plateau region. The mid-cycle or third regime includes the 2005) and (b) shear strain-life approach for superelastic
strain amplitudes between approximately εpr and εAM S , where
Nitinol (Runciman et al., 2011).
the fatigue life constantly decreases as the strain amplitude
increases. This regime corresponds to region B in Fig. 10. The
fatigue behavior in this regime is mainly related to the the behavior in each region is approximately linear in a log–
transition between the linear elastic behaviors in austenite log scale. It is notable that no distinct low-cycle fatigue
phase to the stress-induced martensite phase (i.e. the stress regime as seen in superelastic alloys, can be observed since
plateau of the superelastic stress–strain plot, shown in there is no stress-induced phase transformation in thermally
Fig. 10). The high-cycle or fourth regime corresponds to martensitic Nitinol alloys.
region A in Fig. 10, where εrεpr. It is worth mentioning that
the second turning point in these curves (point PAM in Fig. 9) 3.1.2. Torsional fatigue
may be related to the starting point of the stress-induced Torsional deformation, similar to tension, compression and
martensitic phase (point εAM s in Fig. 10). The third turning bending, is one of the typical loading conditions for Nitinol
point (point Ppr in Fig. 9) may correspond to the “proportional alloys. There are only a limited number of fatigue-related
limit” of this material (point εpr in Fig. 10) as explained by studies and data on torsional fatigue of Nitinol alloys in the
Miyazaki et al. (1999). Pelton et al. (2013) suggested that a present literature (Jensen, 2005; Predki et al., 2006; Runciman
simplistic log–log linear relationship can be used to represent et al., 2011). The results of separate torsional fatigue studies
the fatigue behavior in first, third and fourth regions with β on Nitinol alloys, performed by Jensen (2005) for a martensitic
coefficients that decrease with increasing temperature. The Nitinol in force-controlled condition and Runciman et al.
equation for the fourth regime is very similar to the classical (2011) for a superelastic Nitinol in strain-controlled condition
fatigue for linear elastic metals and the Basquin (1910)’s are presented in Fig. 11(a) and (b), respectively. Unlike
model. The β values calculated for superelastic Nitinol alloys rotating-bending and uniaxial fatigue, the fatigue of Nitinol
from Pelton et al. (2013) study are in the range of 0.3oβo0.38 under strictly torsional deformations is similar to the tradi-
for the first and third regions, which indicates that the rate of tional fatigue curves for other metallic materials. In other
fatigue degradation is almost the same in these regions. words, the very distinct multi-linear curve of shape memory
However, β is between 0.03–0.07 in the high-cycle fatigue, alloys is not observed for the torsional tests in the log-log
indicating a faster degradation of the material with an representation of the test results. The general reduction in
increased strain. the strength of the material is observed clearly under torsion
For thermally martensitic Nitinol alloys, no such distinct for both martensitic and superelastic Nitinol.
regimes may be observed in fatigue data. The data for this Jensen (2005) was among the few researchers who con-
type of Nitinol alloys can be classified into two regions, where sidered the torsional fatigue of Nitinol. He studied the fatigue
238 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

behavior of Ni49.9Ti50.1 alloy under force-controlled torsional there are only a few studies addressing the multiaxial fatigue
tests. The tests continued until the specimen failed or up to behavior of this material. Multiaxial fatigue behavior of
105 cycles as run-out at 23 1C. The stress-life data from these Nitinol alloys may be completely different from uniaxial,
experiments are shown in Fig. 11(a). No clear information on due to the effect of deformation path (i.e. in-phase or out-
the metallurgical phase condition or the transformation of-phase) on the metallurgical phase transformation phe-
temperatures of the material under study were presented; nomenon. There is a need for further investigations to
however, according to the stress–strain curves, it appears to understand and formulate models for fatigue behavior of
be in the martensitic phase. Investigating the fracture surface Nitinol under multiaxial state of deformation.
of the specimens in Jensen’s study, suggests that cracks Jensen (2005) was among the few researchers who studied
initiate and grow on the plane of maximum shear stress the multiaxial fatigue of a martensitic Nitinol alloy. He
and during crack propagation stage it orients to the plane of investigated the biaxial cyclic deformation and fatigue beha-
maximum principal stress. These observations imply that vior of Ni49.9Ti50.1 under force-controlled proportional (i.e. in-
Nitinol, at least in the martensitic phase, is a shear failure- phase) tension–torsion tests. The tests were conducted at
mode material. Understanding the failure mechanism of the 23 1C to specimen failure or run out to 105 cycles. Although
material is necessary to employ an appropriate fatigue model the metallurgical phase condition and the transformation
under multiaxial state of deformation as will be discussed in temperatures of the material were not indicated, the cyclic
the following sections. stress–strain curves appear to follow martensitic Nitinol
Runciman et al. (2011) studied the torsional fatigue beha- behavior. In-phase, tension–torsion loads with combinations
vior of superelastic Ni50.8Ti49.2 with Af ¼ 16 and 22 1C at test of five, fully reversed, axial stress amplitudes (740, 560, 500,
temperatures of 25 and 37 1C, respectively. Tests were con- 250 and 125 MPa) with a fully reversed, shear stress ampli-
tinued to failure or to 107 cycles as run-out. These results are tude of 250 MPa were employed to characterize the in-phase
presented in Fig. 11(b), where the shear strain-life fatigue multiaxial, fatigue behavior of a martensitic alloy. Jensen
results for the superelastic Nitinol under zero mean strain (2005) suggested a relation between equivalent stress and the
can be approximated using a power-law relation. Predki et al. fatigue life of the martensitic Nitinol alloys as presented in
(2006) conducted the comparison of the fatigue life of tubular Eq. (8):
and solid round specimens under force-controlled conditions  
σ eq ¼ 85:766 ln Nf þ 1859:8 ð8Þ
to study the effects of specimen type on torsional fatigue
behavior of superelastic Nitinol alloys. Predki et al. observed where; σ eq is the von Mises equivalent stress amplitude
greater fatigue lives for solid specimens, when compared in MPa.
with those for tubular specimens, which is in agreement with Jensen (2005) also reported the fatigue lives of in-phase
the observations reported for the effects of cross-sectional multiaxial loading for martensitic Nitinol alloys, where the
shape on the torsional fatigue of steels (Shamsaei and shear stress was dominant (i.e. for the combinations of 250
Fatemi, 2009a, 2009b). and 125 MPa axial stress and 250 MPa shear stress), to be
similar to those obtained under pure torsion loading. There-
3.1.3. Multiaxial fatigue fore, no significant effect from axial loading may result in the
A multiaxial state of deformation is very common in engi- presence of considerable torsional loads. Furthermore, biaxial
neering applications, and even under nominal uniaxial load- tension–torsion fatigue tests, in which tension was more
ing situations, multiaxial loading may exist (Fatemi and dominant, were found to be far more damaging and have
Shamsaei, 2011). The multiaxial stresses in critical elements significant reduction in fatigue lives for this loading condition
of components and structures can result from multi- (Jensen, 2005).
directional loading, stress concentrations, or residual stresses Mahtabi and Shamsaei (2015) recently analyzed the torsional
(Shamsaei and McKelvey, 2014). Some studies have evaluated and in-phase multiaxial fatigue data generated by Jensen (2005)
the mechanical loading conditions of many biomedical pro- for martensitic Nitinol. The martensitic Nitinol was found to be
ducts made of Nitinol and confirmed the general state of a shear failure mode material. Therefore, they evaluated
deformation is multiaxial (Allie et al., 2004; Cheng et al., 2006; various classical and critical plane multiaxial fatigue models,
Nikanorov et al., 2008). Fatemi and Shamsaei (2011) have such as von Mises, McDiarmid, and Findley to correlate tor-
discussed several important factors that should be consid- sional and in-phase multiaxial fatigue data and reported
ered under multiaxial deformation. These include damage unsatisfactory results. They also proposed a Fatemi–Socie-type,
mechanisms, which provide the basis for developing the shear stress-based damage parameter, considering the coupling
fatigue damage parameters, non-proportional cyclic harden- effects of shear and normal stresses. The proposed model was
ing/softening and constitutive behavior that may signifi- found to correlate torsional and in-phase multiaxial fatigue
cantly affect the multiaxial fatigue response, and multiaxial data for martensitic Nitinol fairly well.
damage parameters for fatigue life and design durability One of the common methods for multiaxial fatigue life
estimation models. In addition to these important factors, predictions of engineering materials is to use the von Mises
the effects of the multiaxial deformation path (i.e. in-phase equivalent stress/strain to assimilate the multiaxial load
versus out-of-phase) on the phase transformation phenom- parameters to a single parameter and then calculate the
enon should be considered for shape memory alloys such as fatigue strength using the existing uniaxial fatigue data/
Nitinol. properties. Following this approach, Runciman et al. (2011)
Despite the importance of understanding the fatigue conducted experiments on medical grade, superelastic Niti-
behavior of Nitinol under multiaxial state of deformation, nol (i.e. Ni50.8Ti49.2) under torsional strain and compared the
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 239

results to the uniaxial tensile or bending test data to obtain a be appropriate for multiaxial fatigue analysis of Nitinol.
relation for multiaxial fatigue behavior. They proposed an Fatemi–Socie critical plane model (Fatemi and Socie, 1988)
equation that relates the alternating equivalent von Mises appears to be an appropriate fatigue model for shear failure
strain, Δεeq =2, to the fatigue life as expressed by Eq. (9): mode governed materials. In this model, the maximum shear
Δεeq    0:4 strain is the driving parameter for the crack and the max-
¼ 0:25 þ 49:6 Nf ð9Þ imum normal stress on the plane of maximum shear strain, if
2
it is tensile, opens the crack and accelerates crack growth.
Runciman et al. showed that Eq. (9) worked reasonably
The model is a physically based model that is capable of
well for strain ratios in the range  1rRε r0.6, this relation-
considering the constitutive behavior of the material includ-
ship did not correlate with the experimental results for
ing non-proportional cyclic hardening and the failure mode
greater strain ratios (greater mean strain levels). It should
of the material (Fatemi and Shamsaei, 2011; Fatemi and Socie,
be noted that Runciman et al. calculated the equivalent strain
1988; Park and Nelson, 2000).
amplitude using Eq. (10), which does not account for the
Cyclic behavior of Nitinol is also highly dependent on the
effects of the equivalent Poisson’s ratio:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi material phase and its response to different multiaxial cyclic
4 loading. As the multiaxial deformation path may affect the
εeq ¼ ε2t þ ε2s ð10Þ
3 metallurgical phase transformation phenomenon, its effects
In order to consider such an effect, one may use the should be taken into consideration in fatigue life estimations
equivalent von Mises strain equation expressed as: for shape memory alloys. More comprehensive studies
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi regarding the multiaxial fatigue of Nitinol alloys that consider
1
εeq ¼ ð1 þ νÞ2 ε2t þ 3ε2s ð11Þ the effects of multiaxial deformation paths are still required
1þν
to formulate more accurate multiaxial fatigue predictive
Runciman et al. proposed a modified Coffin–Manson equa-
models.
tion for multiaxial fatigue of the superelastic Nitinol. Their
proposed approach relates the number of cycles to failure to
3.1.4. Mean stress/strain effects on fatigue behavior
the equivalent alternating transformation strain, Δεtr =2. Their
In previous sections, the fatigue behavior of Nitinol under
model can be expressed as:
fully reversed loading was discussed. Miyazaki et al. (1986b)
Δεtr indicated that zero mean stress fatigue analysis may be
¼ 61:7 Nf 0:5 ð12Þ
2 suitable for structural materials, but it cannot predict accu-
where the εtr is the equivalent von Mises transformation rately the fatigue behavior of superelastic Nitinol alloys. In
strain and Nf is the number of cycles to failure. The transfor- many applications, Nitinol alloys may be subjected to a mean
mation strain is defined as the difference between total and strain or stress during the durability life of a component.
pure elastic strain. The modified Coffin–Manson relation Nitinol stents, as an example, are subjected to at least two
proposed by Runciman et al., however, seems to work well types of deformations; a semi-constant tension from the
for the torsional fatigue data; unfortunately, it does not artery and a cyclic pulsatile deformation of the artery wall
provide a good correlation with the axial data. Furthermore, caused by the changes in systolic–diastolic blood pressure
the accuracy of the model for multiaxial loading with both (there are, of course, duty cycle deformations that are body
axial and shear strains applied simultaneously to the same location specific that are superimposed upon these and
specimen was not validated. Although the simplified method depending on body location may be of greater significance).
proposed by Runciman et al., can be helpful as initial work for For this reason, non-zero mean stress or strain tests (pulsatile
the multiaxial fatigue of Nitinol, it does not address the type loading) are used to obtain a better evaluation of the
proportional and non-proportional deformation effects that fatigue behavior of Nitinol components in such applications
include a possible non-proportional cyclic hardening in (Duerig et al., 1999; Pelton et al., 2008, 2003; Tolomeo et al.,
superelastic Nitinol. 2000).
In general, classical fatigue criteria, such as von Mises, do Mean stresses affect significantly the fatigue behavior of
not work well for non-proportional (i.e. out-of-phase) loading Nitinol. Fatigue testing of Nitinol alloys under tensile mean
and typically underestimate fatigue damage under such stresses results in a drastic reduction in the fatigue strength
multiaxial loadings (Fatemi and Shamsaei, 2011; Shamsaei in the linear elastic region, as long as the maximum stress is
and Fatemi, 2009b). Furthermore, the accuracy of these not great enough to induce the martensitic phase (Moumni
classical models to work even under multiaxial loading et al., 2005). The S–N curves from identically treated dog-
blocks, consisting of only proportional cycles, has been bone-shape Nitinol specimens under tension-compression
questioned recently (Fatemi and Shamsaei, 2011; Shamsaei demonstrate the reducing effects of tensile mean stresses
et al., 2011). on the fatigue resistance of the superelastic Nitinol alloy as
More observations on the multiaxial fatigue behavior of shown in Fig. 12 (Moumni et al., 2005). As seen, the fatigue
Nitinol can be made regarding the failure mode of the lives under fully reversed stressing (Rσ ¼ 1) are significantly
material. As mentioned previously for fatigue of Nitinol greater than the fatigue lives under pulsating loading (Rσ ¼0).
under torsion, Nitinol appears to be a shear failure mode Another study examining the effects of non-zero mean
governed material. It is known that cracks tend to initiate and stress on the fatigue life of Nitinol was performed by Tabanli
grow around preferred planes, so-called critical planes under et al. (1999). They investigated the fatigue behavior of super-
multiaxial deformations. A critical plane fatigue model, con- elastic Nitinol tubes with Af between 3 and  6 1C at room
sidering the physical nature of the failure in the material may temperature ( 22 1C). The uniaxial tensile experiments were
240 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

Fig. 12 – Mean stress effects on the fatigue of superelastic


Nitinol dogbone specimens (Moumni et al., 2005).
Fig. 14 – Tensile and compressive mean strain effects on
fatigue behavior of the superelastic Nitinol (Elahinia, 2015).

representing mean strain effects on fatigue of Nitinol are the


studies by Tolomeo et al. (2000) for 106 cycles and Pelton et al.
(2003) for 107 cycles of run-out tests, as shown in Fig. 14. In this
plot, the vertical axis is the alternating strain, εa, and the
horizontal axis is the mean strain, εm. According to the
observations reported by Tolomeo et al. (2000), the tolerable
alternating strain for the superelastic Nitinol increases with
increasing mean strain up to a certain mean strain amplitude
Fig. 13 – Stress-life fatigue data indicating the effects of
(e.g. 3%). However, beyond 3% mean strain, the tolerable
mean stress on the fatigue behavior of a superelastic Nitinol
alternating strain decreases as the mean strain increases with-
(Tabanli et al., 1999).
out failure of the specimen.
Analogous to the data from Tolomeo et al. (2000), the
observations by Pelton et al. (2003) confirms similar effects of
performed in the strain-controlled (displacement-controlled) mean strain on the fatigue behavior of superelastic Nitinol
condition at 20 Hz frequency in a non-controlled environment. alloys as can be also seen in Fig. 14. The latter study includes
Different fatigue lives for various mean stress values were the compressive mean strains that are calculated from finite
observed at constant alternating stress amplitude as presented element analysis of the stent-like components. Their findings
in Fig. 13. Data from Tabanli et al. (1999) does not indicate a for superelastic Nitinol indicate that the bearable alternating
specific relation between mean stress and the fatigue life for strain decreases as the mean strain increases up to 1–1.5%.
superelastic Nitinol alloys. In their experiments, specimens E Above this mean strain, however, the alternating strain tends
and A failed in the grip; thus, the actual fatigue lives for these to increase by an increase in the mean strain level. Both
two specimens were higher than the ones reported in Fig. 13. experiments were performed at 37 1C using specimens with
The difference in fatigue life and mean stress level between Af ¼ 28 1C (Tolomeo et al., 2000) and Af ¼ 30 1C (Pelton et al.,
specimen E and other specimens are notable. The mean stress 2003) meaning that all specimens were in the superelastic
in specimen E is much higher than other specimens. The austenitic phase. This behavior can be attributed to the
greater mean stress can cause stress-induced martensitic formation of stress-induced martensitic phase in Nitinol
phase in the material and consequently leads to drastically under greater mean strain values (Pelton et al., 2003). The
greater fatigue strength as compared with its austenitic findings of Pelton and co-workers and Tolomeo et al. on the
counterpart. The contradictory results from Tabanli et al.’s improved fatigue life with greater mean strains was also
study to other studies such as Moumni et al. (2005) may be confirmed by Morgan et al. (2004). They (Morgan et al., 2004)
attributed to the greater testing frequency in Tabanli et al.’s reported greater fatigue lives for superelastic Nitinol speci-
experiment, which may induce latent heat in the specimen. mens with Af ¼ 12 1C by increasing mean strain level as
This latent heat can significantly affect the material properties presented in Fig. 15. It can be seen in this figure that the
and result in different fatigue lives. fatigue life improvement was more significant in the high-
Effects of tensile mean strain on the fatigue of superelastic cycle regime, when compared to the low-cycle fatigue regime.
Nitinol could be beneficial or detrimental, depending on the Similar to the previously mentioned studies, the greater
mean strain level (Pelton et al., 2008). One common method to fatigue lives may be the result of the formation of the
compare the results for different combinations of mean strains stress-induced martensitic phase in presence of greater mean
and strain amplitudes is to compare these combinations at a strains. Some researchers reported greater fatigue lives in
certain run-out life (e.g. 106 or 107 cycles), so-called constant-life superelastic Nitinol for a combination of a low alternating
diagram (Mitchell, 1996). In this method, experimental results strain and a high mean strain (Duerig et al., 1999; Morgan
from different combinations of mean and alternating strain are et al., 2004). Morgan et al. (2004) attributed these longer
plotted and compared. Examples of this method for fatigue lives to the increasing amount of stabilized
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 241

Fig. 17 – Effect of mean shear strain on the torsional fatigue


Fig. 15 – Tensile mean strain effect on the strain-life fatigue of superelastic Nitinol alloys (Runciman et al., 2011).
data of the superelastic Nitinol (Morgan et al., 2004).

et al. (2005), where they related the beneficial effects of the


compressive mean stresses to the closure of microcracks.
Tensile mean stresses seem to be detrimental to the fatigue
resistance of martensitic Nitinol alloys, the more familiar and
expected response, as can be seen in Fig. 16 (Kang et al., 2012).
However, there are insufficient data in the literature to
generalize the effects of stress ratio on the stress-life fatigue
behavior of Nitinol alloys.
The effects of mean shear strain on the fatigue behavior of
the superelastic Nitinol have been experimentally investi-
gated by Runciman et al. (2011) for multiple shear strain
ratios of Rγ ¼γmin/γmax ¼  1, 0, 0.2, and 0.6 under strain-
Fig. 16 – Mean stress effects on the fatigue of martensitic controlled torsional tests. Comparing the results for different
Nitinol (Kang et al., 2012). mean shear strain ratios as presented in Fig. 17, the effect of
mean shear strain on the fatigue resistance of superelastic
Nitinol is not necessarily detrimental. Therefore, the pre-
martensitic Nitinol, which resulted from the non-zero mean sence of mean shear strains may enhance the fatigue life of
strains acting upon the material. Nitinol alloys, which is similar to the effects of tensile mean
Mean stress and mean strain effects on the fatigue strains. This behavior can be seen clearly in Fig. 17, where the
behavior of Nitinol have been the subject of considerable specimens under pulsatile loading (Rγ ¼ 0) endure greater
interest to other researchers as well. Tabanli et al. (2001) also fatigue lives than the specimens under fully reversed cyclic
investigated the effects of mean strain on the fatigue beha- loading (Rγ ¼ 1) with the same alternating shear strains.
vior of superelastic Nitinol. They observed a drastically However, making reliable conclusions on the effects of shear
shorter fatigue life for superelastic Nitinol specimens on the mean strains on the torsional fatigue behavior of Nitinol
stress plateau region as the mean strain level increased. They alloys requires more comprehensive experimental and ana-
considered the sharp edges between different phases, i.e. lytical works to evaluate the possibility of formation of stress-
austenite and deformed martensite, of material (Shaw and induced martensite under different shear load ratios.
Kyriakides, 1997) as a possible cause for shorter fatigue life in
presence of mean strains. 3.1.5. Phase contribution to fatigue behavior
The classical mean stress/strain correction methods such The dependence of the fatigue behavior to different phases of
as Goodman or Soderberg are certainly not applicable for Nitinol was addressed first by Melton and Mercier (1979a). A
superelastic Nitinol alloys (Kugler et al., 2000; Morgan et al., stress-life plot for three different Nitinol alloys for specimens
2004; Tabanli et al., 2001, 1999). This is due mainly to the fact with different Af temperatures is shown in Fig. 4. The first set
that the stress–strain relations of these alloys for loading and of specimens with Af ¼ 10 1C (the austenite finish temperature
unloading directions are not similar (Robertson et al., 2012). is calculated by the approximation; Af ¼Msþ40 1C Robertson
In addition, the fatigue resistance of the Nitinol alloys may et al., 2012) are in the austenitic phase at room temperature
not always decrease as the mean stress or strain increases. and exhibit superelastic behavior. The second set with
Kang et al. (2012) studied the influence of mean stresses Af ¼110 1C are martensitic at the room temperature. Fig. 4
on the fatigue behavior of martensitic Nitinol alloys. Similar indicates clearly that the superelastic specimens have sig-
to other metals, their observations show that compressive nificantly greater fatigue limits in stress-life fatigue, when
mean strain improves the fatigue resistance of martensitic compared with the thermally martensitic specimens. Similar
Nitinol alloys as shown in Fig. 16. Results from Kang et al. results for stress-life fatigue of Nitinol were also reported by
(2012) are also in agreement with the findings of Moumni Kim and Miyazaki (1997) as illustrated in Fig. 6. This greater
242 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

stress-based fatigue life for the superelastic Nitinol can be εAM


S , and martensite finish strain, εF ) the material shows a
AM

attributed to the greater stress carrying capacity of the modulus of elasticity between EA and EM, depending on the
superelastic phase, when compared with the thermally mar- volume fraction of the martensitic phase. This change of the
tensitic Nitinol, as can be seen in Fig. 3. modulus of elasticity in loading and unloading at different
Similarly, the effects of the microstructural phase to the strain levels significantly influences the material cyclic
fatigue behavior of the Nitinol alloys in the strain-based response in the presence of mean strains.
approach can be illustrated by strain-controlled fatigue test- The effects of microstructural phase on the crack growth
ing of identical samples at various test temperatures. Accord- of Nitinol alloys have been also addressed in some studies
ing to the separate strain-controlled experiments conducted (Gollerthan et al., 2009; McKelvey and Ritchie, 2001). For
by Miyazaki et al. (1999) and Kim (2002), the strain-based superelastic Nitinol, as long as the crack nucleation is
fatigue lives of the thermally martensitic Nitinol specimens inhibited, the stress-induced martensite (i.e. region C in
are significantly greater than those of superelastic Nitinol Fig. 10) is the preferred condition, which exhibits greater
specimens, as shown in Fig. 7. Strain-life fatigue data from a crack propagation resistance as compared to the fully auste-
rotating-bending study of Kim and Miyazaki (1997) also nitic phase (i.e. region A in Fig. 10). Thermally martensitic
confirmed greater fatigue lives for thermally martensitic Nitinol alloys usually show drastically superior resistance to
Nitinol specimens. fatigue crack propagation than the superelastic austenitic
The contradictory results in the literature on the fatigue alloys (Holtz et al., 1999; McKelvey and Ritchie, 2001).
resistance of different phases of Nitinol alloys (i.e. thermally
martensitic and superelastic) may be attributed to the fatigue 3.2. Thermo-mechanical fatigue
life approach employed to analyze the data. In summary, the
superelastic Nitinol exhibits greater fatigue resistance in the Nitinol, as well as other shape memory alloys, are used as an
stress-based fatigue approach, whereas martensitic Nitinol actuator in many applications at which they are under
alloys typically have greater lives in the strain-based alternating stresses and strains resulting from thermal
approach. As stated previously, the greater fatigue lives cycling. Therefore, the material is prone to thermo-
observed in stress-based fatigue analysis of Nitinol alloys mechanical fatigue under constant or variable strains/stres-
can be attributed to the greater stress carrying capacity of ses. There are a limited number of studies in the literature
superelastic Nitinol alloys, when compared to martensitic concerning the thermo-mechanical fatigue of shape memory
ones, as can be seen in Fig. 3. alloys (Bertacchini et al., 2009; Bigeon and Morin, 1996;
A complex phase transformation may take place in super- Lagoudas et al., 2009, 2000) and even fewer studies addressing
elastic Nitinol components in presence of mean strains, the thermo-mechanical fatigue behavior of Nitinol (Fumagalli
which can accelerate or decelerate the fatigue crack initiation et al., 2009; Mertmann and Vergani, 2008; Scirè Mammano
process. The effects of metallurgical phase on the fatigue and Dragoni, 2014). Although many researchers have referred
behavior of the superelastic Nitinol in the presence of mean to the failure under coupled thermal and mechanical loads as
strains are reasonably well demonstrated in an article by functional fatigue, this type of failure is, in fact, thermo-
Pelton (2011). As reported, an increase in mean strain in the mechanical fatigue.
austenite phase that is up to the linear elastic limit (i.e. region Scirè Mammano and Dragoni (2014) recently studied the
A in Fig. 10) is detrimental to the fatigue strength. However, thermo-mechanical failure of the Nitinol alloys under various
an increase in tensile mean strain from the linear elastic types of loading; constant stress, constant strain, and con-
“proportional limit” to the strain corresponding to the start of stant stress with limited strain. In their analysis, they have
upper superelastic plateau (i.e. region B in Fig. 10) appears plotted the fatigue life of the specimen (in term of number of
beneficial to the fatigue resistance of superelastic Nitinol cycles to failure) versus a single test parameter such as stress
alloys as long as the total applied strain remains within the or strain, depending on the test constraint. An example plot
superelastic (i.e. recoverable) region (i.e. regions C and linear of the results from their study is shown in Fig. 18(a) for
section of region D in Fig. 10). The formation of stress- constant stress test and constant stress test with 4% limited
induced martensite in the superelastic plateau region is strain. The amplitude of the strain of about 4% in both sets of
responsible for the improved fatigue strength in superelastic test results presented in Fig. 18(a) indicates the similar
Nitinol alloys (Pelton, 2011). Finally, an increase in tensile temperature change in both test conditions. In Fig. 18(b), we
mean strain level in the deformed martensite phase (i.e. have plotted the same sets of data against a Smith–Watson–
beyond the linear section of region D in Fig. 10), which is Topper (SWT)-type parameter that includes both stress and
beyond the superelastic plateau region, results in a reduced strain effects, σεmax. SWT-type parameter was selected
fatigue resistance. because the stress and strain in these tests are almost
Another consequence of stress-induced phase transfor- uncoupled parameters; therefore, the fatigue damage para-
mation in superelastic Nitinol is that the modulus of elasti- meter should be able to capture the effects of both para-
city varies as a function of the volume fraction of each phase meters. The original Smith–Watson–Topper parameter
in the material (Robertson et al., 2012). In Fig. 10, the modulus (Smith et al., 1970) was originally defined as σmaxεa and
of elasticity of the fully austenitic material is denoted as EA. considers the synergistic effects of strain amplitude for mean
As the strain amplitude increases, the volume fraction of the stress effects (Smith et al., 1970). It was used also for multi-
martensite increases until it reaches a fully martensitic axial fatigue analysis (Socie, 1987) as well as load history
region, where the modulus of elasticity becomes EM. For effects (Colin and Fatemi, 2010). As can be seen from Fig. 18
strains between these two limits (i.e. martensite start strain, (b), using a fatigue parameter considering the synergistic
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 243

transformation temperatures of this Nitinol, such as Ms,


and Mf, remained essentially unchanged. Several thermo-
mechanical cyclic heating tests on Nitinol were also con-
ducted recently by Casati et al. (2011) and significant changes
in the transformation temperatures under thermo-
mechanical (i.e. thermal cycling in presence of external
mechanical forces) cycling were observed. Similar results
were reported in a study by Pelton (2011), where several
cycles of heating and cooling were applied to Ni50.5Ti49.5
specimens. Benafan et al. (2014) observed a significant change
in the texture of the martensitic Nitinol under thermal
cycling. However, the change for austenitic Nitinol did not
Nf
seem to be significant.
Although different effects of thermal cycles on transfor-
mation temperatures of Nitinol alloys have been observed, a
stable state of shape memory properties has been reported
after few cycles of thermal loading in almost all studies.
Therefore, a more stable structure and behavior may be
achieved by prior thermal cycling of Nitinol alloys. Micro-
structural rearrangements under thermal cycles may be
responsible for the initial transient behaviors observed for
this material. Nevertheless, more comprehensive experimen-
tal studies are still required to better understand the thermal
fatigue of Nitinol alloys.

Nf
3.3. Functional fatigue
Fig. 18 – Thermo-mechanical fatigue of Nitinol (a) single-
parameter plot (Scirè Mammano and Dragoni, 2014) (b) SWT- Eggeler et al. (2004) first proposed the term “functional
type parameter plot for the same set of the data. fatigue” indicating a decrease in the functionality of the
shape memory alloys during cyclic deformations. The func-
tional property can be the dissipated energy in a pseudo-
elastic damping application or maximum recoverable strain
effects of stress and strain leads to almost identical fit for of the material. Prior to Eggeler et al., other researchers also
both data sets. studied the functional behavior and degradation of the
The change in transformation temperatures of Nitinol Nitinol alloys under cyclic deformation (Melton and Mercier,
under oscillating temperatures with or without external 1979a; Miyazaki et al., 1986b; Sawaguchi et al., 2003); however,
mechanical deformation is sometimes called thermal fatigue. they did not refer to it as “functional fatigue”. Although some
Although this type of the change in the material property of the studies discussed in this section have not referred
does not necessarily lead to actual failure and may not be directly to the change in material properties as “functional
considered as fatigue failure, a brief discussion on this fatigue”, we have categorized them by using the definition
property change in the material is presented in this section. provided by Eggeler et al. (2004).
To date, several studies on thermal fatigue of Nitinol have Miyazaki et al. (1986b) were one of the first to investigate
been conducted and reported generally a reduction in trans- the effects of cyclic deformations on the functionality of
formation temperatures. Miyazaki et al. (Miyazaki et al., superelastic Nitinol alloys. They conducted an experimental
1986a) attributed the decrease in the transformation tem- study on Ni50.6Ti49.4, with different heat treatments from
peratures of Nitinol alloys during thermal cycling to the water-quenching to gradual cooling of annealed materials.
presence of large amount of defects, such as dislocations, Transformation temperatures were measured using the elec-
which resist the transition of the parent austenite phase to trical resistance of the wires. Results of their study show a
the martensitic phase. He et al. (2006) studied the effects of reduction in the stress level corresponding to the starting
thermal cycling on the transformation temperatures of point of the stress-induced martensite region (i.e. martensite
Ni43Ti50Cu7 alloy and found that Ms, Mf and Af tend to start stress, σ AM
S , in Fig. 10) by increasing number of cycles, as
decrease with increasing thermal cycles, while As is prone illustrated in Fig. 19(a). Such changes in the stress–strain
to an increase. The change in transformation temperatures curve of superelastic Nitinol alloys were also reported in
reaches a stable state after  10 cycles of thermal cycling; other studies (Gall et al., 1998; Nayan et al., 2008). Moreover,
thus, more stable Nitinol alloys can be obtained by applying as Miyazaki et al. (1986b) stated, an increase in residual (i.e.
cyclic thermal straining prior to the actual application of the unrecoverable) strain and a reduction in hysteresis loop area
material. (as an indicator of the dissipated energy) are expected by an
Experiments conducted by Urbina et al. (2009) also increase in number of cycles. These changes, which were
revealed a slight reduction in austenite finish temperature, described schematically in Fig. 19(a) and also in Fig. 19(b),
Af, of a Nitinol alloy under thermal cycling. The other were also confirmed by Gall et al. (1998) and Nayan et al.
244 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

Similar to uniaxial loading, the hysteresis loop of both


superelastic and thermally martensitic is subjected to change
in the first few cycles of shear loading (Jensen, 2005; Wang
et al., 2010). The hysteresis loops for the superelastic (Wang
et al., 2010) and thermally martensitic Nitinol (Jensen, 2005)
under force-controlled shear deformation becomes narrower
as the number of cycles increases. This indicates a reduction
in the ability of the superelastic as well as thermally mar-
tensitic Nitinol in absorbing and dissipating energy under
cyclic shear loading.
The functional fatigue under multiaxial cyclic loading is
also a possible failure mode for components made of Nitinol
(Song et al., 2014; Wang et al., 2012). Similar to uniaxial
testing, the capability of the superelastic Nitinol to dissipate
energy decreases for the first few cycles and then reaches a
stable state under multiaxial state of deformation.
Despite the fact that there are some studies in the
literature concerning the “functional fatigue” of Nitinol, this
term is not accepted widely in the fatigue research commu-
nity. This is because all materials are cyclically unstable and
may cyclically soften, harden or remain stable under strain-
controlled fatigue conditions. These changes in the mechan-
ical properties under cyclic deformations are affected by the
material’s microstructure, where dislocations are constantly
in rearrangement. Since the fatigue process is degradation of
Fig. 19 – Schematic representation of the change in the the material under cyclic deformations, and by some
stress–strain curve of superelastic Nitinol (a) cyclic stress– mechanism such as dislocation motion, detwin/twin, etc.,
strain curve for a superelastic and (b) cyclic change in the “bulk” response may not be altered significantly until
starting point of the stress-induced martensitic phase, σ AMS
there is a crack initiated and propagated. All materials
(Miyazaki et al., 1986b). undergo cyclic readjustment in the first few cycles of their
total fatigue life, at a given strain amplitude. This phenom-
enon is very common and does not necessarily cause a failure
in the material. The stability for Nitinol is achieved in the first
(2008). Miyazaki et al. (1986b) attributed the increase in the 100 cycles as reported by many studies (Eggeler et al., 2004;
residual strain to the slip formation under cyclic loading. The Miyazaki et al., 1986b). Since the material quickly reaches its
rate of change in functional parameters of superelastic stable-state and there is no actual failure in the material, not
Nitinol specimens decreases by an increase in number of many researches have considered functional fatigue as a real
cycles that results in a more stable state of functionality type of failure for Nitinol alloys.
(Miyazaki et al., 1986b). Therefore, a more stable superelastic
Nitinol can be expected by prior cyclic training of this shape
memory alloy (Miyazaki et al., 1986b). 4. Failure mechanism and microstructural
Several researches have addressed the degradation of the effects
functional properties of superelastic Nitinol alloys in terms of
the ability of the material to absorb energy (i.e. the area of the Microstructural properties such as grain size, lattice defects,
hysteresis loop) or the evolution of unrecoverable/residual and inclusions are important factors that affect the fatigue
strain that affects the maximum recoverable strain (Eggeler behavior of Nitinol (Eggeler et al., 2004). Wilkes and Liaw
et al., 2004; Scirè Mammano and Dragoni, 2014). All these (2000) reported significant effects of grain size on the fatigue
studies unanimously reported two common behaviors in the behavior of Nitinol alloys. The grain size influences primarily
functional properties of the superelastic Nitinol. First, the the stress concentration on the grain boundaries that may
hysteresis loop of the material becomes narrower as the induce local martensitic phases in the superelastic Nitinol.
number of cycles of loading increases as shown in Fig. 19(a). Stress concentrations at the grain boundaries, eventually lead
Second, the stress–strain response of the material stabilizes to permanent deformation and crack initiation. The localized
after several cycles of loading, i.e. 100–150 cycles (Eggeler strains along the slip bands seem to be reduced for materials
et al., 2004; Frotscher et al., 2011; Gall et al., 1998; Maletta with finer grain size. Moreover, the micro cracks tend to
et al., 2014; Mao et al., 2006; Miyazaki et al., 1986b; Moumni initiate from the locations with more stress concentrations
et al., 2005; Nemat-Nasser and Guo, 2006; Sehitoglu et al., such as inclusions, laminations, and voids/porosities under
2001). Comparable changes for thermally martensitic Nitinol cyclic loadings or thermal cycling. More recently, several
(Kang et al., 2012; Melton and Mercier, 1979a) and mixed- studies have focused on the microstructural aspects of
mode Nitinol (Nayan et al., 2008) were also reported in the fatigue of Nitinol to find a relationship between microstruc-
literature. tural properties and fatigue resistance (Coda, 2014; Launey
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 245

et al., 2014). Launey et al. (2014) investigated the effect of decreasing impurities or providing nano-size Ni-rich particle
impurities by comparing the fatigue behavior of high purity precipitation. They also reported an improved fatigue resis-
grade of Nitinol with those of standard Nitinol. They observed tance of Nitinol alloys by increasing dislocation density.
five times greater fatigue strength for the high purity grade
Nitinol compared with the standard Nitinol in terms of strain
amplitude and for a run-out life of 107 cycles. This can be 5. Treatment and processing effects
attributed to the presence of impurities such as carbide (TiC)
or oxide inclusions that normally serve as locations for One of the key parameters that can influence greatly the
metallurgical-type stress concentration and crack nucleation fatigue resistance of shape memory alloys is the final treat-
(Launey et al., 2014). Such observations generates a need for ment of the material. Based on Transmission Electron Micro-
developing microstructure-sensitive fatigue models for Niti- scopy (TEM) observations, Pelton (2011) concluded
nol alloys. “processing plays a key role in the fatigue behavior of
Microstructural features of failed superelastic Nitinol Nitinol”. In another study, Condorelli et al. (2010) showed
micro stents were studied by Frotscher et al. (2009). They the thermal treatment to have beneficial effects on fatigue
found surface defects, such as TiC particles, pores, laser resistance of Nitinol components.
burrs, and surface cracks, as the main sites for crack initia- Several different thermal and/or mechanical methods are
tion. They also indicated that the fatigue life of strut ele- proposed in the literature to construct a more stable Nitinol
ments in medical stents was controlled by the crack by using different treatment techniques. The combination of
nucleation stage and there was no distinct crack growth cold work and aging treatments may lead to an optimized
stage observed on the fracture surface. In subsequent state of microstructure that results in a more stable shape
research, Frotscher et al. (2011) observed an increase in memory alloy under cyclic deformations (Duerig et al., 1999;
dislocation density in some grains with an increase in the Miyazaki, 1990; Miyazaki et al., 1986b). Combinations of cold
number of cycles. This localized dislocation density may form work and aging treatments result in a minimized amount of
stress-induced martensite in some areas within the material. unresolved strain; therefore, it is a preferred process as
They also reported crack nucleation at a TiC particle under compared to the combination of annealing and aging
cyclic loading. Analogous to the findings of Frotscher et al. treatment.
(2011), Patel (2007) considered the impurities as the main For annealed specimens, Pelton (2011) compared the total
source of crack initiation and introduced the particle mor- fatigue lives under uniaxial strain-controlled with 1.5% mean
phology as a controlling parameter on the microcrack propa- strain and 0.25% strain amplitude at room temperature
gation behavior of the superelastic Nitinol. Suggestions are (T¼ 22 1C). He demonstrated an enhanced fatigue resistance
made in the literature to correlate the fatigue limit of Nitinol by increasing annealing time and temperature for Nitinol
with the size of the extreme inclusions in the material alloys with Af ¼  15 1C (i.e. austenitic phase in room
(Urbano et al., 2013). The material with a larger inclusion size temperature).
tends to have a lower fatigue limit as anticipated from a
generalized local strain-based fatigue notch analysis of such
micro-discontinuities (Mitchell, 1979, 1977). 6. Other parameters
Delville et al. (2011) indicated that the finer microstruc-
tures for Nitinol wires result in greater resistance to disloca- In addition to the influential parameters that previously
tion slip, which is typical in Nitinol alloys with larger grain discussed, other parameters, such as test temperature, sur-
size. They defined a grain diameter less than 100 nm as a fine face finish, specimen geometry, and cyclic frequency (rota-
grain and that with 200 nm of diameter or larger as a coarse tional speed in rotating-bending test), may also influence the
grain. They also attributed the cumulative plastic strain and fatigue behavior of Nitinol alloys. Effects of these parameters
functional fatigue to the dislocation slip phenomenon under on the fatigue behavior of Nitinol have been investigated in
cyclic deformations. Mao et al. (2006) also related the func- the literature (Norwich and Fasching, 2009; Reinoehl et al.,
tional fatigue of superelastic Nitinol, in the form of reduction 2000; Sawaguchi et al., 2003). One of the parameters that
in modulus of elasticity, to the grain orientation and the greatly affects the fatigue resistance of Nitinol components is
stabilization of martensitic twins in the material as well as the test temperature (Miyazaki et al., 1999, 1988; Pelton et al.,
the formation of defects and dislocations through cyclic 2013). Some studies on the effects of test temperature on
loading. They reported the reduction in the stress plateau to fatigue behavior of Nitinol presented in Figs. 7 and 8 have
be dependent on the grain reorientation. been discussed in the previous sections. As can be seen in
Nayan et al. (2008) reported that progressive accumulation Fig. 7, the effect of test temperature is more significant for the
of stress-induced martensite in the superelastic Nitinol alloy test temperatures below the austenite finish temperature, Af,
to be a cause of the fatigue failure. Zurbitu et al. (2010) where the material is not in the superelastic phase (Miyazaki
suggested that the accumulation of the dislocation density et al., 1999).
is highly dependent on the loading rate. Their conclusion These data from Miyazaki et al. (1999) on the fatigue
was based on the comparison of the reduction in the critical behavior of Nitinol alloys under strain-controlled test condi-
stress for inducing martensitic phase under impact and cyclic tion, presented in Fig. 7(a), indicates a decrease in the fatigue
fatigue loads. Finally, Kollerov et al. (2013) suggested that life as the test temperature increases. Similar results were
the low-cycle fatigue of superelastic Nitinol alloys can be obtained in a more recent study by Pelton et al. (2013) for
increased by microstructural enhancement such as superelastic Nitinol. For temperatures above the austenite
246 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

Fig. 20 – Effects of test temperature on the tolerable strain Fig. 21 – Effects of surface finish quality on the fatigue
amplitude of superelastic Nitinol alloy at 107 cycles. behavior of the superelastic Nitinol (Pelton et al., 2013).
ΔT ¼ T Af, where T is test temperature and Af is the
austenite finish temperature (Pelton et al., 2013).
etching are confirmed in other studies as well (Patel and
Gordon, 2006; Condorelli et al., 2010).
finish temperature, and in strain-based fatigue analysis, the Introducing additive materials, such as Cu or TiC to typical
greater temperature may result in lower fatigue limits (Pelton binary Nitinol alloys may not significantly affect the fatigue
et al., 2013). The difference between the test temperature and behavior of these alloys (Miyazaki et al., 1999; Vaidyanathan
the austenite finish temperature of the Nitinol has been et al., 2000). Vaidyanathan et al. (2000) studied the effects of
reported to be a more important factor than the absolute TiC on the fatigue crack growth properties of Nitinol by using
test temperature (Pelton et al., 2004). Pelton et al. (2013) two different contents of TiC as the reinforcement; 10% and
investigated the effect of relative test temperatures, ΔT, 20%. Although the uniaxial stress–strain behaviors of the
which is the difference between test temperature and the specimens were quite different, Vaidyanathan and co-
austenite finish temperature, of medical grade Nitinol workers concluded that the overall crack propagation beha-
(Ni50.8Ti49.2) for a 107 cycles run-out. Their results, shown in vior of the reinforced shape memory alloys were similar to
Fig. 20, indicate an increase in fatigue limit for 107 cycles the original alloy and there was no major change correspond-
when the test temperature is closer to Af but a decrease in ing to the amount of TiC. However, a slight reduction in the
fatigue limit, when the test temperature is increased above crack initiation of Nitinol was noted by adding 10% Cu to the
the austenite finish temperature. However, no effects of test composition of this alloy (Miyazaki et al., 1999).
temperature on fatigue resistance of Nitinol can be observed Takeda et al. (2013) implanted nitrogen ions into
for relative test temperatures lower than 40 1C, which is Ni50.85Ti49.15 wires and investigated its effect on different
approximately equal to the martensite start temperature, Ms. functional parameters. They found that the austenite finish
A better understanding of effects of test temperature on temperature, Af, increases and the upper plateau stress
fatigue resistance of Nitinol in mixed-mode and martensitic decreases by increasing the nitrogen ion content and no
metallurgical phases still requires conducting more compre- further superelastic behavior of the material was observed.
hensive investigations. In strain-based fatigue analysis, higher content of nitrogen
Fatigue cracks typically start from impurities on the sur- ions leads to longer fatigue lives in high-cycle regime,
face or close to the surface in metallic materials; therefore, whereas no significant effect was observed in low-cycle
surface condition may affect significantly the crack initiation fatigue regime (i.e. εa44%). Longer fatigue lives in high-cycle
life. Generally, Nitinol products are available in various sur- regime can be a result of the increased Af due to the presence
face finishes such as eletropolished, mechanically polished, of nitrogen ions.. However, these conclusions may not be
etched, dark and black oxide. Pelton et al. (2013) have valid in the stress-based fatigue analysis, where the stress
summarized the results from different types of surface finish response of the material significantly drops by adding the
(i.e. black oxide and bright/mechanically polished surface) for nitrogen ion content.
a similar composition and treatment of Nitinol as shown in Since most of the experiments are performed on wire
Fig. 21. The heat treated wires have an approximately a specimens, the effects of the wire diameter on the fatigue
300 nm thick layer of oxide, so called black oxide. A bright behavior of Nitinol alloys are also investigated in several
surface wire is mechanically polished after drawing (Pelton studies (Norwich and Fasching, 2009; Pruett et al., 1997;
et al., 2013). Mechanical polishing results in removing the Sawaguchi et al., 2003). Sawaguchi et al. (2003) studied the
black oxide surface as well as drawing marks, and therefore, effects of the wire diameter on the fatigue behavior of the
provides a smoother surface finish. As seen in Fig. 21, the superelastic Nitinol alloys and found slightly lower fatigue
effect of the surface finish is more significant in the high- lives for larger wire diameters. Norwich and Fasching (2009)
cycle fatigue than low-cycle fatigue regime. The bright sur- also reported the effects of Nitinol wire diameter on the
face (mechanically polished) specimens show greater fatigue fatigue behavior as presented in Fig. 22. As can be seen from
strength in the high-cycle fatigue regime as compared to this figure, greater fatigue strengths resulted from smaller
specimens with black surface finish. Beneficial effects of finer diameter wires of superelastic Nitinol alloys. They also
surface finishing techniques such as electropolishing and proposed a power-law equation to illustrate the low-cycle
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 247

cycle at body specific locations are displacement or strain-


based cyclic deformations histories. Thus, the emphasis to
date in the fatigue analysis of Nitinol has been on strain (i.e.
displacement)-controlled testing and the strain-life approach.
There is also a need for a more thorough understanding of
the stress-life fatigue behavior of Nitinol as its application is
extended to bio-implants and other fields of engineering that
may require a force carrying tolerance.
In order to obtain stress-life fatigue data, different test
techniques can be employed. In other words, force (stress)-
controlled tests can be conducted using a classical rotating
Fig. 22 – Effects of wire diameter on the fatigue behavior of beam machine or a closed-loop uniaxial fatigue test setup,
the superelastic Nitinol (Norwich and Fasching, 2009). where both methods require dogbone specimens. In rotating
beam testing technique, the specimen can be under a four-
point bending moment at the critical section and a motor
fatigue behavior of dentistry instruments made of Nitinol. rotates the specimen about its longitudinal axis at a desired
Additional results are required to generalize the effects of frequency. Similar to the rotating-bending test, in rotating
specimen size on the fatigue behavior of Nitinol alloys. beam test the outer layer of the material experiences the
However, lower fatigue lives for larger specimens may be maximum stress/strain. Therefore, the surface finish is a very
expected because more impurities typically exist in larger important factor, governing the fatigue behavior in these
cross-sectional areas. This important factor should be con- tests. However, unlike the rotary-bending test that is a
sidered and accounted for when a larger instrument is strain-controlled test setup, the rotating beam fatigue test is
designed based on the fatigue data from Nitinol wires. a force (i.e. stress)-controlled test setup. Due to the distribu-
Another important parameter affecting the fatigue beha- tion of the stress and consequently significant contribution of
vior of Nitinol alloys is the rotational speed (i.e. frequency) of the surface quality to the fatigue life in rotating beam
the rotating-bending test. This parameter is considered technique, this method may not capture the effects of sub-
mainly for the endodontic files in the literature (Gabel et al., surface impurities and other microstructural aspects within
1999; Gambarini, 2001; MuKhlif and Al-Azzawi, 2014). Results the material. In this regard, uniaxial fatigue testing by a
from a study by Gabel et al. (1999) indicate a significant closed-loop machine may be more appropriate.
increase in failure probability (  4  ) by increasing the rota- In contrast to rotating beam testing, the stress distribution
tional speed by a factor of two. Considerable reductions in is uniform over the cross section of the specimen in closed-
strain-based fatigue lives of Nitinol specimens that were loop uniaxial fatigue machines. Therefore, the effects of
tested in un-controlled air environment, with an increased subsurface impurities and microstructural features will be
test frequency were reported (Eggeler et al., 2004; Tobushi accounted for in fatigue behavior assessment in such an
et al., 1997). However, no such differences in fatigue lives experiment. However, more challenges may be involved with
were observed for specimens under controlled environments, fatigue testing of Nitinol alloys using uniaxial fatigue
i.e. maintaining constant temperatures using a water bath machines. For instance, test frequency is an important factor
(Tobushi et al., 1997). These observations suggest that in uniaxial fatigue testing of larger samples due to the heat
frequency-induced hysteresis heating is responsible for generated during straining and un-straining (Lim and
shorter fatigue lives of Nitinol in higher frequencies. As McDowell, 1999). Greater test frequencies may cause hyster-
stated previously, the relative test temperature may greatly esis heat generation in the specimen that may significantly
affect the fatigue behavior of Nitinol alloys. influence the mechanical behavior of the material leading to
Zurbitu et al. (2010) investigated the effects of very high altered fatigue behavior. Thus, careful attention should be
strain rates (i.e. impact) on fatigue behavior of superelastic taken to test frequency in uniaxial fatigue testing of dogbone
Nitinol alloys. They found the critical stress for inducing the specimens to maintain a constant temperature in the mate-
martensitic phase (i.e. stress plateau) reduced with cyclic rial and avoid undesired effects of self-heating.
impact forces. They also reported more reduction in the Appropriate gripping of the sample is another challenge
critical stress with increasing strain rates. Additionally, fati- since slip in the grips was observed in (McNaney et al., 2003)
gue resistance of the rotary instruments after clinical usage due to the stress-induced phase transformation in Nitinol.
has been investigated in a study of Nitinol by Bahia and Furthermore, machining Nitinol rods to make dogbone fati-
Buono (2005). Their results indicate that the fatigue resistance gue specimens also requires careful attention since the
of such rotary instrument decreased by previous usage therby material is a hard, intermetallic compound and its mechan-
implying that fatigue damage of Nitinol is progressive and ical properties are highly sensitive to processing (Wu et al.,
irreversible in nature (Bahia and Buono, 2005). 1999). Thus, coolant must be used generously while machin-
ing Nitinol specimens.
In order to have a set of comparable test results, all the
7. Future works and challenges experiments should be conducted in similar test conditions. In
other words, the test temperature and the difference between
In most biomedical applications of Nitinol alloys, such as the test temperature and the austenite finish temperature, Af,
endovascular stents, abdominal aortic devices, etc., the duty should be held constant (Pelton et al., 2013). Unlike the wire
248 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

specimens in the rotating-bending fatigue test, performing


uniaxial strain-controlled fatigue tests in a controlled environ-
ment, such as a constant temperature bath, may not be easy.
As with most engineering alloys, the fatigue of Nitinol is
highly dependent on the microstructural aspects of the
material, such as impurities, inclusion, voids, grain size,
orientation, etc. More comprehensive research is indeed
essential to obtain the effects of such microstructural fea-
tures on the fatigue behavior. Understanding the effects of
microstructural impurities on fatigue behavior of Nitinol will
help the development of microstructure-sensitive fatigue
models for these alloys. A microstructure-sensitive fatigue
model can be also used to estimate the scatter in fatigue data
of Nitinol alloys at a certain strain/stress amplitude. In order
to validate/calibrate a comprehensive microstructure-
sensitive fatigue model, an extensive amount of experimen-
tal and computational research is needed, involving Nitinol
Fig. 23 – Two schematic variable amplitude load histories: (a)
specimens with different levels of impurities, to understand
large-amplitude cycle followed by a small-amplitude and (b)
the synergistic effects of multiple microstructural features.
small-amplitude cycle followed by a large-amplitude cycle
Due to the dependency of the fatigue behavior of super-
and their corresponding material response.
elastic Nitinol alloys to the stress-induced phase transforma-
tion phenomenon, the classical fatigue theories for mean
stress/strain corrections, such as Goodman relationship, should the load history in Fig. 23(b), the large cycle occurs after
not be used. Currently, there is no mean stress/strain correction unloading of a small cycle. Basically, both histories are
model for Nitinol alloys in the open literature. Additionally, comprised of two deformation cycles, i.e. small and large
there is a lack of comprehensive mean stress/strain data to amplitudes, but in different sequence. Using rainflow cycle
understand the effects of mean stress/strain in the different counting technique, similar cycles, mean stresses/strains and
metallurgical phases of Nitinol. Thus, more experimental data stress/strain amplitudes will be counted for both histories.
including various combinations of alternating and mean stres- Therefore, the same fatigue lives are expected for both
ses/strains should be generated to better understand the effects histories in Fig. 23. However, the smaller counted cycle in
of mean stress/strain on fatigue behavior of these alloys. It is Fig. 23(a) is on the lower stress plateau (i.e. unloading),
worth mentioning that developing a mean strain/stress correc- whereas that in Fig. 23(b) is on the upper stress plateau (i.e.
tion model for Nitinol is important since the material exhibits a loading). The fatigue damage caused by the small counted
pronounced asymmetry in tensile and compressive stress/ cycle may be different for the load history due to the different
strain response. As mentioned earlier, due to the tension- volume fraction of the martensitic and austenitic phases in
compression asymmetry in deformation, there would be either loading and unloading reversals. Therefore, the cycle count-
a compressive mean stress (in strain-controlled fatigue tests) or ing technique must be deformation path-sensitive to capture
a tensile mean strain (in force-controlled fatigue tests), specifi- the volume fraction of each phase for each counted cycle.
cally in low-cycle fatigue regime (i.e. large deformation). There- Such a cycle counting method for shape memory alloys are
fore, developing a mean strain/stress model for Nitinol may yet to be developed.
facilitate better interpretations of fatigue test data as mean Under variable amplitude state of loading, there is also a
strain/stress responses typically exist in presence of phase need for a cumulative damage rule to calculate the overall
transformation, even in fully reversed load/deformation tests. damage and total fatigue life. Due to the dependency of the
Although the more realistic deformation conditions in fatigue behavior of Nitinol alloys to the microstructural phase
many applications is variable amplitude and/or multiaxial transformation, the sequence effects on the phase transfor-
there is no comprehensive study in the existing literature mation and fatigue behavior should be clearly understood
concerning the fatigue behavior of Nitinol alloys under vari- and incorporated into the cumulative damage rule. For
able amplitude or multiaxial state of loading. A cycle count- example Fig. 24 displays two histories that consist of one
ing method and a cumulative damage rule are the two sequence with a small-amplitude block followed by a large-
inevitable elements of fatigue analysis under variable ampli- amplitude block (i.e. Fig. 24(a)) and the other sequence with a
tude loading. Similar to other materials, a standard cycle large-amplitude block followed by a small-amplitude block
counting method such as the method introduced in ASTM (i.e. Fig. 24(b)). The material under the small-amplitude block
E1049 Standard (ASTM E1049-85, 2011), more commonly is an austenitic Nitinol, where a significant portion of the
called the “rainflow” technique, is needed to identify material has transformed to the martensitic phase for the
deformation-time reversals within a variable amplitude larger amplitude block (see Fig. 24(c)). As the fatigue behavior
block. and mechanical properties will be different for each material,
To illustrate the challenge of cycle counting for Nitinol, the deformation sequence matters and the linear cumulative
two different strain histories are designed as displayed in damage rule presumably may not be appropriate for damage
Fig. 23. The history in Fig. 23(a) has a large cycle in deforma- accumulation in this material. Some studies have suggested
tion path followed by a small cycle in unloading, whereas in the linear damage rule and an appropriate strain–stress
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 249

Fig. 24 – Load sequence consideration for variable amplitude fatigue analysis of Nitinol alloys.

damage model such as Smith–Watson–Topper (Colin and loading, the multiaxial load path non-proportionality, alternation
Fatemi, 2010) or Fatemi–Socie (Shamsaei et al., 2011, 2010). and sequence effects on deformation, fatigue and phase trans-
The stress term in such damage models will capture the formation are to be also studied in the future.
sequence effects; therefore, fatigue life predictions using One of the major challenges in multiaxial fatigue experi-
linear damage rule are reported to be in good agreement mentation of Nitinol is to manufacture tubular specimens.
with experimentally observed fatigue lives for low-high, high- Thin-walled tubular specimens are appropriate for multiaxial
low, and periodic overloading (Colin and Fatemi, 2010) as well fatigue testing since there is minimal shear stress gradient
as random and multiaxial loading (Colin and Fatemi, 2010; under torsion in these specimens (ASTM E2207-08(2013)e1,
Shamsaei et al., 2011, 2010). However, such an approach may 2008). However, machining and manufacturing Nitinol to
not work for Nitinol because its deformation is pseudo-elastic thin-walled tubular specimens is very difficult (Wu et al.,
and no history effects will be reflected in the stress response. 1999) and careful attention should be taken to keep the
Therefore, further investigations to understand the fatigue material properties unchanged.
behavior of Nitinol under variable amplitude deformations is In order to extend the application of Nitinol in biomedical
essential to develop/calibrate a proper cumulative damage industry even further, porous Nitinol parts may be fabricated
model that accurately captures the sequence and phase to alter the mechanical properties, such as modulus of
transformation effects on fatigue life. In addition, under elasticity, to match with those of biological materials such as
variable amplitude forces/deformations, mean strains/stres- bone (Elahinia et al., 2012). More importantly, additive manu-
ses may play a key role on the fatigue behavior, indicating a facturing techniques can be used to optimize the pore size and
need for a mean strain/stress correction model to facilitate distribution and build patient-specific implants. Additive man-
variable amplitude fatigue data analysis. ufacturing also provides the opportunity to fabricate parts
The multiaxial loading in critical elements of components with very complex geometries. Therefore, there are many
and structures arises not only from multidirectional loading, ongoing research projects (Haberland et al., 2014, 2013) to
stress concentrations, or residual stresses, but also from the facilitate and optimize the additive manufacturing process
underlying microstructural heterogeneities. Multiaxial fatigue for Nitinol components in biomedical applications.
behavior of Nitinol alloys is more complex than other metallic Employing additively manufactured Nitinol parts for med-
materials due to their dependency on the loading path (i.e. in- ical applications involves additional challenges for fatigue life
phase versus out-of-phase) and on the microstructural phase modeling and prediction. These challenges arise from several
transformation behavior that is also a path dependent phenom- process parameters involved in the additive manufacturing
enon. A systematic study to understand the multiaxial fatigue process such as laser power, laser transverse speed, powder
behavior of Nitinol alloys should address several important feed rate, layer thickness, scanning pattern, building orienta-
factors including damage mechanism, deformation behavior, tion and more (Shamsaei et al., 2015). Various combinations
and damage parameters (Fatemi and Shamsaei, 2011), as well of these parameters will cause variations in the thermal
as the load path effects on fatigue behavior and phase history, in particular cooling rate and thermal gradients,
transformation. during fabrication, which may affect microstructural features
Understanding the failure mechanism (i.e. shear or tensile and mechanical properties of the product (Shamsaei et al.,
failure mode) of the material provides insights for developing an 2015; Thompson et al., 2015). The effect of thermal history on
appropriate damage parameter. Cyclic deformation may also the mechanical and microstructural properties in general and
influence the multiaxial fatigue behavior as some materials fatigue behavior in particular is even more noticeable for
exhibit an additional cyclic hardening under non-proportional additively manufactured Nitinol parts due to the strong
multiaxial loads. Therefore, cyclic response of Nitinol to propor- sensitivity of the transformation temperatures to the thermal
tional and non-proportional multiaxial load paths needs to be history during (Haberland et al., 2014) and after manufactur-
well understood and amount of possible hardening/softening ing (i.e. heat treatment) (Pelton et al., 2000). This may be
should be determined for multiaxial fatigue modeling. Thus, a explained by re-melting of previously deposited layers and
comprehensive study of multiaxial and load non-proportionality frequent reheating cycles of the material during part fabrica-
effects on deformation, fatigue, and phase transformation is tion, affecting the thermal history by aging and tempering,
needed. Under a more general variable amplitude multiaxial and resulting in potential changes in characteristic
250 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

temperatures. Moreover, presence of residual stresses and 5- For a wide range of strain amplitudes (  2–7%), which
unwanted defects (i.e. pores and particles) inherent to addi- corresponds to the stress plateau region in the stress–
tive manufacturing (Rangaswamy et al., 2005) may influence strain diagram for the superelastic Nitinol, the fatigue
the fatigue lifetime of these products. Geometry, laser pro- life is nearly constant in strain-based fatigue analysis.
cess parameters, and scanning pattern may change the Therefore, the stress component in the stress plateau
amount and distribution of the residual stress within the region seems to have more dominant effects than the
additively manufactured part (Rangaswamy et al., 2005). strain component on the fatigue behavior. Due to this
Despite the increasing interest in using additive manufac- reason and the lack of stress-based fatigue data in the
turing techniques to produce Nitinol parts and the impor- literature, there is a need for more stress-life fatigue
tance of fatigue analysis for these products, the current analysis of Nitinol in order to obtain a better under-
literature suffers from the lack of fatigue data for additively standing of the fatigue behavior in force-controlled
manufactured Nitinol. Therefore, there is a critical need for conditions.
systematic studies to determine the synergistic effects of 6- The strain-life fatigue data for superelastic Nitinol alloys
various additive manufacturing process (e.g. laser power, can be divided into different linear sections in the log–log
powder feed rate, etc.) and design (e.g. layer thickness, plot, which correspond to different fatigue regimes.
scanning pattern, building orientation, etc.) parameters on These fatigue regimes may be generally classified as very
thermal history that influence the microstructural evolution low-cycle, low-cycle, mid-cycle and high-cycle regimes.
during solidification. As a result, distinct microstructural Each regime may correspond to different regions in the
properties, porosity, residual stresses, and characteristic stress–strain response of the material such as the auste-
temperatures may be obtained (Shamsaei et al., 2015), ulti- nite region, transition region, stress plateau (transforma-
mately affecting the mechanical properties and fatigue beha- tion region) and fully martensitic region.
vior of additively manufactured Nitinol. 7- Test temperature and transformation temperatures of
different compositions of Nitinol can highly influence the
fatigue resistance of the material. In strain-life fatigue
8. Conclusions analysis of Nitinol alloys and for test temperatures above
austenite finish level, where the material is superelastic,
A review on the fatigue behavior of Nitinol alloy was pre- the fatigue resistance decreases by increasing the test
sented and some important factors that influence the fatigue temperature. However, a reverse behavior is observed in
resistance were discussed. Based on the current state of the stress-life analysis, as the fatigue resistance of Nitinol
studies in the field, the following conclusions are presented: increases with increasing test temperature. This increase
can be mainly attributed to the increase in the stress
1- Fatigue behavior of Nitinol is more complex than other plateau of Nitinol in higher temperatures.
metallic materials. These complexities arise mainly from 8- Both uniaxial stress-life and strain-life fatigue analysis
the unique properties of Nitinol such as superelasticity of Nitinol alloys indicate a multilinear relationship
and shape memory characteristics. From the materials between the stress/strain amplitude and the fatigue life
point of view, the complexities can be related to the in a log–log plot. However, not such a conclusion can be
phase transformation and the structural phase of the made under torsion as there are only a few studies
material. concerning the torsional fatigue behavior of Nitinol in
2- In majority of current applications, such as biomedical, the literature.
superelastic Nitinol is mainly employed where crack 9- In contrast with most metals, the presence of tensile
initiation dominates the fatigue behavior. In today’s appli- mean strains may lead to a greater fatigue life for super-
cations of Nitinol, fatigue crack growth stage is relatively elastic Nitinol alloys. Experimental results in the litera-
short and may not contribute significantly to the total ture have revealed that for a certain range of strains,
fatigue life. However, the fatigue crack growth stage may tensile mean strains may enhance the fatigue strength of
be longer for Nitinol components such as those used in Nitinol alloys due to the formation of stress-induced
aerospace industry. martensite. The same behavior may exist for Nitinol
3- Better understanding of the deformation behavior of under non-zero torsional mean strains.
Nitinol is essential for an accurate fatigue evaluation. 10- Functionality of the Nitinol alloys under cyclic loads
The deformation models for Nitinol alloys under cyclic decreases. However, the functional parameter reaches a
uniaxial or multiaxial loading should include the effects stable state after a few numbers of cycles of loading (e.g.
of the phase transformation under various combinations 100–150 cycles). Such reductions in some functional
of mechanical and/or thermal loading conditions. properties of Nitinol may be attributed to the activation
4- Although greater fatigue resistance for thermally mar- of slip systems and/or formation of localized residual
tensitic Nitinol is observed typically in strain-based martensitic phase in the material.
approaches, superior fatigue resistance for superelastic 11- The transformation temperatures of Nitinol alloys are
Nitinol alloys can be expected in stress-based fatigue subjected to change under thermal cycling, with or without
analysis. Therefore, the fatigue analysis approach for applied external mechanical forces. This change primarily
different applications should be selected based on the appears as a reduction in transformation temperatures. A
loading or deformation condition (i.e. force-controlled or stable state of transformation temperatures can be
strain-controlled) of the real-world application. obtained after a limited number of thermal cycling.
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 251

12- Fatigue resistance of the Nitinol is highly dependent on Casati, R., Passaretti, F., Tuissi, A., 2011. Effect of electrical heating
the microstructural factors, such as impurities, inclu- conditions on functional fatigue of thin NiTi wire for shape
sions, grain size, and grain orientations. As suggested in memory actuators. Procedia Eng. 10, 3423–3428.
the literature, the fatigue resistance of Nitinol correlates Cheng, C.P., Wilson, N.M., Hallett, R.L., Herfkens, R.J., Taylor, C.A.,
2006. In vivo MR angiographic quantification of axial and
well with the dimension of the largest inclusion, and
twisting deformations of the superficial femoral artery
therefore, can be improved by microstructural enhance-
resulting from maximum hip and knee Flexion. J. Vasc. Interv.
ments such as reducing the grain size or impurity
Radiol. 17, 979–987.
contents. Coda, A., 2014. Progress on the correlation between inclusions
13- In a non-controlled test environment, higher test fre- and fatigue behavior in NiTi shape memory alloys for
quencies, which result in an increase in the test tem- biomedical applications: refinement of the statistical
perature, may cause a reduction in the strain-based approach. In: Presented at the International Conference on
fatigue lives of Nitinol wires. However, in controlled test Shape Memory and Superelastic Technologies (SMST) (May
environments (i.e. constant-temperature tests), effects of 12–16, 2014), ASM.
test frequency on the fatigue of Nitinol wires are not Coffin, L.F., Tavernelli, J.F., 1959. The cyclic straining and fatigue of
significant. metals. Trans. Metall. Soc. AIME 215, 794–807.
Colin, J., Fatemi, A., 2010. Variable amplitude cyclic deformation
14- Based on the limited data in the literature, various
and fatigue behaviour of stainless steel 304L including step,
parameters such as specimen dimensions, surface finish,
periodic, and random loadings. Fatigue Fract. Eng. Mater.
processing and treatments may affect the fatigue resis- Struct. 33, 205–220.
tance of Nitinol alloys. Therefore, more studies are Condorelli, G.G., Bonaccorso, A., Smecca, E., Schäfer, E.,
needed to obtain the effects of these parameters on the Cantatore, G., Tripi, T.R., 2010. Improvement of the fatigue
fatigue behavior of these alloys. resistance of NiTi endodontic files by surface and bulk
modifications. Int. Endod. J. 43, 866–873.
Delville, R., Malard, B., Pilch, J., Sittner, P., Schryvers, D., 2011.
Transmission electron microscopy investigation of dislocation
slip during superelastic cycling of Ni–Ti wires. Int. J. Plast. 27,
r e f e r e n c e s
282–297.
Dowling, N.E., 2009. Mean stress effects in strain-life fatigue.
Fatigue Fract. Eng. Mater. Struct. 32, 1004–1019.
Allie, D.E., Hebert, C.J., Walker, C.M., 2004. Nitinol stent fractures Duerig, T.W., Pelton, A., Stöckel, D., 1999. An overview of Nitinol
in the SFA. Endovasc. Today 7, 22–34. medical applications. Mater. Sci. Eng., A 273, 149–160.
ASTM E466-07, 2007. Practice for Conducting Force Controlled Eggeler, G., Hornbogen, E., Yawny, A., Heckmann, A., Wagner, M.,
Constant Amplitude Axial Fatigue Tests of Metallic Materials 2004. Structural and functional fatigue of NiTi shape memory
(No. E466-07), E08. ASTM International, West Conshohocken, PA. alloys. Mater. Sci. Eng., A 378, 24–33.
ASTM E606/E606M-12, 2012. Test Method for Strain-controlled El Feninat, F., Laroche, G., Fiset, M., Mantovani, D., 2002. Shape
Fatigue Testing (No. E606-/E606M-12), E08. ASTM memory materials for biomedical applications. Adv. Eng.
International, West Conshohocken, PA. Mater. 4, 91.
ASTM E1049-85, 2011. Practices for Cycle Counting in Fatigue Elahinia, M., 2015. Shape Memory Alloy Actuators: Design,
Analysis (No. E1049-85(2011)e1), E08. ASTM International. Fabrication and Experimental Evaluation. John Wiley & Sons
ASTM E2207-08(2013)e1, 2008. Practice for Strain-controlled Axial- Inc..
torsional Fatigue Testing with Thin-walled Tubular Specimens Elahinia, M.H., Hashemi, M., Tabesh, M., Bhaduri, S.B., 2012.
(No. E2207-08), E08. ASTM International, West Conshohocken, PA. Manufacturing and processing of NiTi implants: a review.
ASTM E2948-14, 2014. Standard Test Method for Conducting Prog. Mater. Sci. 57, 911–946.
Rotating Bending Fatigue Tests of Solid Round Fine Wire (No. Fatemi, A., Shamsaei, N., 2011. Multiaxial fatigue: an overview
E2948-14), E08. ASTM International, West Conshohocken, PA. and some approximation models for life estimation. Int. J.
Bahia, M.G.A., Buono, V.T.L., 2005. Decrease in the fatigue Fatigue 33, 948–958.
resistance of nickel–titanium rotary instruments after clinical Fatemi, A., Socie, D.F., 1988. A critical plane approach to
use in curved root canals. Oral Surg. Oral Med. Oral Pathol. multiaxial fatigue damage including out-of-phase loading.
Oral Radiol. Endodontol. 100, 249–255. Fatigue Fract. Eng. Mater. Struct. 11, 149–165.
Basquin, O.H., 1910. The exponential law of endurance tests. Proc. Frotscher, M., Nörtershäuser, P., Somsen, C., Neuking, K.,
ASTM, 625–630. Böckmann, R., Eggeler, G., 2009. Microstructure and structural
Benafan, O., Noebe, R.D., Padula II, S.A., Brown, D.W., Vogel, S., fatigue of ultra-fine grained NiTi-stents. Mater. Sci. Eng., A
Vaidyanathan, R., 2014. Thermomechanical cycling of a NiTi 503, 96–98.
shape memory alloy-macroscopic response and Frotscher, M., Wu, S., Simon, T., Somsen, C., Dlouhy, A., Eggeler,
microstructural evolution. Int. J. Plast. 56, 99–118. G., 2011. Elementary deformation and damage mechanisms
Bertacchini, O.W., Lagoudas, D.C., Patoor, E., 2009. during fatigue of pseudoelastic NiTi microstents. Adv. Eng.
Thermomechanical transformation fatigue of TiNiCu SMA Mater. 13, B181–B186.
actuators under a corrosive environment—Part I: Fumagalli, L., Butera, F., Coda, A., 2009. SmartFlexs NiTi wires for
Experimental results. Int. J. Fatigue 31, 1571–1578. shape memory actuators. J. Mater. Eng. Perform. 18, 691–695.
Bigeon, M.J., Morin, M., 1996. Thermomechanical study of the Gabel, W.P., Hoen, M., Robert Steiman, H., Pink, F.E., Dietz, R.,
stress assisted two way memory effect fatigue in TiNi and 1999. Effect of rotational speed on nickel–titanium file
CuZnAl wires. Scr. Mater. 35, 1373–1378. distortion. J. Endod. 25, 752–754.
Carpenter, B., Lyons, J., 2001. EO-1 Technology Validation Report: Gall, K., Sehitoglu, H., Chumlyakov, Y.I., Kireeva, I.V., 1998.
Lightweight Flexible Solar Array Experiment. NASAGSFC (last Pseudoelastic cyclic stress–strain response of over-aged single
update August 8). crystal Ti–50.8 at% Ni. Scr. Mater. 40, 7–12.
252 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

Gall, K., Sehitoglu, H., 1999. The role of texture in tension– Lim, T.J., McDowell, D.L., 1999. Mechanical behavior of an Ni–Ti
compression asymmetry in polycrystalline NiTi. Int. J. Plast. shape memory alloy under axial-torsional proportional and
15, 69–92. nonproportional loading. J. Eng. Mater. Technol. 121, 9–18.
Gambarini, G., 2001. Cyclic fatigue of nickel–titanium rotary Mahtabi, M.J., Shamsaei, N., 2015. Multiaxial fatigue modeling for
instruments after clinical use with low-and high-torque Nitinol shape memory alloys under in-phase loading. under
endodontic motors. J. Endod. 27, 772–774. review.
Godard, O.J., Lagoudas, M.Z., Lagoudas, D.C., 2003. Design of space Maletta, C., Sgambitterra, E., Furgiuele, F., Casati, R., Tuissi, A.,
systems using shape memory alloys. In: Smart Structures and 2012. Fatigue of pseudoelastic NiTi within the stress-induced
Materials. International Society for Optics and Photonics, transformation regime: a modified Coffin–Manson approach.
pp. 545–558. Smart Mater. Struct. 21, 112001.
Gollerthan, S., Young, M.L., Baruj, A., Frenzel, J., Schmahl, W.W., Maletta, C., Sgambitterra, E., Furgiuele, F., Casati, R., Tuissi, A.,
Eggeler, G., 2009. Fracture mechanics and microstructure in
2014. Fatigue properties of a pseudoelastic NiTi alloy: Strain
NiTi shape memory alloys. Acta Mater. 57, 1015–1025.
ratcheting and hysteresis under cyclic tensile loading. Int. J.
Haberland, C., Elahinia, M., Walker, J., Meier, H., Frenzel, J., 2013.
Fatigue 66, 78–85.
Additive Manufacturing of Shape Memory Devices and
Manson, S.S., Dolan, T.J., 1966. Thermal stress and low cycle
Pseudoelastic Components.
fatigue. J. Appl. Mech. 33, 957.
Haberland, C., Elahinia, M., Walker, J.M., Meier, H., Frenzel, J.,
Mao, S., Han, X., Wu, M.H., Zhang, Z., Hao, F., Liu, D., Zhang, Y.,
2014. On the development of high quality NiTi shape memory
Hou, B., 2006. Effect of cyclic loading on apparent Young’s
and pseudoelastic parts by additive manufacturing. Smart
modulus and critical stress in nano-subgrained superelastic
Mater. Struct. 23, 104002.
Hartl, D.J., Lagoudas, D.C., 2007. Aerospace applications of shape NiTi shape memory alloys. Mater. Trans. 47, 735.
memory alloys. Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng. McKelvey, A.L., Ritchie, R.O., 2001. Fatigue-crack growth behavior
221, 535–552. in the superelastic and shape-memory alloy Nitinol. Metall.
He, X., Zhao, L., Wang, X., Zhang, R., Li, M., 2006. Transformation Mater. Trans. A 32, 731–743.
behaviour with thermal cycling in Ti50Ni43Cu7 shape McNaney, J.M., Imbeni, V., Jung, Y., Papadopoulos, P., Ritchie, R.O.,
memory alloy. Mater. Sci. Eng., A 427, 327–330. 2003. An experimental study of the superelastic effect in a
Holtz, R.L., Sadananda, K., Imam, M.A., 1999. Fatigue thresholds shape-memory Nitinol alloy under biaxial loading. Mech.
of Ni–Ti alloy near the shape memory transition temperature. Mater. 35, 969–986.
Int. J. Fatigue 21 (Suppl. 1), S137–S145. McNichols, J.L., Brookes, P.C., Cory, J.S., 1981. NiTi fatigue
Huett, B., Willey, C., 2000. Design and development of miniature behavior. J. Appl. Phys. 52, 7442–7444.
mechanisms for small spacecraft. AIAA/USU Conf. Small Melton, K.N., Mercier, O., 1979a. Fatigue of NiTi thermoelastic
Satell.. martensites. Acta Metall. 27, 137–144.
Ince, A., Glinka, G., 2011. A modification of Morrow and Smith– Melton, K.N., Mercier, O., 1979b. Fatigue of NiTi and CuZnAl shape
Watson–Topper mean stress correction models. Fatigue Fract. memory alloys. Strength Met. Alloys, 1243–1248.
Eng. Mater. Struct. 34, 854–867. Mertmann, M., Vergani, G., 2008. Design and application of shape
Jensen, D.M., 2005. Biaxial Fatigue Behavior of NiTi Shape Memory memory actuators. Eur. Phys. J. Spec. Top. 158, 221–230.
Alloy. Air Force Institute of Technology, Wright-Patterson Air Mitchell, M.R., 1977. Review of the mechanical properties of cast
Force Base, Ohio. steels with emphasis on fatigue behavior and the influence of
Kang, G., Kan, Q., Yu, C., Song, D., Liu, Y., 2012. Whole-life microdiscontinuities. J. Eng. Mater. Technol. 99, 329–343.
transformation ratchetting and fatigue of super-elastic NiTi Mitchell, M.R., 1979. A Unified Predictive Technique for the
Alloy under uniaxial stress-controlled cyclic loading. Mater. Fatigue Resistance of Cast Ferrous-based Metals and High
Sci. Eng., A 535, 228–234. Hardness Wrought Steels (SAE Technical Paper No. 790890).
Kim, Y., 2002. Fatigue properties of the Ti–Ni base shape memory SAE International, Warrendale, PA.
alloy wire. Mater. Trans. 43, 1703–1706. Mitchell, M.R., 1996. Fundamentals of Modern Fatigue Analysis
Kim, Y.S., Miyazaki, S., 1997. Fatigue properties of Ti–50.9 at% Ni for Design.
shape memory wires. In: Pelton, A.R., Hodgson Russell, S.M., Miyazaki, S., 1990. Thermal and stress cycling effects and fatigue
Duerig, T. (Eds.), Proc. Second Int. Conf. Shape Mem.
properties of Ni–Ti alloys. Eng. Asp. Shape Mem. Alloys, 394.
Superelastic Technol. SMST Pac. Grove MIAS SMST-97, Miyazaki, S., Igo, Y., Otsuka, K., 1986a. Effect of thermal cycling on
pp. 473–477.
the transformation temperatures of Ti–Ni alloys. Acta Metall.
Kollerov, M., Lukina, E., Gusev, D., Mason, P., Wagstaff, P., 2013.
34, 2045–2051.
Influence of the structure on the strain-controlled fatigue of
Miyazaki, S., Imai, T., Igo, Y., Otsuka, K., 1986b. Effect of cyclic
Nitinol. Mater. Sci. Forum 738-739, 316–320.
deformation on the pseudoelasticity characteristics of Ti–Ni
Kugler, C., Matson, D., Perry, K., 2000. Non-zero mean fatigue test
alloys. Metall. Trans. A 17, 115–120.
protocol for NiTi. In: Russell, S.M., Pelton, A.R. (Eds.), SMST-
Miyazaki, S., Mizukoshi, K., Ueki, T., Sakuma, T., Liu, Y., 1999.
2000: Proceedings of the International Conference on Shape
Fatigue life of Ti–50 at% Ni and Ti–40Ni–10Cu (at%) shape
Memory and Superelastic Technologies, Pacific Grove, CA,
memory alloy wires. Mater. Sci. Eng., A 273–275, 658–663.
USA, p. 409.
Miyazaki, S., Sugaya, Y., Otsuka, K., 1988. Effects of various factors
Lagoudas, D.C., Li, C., Miller, D.A., Rong, L., 2000.
Thermomechanical Transformation Fatigue of SMA Actuators, on fatigue life of Ti–Ni alloys. In: Proceedings of the MRS
pp. 420–429. International Meeting on Advanced Materials. pp. 251–256.
Lagoudas, D.C., Miller, D.A., Rong, L., Kumar, P.K., 2009. Mohd Jani, J., Leary, M., Subic, A., Gibson, M.A., 2014. A review of
Thermomechanical fatigue of shape memory alloys. Smart shape memory alloy research, applications and opportunities.
Mater. Struct. 18, 085021. Mater. Des. 56, 1078–1113.
Launey, M., Robertson, S.W., Vien, L., Senthilnathan, K., Morgan, N.B., Painter, J., Moffat, A., 2004. Mean strain effects and
Chintapalli, P., Pelton, A.R., 2014. Influence of microstructural microstructural observations during in vitro fatigue testing of
purity on the bending fatigue behavior of VAR-melted NiTi. In: SMST-2003. Proceedings of the International
superelastic Nitinol. J. Mech. Behav. Biomed. Mater. 34, Conference on Shape Memory and Superelastic Technologies.
181–186. SMST Society, Pacific Grove, CA, pp. 303–310.
journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254 253

Moumni, Z., Herpen, A.V., Riberty, P., 2005. Fatigue analysis of Predki, W., Klönne, M., Knopik, A., 2006. Cyclic torsional loading
shape memory alloys: energy approach. Smart Mater. Struct. of pseudoelastic NiTi shape memory alloys: damping and
14, S287. fatigue failure. Mater. Sci. Eng., A 417, 182–189.
Moumni, Z., Zaki, W., Maitournam, H., 2009. Cyclic behavior and Pruett, J.P., Clement, D.J., Carnes Jr., D.L., 1997. Cyclic fatigue
energy approach to the fatigue of shape memory alloys. J. testing of nickel–titanium endodontic instruments. J. Endod.
Mech. Mater. Struct. 4, 395–411. 23, 77–85.
MuKhlif, T.A., Al-Azzawi, A.-K.J., 2014. The effect of curvature Rangaswamy, P., Griffith, M.L., Prime, M.B., Holden, T.M., Rogge, R.
angle and rotational speed on the cyclic fatigue of three types B., Edwards, J.M., Sebring, R.J., 2005. Residual stresses in
of rotary instrument (In vitro): comparative study. J. Baghdad LENSs components using neutron diffraction and contour
Coll. Dent. 25, 38–42. method A, Measurement and Interpretation of Internal/
Nayan, N., Roy, D., Buravalla, V., Ramamurty, U., 2008. Unnotched Residual Stresses. Mater. Sci. Eng. 399, 72–83.
fatigue behavior of an austenitic Ni–Ti shape memory alloy. Reinoehl, M., Bradley, D., Bouthot, R., Proft, J., 2000. The influence of
Mater. Sci. Eng., A 497, 333–340. melt practice on final fatigue properties of superelastic NiTi
Nemat-Nasser, S., Guo, W.-G., 2006. Superelastic and cyclic wires. In: Proceedings of the International Conference on Shape
response of NiTi SMA at various strain rates and Memory and Superelastic Technologies, April, pp. 397–403.
temperatures. Mech. Mater. 38, 463–474. Robertson, S.W., Pelton, A.R., Ritchie, R.O., 2012. Mechanical
Nikanorov, A., Smouse, H.B., Osman, K., Bialas, M., Shrivastava, fatigue and fracture of Nitinol. Int. Mater. Rev. 57, 1–37.
S., Schwartz, L.B., 2008. Fracture of self-expanding nitinol Robertson, S.W., Ritchie, R.O., 2007. In vitro fatigue–crack growth
stents stressed in vitro under simulated intravascular and fracture toughness behavior of thin-walled superelastic
conditions. J. Vasc. Surg. 48, 435–440. Nitinol tube for endovascular stents: a basis for defining the
Norwich, D.W., 2014. A comparison of zero mean strain rotating effect of crack-like defects. Biomaterials 28, 700–709.
beam fatigue test methods for Nitinol wire. J. Mater. Eng. Runciman, A., Xu, D., Pelton, A.R., Ritchie, R.O., 2011. An
Perform. 23, 2515–2522. equivalent strain/Coffin–Manson approach to multiaxial
Norwich, D.W., Fasching, A., 2009. A study of the effect of fatigue and life prediction in superelastic Nitinol medical
diameter on the fatigue properties of NiTi wire. J. Mater. Eng. devices. Biomaterials 32, 4987–4993.
Saiidi, M.S., Wang, H., 2006. Exploratory study of seismic response
Perform. 18, 558–562.
Park, J., Nelson, D., 2000. Evaluation of an energy-based approach of concrete columns with shape memory alloys
reinforcement. Struct. J. 103.
and a critical plane approach for predicting constant
Sawaguchi, T.A., Kausträter, G., Yawny, A., Wagner, M., Eggeler, G.,
amplitude multiaxial fatigue life. Int. J. Fatigue 22, 23–39.
2003. Crack initiation and propagation in 50.9 at. pct Ni–Ti
Patel, M.M., Gordon, R., 2006. An investigation of diverse surface
pseudoelastic shape-memory wires in bending-rotation
finishes on fatigue properties of superelastic nitinol wire. In:
fatigue. Metall. Mater. Trans. A 34, 2847–2860.
International Conference on Shape Memory and Superelastic
Scirè Mammano, G., Dragoni, E., 2014. Functional fatigue of Ni–Ti
Technologies (SMST).
shape memory wires under various loading conditions SI:
Patel, M.M., 2007. Characterizing fatigue response of nickel–
Theory and experiments in fatigue lifetime assessment. Int. J.
titanium alloys by rotary beam testing. ASTM Int. 4, 1–11.
Fatigue 69, 71–83.
Pelton, A.R., 2011. Nitinol fatigue: a review of microstructures and
Sehitoglu, H., Anderson, R., Karaman, I., Gall, K., Chumlyakov, Y.,
mechanisms. J. Mater. Eng. Perform. 20, 613–617.
2001. Cyclic deformation behavior of single crystal NiTi. Mater.
Pelton, A.R., Dicello, J., Miyazaki, S., 2000. Optimisation of
Sci. Eng., A 314, 67–74.
processing and properties of medical grade Nitinol wire.
Shamsaei, N., Fatemi, A., 2009a. Deformation and fatigue
Minim. Invasive Ther. Allied Technol. 9, 107–118.
behaviors of case-hardened steels in torsion: experiments and
Pelton, A.R., Duerig, T.W., Stöckel, D., 2004. A guide to shape
predictions. Int. J. Fatigue 31, 1386–1396.
memory and superelasticity in Nitinol medical devices.
Shamsaei, N., Fatemi, A., 2009b. Effect of hardness on multiaxial
Minim. Invasive Ther. Allied Technol. 13, 218–221.
fatigue behaviour and some simple approximations for steels.
Pelton, A.R., Fino-Decker, J., Vien, L., Bonsignore, C., Saffari, P.,
Fatigue Fract. Eng. Mater. Struct. 32, 631–646.
Launey, M., Mitchell, M.R., 2013. Rotary-bending fatigue
Shamsaei, N., Fatemi, A., Socie, D.F., 2010. Multiaxial cyclic
characteristics of medical-grade Nitinol wire. J. Mech. Behav. deformation and non-proportional hardening employing
Biomed. Mater. 27, 19–32. discriminating load paths. Int. J. Plast. 26, 1680–1701.
Pelton, A.R., Gong, X.-Y., Duerig, T., 2003. Fatigue testing of Shamsaei, N., Fatemi, A., Socie, D.F., 2011. Multiaxial fatigue
diamond-shaped specimens. In: Medical Device Materials- evaluation using discriminating strain paths. Int. J. Fatigue 33,
Proceedings of the Materials and Processes for Medical 597–609.
Devices Conference, pp. 199–204. Shamsaei, N., McKelvey, S.A., 2014. Multiaxial life predictions in
Pelton, A.R., Schroeder, V., Mitchell, M.R., Gong, X.-Y., Barney, M., absence of any fatigue properties Multiaxial Fatigue. Int. J.
Robertson, S.W., 2008. Fatigue and durability of Nitinol stents. Fatigue 2013 (67), 62–72.
J. Mech. Behav. Biomed. Mater. 1, 153–164. Shamsaei, N., Yadollahi, A., Bian, L., Thompson, S.M., 2015. An
Peng, F., Jiang, X.-X., Hu, Y.-R., Ng, A., 2005. Application of shape overview of direct laser deposition for additive manufacturing;
memory alloy actuators in active shape control of inflatable Part II: Mechanical behavior, process parameter optimization
space structures. In: Aerospace Conference, 2005 IEEE. IEEE, and control. Addit. Manuf., under review for Publication.
pp. 1–10. Shaw, J.A., Kyriakides, S., 1997. On the nucleation and
Plotino, G., Grande, N.M., Cordaro, M., Testarelli, L., Gambarini, G., propagation of phase transformation fronts in a NiTi alloy.
2009. A review of cyclic fatigue testing of nickel–titanium Acta Mater. 45, 683–700.
rotary instruments. J. Endod. 35, 1469–1476. Smith, K.N., Topper, T.H., Watson, P., 1970. A stress–strain
Prahlad, H., Chopra, I., 2001. Design of a variable twist tilt-rotor function for the fatigue of metals (stress–strain function for
blade using shape memory alloy (SMA) actuators. In: SPIE’s metal fatigue including mean stress effect). J. Mater. 5,
Eighth Annual International Symposium on Smart Structures 767–778.
and Materials. International Society for Optics and Photonics, Socie, D., 1987. Multiaxial fatigue damage models. J. Eng. Mater.
pp. 46–59. Technol. 109, 293–298.
254 journal of the mechanical behavior of biomedical materials 50 (2015) 228 –254

Song, D., Kang, G., Kan, Q., Yu, C., Zhang, C., 2014. Non- memory alloys under biaxial proportional and non-
proportional multiaxial transformation ratchetting of super- proportional cyclic loadings. Mech. Mater. 42, 365–373.
elastic NiTi shape memory alloy: experimental observations. Wang, X.M., Zhou, Q.T., Liu, H., Deng, C.H., Yue, Z.F., 2012.
Mech. Mater. 70, 94–105. Experimental study of the biaxial cyclic behavior of thin-wall
Song, G., Ma, N., Li, H.-N., 2006. Applications of shape memory tubes of NiTi shape memory alloys. Metall. Mater. Trans. A 43,
alloys in civil structures. Eng. Struct. 28, 1266–1274. 4123–4128.
Stephens, R.I., Fatemi, A., Stephens, R.R., Fuchs, H.O., 2000. Metal Wilkes, K.E., Liaw, P.K., 2000. The fatigue behavior of shape-
Fatigue in Engineering, second ed. Wiley. memory alloys. JOM 52, 45–51.
Subramaniam, A., 1998. Fatigue Behavior of Copper Zinc Wu, S.K., Lin, H.C., Chen, C.C., 1999. A study on the machinability
Aluminum Shape Memory Alloys Master of Science. of a Ti49.6Ni50.4 shape memory alloy. Mater. Lett. 40, 27–32.
University of Manitoba. Zurbitu, J., Santamarta, R., Picornell, C., Gan, W.M., Brokmeier, H.-
Tabanli, R.M., Simha, N.K., Berg, B.T., 1999. Mean stress effects on G., Aurrekoetxea, J., 2010. Impact fatigue behavior of
fatigue of NiTi. Mater. Sci. Eng., A 273–275, 644–648. superelastic NiTi shape memory alloy wires. Mater. Sci. Eng.,
Tabanli, R.M., Simha, N.K., Berg, B.T., 2001. Mean strain effects on A 528, 764–769.
the fatigue properties of superelastic NiTi. Metall. Mater.
Trans. A 32, 1866–1869.
Takeda, K., Mitsui, K., Tobushi, H., Levintant-Zayonts, N.,
Glossary
Kucharski, S., 2013. Influence of nitrogen ion implantation on
deformation and fatigue properties of TiNi shape memory
Austenite: Stable phase of Nitinol having a simple cubic crystal
alloy wire. Arch. Mech. 65.
structure. Austenite is the parent phase of Nitinol and exists
Thompson, S.M., Bian, L., Shamsaei, N., Yadollahi, A., 2015. An
at temperatures above, what is called, the austenite finish
overview of direct laser deposition for additive manufacturing;
temperature, Af.
Part I: Transport phenomena, modeling and diagnostics.
Force-controlled test: A cyclic fatigue test under fixed force
Addit. Manuf. Manuf., under review.
amplitudes, regardless of displacement required to achieve
Tobushi, H., Hachisuka, T., Hashimoto, T., Yamada, S., 1998. Cyclic
those forces; also called stress-controlled test.
deformation and fatigue of a TiNi shape-memory alloy wire
Martensite: Stable phase of Nitinol with a monoclinic crystal
subjected to rotating bending. J. Eng. Mater. Technol. 120,
structure at low temperatures. Martensite is the daughter
64–70.
phase of Nitinol and generally exists at temperatures below,
Tobushi, H., Hachisuka, T., Yamada, S., Lin, P.-H., 1997. Rotating-
what is called, the martensite finish temperature, Mf.
bending fatigue of a TiNi shape-memory alloy wire. Mech. Mixed-phase Nitinol: Nitinol with both martensitic and austenitic
Mater. 26, 35–42. phases. This is mainly referred to a Nitinol at a temperature
Tolomeo, D., Davidson, S., Santinoranont, M., 2000. Cyclic above the martensite start temperature, Ms and below the
properties of superelastic Nitinol: design implications. In: austenite start temperature, As.
Russell, S.M., Pelton, A.R. (Eds.), SMST-2000: Proceedings of the Shape memory behavior: Ability of a thermally martensitic
International Conference on Shape Memory and Superelastic Nitinol to recover a pre-trained shape upon heating to a
Technologies, Pacific Grove, CA, USA. pp. 471–476. temperature above the austenite finish temperature, Af.
Urbano, M.F., Coda, A., Beretta, S., Cadelli, A., Sczerzenie, F., 2013. Shape memory Nitinol: The thermally martensitic Nitinol that
The effect of inclusions on fatigue properties for Nitinol. In: refers to a Nitinol at a temperature below the martensite
Mitchell, M.R., Smith, S.W., Woods, T., Berg, B. (Eds.), finish temperature, Mf.
Fatigue and Fracture Metallic Medical Materials and Strain (Displacement)-controlled test: A cyclic fatigue test per-
Devices. ASTM International, 100 Barr Harbor Drive, PO formed under fixed strain amplitudes, regardless of force
Box C700, West Conshohocken, PA 19428-2959, pp. 18–34. required to achieve those displacements. It should be noted,
Urbina, C., De la Flor, S., Ferrando, F., 2009. Effect of thermal “displacement”, per se, is not strain precisely in many cases.
cycling on the thermomechanical behaviour of NiTi shape Stress-induced martensite: A martensitic phase of Nitinol that
memory alloys. Mater. Sci. Eng., A 501, 197–206. forms from the austenite phase in superelastic Nitinol, when
Vaidyanathan, R., Dunand, D.C., Ramamurty, U., 2000. Fatigue the stress reaches a certain level.
crack-growth in shape-memory NiTi and NiTi–TiC composites. Superelasticity (pseudo elasticity): Ability of Nitinol to recover its
Mater. Sci. Eng., A 289, 208–216. original shape upon unloading to zero strain, even after
Van Humbeeck, J., 1991. Cycling effects, fatigue and degradation undergoing large strains, e.g.  10%.
of shape memory alloys. J. Phys. IV 01 C4-189–C4-197. Thermally martensitic Nitinol: The thermally martensitic Nitinol
Wang, X.M., Wang, Y.F., Lu, Z.Z., Deng, C.H., Yue, Z.F., 2010. An refers to Nitinol at a temperature below the martensite finish
experimental study of the superelastic behavior in NiTi shape temperature, Mf. also known as shape memory Nitinol.

You might also like