Sunteți pe pagina 1din 26

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/223480162

A finite-deformation-based phenomenological theory for shape-memory


alloys

Article  in  International Journal of Plasticity · August 2010


DOI: 10.1016/j.ijplas.2009.12.004

CITATIONS READS

48 56

1 author:

Prakash Thamburaja
Universiti Kebangsaan Malaysia
34 PUBLICATIONS   772 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Computational modeling of deformation of metallic glasses View project

All content following this page was uploaded by Prakash Thamburaja on 14 December 2017.

The user has requested enhancement of the downloaded file.


International Journal of Plasticity 26 (2010) 1195–1219

Contents lists available at ScienceDirect

International Journal of Plasticity


journal homepage: www.elsevier.com/locate/ijplas

A finite-deformation-based phenomenological theory for


shape-memory alloys
P. Thamburaja *
Department of Mechanical Engineering, National University of Singapore, Singapore 117576, Singapore

a r t i c l e i n f o a b s t r a c t

Article history: In this work we develop a finite-deformation-based, thermo-mechanically-coupled and


Received 24 September 2009 non-local phenomenological theory for polycrystalline shape-memory alloys (SMAs) capa-
Received in final revised form 10 December ble of undergoing austenite $ martensite phase transformations. The constitutive model is
2009
developed in the isotropic plasticity setting using standard balance laws, thermodynamic
Available online 4 January 2010
laws and the theory of micro-force balance (Fried and Gurtin, 1994). The constitutive
model is then implemented in the ABAQUS/Explicit (2009) finite-element program by writ-
Keywords:
ing a user-material subroutine. Material parameters in the constitutive model were fitted
A. Shape-memory alloys
B. Constitutive behavior
to a set of superelastic experiments conducted by Thamburaja and Anand (2001) on a poly-
Plasticity crystalline rod Ti–Ni. With the material parameters calibrated, we show that the experi-
C. Finite elements mental stress-biased strain–temperature-cycling and shape-memory effect responses are
qualitatively well-reproduced by the constitutive model and the numerical simulations.
We also show the capability of our constitutive mode in studying the response of SMAs
undergoing coupled thermo-mechanical loading and also multi-axial loading conditions
by studying the deformation behavior of a stent unit cell.
Finally, with the aid of finite-element simulations we also show that our non-local con-
stitutive theory is able to accurately determine the position and motion of austenite–mar-
tensite interfaces during phase transformations regardless of mesh density and without the
aid of jump conditions.
Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction

Shape-memory alloys (SMA) are a class of metallic alloys which have the ability to exist in different phases depending on
the temperature and/or stress state i.e. its current state is very sensitive to stress states and temperature changes. At high
temperatures and/or low stresses, SMAs will exist in the high symmetry austenite phase. However, at low temperatures
and/or high stresses SMAs will exist in the low-symmetry martensite phase. SMAs can also exist in a metastable state termed
as the rhombohedral phase or R-phase (Shaw and Kyriakides, 1995; Otsuka and Wayman, 1999). Of the multitude of smart/
functional materials being used in the recent years, polycrystalline shape-memory alloys (e.g. Ti–Ni, Cu–Al–Ni, Cu–Zn–Al,
Au–Cd systems etc.) remain one of the more popular choices with its variety of applications in the Micro-Electro-Mechanical
Systems (MEMS), biomedical, aerospace and civil structures area (Aizawa et al., 1998; Duerig et al., 1999; Machado and Savi,
2003; Seelecke and Mueller, 2004).
The phase transformation between the austenite phase and the martensite phase in SMAs is classified as a first-order
phase transformation. Austenite $ martensite phase transformations in SMAs are also diffusionless and reversible. Due to
its ability in undergoing reversible first-order phase transformation, SMAs exhibit exotic behaviors such as (a) superelasticity

* Corresponding author. Fax: +65 6779 6559.


E-mail address: mpept@nus.edu.sg

0749-6419/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2009.12.004
1196 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

or pseudoelasticity by transformation, and (b) the shape-memory effect. For a detailed discussion regarding superelasticity
and the shape-memory effect, please refer to the works of Shaw and Kyriakides (1995) and Otsuka and Wayman (1999).
Although austenite $ martensite transformations in SMAs are reversible, there is hysteresis that accompanies these types
of phase transformations. The mechanism responsible for the exhibited hysteresis is the motion of interfaces between the
two phases.
There is a world-wide activity in the development of models to describe the deformation behavior of SMAs. Concentrating
on the continuum-mechanics-based constitutive theories, these models can broadly be classified under two categories: (1)
phenomenologically-based models, and (2) crystal-plasticity-based constitutive theories. Some examples of phenomenolog-
ically-based theories include the works of Liang and Rogers (1990), Brinson (1993), Sun and Hwang (1993), Abeyaratne et al.
(1994), Boyd and Lagoudas, 1994, 1996, Auricchio and Taylor (1997), Lim and McDowell (1999), Mueller and Bruhns (2006),
Helm (2007), Ziolkowski (2007), Moumni et al. (2008) and Reese and Christ (2008). The constitutive models of Patoor et al.
(1996), Lu and Weng (1998), Fang et al. (1999), Gall and Sehitoglu (1999), Thamburaja and Anand (2001), Anand and Gurtin
(2003a), Jung et al. (2004), Peng et al. (2008) and Wang et al. (2008) are examples of crystal-mechanics-based SMA consti-
tutive models.1 Although crystal-plasticity-based theories are more physical i.e. it takes into account the crystallographic infor-
mation and texture effects, phenomenological models are still very useful in determining the deformation behavior of SMAs
because: (a) phenomenological models are easier to implement numerically, and (b) simulations using phenomenological mod-
els run much faster. This makes the usage of phenomenological models very relevant in the design of especially complex com-
ponents utilizing the exotic behavior exhibited by SMAs. Some recent work regarding the use of phenomenologically-based
SMA models to study the deformation behavior of biomedical structures such as orthodontic wires and arterial stents can be
found in Auricchio (2001) and Zaki and Moumni (2007), respectively. Since these SMA structures have the ability to undergo
relatively large deformations during its deployment, it is of paramount importance to develop a theory within a finite-deforma-
tion setting (cf. Auricchio and Taylor, 1997; Auricchio, 2001) to accurately model the response of these structures when
deformed.
As also mentioned previously, the behavior of SMAs is very sensitive to changes in temperature. Since the numerical
implementation of the phenomenologically-based constitutive models described above were not performed in a thermo-
mechanically-coupled setting, the response of structures and components which have spatially-varying temperature fields
will not be accurately modeled. Furthermore, the above mentioned phenomenological models are derived based on local the-
ories i.e. there is no material length scale in the constitutive equations. Hence, these models are unable to track the propa-
gation of austenite–martensite interfaces during phase transformation without the aid of jump conditions as the calculations
are heavily dependent on mesh size or grid size.2 A comprehensive theory for SMAs should also be able to model the coupled
thermo-mechanical response of SMAs during deformation and also accurately track the motion of the austenite–martensite
interfaces.
Therefore, the main tasks of this present work are to: (a) develop a finite-deformation-based phenomenological theory to
describe the coupled thermo-mechanical behavior of SMAs,3 and (b) develop a non-local-based constitutive model which will
minimize the effect of mesh density when tracking of the austenite–martensite interface motion during phase transformation.
The plan of this paper is as follows: In Section 2, we will develop a finite-deformation-based phenomenological consti-
tutive model based on isotropic plasticity theory. The derivation of this constitutive model is performed with the aid of stan-
dard balance laws, thermodynamic laws and the theory of micro-force balance (Fried and Gurtin, 1994). We have also
implemented our constitutive model in the ABAQUS/Explicit (2009) finite-element program. Algorithmic details of the
time-integration procedure used to implement the model are given in Appendix A. In Section 3, we fit the material param-
eters in the constitutive model to the superelastic experiments of Thamburaja and Anand (2001) conducted on a polycrys-
talline rod Ti–Ni. With the material parameters calibrated, we proceed to show that our constitutive model is able to
reproduce other exotic shape-memory alloy behaviors such as stress-biased strain–temperature-cycling and the austen-
ite ! martensite ! austenite shape-memory effect. Here we also demonstrate the capability of our constitutive model in
performing coupled thermo-mechanical analysis/simulations and qualitatively analyzing the multi-axial deformation
behavior of a stent unit cell. In Section 4, we showcase the non-local version of our theory by studying the austenite–mar-
tensite interface propagation during superelastic deformation. We show that our theory is able to heavily minimize the ef-
fect of mesh density in the tracking of the austenite–martensite interface motion during superelastic deformation. We
conclude and provide directions for future research in Section 5.

2. Constitutive model

In this section we construct a constitutive model for shape-memory alloys capable of undergoing austenite $ martensite
phase transformation using finite-deformation-based isotropic plasticity theory. To construct the constitutive theory, we

1
There are also detailed micromechanics-based constitutive models developed by Levitas and co-workers to describe various phenomena exhibited by
shape-memory alloys. The goal of the present work is to develop the simplest phenomenologically-based finite-deformation-based constitutive model to
describe the deformation behavior of polycrystalline SMAs under thermo-mechanical loading conditions.
2
The present discussions regarding the mesh sensitivity on the prediction of the austenite–martensite interface positions were made under the assumption
of isothermal conditions.
3
We have also developed a small-strain-based phenomenological theory for polycrystalline SMAs in Thamburaja and Nikabdullah (2009).
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1197

focus on an arbitrary subregion R of a continuum body in the reference configuration with n denoting the outward unit nor-
mal of the body’s boundary specified by @R. All the balance and imbalance equations in this work are formulated in the ref-
erence configuration.
The list of the governing variables in the constitutive model are4: (i) The Helmholtz free energy per unit reference volume,
w. (ii) The Cauchy stress tensor, T. (iii) The total deformation gradient tensor, F with det F > 0. (iv) The absolute temperature, h.
(v) The inelastic distortion tensor, Fp with det Fp > 0. It measures the cumulative deformation due to austenite $ martensite
phase transformations. (vi) The elastic distortion tensor, Fe with det Fe > 0. It describes the elastic stretches that gives rise to
1
the Cauchy stress T. From the theory of Kröner (1960) and Lee (1969)), the elastic distortion tensor is given by Fe ¼ FFp .
(vii) The total martensite volume fraction, n with 0 6 n 6 1.

2.1. Motion, kinematics and flow rule

Let y denote the position vector of a material point in the current configuration. The velocity vector of the material point is
then given by v ¼ y.
_ Hence, the material time derivative of the deformation gradient tensor F can be written as

F_  rv ¼ LF ð1Þ
e p
where L represents the total velocity gradient. Using the definition F ¼ F F , we can further write the total velocity gradient
as
1
L ¼ Le þ Fe Lp Fe ð2Þ
1 1 >
where Le ¼ F_ e Fe and Lp ¼ F_ p Fp represent the elastic and inelastic velocity gradients, respectively. With Ce ¼ Fe Fe , we de-
fine a measure for the elastic strain, Ee by
>
Ee ¼ ð1=2ÞðCe  1Þ ! E_ e ¼ Fe ðsym Le ÞFe : ð3Þ
In formulating the flow rule, we make two key assumptions: (1) from the work of Gurtin and Anand (2005), the inelastic
velocity gradient, Lp is taken to be spinless i.e. skw Lp ¼ 0, and (2) small volume changes accompanying austenite $ martens-
ite phase transformations are neglected (Sun and Hwang, 1993) i.e. the inelastic velocity gradient, Lp is purely deviatoric.
With these assumptions, we generalize the work of Sun and Hwang (1993) to a finite-deformation setting and set the
inelastic velocity gradient to be

X
2
Lp ¼ kð1 þ a/Þ n_ i Ni : ð4Þ
i¼1

Here n_ 1 P 0 and n_ 2 6 0 denote the martensitic transformation rates. Forward transformation occurs if n_ 1 > 0 whereas re-
verse transformation occurs if n_ 2 < 0. The flow direction tensors Ni with i ¼ 1; 2 are restricted by Ni ¼ N>i and trace Ni ¼ 0.
We define N1 and N2 as the forward transformation and reverse transformation flow direction, respectively. Following Auric-
chio et al. (1997), the total rate of martensitic volume fraction change is given by

X
2
n_ ¼ n_ i ¼ n_ 1 þ n_ 2 : ð5Þ
i¼1

A transformation from austenite to martensite takes place when n_ > 0. Conversely, a transformation from martensite to
austenite takes place when n_ < 0. No net phase transformation occurs if n_ ¼ 0.
With k > 0 being a constant of proportionality, we will further enforce jNi j ¼ T for each i where T > 0 denotes the trans-
formation strain due to austenite–martensite phase transformation (to be determined experimentally). Furthermore, guided
by the work of Orgeas and Favier (1998) we have the scalar / with 1 6 / 6 1 to represent the J 3 i.e. the third stress-invari-
ant measure with the dimensionless constitutive parameter a (to be calibrated from experiments) controlling the extent of
the tension–compression asymmetry exhibited by SMAs during deformation. The functional form for / will be described
later.
From microscopic considerations, the reverse transformation is the crystallographic recovery of the deformation which is
induced during forward transformation i.e. the reverse transformation is restricted by the forward transformation history.
The deformation experienced due to austenite ! martensite (forward) phase transformation can be completely recovered
by the reversal of the forward loading history. Therefore, with N1 to be defined we take the flow direction N2 to be given
by (Sun and Hwang, 1993; Boyd and Lagoudas, 1996):

4
Notation: The terms Div, r and r2 denote the referential divergence, gradient and Laplacian operators, respectively. All the tensorial variables in this work
are second-order tensors unless stated otherwise. For a tensor B; B> denotes its transpose. We also write trace B for the trace of the tensor B. The determinant of
tensor B is denoted by det B. The symmetric portion of tensor B is denoted by sym B  ð1=2ÞðB þ B> Þ. The skew-symmetric portion of tensor B is denoted by
skw B  ð1=2ÞðB  B> Þ. The deviatoric portion of tensor B is denoted by B0  B  ð1=3Þðtrace BÞ1. The symmetric and deviatoric portion of tensor B is denoted by
sym B0 . The scalar product of two tensors A and B is denoted by A  B ¼ trace ðB> AÞ. The scalar product of two vectors u and v is also denoted by u  v . The
magnitude of vector u and tensor B is denoted by juj and jBj, respectively. The second-order identity tensor is denoted by 1.
1198 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

 
B
N2 ¼ T with B_ ¼ Lp :
jBj
Here B is a tensor defined in the configuration determined by Fp i.e. the relaxed configuration.

2.2. Micro-force balance

For the martensite volume fraction n, the micro-force system which describe the forces that perform work associated to
austenite $ martensite phase transformations consist of: (a) the micro-traction vector, c, measured per unit area in the ref-
erence configuration; (b) the scalar internal micro-force, pint , measured per unit volume in the reference configuration; and (c)
the scalar external micro-force, pext , measured per unit volume in the reference configuration. Following the work of Fried and
Gurtin (1994), we write the corresponding micro-force balance equation associated with these micro-force systems as
Z Z Z
c  n dA þ pext dV ¼ pint dV ð6Þ
@R R R

where dA and dV denotes the area and volume integral in the reference configuration, respectively. Applying the divergence
law on Eq. (6) and localizing the result within R results in
Div c  pint þ pext ¼ 0: ð7Þ
We take c; pint and pext to be functions of the variables n; n_ and rn:
_ rnÞ;
c ¼ ^cðn; n; _ rnÞ and pext ¼ p
pint ¼ p^ int ðn; n; _ rnÞ:
^ ext ðn; n;

2.3. Balance of linear momentum

The balance of linear momentum is given by


Z Z
Sn dA þ b dV ¼ o ð8Þ
@R R

where S  ðdet FÞTF> denotes the First Piola–Kirchoff stress tensor and b the macroscopic body force vector per unit refer-
ence volume. Inertial forces are also included in the body force b. Using the divergence law on Eq. (8) and localizing the result
within R yields
Div S þ b ¼ o: ð9Þ

2.4. Balance of angular momentum

The balance of angular momentum is written as


Z Z
y  Sn dA þ y  b dV ¼ o: ð10Þ
@R R

Applying the divergence law on Eq. (10) and localizing the result within R while using Eq. (9) yields

SF> ¼ FS> : ð11Þ


> >
Substituting S ¼ ðdet FÞTF into Eq. (11) results in T ¼ T i.e. the Cauchy stress is symmetric.

2.5. Balance of energy

The first law of thermodynamics (the balance of energy) is stated as


Z Z Z
d
½Sn  v þ ðc  nÞn_  q  n dA þ ðb  v þ pext n_ þ rÞ dV ¼  dV ð12Þ
@R R dt R

where  is the internal energy per unit reference volume. Here q is the heat flux vector measured per unit area in the reference
configuration and r is the heat supply per unit reference volume. Applying the divergence law on Eq. (12) and localizing the
result within R while using Eqs. (1), (7) and (9) yields

SF>  L þ c  rn_ þ pint n_  Divq þ r ¼ _ : ð13Þ


>
Assuming that the external micro-force vanishes i.e. pext ¼ 0, substituting S  ðdet FÞTF , Eqs. (2), (3) and (7) into Eq.
(13) while using the result of Eq. (11) yields

T  E_ e þ T  Lp þ c  rn_ þ ðDiv cÞn_  Div q þ r ¼ _ ð14Þ


P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1199

where
1 >
T ¼ ðdet FÞFe TFe and T ¼ Ce T ð15Þ
5 
denote frame-invariant measures of stresses. Note that T is symmetric whereas T is generally not symmetric.

2.6. Entropy imbalance

The second law of thermodynamics is written as


Z Z Z
d q r
gdV P   n dA þ dV ð16Þ
dt R @R h R h
with g representing the entropy per unit reference volume. Using the divergence law on Eq. (16) and localizing the result
within R yields
q
g_ h þ Div q   rh  r P 0: ð17Þ
h
The Helmholtz free energy per unit reference volume, w, is defined as
_
w ¼   gh ! w_ ¼ _  g_ h  gh: ð18Þ
Denoting m ¼ rn, we use the functional expression for the free energy density of a shape-memory alloy (Helm and Hau-
pt, 2003) and augment it with a gradient energy (Fried and Gurtin, 1994) viz.

^ e ; m; n; hÞ ! w_ ¼ @w  E_ e þ @w  m
w ¼ wðE _ þ
@w _ @w _
nþ h: ð19Þ
@Ee @m @n @h
Substituting Eqs. (18) and (19) into Eq. (14) yields
     
@w @w _ @w
T  e  E_ e  g þ hþ c _ þ C ¼ g_ h
m ð20Þ
@E @h @m
where
@w _
C  T  Lp þ ðDivcÞn_  n  Divq þ r: ð21Þ
@n
Further substitution of Eq. (20) into inequality (17) results in the dissipation inequality:
     
@w @w _ @w
T  e  E_ e  g þ hþ c _ þPP0
m ð22Þ
@E @h @m
where
@w _ q
P  T  Lp þ ðDiv cÞn_  n   rh: ð23Þ
@n h
From rational thermodynamic arguments, inequality (22) yields
@w @w @w
T ¼ ; g¼ and c ¼ : ð24Þ
@Ee @h @m
Eq. (24)1, (24)2 and (24)3 are the constitutive equations for the stress, entropy and the micro-traction vector, respectively.

2.7. Phase transformation criteria and Fourier’s law

Using Eq. (24), we obtain the reduced dissipation inequality from inequality (22):
@w _ q
P  T  Lp þ ðDivcÞn_  n   rh P 0: ð25Þ
@n h
Here P represents the total dissipation and it is always non-negative. Recall that each Ni is defined to be symmetric and
deviatoric. Substituting Eqs. (4) and (5) into Eq. (25), and assuming each dissipative mechanism to be strongly dissipative
(Anand and Gurtin, 2003a,b) yields:

5
Frame-invariance (Anand and Gurtin, 2003b): Let x and z denote the position of the material point in the reference configuration and relaxed configuration
i.e. the configuration determined by Fp , respectively. As mentioned previously, y denotes the position of the material point in the current configuration. Now,
consider the transformations of the form x ! x; z ! z and y ! Q ðtÞy þ aðtÞ where t denotes time, Q ðtÞ is a proper orthogonal rotation tensor and aðtÞ a
translational vector. The reference and relaxed configurations are independent of the choice of such changes in frame. Under changes in frame of the form given
above, the variables: (a) F ! QF, (b) Fp ! Fp , (c) Lp ! Lp , (d) Fe ! QFe , (e) Ce ! Ce since Fe ! QFe , (f) Ee ! Ee since Ce ! Ce , (g) T ! T since
det F ! det F; Fe ! QFe and T ! QTQ > , (h) T ! T since Ce ! Ce and T ! T , (i) B ! B since B is defined in the relaxed configuration, (j) B_ ! B_ since B_ ¼ Lp and
Lp ! Lp , (k) c ! c and q ! q since the micro-traction vector and the heat flux vector are referentially-defined, and finally (l)
w ! w; pint ! pint ; pext ! pext ; h ! h; n ! n and  !  since these variables are scalar fields.
1200 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

f1 n_ 1 > 0 whenever n_ 1 –0 ð26Þ


 
where f1  kð1 þ a/Þ½symT0  N1  þ Divc  @w
@n
denotes the driving force for forward transformation,

f2 n_ 2 > 0 whenever n_ 2 –0 ð27Þ


 
where f2  kð1 þ a/Þ½symT0  N2  þ Divc  @w
@n
denotes the driving force for reverse transformation, and finally

q
  rh > 0 whenever rh–o: ð28Þ
h
The inequalities (26)–(28) are assumed to be obeyed at all times so that the reduced dissipation inequality (25) is con-
currently satisfied.
To satisfy inequality (26) under the assumption of rate-independence, we choose an expression for f1 as follows:

f1 ¼ fc1 ðsignðn_ 1 ÞÞ ! f1 ¼ fc1 whenever n_ 1 > 0 ð29Þ


where fc1 ¼ ^f c1 ðhÞ > 0 represents the critical resistance to forward transformation.
Similarly, to satisfy inequality (27) under the assumption of rate-independence, we choose an expression for f2 as follows:

f2 ¼ fc2 ðsignðn_ 2 ÞÞ ! f2 ¼ fc2 whenever n_ 2 < 0 ð30Þ


where fc2 ¼ ^f c2 ðhÞ > 0 represents the critical resistance to reverse transformation.
Eqs. (29) and (30) are the criteria for forward and reverse phase transformations, respectively. Generally, we can have
fc1 –fc2 .
Finally, assuming the material obeys Fourier’s law of heat conduction we enforce

q ¼ kth rh ð31Þ
^ ðn; hÞ > 0 denotes the coefficient of thermal conductivity. For simplicity, we will as-
to satisfy inequality (28) where kth ¼ kth
sume the coefficient of thermal conductivity to be equal and constant at all times regardless of martensite volume fraction
and temperature.

2.8. Free energy density and specific constitutive functions

The free-energy density of the material is taken to contain the free energy of a conventional shape-memory alloy (Helm
and Haupt, 2003) augmented with a gradient energy (Fried and Gurtin, 1994). We take the free energy per unit reference vol-
ume, w to be in the separable form

w ¼ we ðEe ; hÞ þ wg ðmÞ þ wn ðn; hÞ þ wh ðhÞ where ð32Þ


 2 1
we ðEe ; hÞ ¼ lEe0  þ jðtraceEe Þ2  3jath ðh  ho ÞðtraceEe Þ; wg ðmÞ ¼ sn jmj2 ; ð33Þ
2
kT 1
wn ðn; hÞ ¼ ðh  hT Þn þ hn2 and wh ðhÞ ¼ cðh  ho Þ  ch logðh=ho Þ: ð34Þ
hT 2

Here we denotes the classical thermo-elastic free energy density with l ¼ l ^ ðn; hÞ; j ¼ j
^ ðn; hÞ and ath ¼ a
^ th ðn; hÞ denoting
the shear modulus, bulk modulus and the coefficient of thermal expansion, respectively.
Following the work of Fried and Gurtin (1994), we introduce an isotropic gradient energy wg where sn ¼ ^sn ðn; hÞ P 0 de-
notes a material parameter with units of energy per unit length. The gradient energy acts to penalize to presence of austen-
ite–martensite interfaces. Hence, due to the gradient energy there exists an intrinsic material length scale in the constitutive
model. As a first-cut assumption, we will treat sn as a constant.
The austenite $ martensite phase transformation energy is denoted by wn where kT and hT represents the latent heat re-
leased/absorbed (units of energy per unit volume) during the austenite $ martensite phase transformation and the phase
equilibrium temperature, respectively.6 The energetic interaction coefficient, h has units of energy per unit volume.
Finally, wh represents the purely thermal portion of the free energy with c ¼ ^cðn; hÞ being the specific heat per unit refer-
ence volume. Following the modeling assumptions adopted by Abeyaratne and Knowles (1993), we will suppress the depen-
dence of l; j; ath and c on the martensite volume fraction and temperature, and treat them as constants.

2.9. Constitutive equation for the stress, entropy and micro-traction vector

Substituting Eq. (32) into Eq. (24)1 yields the constitutive equation for the stress:

T ¼ 2lEe0 þ j½traceEe  3ath ðh  ho Þ1: ð35Þ

6
In formulating the phase transformation free energy, we are guided by the one-dimensional model of Abeyaratne and Knowles (1993).
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1201

The constitutive equation for the entropy density and micro-traction vector are respectively given substituting Eq. (32)
into Eq. (24)2 and (24)3:
g ¼ c logðh=ho Þ þ 3jath ðtraceEe Þ  ðkT =hT Þn and c ¼ sn ðrnÞ: ð36Þ

2.10. Flow direction N1 and the J 3 parameter

Since the material is assumed to be elastically-isotropic, we can conclude from Eq. (35) that Ee and T are co-axial. Thus,
the stress tensor T  Ce T is also symmetric. Using this result and substituting Eqs. (32) and (36)2 into the expressions for the
driving force for phase transformation yield:
kT
f1 ¼ kð1 þ a/ÞðT0  N1 Þ þ sn ðr2 nÞ  ðh  hT Þ  hn; ð37Þ
hT
kT
f2 ¼ kð1 þ a/ÞðT0  N2 Þ þ sn ðr2 nÞ  ðh  hT Þ  hn: ð38Þ
hT
From the expression for the driving force shown in Eqs. (37) and (38), we can see that the transformation between the
austenite and martensite phase is affected by local terms (due to the stress and phase transformation energy) and non-local
terms (due to the gradient energy).
During forward transformation i.e. n_ 1 –0, substituting Eq. (37) into Eq. (29) results in
kT
kð1 þ a/ÞðT0  N1 Þ ¼ ðh  hT Þ þ hn  sn ðr2 nÞ þ fc1 ðsignðn_ 1 ÞÞ: ð39Þ
hT
To satisfy Eq. (38), we take

kT
ðT Þ2 ½kð1 þ a/ÞT0 ¼ ðh  hT Þ þ hn  sn ðr2 nÞ þ fc1 ðsignðn_ 1 ÞÞ N1 ð40Þ
hT
since jN1 j ¼ T . Taking the magnitude on both sides of Eq. (40) yields
( )
T0 kT
N1 ¼ T  T ð1 þ a/Þ þ sn ðr2 nÞ 
! f1 ¼ r ðh  hT Þ  hn ð41Þ
jT0 j hT

where r
 ¼ kjT0 j represents an equivalent stress. From the work of Orgeas and Favier (1998), the J 3 parameter is then given by
pffiffiffi
/¼ 6½N1  N21 ðT Þ3 :

2.11. Flow rule: revisited

During forward transformation i.e. n_ 1 –0 and n_ 2 ¼ 0, we define


pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi
T ð1 þ a/Þn_ 1  2=3jsymLp j ! k ¼ 3=2 and r ¼ 3=2jT0 j:
Therefore r
 denotes the equivalent tensile stress or Mises stress. The final form for the inelastic strain-rate i.e the flow rule is
then given by
pffiffiffiffiffiffiffiffi X
2
Lp ¼ 3=2ð1 þ a/Þ n_ i Ni : ð42Þ
i¼1

2.12. Conditions on the driving forces and phase transformation rates

The driving forces f1 and f2 are defined to be within the ranges:


f1 6 fc1 for 0 6 n < 1; f 2 P fc2 for 0 < n 6 1:
For n ¼ 1, the driving force for forward transformation is defined for all values of f1 . For n ¼ 0, the driving force for reverse
transformation is defined for all values of f2 .
In a rate-independent theory, the variables ff1 ; n_ 1 g and ff2 ; n_ 2 g must satisfy the following conditions:

 Elastic range conditions: If f1 –fc1 , then n_ 1 ¼ 0. If f2 –  fc2 , then n_ 2 ¼ 0.


 Forward transformation: If 0 6 n < 1 and f1 ¼ fc1 , then
_
n_ 1 ðf1  fc1 Þ ¼ 0: ð43Þ
 Reverse transformation: If 0 < n 6 1 and f2 ¼ fc2 , then
1202 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

_
n_ 2 ðf2 þ fc2 Þ ¼ 0: ð44Þ
 End conditions: If n ¼ 1 and f1 ¼ fc1 , then n_ 1 ¼ 0. If n ¼ 0 and f2 ¼ fc2 , then n_ 2 ¼ 0.

Eqs. (43) and (44) are the consistency conditions for forward and reverse phase transformation, respectively. The consis-
tency conditions are used to determine the transformation rates n_ 1 and n_ 2 .

2.13. Balance of energy: revisited

Substituting Eqs. (31), (32), (35), (36)1 and (36)2 into Eq. (20) yields

kT
T0  Lp þ sn ðr2 nÞn_  ðh  hT Þn_ þ kth ðr2 hÞ þ r ¼ g_ h: ð45Þ
hT
Taking the time-derivative of Eq. (36)1 results in
c  kT _
g_ ¼ h_ þ 3jath ðtraceE_ e Þ  n: ð46Þ
h hT
The evolution equation for the temperature is given by substituting Eqs. (5), (42) and (46) into Eq. (45):

X
2
ch_ ¼ kth ðr2 hÞ þ ðkT =hT Þhn_  3jath ðtraceE_ e Þh þ fi n_ i þ r: ð47Þ
i¼1

On the right-hand side of Eq. (47), the first term describes the heat conduction contribution, the second and third term
describes the heat source due to entropic contributions, and the fourth term describes the dissipation due to the motion of
austenite–martensite interfaces.
The list of constitutive parameters/functions that needed to calibrated/specified are

fl; j; ath ; sn ; kT ; hT ; h; c; a; T ; fc1 ; fc2 ; kth ; rg:


A time-integration procedure based on the isotropic-plasticity-based constitutive model for shape-memory alloys has
been developed and implemented in the ABAQUS/Explicit (Abaqus reference manuals, 2009) finite-element program by
writing a user-material subroutine. Algorithmic details for the time-integration procedure used to implement the model
in the finite-element code are given in Appendix A.

3. Experiments and finite-element simulations

To calibrate the material parameters in the constitutive model, we fit the constitutive theory to the stress–strain curves
obtained from simple tension and simple compression superelastic experiments conducted on a polycrystalline rod Ti–Ni at
a temperature of 298 K (Thamburaja and Anand, 2001). Isothermal conditions were maintained by conducting the experi-
ments under very low strain-rates. With hmf and has denoting the martensite finish temperature and the austenite start tem-
perature, respectively, the transformation temperatures for the polycrystalline Ti–Ni material were determined to be

Stress Stress

austenite to martensite

martensite to austenite

Strain Strain
Actual behavior Idealized behavior
(a) (b)
Fig. 1. The schematic stress–strain response of a polycrystalline shape-memory alloys undergoing superelastic deformation under uniaxial stress states.
Also shown is the idealized stress–strain response used to calibrate the material parameters in the constitutive model.
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1203

hms ¼ 251:3 K;hmf ¼ 213:0 K;has ¼ 260:3 K and haf ¼ 268:5 K. Hence the material is initially in the fully-austenitic phase at
temperature 298 K.
Fig. 1a shows a schematic stress–strain response of an actual polycrystalline shape-memory alloy undergoing superelastic
deformation under uniaxial loading. The fitting of the constitutive parameters were performed on the idealized version of
the actual experimental stress–strain response cf. Fig. 1b. The idealized stress–strain response contains the key features
of superelastic deformation, namely: (1) Initial loading which causes the elastic deformation of the austenitic material;

(a)

800 1000

700 900
EXPERIMENT EXPERIMENT
800 SIMULATION
600 SIMULATION
STRESS [MPa]

STRESS [MPa]

700
500
600
400 500

300 400
300
200
200
100
100
0 0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0 0.01 0.02 0.03 0.04 0.05 0.06
STRAIN STRAIN
(b) (c)

500 1000
450 900
EXPERIMENT
SHEAR STRESS [MPa]

TENSION
400 SIMULATION 800
COMPRESSION
STRESS [MPa]

350 700
300 600
250 500
200 400
150 300
100 200
50 100
0 0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
SHEAR STRAIN STRAIN
(d) (e)

Fig. 2. (a) An initially-undeformed single ABAQUS C3D8R continuum-three-dimensional brick element. Numerical and experimental superelastic stress–
strain curves in (b) simple tension, (c) simple compression and (d) simple shear. The data from the tension and compression experiments were used to fit
the material parameters. The simulated shear stress-shear strain response corresponds to an independent prediction. (e) Comparison of the stress–strain
response from the tension and compression simulations to demonstrate the tension–compression asymmetry.
1204 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

(2) continued loading which results in a phase transformation from austenite ! martensite; (3) further loading which will
cause the elastic deformation of the martensitic material; (4) reverse loading which causes the elastic unloading of the

Table 1
Material parameters for the polycrystalline rod Ti–Ni.

l ¼ 23:31 GPa j ¼ 60:78 GPa ath ¼ 10  106 K1 T ¼ 0:046


a ¼ 0:13 h ¼ 0 J=m3 c ¼ 2:08 MJ=Km3 kth ¼ 18 W=m K
hT ¼ 255:8 K kT ¼ 97 MJ=m3 fc ¼ 7:8 MJ=m3 sn ¼ 0 J=m
r ¼ 0 W=m3

(a) 900

800
TENSION : 308 K

700 TENSION : 298 K


TENSION : 288 K

600
STRESS [MPa]

500

400

300

200

100

0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
STRAIN

(b) 1200

1000 COMPRESSION : 308 K


COMPRESSION : 298 K
COMPRESSION : 288 K

800
STRESS [MPa]

600

400

200

0
0 0.01 0.02 0.03 0.04 0.05 0.06
STRAIN

Fig. 3. Simulated superelastic stress–strain responses in (a) simple tension and (b) simple compression at ambient temperatures of 288 K, 298 K and 308 K
under isothermal conditions.
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1205

martensitic material; (5) continued reverse loading which results in the phase transformation from martensite ! austenite;
and (6) further reverse loading which results in the elastic unloading of the austenitic material.
As a first-cut assumption, we will take the critical resistances to forward and reverse transformation to be equal, constant
and suppress its dependence on temperature i.e. fc1 ¼ fc2 ¼ fc where fc > 0 has units of energy per unit volume. Since the size
of the test specimens are several orders of magnitude larger than the thickness of the austenite–martensite interface, we will
ignore the effect of the gradient energy in our calculations as a first-cut assumption i.e. we set sn ¼ 0 J=m. The influence of
the gradient energy on the deformation behavior of shape-memory alloys undergoing austenite $ martensite phase trans-
formations will be investigated in Section 4. In the spirit of modeling elastic-perfectly inelastic materials, we set the ener-
getic interaction coefficient, h ¼ 0 J=m3 . Finally the heat supply per unit volume, r, is ignored in all of our calculations by
setting it to be zero.

(a) 0.09

0.08 TENSION : 50 MPa


TENSION : 100 MPa
0.07
TENSION : 150 MPa

0.06

0.05
STRAIN

0.04

0.03

0.02

0.01

220 240 260 280 300 320


TEMPERATURE [K]

(b) 0.07

COMPRESSION : 50 MPa
0.06
COMPRESSION : 100 MPa
COMPRESSION : 150 MPa
0.05

0.04
STRAIN

0.03

0.02

0.01

220 240 260 280 300 320


TEMPERATURE [K]

Fig. 4. Strain vs. temperature response obtained from the strain–temperature-cycling simulations conducted under constant (a) tensile and (b) compressive
stresses of 50 MPa, 100 MPa and 150 MPa. The thermal cycling is conducted between temperatures of 220 K and 320 K.
1206 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

500

400

STRESS [MPa]
300 TENSION
COMPRESSION
200

100

0
265
0.06
270 0.05
0.04
TEMPERATURE [K] 275 0.03
0.02
0.01 STRAIN
280 0

Fig. 5. Stress–strain–temperature response obtained from the shape-memory effect simulations conducted in simple tension and simple compression. An
isothermal stress–strain response from straining occurs at a temperature of 265 K. Following this, an increase in temperature to 280 K takes place.

All the finite-element simulations in this work were conducted using a single ABAQUS C3D8R continuum-three-dimen-
sional brick element shown in Fig. 2a under unless stated otherwise. This element is meshed using eight corner nodes and is
integrated numerically using a reduced integration scheme.
The material is taken to be fully-austenitic at the beginning of each simulation. The constitutive parameters were deter-
mined by fitting the model to the simple tension and compression experiments using a similar methodology outlined in
Thamburaja and Anand (2001). Using the material parameters listed out in Table 17 the isothermal stress–strain curves at
a temperature of 298 K obtained from the simple tension and simple compression finite-element simulations are plotted in
Fig. 2b and c, respectively. The fit from the numerical simulations are in good accord with the experimental stress–strain
responses.
With the constitutive parameters calibrated, a finite-element simulation in simple shear was performed and the resulting
isothermal shear stress-shear strain response at a temperature of 298 K is plotted in Fig. 2d. The experimental shear stress-
shear strain curve is well-predicted by the constitutive model.
The numerical stress–strain curves obtained from the simple tension and simple compression simulations conducted
above are repeatedly plotted in Fig. 2e for comparison. As shown by these stress–strain responses, the present constitutive
theory is able to model the tension–compression asymmetry exhibited by polycrystalline rod Ti–Ni namely: (i) the stress
level required to nucleate the martensitic phase from the parent austenitic phase is considerably higher in compression than
in tension; (ii) the transformation strain measured in compression is smaller than that in tension; and (iii) the hysteresis
loop generated in compression is wider (measured along the stress axis) than the hysteresis loop generated in tension. These
features in the stress–strain responses are exhibited due to the influence of the J 3 parameter.
With the model calibrated to have values as shown in Table 1, we perform a set of superelastic simulations in simple ten-
sion and simple compression under isothermal conditions at two other ambient temperatures: 288 K and 308 K. The stress–
strain curves from these finite-element simulations are plotted in Fig. 3a and b together with the simple tension and simple
compression stress–strain curves obtained from the simulations conducted at an ambient temperature of 298 K (as also
shown in Fig. 2b and c). The stress–strain responses plotted in Fig. 3a and b show that the stress required to induce austenite
to martensite transformation or martensite to austenite transformation increases with increasing ambient/test temperature.
This concurs very well with experimental findings (e.g. Thamburaja and Anand, 2003).
Next we conduct a series of strain–temperature-cycling simulations which can be described as follows: At a temperature
of 320 K, the material is first pre-stressed to a desired stress level. With the pre-stress maintained, the temperature of the
material is reduced to 220 K and then increased back again to 320 K. Depending on the pre-stress level, a transformation
from the austenite to martensite phase will occur at a particular temperature as result of a sufficient reduction in temper-
ature. At this point, a sufficient increase in temperature will then cause a transformation from the martensite to austenite
phase to take place at another critical temperature.
In our finite-element simulations, we choose three different pre-stress levels under simple tension and simple compres-
sion conditions: 50 MPa, 100 MPa and 150 MPa. Fig. 4a shows the strain versus temperature responses for the strain–tem-
perature-cycling simulations conducted under the aforementioned tensile stress levels. The strain versus temperature curves

7
The austenitic phase values are chosen for the material parameters fl; j; ath ; c; kth g. As mentioned previously, we have assumed no mismatches between
the austenite and martensite phase material parameters for simplicity.
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1207

for the strain–temperature-cycling finite-element simulations performed under the aforementioned compressive stress lev-
els are plotted in Fig. 4b. From the simulation results shown in Fig. 4a and b, we can conclude the following trends: with
increasing pre-stress level, a phase transformation from austenite to martensite or martensite to austenite occurs at a higher
temperature. The results shown in Fig. 4a and b also qualitatively reproduce the strain–temperature-cycling experimental
data shown in Thamburaja and Anand (2003). During the strain–temperature-cycling, note that the transformation (actua-
tion) strain obtained due to the tensile pre-stress is higher to that obtained using a compressive pre-stress. This is again due
to the introduction of the J 3 parameter which takes different values under tensile or compressive stress states. Another point
to note is that despite the amount of pre-stress and the sign of the pre-stress i.e. tensile or compressive, the temperature at
which the martensite to austenite transformation occurs is always approximately 42 K higher than the temperature at which
the austenite to martensite transformation takes place.
To simulate the one-way austenite ! martensite ! austenite shape-memory effect, we perform the following finite-ele-
ment calculations: With the temperature of the material initially at ho ¼ 265 K where hms < 265 K < haf , we perform an
isothermal simple tension and simple compression simulation to cause a transformation from austenite to martensite. When
the material is fully-martensitic at the completion of the forward loading process, a reverse loading process to an unstressed
state occurs through an elastic unloading of the material. With the applied stress in the material maintained at zero, the
temperature of the material is then raised to 280 K. The stress–strain–temperature response from these finite-element
simulations are plotted in Fig. 5. Note that once the reverse loading process to an unstressed state has taken place, a residual
strain will exist in the material as the temperature is not sufficiently high enough for reverse transformation from martensite

(a)
5 mm

3
2

1 40 mm

1 mm

(b) 900

800 −4 s −1 (Simulation A)
STRAIN−RATE = 1 x 10
−4 s−1 (Simulation B)
e
STRAIN−RATE = 5 x 10
700

600 d
STRESS [MPa]

c
500 b

400

300

200

100 g f
h
a/i
0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
STRAIN

Fig. 6. (a) Initially-undeformed finite-element mesh of an SMA sheet with dimensions of 5 mm by 40 mm by 1 mm meshed using 200 ABAQUS C3D8RT
elements. (b) Simulated tensile superelastic stress–strain response of the sheet shown in Fig. 6a conducted under a strain-rate of 1  104 s1 (Simulation
A) and 5  104 s1 (Simulation B). Both these simulations were performed using a fully-coupled thermo-mechanical analysis.
1208 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

to austenite to take place. However, as shown in Fig. 5, the residual strain obtained from both these simulations during the
deformation process at temperature ho will be fully recovered once the temperature is raised above 276:6 K > haf . Therefore,
our constitutive model is able to qualitatively reproduce the one-way austenite ! martensite ! austenite shape-memory
effect to good accord.
To investigate the non-isothermal behavior of shape-memory alloys during tensile superelastic deformation, we perform
the following fully-coupled thermo-mechanical simulations: At an initial temperature of 298 K, an initially-undeformed
shape-memory alloy sheet with dimensions of 5 mm by 20 mm by 1 mm is meshed using 200 ABAQUS C3D8RT elements
as shown in Fig. 6a. Each C3D8RT element has displacement and temperature degrees of freedom. The nodes on both ends
of the specimen in the 1-3 plane act as grip sections, and their temperature is kept fixed at 298 K throughout the duration of
the simulations i.e. the grips serve as a constant temperature bath. One grip section is constrained against motion along
direction-2 and the other grip section is deformed along direction-2 under strain-rates of 1  104 s1 (Simulation A) and
5  104 s1 (Simulation B). Furthermore, heat convection from the outer surfaces of the sheet to the ambient environment
(still air) is taken into account by setting the surface film heat transfer coefficient to be 12 W=m2 K.
The stress–strain response from these two simulations using the initially-undeformed finite-element mesh shown in
Fig. 6a are plotted in Fig. 6b. The contours of the martensite volume fraction in the sheet specimen obtained from Simulation
B keyed to different points on its corresponding stress–strain curve is shown in Fig. 7. Due to the boundary conditions

Martensite volume fraction


1.00
0.90
0.80
0.70
0.60
0.50
0.40
0.30
0.20
0.10
0.00

3 2
(a) (b)

(c) (d)

(e) (f)

(g) (h)

(i)
Fig. 7. Contours of the martensite volume fraction keyed to various points of the stress–strain curve obtained from Simulation B as shown in Fig. 6b. Due to
the boundary conditions, both the forward and reverse phase transformations initiate from the ends and move towards the center of the specimen.
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1209

imposed on the specimen as explained above, the austenite–martensite phase boundaries propagate from the grip sections
to the specimen’s center during the forward loading and reverse loading process. The contour plots presented in Fig. 7 show
the possibility of multiple austenite–martensite phase transformation fronts propagating in the specimen during superelas-
tic deformation (cf. Shaw and Kyriakides, 1997).
Referring back to the stress–strain curves plotted in Fig. 6b, Simulation B exhibits these following trends compared to
Simulation A: (a) a wider hysteresis loop (measured along the stress axis), (b) a significantly larger hardening in the
stress–strain response during the forward loading process; and (c) a significantly larger softening in the stress–strain re-
sponse during the reverse loading process.
The causes for these observed trends are as follows: Simulation A was conducted at a deformation rate which results in a
nearly isothermal response i.e. the austenite $ martensite phase transformations occur at nearly constant stress plateaus.
However, Fig. 8 shows the contours of the temperature in the sheet specimen obtained from Simulation B keyed to different
points on its corresponding stress–strain response shown in Fig. 6b i.e. Simulation B was conducted at a strain-rate which
results in a non-isothermal temperature field in the specimen during phase transformations. As shown in Fig. 8, the temper-
ature in the mid-section of the sheet increases by as much as 16 K above the ambient temperature (298 K) during the for-
ward loading process. At a strain-rate of 5  104 s1 the heat generated due to the release of the latent heat and mechanical

Temperature [K]
319.0
315.3
311.6
307.9
304.2
300.5
296.8
293.1
289.4
285.7
282.0

3 2
(a) (b)

(c) (d)

(e) (f)

(g) (h)

(i)
Fig. 8. Contours of the temperature keyed to various points of the stress–strain curve obtained from Simulation B as shown in Fig. 6b. During forward
loading, the temperature in the specimen increases by approximately 16 K above the ambient temperature of 298 K due to the release of latent heat as a
result of the austenite to martensite phase transformation. During reverse loading, the temperature in the specimen decreases by approximately 14 K below
the ambient temperature of 298 K due to the absorption of latent heat as a result of the martensite to austenite phase transformation.
1210 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

dissipation is not conducted and convected out of the specimen quickly enough, and this results in the increase of the spec-
imen temperature with respect to the ambient temperature. Thus, it is the increase in temperature which causes the hard-
ening in the stress–strain response during the forward loading process as shown in Fig. 6b.
Conversely, as also shown in Fig. 8 the temperature in the mid-section of the sheet decreases to below the ambient tem-
perature during the reverse loading process. At this deformation-rate i.e. 5  104 s1 the heat loss in the specimen due to
the absorption of the latent heat outweighs the amount of heat conducted and convected into the specimen, and hence
causes a reduction in the specimen’s mid-section temperature by as much as 14 K below the ambient temperature
(298 K). Therefore, it is this decrease in temperature which causes the softening in the stress–strain response during the re-
verse loading process as shown in Fig. 6b.
As mentioned previously, finite-deformation-based constitutive models have great utility in studying the deformation
behavior of flexible structures experiencing large deformations (Auricchio and Taylor, 1997; Auricchio, 2001). To exhibit
the capability of our constitutive model, we model the deformation behavior of a stent unit cell with its initial geometry
shown in Fig. 9a (Pan et al., 2007). Fig. 9b shows the initially-undeformed finite-element mesh of the stent unit cell shown
in Fig. 9a using 1304 ABAQUS C3D8R elements. The boundary conditions for the stent unit cell are as follows: the nodes on
the bottom face are prevented from motion whereas a displacement profile along direction-2 is prescribed for all the nodes
on the top face. All the nodes which make up the stent section have an initial temperature of 298 K and the finite-element
simulation was conducted under isothermal conditions using the material parameters listed in Table 1.
Fig. 10 shows the force vs. displacement curve obtained from the loading–unloading simulation performed on the stent
unit cell. The total force is obtained from the summation of the reaction forces along direction-2 for the nodes making up the
top surface. The force vs. displacement response resembles the typical uniaxial superelastic stress–strain curves exhibited by
shape-memory alloys. From the conducted numerical simulations, we can determine the contours of the martensite volume
fraction within the stent unit cell at any given point of the deformation process. Fig. 10a and b respectively show the mar-
tensite volume fraction contours keyed to points a and b on the force vs. displacement curve plotted in Fig. 10. As expected,
the stent unit cell will revert back to the fully-austenitic state upon the full reversal of the deformation back to the initial
state (cf. Fig. 10a). The numerical simulation is also able to predict the regions with the highest concentration of martensite
volume fraction within the stent unit cell in the deformed state (cf. Fig. 10b).

(a) Grip section


1.50

8.00
11.00 R 2.00

R 0.50
0.50

R 2.00

R 0.90

2
unit=mm

Stent section
1

Grip section 10.00

(b)
Top face

3
2

Bottom face

Fig. 9. (a) Specimen geometry for a stent unit cell taken from the work of Pan et al. (2007). The stent has a thickness of 0.38 mm. (b) Undeformed mesh of
the tested section of the stent unit cell using 1304 ABAQUS C3D8R elements. Direction-2 denotes the loading axis.
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1211

20

18
FEM SIMULATION
16 b

14

FORCE [N]
12

10

0
a0 0.5 1 1.5 2 2.5 3
DISPLACEMENT [mm]
Martensite volume fraction
1.00
0.90
0.80
0.70
0.60
0.50
0.40
0.30
0.20
0.10
0.00

3 1
(a) (b)
Fig. 10. The numerically-determined force vs. displacement response of the stent unit cell shown in Fig. 9a under superelastic deformation. Contours of
martensite volume fraction within the stent unit cell (a) in the initial state and upon full reversal of the deformation to the initial position i.e. point a on the
force vs. displacement curve shown above, and (b) keyed to point b on the force vs. displacement curve shown above.

4. Phase boundary propagation during superelastic deformation

To study the austenite–martensite phase boundary propagation in a shape-memory alloy undergoing superelastic
deformation, we perform the following simulations: Consider an initially-austenitic cuboidal section as shown in
Fig. 11a with an initially-undeformed dimensions of 2 mm by 10 lm by 10 lm measured along direction-1, direction-2
and direction-3, respectively. This cuboidal section is meshed using 100 and 600 ABAQUS C3D8R elements along direc-
tion-1. The nodes on the two ends of the cuboidal specimen in the 2–3 plane serve as the grip sections (grip section A
and B). Sites of geometrical imperfection have also been introduced in both the grip sections. Grip section A is prevented
from motion along direction-1 whereas a deformation profile is imparted on grip section B along direction-1. The finite-
element simulations were performed under isothermal conditions with the specimen temperature maintained at 298 K
throughout the duration of the simulations. Finally, for these set of simulations we use the values for the material param-
eters listed in Table 1 except for: (1) a very small energetic interaction coefficient h of 0:24 MJ=m3 introduced to accel-
erate the localization process, and (2) the value of sn to be set at 1:25  103 J=m which will correspond to an
experimentally-determined austenite–martensite interface thickness of approximately 100 l (Sun and Li, 2002).8 For
the two different mesh densities described above, we have also performed simulations with sn ¼ 0 J=m which collapses
our constitutive model to a local theory.

8
Although the interface energy used in this work is about two orders of magnitude larger than that calculated by Levitas et al. (2003), the main purpose of
this Section is to show that the non-local version of our phenomenological theory is able to eliminate the dependence of the austenite–martensite interface
thickness on mesh density.
1212 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

(a)
3

Grip section B

1 2

Grip section A

(b) 700

100 elements : Local


600 600 elements : Local
100 elements : Non-local
600 elements : Non-local
500
STRESS [MPa]

a
400

300

200
b

100

0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
STRAIN

Fig. 11. (a) A cuboidal specimen with initially-undeformed dimensions of 2 mm by 10 lm by 10 lm measured along direction-1, direction-2 and direction-
3, respectively. (b) The simulated tensile stress–strain curves using the local ðsn ¼ 0 J=mÞ vs. the non-local ðsn ¼ 1:25  103 J=mÞ version of theory. The
simulations were conducted on the cuboidal specimen shown in Fig. 11a meshed uniformly along direction-1 using 100 and 600 ABAQUS C3D8R elements.
All the calculations were conducted at a temperature of 298 K under isothermal conditions.

Fig. 11b shows the calculated superelastic tensile stress–strain curves for the above described simulations. The stress–
strain responses for the finite-element simulations conducted on the cuboidal section shown in Fig. 11a meshed using
the two different mesh densities for both the local vs. non-local version of the constitutive theory are equivalent.
Fig. 12a and b show the contours of the martensite volume fraction for the simple tension simulation using the local the-
ory i.e. sn ¼ 0 J=m conducted on the cuboidal specimen section shown in Fig. 11a meshed using 100 and 600 ABAQUS C3D8R
elements keyed to points a and b, respectively on the corresponding stress–strain curves plotted in Fig. 11b. These contour
plots are drawn in the reference i.e. undeformed configuration. During the forward loading process, the austenite–martensite
interface9 propagates from grip section A towards grip section B where the austenitic phase is gradually transforming into the
martensitic phase. However, during the reverse loading process the austenite–martensite interface propagates from grip section
B towards grip section A where the martensitic phase is now gradually transforming back into the austenitic phase. This prop-
agation pattern is enforced numerically through the sites of geometrical imperfections introduced into the specimen at the grip
ends.
During forward transformation, Fig. 12a shows that the austenite–martensite interface thickness predicted using the two
different mesh densities to be significantly different. Similarly, during reverse transformation, Fig. 12b shows that the aus-
tenite–martensite interface thickness predicted using the two different mesh densities to also be significantly different.
Hence the local theory ðsn ¼ 0 J=mÞ is predicting different austenite–martensite interface thickness with varying mesh den-
sity i.e. its calculated position is heavily dependent on the mesh density.
From the simulations conducted using the non-local theory i.e. sn ¼ 1:25  103 J=m, Fig. 13a and b show the contours
of the martensite volume fraction obtained from the simple tension simulation conducted on the cuboidal specimen sec-
tion shown in Fig. 11a meshed using 100 and 600 ABAQUS C3D8R elements keyed to points a and b, respectively, on the

9
The austenite–martensite interface is defined as a region where a mixture of the austenite and martensite phases are present.
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1213

Martensite volume fraction


1.000
0.917
0.833
0.750
0.667
0.583
0.500
0.417
0.333
0.250
0.167
0.083
0.000

100 elements

(a) 600 elements

3 1

100 elements

600 elements
(b)
Fig. 12. The contours of the martensite volume fraction keyed to points a and b on the stress–strain curve shown in Fig. 11b during (a) forward
transformation, and (b) reverse transformation obtained from the simulations conducted using the local version of the theory with sn ¼ 0 J=m. The contours
are plotted on the undeformed mesh.

corresponding stress–strain curves plotted in Fig. 11b. During forward transformation, Fig. 13a shows that the austenite–
martensite interface thickness predicted from the simulations conducted using the two different mesh densities to be
approximately equal. Hence the non-local version of the constitutive theory is predicting equal austenite–martensite inter-
face thickness during forward transformation regardless of mesh density. Similarly, during reverse transformation, Fig. 13b
shows that the austenite–martensite interface thickness predicted from the simulations conducted using the two different
mesh densities to also be approximately equal. Thus, the non-local version of the constitutive theory is also predicting
equal austenite–martensite interface thickness during reverse transformation independent of mesh density. From the re-
sults shown in Fig. 13 we can see that at a given deformation level during superelasticity, the non-local version of the
theory predicts the same position for the austenite–martensite interface regardless of mesh density. A more detailed com-
parison of the martensite volume fraction contours in the vicinity of the austenite–martensite interface is shown in
Fig. 14.
Also note that from the non-local version of the theory, the simulated austenite–martensite interface thicknesses
shown in Fig. 14 during reverse loading is larger than the calculated austenite–martensite interface thicknesses during
forward loading. This trend can be explained as follows: From Fig. 11b, we can see that the stress to cause forward
1214 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

Martensite volume fraction


1.000
0.917
0.833
0.750
0.667
0.583
0.500
0.417
0.333
0.250
0.167
0.083
0.000

100 elements

(a) 600 elements

3 1

100 elements

600 elements
(b)
Fig. 13. The contours of the martensite volume fraction keyed to points a and b on the stress–strain curve shown in Fig. 11b during (a) forward
transformation, and (b) reverse transformation obtained from the simulations conducted using the non-local version of the theory with
sn ¼ 1:25  103 J=m. The contours are plotted on the undeformed mesh.

transformation is larger than the stress to cause reverse transformation. Hence the resistance to forward transformation,
sþ is larger than the resistance to reverse transformation, s . Since the thickness of the austenite–martensite interface
scales with s1=2þ and s1=2
 during forward and reverse transformation, respectively, it follows that the thickness of
the austenite–martensite interface is larger during the reverse transformation process as sn and h are treated as con-
stants at all times.
It is a well-known fact that local theories are not able to eliminate mesh sensitivity in regards to the prediction of interface
thicknesses during deformation.10 From the contour plots shown in Fig. 14 we can conclude that for the mesh densities used
for the calculations in this section, a non-zero gradient energy i.e. a non-local theory with sn –0 heavily minimizes the effect of
mesh density on the prediction of the austenite–martensite interface thicknesses. Hence the position of the austenite–mar-
tensite interface region can now be accurately tracked during phase transformations without the aid of jump conditions.

5. Conclusion

A thermo-mechanically-coupled, non-local and isotropic-plasticity-based constitutive model for shape-memory alloys


capable of undergoing austenite $ martensite phase transformations has been developed with the aid of standard balance
laws, thermodynamic laws and the principle of micro-force balance (Fried and Gurtin, 1994). The constitutive model has
been implemented in the ABAQUS/Explicit (2009) finite-element program by writing a user-material subroutine.

10
For instance, see the recent works of Borg (2007) and Lele and Anand (2009) on shear banding in metallic alloys.
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1215

Martensite vol. frac.


1.000
0.833
0.667
0.500
0.333 100 elements (forward loading) 100 elements (reverse loading)
0.167
0.000
vs. vs.

600 elements (forward loading) 600 elements (reverse loading)


(a)

100 elements (forward loading) 100 elements (reverse loading)

vs. vs.

600 elements (forward loading) 600 elements (reverse loading)


(b)
Fig. 14. The contours of the martensite volume fraction in the vicinity of the austenite–martensite interface obtained from the simulations conducting
using the (a) local version of the constitutive theory i.e. sn ¼ 0 J=m, and (b) non-local version of the constitutive theory i.e. sn ¼ 1:25  103 J=m.

By calibrating the material parameters in the constitutive model to the superelastic stress–strain responses of represen-
tative physical experiments i.e. tension, compression and simple shear, exotic behavior exhibited by SMAs such as the stress-
biased strain–temperature-cycling and the shape-memory effect are also qualitatively well-reproduced by the theory. We
have also shown that our present theory is able to model the coupled thermo-mechanical response during superelastic
deformation and also qualitatively model the reversible behavior of a stent unit cell undergoing superelastic response.
Finally, we also show that the non-local version of our theory i.e. with sn –0 allows the accurate tracking of the austenite–
martensite interface motion during superelastic deformation independent of mesh density. Hence the exact position(s) of
the austenite–martensite interface can be determined without the aid of jump-conditions.
Some directions for future work include: (a) the prediction of complicated multi-axial and coupled thermo-mechanical
experiments (e.g. Tokuda et al. (1999, 2002)) using our developed constitutive model, and (b) extending the theory to model
martensitic reorientation/detwinning which may occur during superelasticity under non-proportional loading conditions.

Acknowledgements

The financial support for this work was provided by the Ministry of Science, Technology and Innovation, Malaysia under
Grant 03-01-02-SF0257. The ABAQUS finite-element software was made available under an academic license from HKS, Inc.
Pawtucket, R.I.

Appendix A. Time-integration procedure

In this appendix we summarize the explicit time-integration procedure that we have developed for our constitutive mod-
el presented in Section 2. With t denoting the current time, Dt is an infinitesimal time increment, and s ¼ t þ Dt. The algo-
rithm is as follows:
1216 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

Given: (1) fFðtÞ; FðsÞ; hðtÞ; hðsÞg11; (2) fTðtÞ; Fp ðtÞg; (3) fBðtÞ; N1 ðtÞ; N2 ðtÞ; /ðtÞg; (4) the martensite volume fraction nðtÞ. At
time t ¼ 0, we initialize BðtÞ ¼ 0.
Calculate: (a) fTðsÞ; Fp ðsÞg, (b) fBðsÞ; N1 ðsÞ; N2 ðsÞ; /ðsÞg, (c) the martensite volume fraction nðsÞ, and march forward in
time.

The steps used in the calculation procedure are:

Step 1. Calculate the trial elastic strain Ee ðsÞtrial :

Fe ðsÞtrial ¼ FðsÞFp ðtÞ1 ;


Ce ðsÞtrial ¼ ðFe ðsÞtrial Þ> Fe ðsÞtrial ;
Ee ðsÞtrial ¼ ð1=2ÞðCe ðsÞtrial  1Þ:

Step 2. Calculate the trial stress T ðsÞtrial :

T ðsÞtrial ¼ 2lEe0 ðsÞtrial þ j½traceEe ðsÞtrial  3ath ðhðsÞ  ho Þ1: ð48Þ

Step 3. Calculate the trial driving forces fi ðsÞtrial . In our explicit numerical algorithm presented here, we approximate12

Ni ðsÞ Ni ðtÞ; /ðsÞ /ðtÞ and r2 nðsÞ r2 nðtÞ: ð49Þ



For infinitesimal elastic stretches, we can also use the approximation TðsÞ T ðsÞ. Hence, the trial driving forces for phase
transformation are then given by
rffiffiffi
trial 3 kT
fi ðsÞ ¼ ð1 þ a/ðtÞÞ½T0 ðsÞtrial  Ni ðtÞ þ sn ½r2 nðtÞ  ðhðsÞ  hT Þ  hnðtÞ: ð50Þ
2 hT

Step 4. Determine the set PA of potentially active transformation systems which satisfy

f1 ðsÞtrial  fc > 0 and 0 6 nðtÞ < 1


for forward transformation, and

f2 ðsÞtrial þ fc < 0 and 0 < nðtÞ 6 1


for reverse transformation.

Step 5. Using the approximations given in Eq. (49), we calculate


( rffiffiffi )
3 X
Fp ðsÞ ¼ 1þ ð1 þ a/ðtÞÞ Dnj Nj ðtÞ Fp ðtÞ where j ¼ 1; . . . ; Q: ð51Þ
2 j2PA

Here Q 6 2 is the total number of potential transformation systems. Of the Q potentially active systems in the set PA,
only a subset A with elements M 6 Q , may actually be active (non-zero increments). This set is determined in an iterative
fashion described below.
During phase transformation, the active transformation systems must satisfy the consistency conditions
f1 ðsÞ  fc ¼ 0 and=or f 2 ðsÞ þ fc ¼ 0 ð52Þ
for forward transformation and/or reverse transformation, respectively. Using Eqs. (48)–(51), it is straightforward to show
that
X
fi ðsÞ ¼ fi ðsÞtrial  Dnj ½3lð1 þ a/ðtÞÞ2 symðCe ðsÞtrial Nj ðtÞÞ  Ni ðtÞ þ hdij  ð53Þ
j2PA

where dij is the Kronecker delta. Substituting Eq. (53) into the consistency conditions (52) give

11
The quantities FðtÞ; FðsÞ; hðtÞ and hðsÞ are inputs provided by the ABAQUS (2009) finite-element program.
12
The Laplacian of the martensite volume fraction is calculated via a finite-difference scheme. At the free boundaries, we assume a Neumann-type boundary
condition for the martensite volume fraction i.e rn  n ¼ 0.
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1217

X
Aij xj ¼ bi ; i 2 PA; ð54Þ
j2PA

with

Aij ¼ 3lð1 þ a/ðtÞÞ2 Ni ðtÞ  symðCe ðsÞtrial Nj ðtÞÞ þ hdij :


We also have

b1 ¼ f1 ðsÞtrial  fc > 0 and x1  Dn1 > 0


for forward transformation, and

b2 ¼ f2 ðsÞtrial þ fc < 0 and x2  Dn2 < 0


for reverse transformation.
Eq. (54) is a system of linear equations for the transformation increments xj  Dnj (for j 2 PA). Assuming the matrix A to
be invertible i.e. non-singular, the transformation rates are determined by

x ¼ A1 b ð55Þ
1
where A is the inverse matrix of the matrix A. If x1 6 0 when b1 > 0 (forward transformation), then this system is inactive
and it is removed from the set of potentially active systems PA and a new matrix A is calculated. Similarly, if x2 P 0 when
b2 < 0 (reverse transformation), then this system is also inactive and it is not included in the set PA used to determine the
new matrix A. The final size of the matrix A is M  M.

Step 6. Update the martensite volume fraction:


X
nðsÞ ¼ nðtÞ þ Dni ; i 2 A:
i

If nðsÞ > 1, then set nðsÞ ¼ 1. If nðsÞ < 0, then set nðsÞ ¼ 0.

Step 7. Update the inelastic deformation gradient Fp ðsÞ:


( rffiffiffi )
p 3 X
F ð sÞ ¼ 1þ ð1 þ a/ðtÞÞ Dnj Nj ðtÞ Fp ðtÞ:
2 j2A

Step 8. Compute the elastic strain Ee ðsÞ and the stress T ðsÞ:

Fe ðsÞ ¼ FðsÞFp ðsÞ1 ;


Ce ðsÞ ¼ Fe ðsÞ> Fe ðsÞ;
Ee ðsÞ ¼ ð1=2ÞðCe ðsÞ  1Þ;
T ðsÞ ¼ 2lEe0 ðsÞ þ j½traceEe ðsÞ  3ath ðhðsÞ  ho Þ1:

Step 9. Update the flow direction N1 ðsÞ and the J-3 parameter /ðsÞ:
  pffiffiffi
T0 ðsÞ
N1 ðsÞ ¼ T and /ðsÞ ¼ 6½N1 ðsÞ  ðN1 ðsÞÞ2 ðT Þ3 :
jT0 ðsÞj

Step 10. Update the tensors BðsÞ and N2 ðsÞ:


rffiffiffi

3 X BðsÞ
BðsÞ ¼ BðtÞ þ ð1 þ a/ðtÞÞ Dnj Nj ðtÞ and N2 ðsÞ ¼ T :
2 j2A
jBðsÞj

Step 11. Calculate the driving forces fi ðsÞ:


rffiffiffi
3 kT
fi ðsÞ ¼ ð1 þ a/ðsÞÞ T0 ðsÞ  Ni ðsÞ þ sn ðr2 nðsÞÞ  ðhðsÞ  hT Þ  hnðsÞ:
2 hT

Step 12. Calculate the inelastic work increment Dxp :


1218 P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219

  X X
hT
Dxp ¼ hðsÞ Dni  3jath hðsÞðtraceðDEe ÞÞ þ fi ðsÞDni
hT i2A i2A

where DEe ¼ Ee ðsÞ  Ee ðtÞ; Ee ðtÞ ¼ ð1=2ÞðCe ðtÞ  1Þ and Ce ðtÞ ¼ Fe ðtÞ> Fe ðtÞ. The inelastic work increment is treated as the heat
source which causes heating/cooling at a material point during deformation.

Step 13. Calculate the Cauchy stress TðsÞ:

TðsÞ ¼ ½det FðsÞ1 Fe ðsÞT ðsÞFe ðsÞ> :

References

Abaqus reference manuals, 2009. SIMULIA Inc., Providence, R.I.


Abeyaratne, R., Knowles, J., 1993. A continuum model for a themoelastic solid capable of undergoing phase transitions. Journal of the Mechanics and Physics
of Solids 41, 541–571.
Abeyaratne, R., Kim, S., Knowles, J., 1994. A one-dimensional continuum model for shape-memory alloys. International Journal of Solids and Structures 31,
2229–2249.
Aizawa, S., Kakizawa, T., Higashino, M., 1998. Case studies of smart materials for civil structures. Smart Materials and Structures 7, 617–626.
Anand, L., Gurtin, M., 2003b. A theory of amorphous solids undergoing large deformations, with applications to polymeric glasses. International Journal of
Solids and Structures 40, 1465–1487.
Anand, L., Gurtin, M., 2003a. Thermal effects in the superelasticity of crystalline shape-memory materials. Journal of the Mechanics and Physics of Solids 51,
1015–1058.
Auricchio, F., 2001. A robust integration-algorithm for a finite-strain shape-memory alloy superelastic model. International Journal of Plasticity 17, 971–
990.
Auricchio, F., Taylor, R., 1997. Shape-memory alloys: modelling and numerical simulations of the finite-strain superelastic behavior. Computer Methods in
Applied Mechanics and Engineering 143, 175–194.
Auricchio, F., Taylor, R., Lubliner, J., 1997. Shape-memory alloys: macromodelling and numerical simulations of the superelastic behavior. Computer
Methods in Applied Mechanics and Engineering 146, 281–312.
Borg, U., 2007. Strain gradient crystal plasticity effects on flow localization. International Journal of Plasticity 23, 1400–1416.
Boyd, J., Lagoudas, D., 1994. A thermodynamical constitutive model for the shape memory effect due to transformation and reorientation. Proceedings of
SPIE – the International Society for Optical Engineering 2189, 276–288.
Boyd, J., Lagoudas, D., 1996. A thermodynamical model for shape memory materials I. The monolithic shape memory alloy. International Journal of Plasticity
12, 805–842.
Brinson, L., 1993. One-dimensional constitutive behavior of shape-memory alloys: thermomechanical derivation with non-constant material functions and
redefined martensite internal variable. Journal of Intelligent Material Systems and Structures 4, 229–242.
Duerig, T., Pelton, A., Stoeckel, D., 1999. An overview of nitinol medical applications. Materials Science and Engineering A, 149–160.
Fang, D., Lu, W., Yan, W., Inoue, T., Hwang, K., 1999. Stress–strain relation of cualni sma single crystal under biaxial loading – consitutive model and
experiments. Acta Materialia 47, 269–280.
Fried, E., Gurtin, M., 1994. Dynamic solid–solid transitions with phase characterized by an order parameter. Physica D 72, 287–308.
Gall, K., Sehitoglu, H., 1999. The role of texture in tension–compression asymmetry in polycrystalline Ni–Ti. International Journal of Plasticity 15, 69–92.
Gurtin, M., Anand, L., 2005. The decomposition f ¼ f e f p , material symmetry, and plastic irrotationality for solids that are isotropic–viscoplastic or
amorphous. International Journal of Plasticity 21, 1686–1719.
Helm, D., 2007. Thermomechanics of martensitic phase transitions in shape memory alloys. Journal of Mechanics of Materials and Structures 2, 87–112.
Helm, D., Haupt, P., 2003. Shape memory behaviour: modelling within continuum thermomechanics. International Journal of Solids and Structures 40, 827–
849.
Idesman, A., Cho, J., Levitas, V., 2008. Finite element modeling of dynamics of martensitic phase transitions. Applied Physics Letters 93, 043102.
Jung, Y., Papadopoulos, P., Ritchie, R., 2004. Constitutive modelling and numerical simulation of multivariant phase transformation in superelastic shape-
memory alloys. International Journal for Numerical Methods in Engineering 60, 429–460.
Kröner, E., 1960. Allgemeine kontinuumstheorie der versetzungen und eigenspannungen. Archive of Rational Mechanics and Analysis 4, 273–334.
Lee, E., 1969. Elastic plastic deformation at finite strain. ASME Journal of Applied Mechanics 36, 1–6.
Lele, S., Anand, L., 2009. A large-deformation strain-gradient theory for isotropic viscoplastic materials. International Journal of Plasticity 25, 420–453.
Levitas, V., Lee, D., 2007. Athermal resistance to an interface motion in phase field theory of microstructure evolution. Physical Review Letter 99, 245701.
Levitas, V., Preston, D., 2002b. Three-dimensional Landau theory for multivariant stress-induced martensitic phase transformations. II. Multivariant phase
transformations and stress space analysis. Physical Review B 66, 134207.
Levitas, V., Preston, D., 2002a. Three-dimensional landau theory for multivariant stress-induced martensitic phase transformations I. Austenite–martensite.
Physical Review B 66, 134206.
Levitas, V., Preston, D., Lee, D., 2003. Three-dimensional landau theory for multivariant stress-induced martensitic phase transformations III. Alternative
potentials, critical nuclei, kink solutions, and dislocation theory. Physical Review B 68, 134201.
Levitas, V., Lee, D., Preston, D., 2009. Interface propagation and microstructure evolution in phase field models of stress-induced martensitic phase
transformations. International Journal of Plasticity. doi:10.1016/j.ijplas.2009.08.003.
Levitas, V., Levin, V., Zingerman, K., Freiman, E., 2009. Displacive phase transitions at large strains: phase-field theory and simulations. Physical Review
Letters 103, 025702.
Liang, C., Rogers, C., 1990. One-dimensional thermomechanical constitutive relations for shape memory materials. Journal of Intelligent Material Systems
and Structures 1, 207–234.
Lim, T., McDowell, D., 1999. Mechanical behavior of a Ni–Ti shape memory alloy under axial–torsional proportional and nonproportional loading. Journal of
Engineering Materials and Technology – Transactions of the ASME 121, 9–18.
Lu, Z., Weng, G., 1998. A self-consistent model for the stress–strain behavior of shape-memory alloy polycrystals. Acta Materialia 46, 5423–5433.
Machado, L., Savi, M., 2003. Medical applications of shape-memory alloys. Brazillian Journal of Medicine and Biological Research 36, 683–691.
Moumni, Z., Zaki, W., Nguyen, Q., 2008. Theoretical and numerical modeling of solidsolid phase change: application to the description of the
thermomechanical behavior of shape memory alloys. International Journal of Plasticity 24, 614–615.
Mueller, C., Bruhns, O., 2006. A thermodynamic finite-strain model for pseudoelastic shape memory alloys. International Journal of Plasticity 22, 1658–
1682.
Orgeas, L., Favier, D., 1998. Stress-induced martensitic transformation of Ni–Ti alloy in isothermal shear, tension and compression. Acta Materialia 46, 5579–
5591.
Otsuka, K., Wayman, C., 1999. Shape Memory Materials. Cambridge University Press.
P. Thamburaja / International Journal of Plasticity 26 (2010) 1195–1219 1219

Pan, H., Thamburaja, P., Chau, F., 2007. An isotropic-plasticity-based constitutive model for martensitic reorientation and shape-memory effect in shape-
memory alloys. International Journal of Solids and Structures 44, 7688–7712.
Patoor, E., Eberhardt, A., Berveiller, M., 1996. Micromechanical modelling of superelasticity in shape memory alloys. Journal de Physique IV 6, 277–292.
Peng, X., Pi, W., Fan, J., 2008. A microstructure-based constitutive model for the pseudoelastic behavior of NiTi smas. International Journal of Plasticity 24,
966–990.
Reese, S., Christ, D., 2008. Finite deformation pseudo-elasticity of shape memory alloys constitutive modelling and finite element implementation.
International Journal of Plasticity 24, 455–482.
Seelecke, S., Mueller, I., 2004. Shape memory alloy actuators in smart materials: modeling and simulation. Applied Mechanics Reviews 57, 23–46.
Shaw, J., Kyriakides, S., 1995. Thermomechanical aspects of NiTi. Journal of the Mechanics and Physics of Solids 43, 1243–1281.
Shaw, J., Kyriakides, S., 1997. On the nucleation and propagation of phase transformation fronts in a NiTi alloy. Acta Materialia 45, 683–700.
Sun, Q., Hwang, K., 1993. Micromechanics modelling for the constitutive behavior of polycrystalline shape memory alloys – I. Derivation of general
relations. Journal of the Mechanics and Physics of Solids 41, 1–17.
Sun, Q., Li, Z., 2002. Phase transformation in superelastic NiTi polycrystalline micro-tubes under tension and torsion – from localization to homogeneous
deformation. International Journal of Solids and Structures 39, 3797–3809.
Thamburaja, P., Anand, L., 2001. Polycrystalline shape-memory materials: effect of crystallographic texture. Journal of the Mechanics and Physics of Solids
49, 709–737.
Thamburaja, P., Anand, L., 2003. Thermo-mechanically coupled superelastic response of initially-textured Ti–Ni sheet. Acta Materialia 51, 325–338.
Thamburaja, P., Nikabdullah, N., 2009. A macroscopic constitutive model for shape-memory alloys: theory and finite-element simulations. Computer
Methods in Applied Mechanics and Engineering 198, 1074–1086.
Tokuda, M., Ye, M., Takakura, M., Sittner, P., 1999. Thermomechanical behavior of shape memory alloy under complex loading conditions. International
Journal of Plasticity 15, 223–239.
Tokuda, M., Sittner, P., Takakura, M., Haze, M., 2002. Multi-axial constitutive equations of polycrystalline shape memory alloy experimental background.
JSME International Journal Series A 45, 276–281.
Wang, X., Xu, B., Yue, Z., 2008. Micromechanical modelling of the effect of plastic deformation on the mechanical behaviour in pseudoelastic shape memory
alloys. International Journal of Plasticity 24, 1307–1332.
Zaki, W., Moumni, Z., 2007. A three-dimensional model of the thermomechanical behavior of shape memory alloys. Journal of the Mechanics and Physics of
Solids 55, 2455–2490.
Ziolkowski, A., 2007. Three-dimensional phenomenological thermodynamic model of pseudoelasticity of shape memory alloys at finite strains. Continuum
Mechanics and Thermodynamics 19, 379–398.

View publication stats

S-ar putea să vă placă și