You are on page 1of 14

Acta Mechanica Solida Sinica, Vol. 25, No.

3, June, 2012 ISSN 0894-9166


Published by AMSS Press, Wuhan, China

A CONSTITUTIVE MODEL FOR TRANSFORMATION,


REORIENTATION AND PLASTIC DEFORMATION OF
SHAPE MEMORY ALLOYS

Xianghe Peng1 Bin Chen1 Xiang Chen1 Jun Wang2 Huyi Wang2
1
( Department of Engineering Mechanics, Chongqing University, Chongqing 400044, China; State Key
Laboratory of Coal Mine Disaster Dynamics and Control, Chongqing University, Chongqing 400044, China)
(2 China Academy of Engineering Physics, Mianyang 621900, China)

Received 7 February 2012, revision received 28 March 2012

ABSTRACT A constitutive model is developed for the transformation, reorientation and plastic
deformation of shape memory alloys (SMAs). It is based on the concept that an SMA is a mixture
composed of austenite and martensite, the volume fraction of each phase is transformable with
the change of applied thermal-mechanical loading, and the constitutive behavior of the SMA is
the combination of the individual behavior of its two phases. The deformation of the martensite
is separated into elastic, thermal, reorientation and plastic parts, and that of the austenite is
separated into elastic, thermal and plastic parts. Making use of the Tanaka’s transformation rule
modified by taking into account the effect of plastic deformation, the constitutive model of the
SMA is obtained. The ferroelasticity, pseudoelasticity and shape memory effect of SMA Au-47.5
at.%Cd, and the pseudoelasticity and shape memory effect as well as plastic deformation and its
effect of an NiTi SMA, are analyzed and compared with experimental results.

KEY WORDS shape memory alloys, two-phase mixture, transformation, reorientation, plasticity,
constitutive model

I. INTRODUCTION
Because of the unique properties, such as pseudoelasticity (PE) and shape memory effects (SME),
shape memory alloys (SMAs) have been successfully used in various fields, such as aerospace, medical,
mechanical, chemical, electronic, transportation and energy industries, etc. Meanwhile, the interest
in the research of the thermal-mechanical behavior of SMAs has been rapidly growing. Besides the
conventional properties of SMAs, plastic behavior during or after forward martensitic transformation
and reorientation as well as its effects has attracted extensive attention.
It has been found that the plastic deformation after forward transformation could markedly stabilize
the deformed martensite and strongly affect the inverse transformation. A typical example is that when
SMA Ni47 Ti44 Nb9 was subjected to tensile strain of 16%, the austenite transformation start temperature
As and the austenite transformation finish temperature Af could increase from −65 ◦ C and −25 ◦ C to
over 40 ◦ C and around 70 ◦ C, respectively. With this property, the deformed parts made of the material at
low temperature can be stored at room temperature, instead of being stored at a low temperature, before
assembled and recovered by heating[1, 2] . In order to explain this phenomenon, Piao[3] investigated the
 Corresponding author. E-mail: xhpeng@cqu.edu.cn
 Project supported by the National Natural Science Foundation of China (No. 10976032).
· 286 · ACTA MECHANICA SOLIDA SINICA 2012

transformation behavior of several SMAs, and speculated that the elastic strain energy stored during the
forward transformation could be relaxed by plastic deformation. On the other hand, the failure of SMAs
is also important for the application of SMAs[4] . In practical applications, stress concentration may
inevitably exist in SMAs and SMA structures, which may induce local plastic deformation, therefore, the
interaction between irreversible plastic strain and recoverable transformation strain should be considered.
Strnade et al.[5] , Sehitoglu et al.[6] , and Gall and Maier[7] investigated plastic strain in stress induced
phase transformation of NiTi alloys and its effect on the following inverse transformation. Brinson et
al.[8] , and Miller and Lagoudas[9] studied the influence of plastic strain on the two-way shape memory
effect. In order to gain an insight into the effects of plastic deformation and the failure mechanism of
SMAs, it is important to investigate plastic deformation and its interaction with transformation. It
was reported by McKelvey and Ritchie[10] that plastic deformation after forward transformation could
stabilize the deformed martensite and increase the resistance to inverse transformation. It is known
that plastic deformation after the stress-induced martensitic transformation can be attributed to the
dislocation slip in the martensite[11, 12] . Wang et al.[4] performed recently tensile experiments on an
NiTi SMA with different strain amplitudes, and investigated the effects of plastic deformation after
phase transformation as well as plastic strain cycles on the phase transformation stress, the residual
and recoverable strain and the dissipated and recoverable energy density.
Constitutive modeling of SMAs has been an important topic of research and many significant pro-
gresses have been made during the past two decades. Many micromechanics or micro-macro models have
been proposed[13–20] , which are useful to gain insight into the mechanism and illustrate the fundamental
features of material behavior, but most of them are not easy to be applied in the analysis of practical
problems. Many phenomenological models have also been proposed[21–23] , in most of which the behavior
of SMAs is often described with strain, stress, temperature, entropy and a set of internal state variables,
e.g. martensite volume fraction, ξ, or macroscopic transformation strain. Tanaka et al.[24] proposed a set
of exponential functions to describe respectively the forward and inverse transformations, making use of
von-Mises equivalent stress and temperature. Liang and Rogers[25] proposed an evolution of ξ with cosine
functions. Brinson[26] suggested a different method to describe the transformation kinetics, in which ξ is
divided into two parts, determined by temperature and stress respectively. Peng et al.[27–29] proposed the
two-phase mixture models for SMAs, in which the Tanaka’s transformation rule was adopted. Leclercq
and Lexcellent[30] proposed a model, in which the volume fractions of self-accommodation (pure thermal
effect) and oriented (stress-induced) product phases were introduced as two internal variables. Popov
and Lagoudas[31] presented a model taking into account both the direct conversion of the austenite to
the detwinned martensite and the detwinning of self-accommodated martensite. In the model by Bo
and Lagoudas[32], a set of internal variables were adopted, including the martensite volume fraction,
the macroscopic transformation strain, the back stress and the drag stress due to both martensitic
transformation and its interaction with plastic strain. Panico and Brinson[33] also proposed a model,
which could account for the evolution of both twinned and detwinned martensite. In the model by
Dong and Sun[34] and He and Sun[35] , the pattern and evolution of domain were considered. In the
model proposed by Souza et al.[36] , both SME and PE in a small strain regime could be captured,
and the macroscopic kinetics consequences of stress-induced transformation could be described with a
transformation strain tensor. Kan et al.[37] developed a pseudoelastic model considering plasticity. Peng
et al.[38] discussed the stress-induced transformation and the microstructure evolution in a Cu-based
SMA.
Although great progress has been made in constitutive modeling for SMAs, the modeling involving
plastic strain and its effect on the following inverse transformation has less been investigated. In this
article, a two-phase mixture model is developed for the transformation, reorientation and plastic behavior
of SMAs based on a thermomechanically consistent mechanical model, and making use of the Voigt
estimation and the Tanaka’s transformation rule modified by introducing the effect of the maximum
equivalent plastic strain. The corresponding numerical algorithm is developed and the constitutive
behavior of SMAs, including FE, PE, SME and plasticity as well as their effects, is described and
compared with experimental results.
Vol. 25, No. 3 Xianghe Peng et al.: A Constitutive Model of Shape Memory Alloys · 287 ·

II. A TWO-PHASE MIXTURE MODEL FOR SMAS


An SMA is considered as a two-phase mixture composed of martensite and austenite, and the property
of the material is the combination of that of its two phases. When an SMA is subjected to thermal-
mechanical loading, the property and the volume fraction of each phase may change, which determines
the amazing properties of the material[27] .
In order to validate the main property of an SMA, such as FE, PE, SME, plasticity and their effects,
we assume that, in the interested ranges of stress and temperature, the austenite, as the parent phase,
should be thermal-elastoplastic; while the deformation in the martensite consists of thermal, elastic,
reorientation (denoting the reorientation or detwinning of the martensite variants) and plastic parts.
As an initial research, for simplicity and easy application, the following assumptions are adopted:
(a) The transformation lattice volume change is ignored[13, 14] , and the two phases are inelastically
incompressible;
(b) The material of each phase is initially isotropic;
(c) The Voigt estimation may provide acceptable approximation for the response of the mixture,
therefore, the strain in each phase is identical to the overall strain in the mixture, i.e., ε = εm = εa ,
where ε, εm and εa denote the strains in the mixture, the martensite and the austenite, respectively.
It should be emphasized that the reorientation deformation mentioned above corresponds to the
deformation caused by the reorientation (or detwinning) of martensite variants, instead of transforma-
tion. As was suggested by Leclercq and Lexcellent[30] and Popov and Lagoudas[31], we distinguish the
reorientation deformation from transformation, because pure forward/inverse martensitic transforma-
tion (such as temperature-induced forward and inverse transformation or ‘self-accommodation due to
pure thermal effect’) does not necessarily result in significant macroscopic deformation, on the other
hand, pure reorientation (or purely stress-induced) deformation is not necessarily related to macroscopic
transformation (such as ferroelastic deformation at a sufficient low temperature when stress can not
induce any macroscopic transformation).
Letting εm consist of elastic strain εme , thermal strain εmθ , reorientation strain εmr , and plastic
strain εp ; and εa consist of elastic strain εae , thermal strain εaθ and plastic strain εp , respectively, i.e.,
in the case of small deformation,
ε = εme + εmr + εp + εmθ , ε = εae + εp + εaθ (1)
where
εmθ = αm (θ)(θ − θ0 )I 2 , εaθ = αa (θ)(θ − θ0 )I 2 (2)
m a
α and α are respectively the linear thermal expansion coefficients of the martensite and the austenite,
θ and θ0 the current temperature and the reference temperature, and I 2 is the identity tensor of rank
two.
It should be stressed that in Eq.(1) we use the same plastic strain εp in the expressions for both
phases. It is based on the fact that the plastic strain is irreversible and irrecoverable, once plastic
deformation takes place in one phase, it would not only exist in this phase but would be taken over by
the other phase after transformation. For example, the plastic strain occurring in the martensite would
be taken over by its parent phase after inverse transformation; and vise versa.
The combination of Eq.(1) with
1
ε = e + εkk I 2 (3)
3
where e is the deviator of ε and εkk = ε : I 2 is the volumetric strain, yields
e = eme + emr + ep , e = eae + ep
(4)
εkk = εme mθ
kk + εkk , εkk = εae aθ
kk + εkk

where eme and emr are the elastic deviatoric strain and reorientation strain in the martensite, εme kk
and εmθ
kk are respectively the elastic bulk strain and thermal strain in the martensite; e
ae
is the elastic
deviatoric strain in the austenite, εae aθ
kk and εkk are the elastic bulk strain and thermal strain in the
p
austenite, and e is the plastic strain in both phases.
Making use of the Voigt estimation, we have
σ = ξσ m + (1 − ξ)σ a (5)
· 288 · ACTA MECHANICA SOLIDA SINICA 2012

where ξ is the volume fraction of martensite, and σ, σ m and σ a are the stresses in the material, the
martensite and the austenite, respectively. Making use of σ = s + σkk I 2 /3, where s and σkk = σ : I 2
are the deviator and the bulk component of σ, one obtains the following relations from Eq.(5):
s = ξsm + (1 − ξ)sa , m
σkk = ξσkk a
+ (1 − ξ)σkk (6)
m m a a
where s and σkkare the deviatoric stress tensor and bulk stress in the martensite, and s and σkk
are those in the austenite.
The deviatoric stress in the martensite can be determined with
sm = 2Gm (θ)eme = 2Gm (θ)(e − emr − ep ) (7)
where Gm is the elastic shear modulus. The relationship between the reorientation strain, emr , and
the stress in the martensite, sm , can be obtained with the thermomechanically consistent mechanical
model shown in Fig.1[39] . It consists of l branches, in which the jth branch consists of a spring Djm
(with stiffness Djm ) and a joint-like component bm m
j (with reorientation damping-like coefficient bj ). The
introduction of the joint-like components is based on the following considerations: (a) To distinguish
reorientation deformation from plastic or viscous deformation, and (b) To express the deformation
due to the orientation of martensite variants in a visual manner. bm j (j = 1, 2, · · · , l) describes the
dissipation during reorientation, the change of the displacement of bm j , dp
m(j)
, and the corresponding
m(j) m(j) m(j)
generalized force, P , satisfies dissipation inequality P : dp ≥ 0 (j = 1, 2, · · · , l). The energy
stored in Djm corresponds to that stored in the microstress fields determined by the respective pattern
of defects on the microlevel, such as inhomogeneous microstructures due to different deformation in
different martensite variants, and the interfaces between variants and between grain boundaries, etc.
From Fig.1, we have
l l
sm = P m(j) , dsm = dP m(j) (8)
j=1 j=1

Fig. 1. Mechanical model for reorientation response of martensite.

The response of Djm can be expressed as


P m(j) = Djm (θ)(emr − pm(j) ) (j = 1, 2, · · · , l) (9)
We simply assume that P m(j) is proportional to the derivative of pm(j) with respect to the generalized
time for reorientation deformation, z mr, i.e.,
dpm(j)
P m(j) = bm
j (θ) (j = 1, 2, · · · , l) (10)
dz mr
where
dζ mr √
dz mr = mr mr
, dζ mr = demr : demr (11)
f (z )
in which f mr (z mr ) is related to the change in the material properties that may occur during reorientation.
If dz mr = 0, the combination of Eqs.(10) and (11) yields:
(Djm ) m(j)
dP m(j) = Djm (θ)demr − λm
j P
m(j)
dz mr + P dθ (12)
Djm
Vol. 25, No. 3 Xianghe Peng et al.: A Constitutive Model of Shape Memory Alloys · 289 ·

where λm m m
j = Dj /bj . Substituting Eq.(12) into Eq.(8), one has

dsm = Amr demr − B mr dz mr + Dmr dθ (13)

where
l
 l
 l  m 
mr mr m(j) mr Dj m(j)
A = Djm (θ), B = λm
j P , D = m P (14)
j=1 j=1 j=1
D j

The reorientation strain should reach a maximum value once the reorientation (or detwinning) of
martensite variants finishes, then the reorientation strain stops developing even if the applied stress
increases, indicating that there should exist a limit emr 0 , with which the reorientation strain can be
expressed as  mr
de , if emr mr
eq ≤ e0
< demr >= (15)
0, if eeq > emr
mr
0

where emr
eq = emr : emr . Using the above defined < demr > to replace demr in the differential form of
Eq.(7), one has

Gm
dsm = 2Gm (θ) [de− < demr > −dep ] + m sm dθ (16)
G
The relationship between the plastic strain ep and the applied stress deviator s can be obtained with
the thermomechanically consistent mechanical model shown in Fig.2[39] . It consists of n branches, in
which the jth branch consists of the spring Cj (with stiffness Cj ) and dashpot-like block aj (with plastic
damping-like coefficient aj ). aj (j = 1, 2, · · · , n) describes the dissipation during plastic deformation,
the change in the displacement of aj , dp(j) , and the corresponding generalized force, Q(j) , satisfies Q(j) :
dp(j) ≥ 0 ( j = 1, 2, · · · , n). The energy stored in Cj corresponds to that stored in the microstress fields
determined by the respective pattern of defects on the microlevel, such as dislocations and dislocation
substructures, etc. From Fig.2, we have
n
 n

s= Q(j) , ds = dQ(j) (17)
j=1 j=1

Fig. 2. Mechanical model for plastic response of SMA.

The response of Cj follows:

Q(j) = Cj (θ)(ep − q (j) ) (j = 1, 2, · · · , n) (18)

We simply assume that Q(j) is proportional to the derivative of q (j) with respect to the generalized
time for plastic deformation, z p , i.e.,

dq (j)
Q(j) = aj (θ) (j = 1, 2, · · · , n) (19)
dz p
where
dζ p √
dz p = , dζ p = dep : dep (20)
f p (z p )
· 290 · ACTA MECHANICA SOLIDA SINICA 2012

f p (z p ) describes the hardening (or softening) induced by plastic deformation. From Eqs.(18) and (19)
we have
(Cj ) (j)
dQ(j) = Cj (θ)dep − αj Q(j) dz p + Q dθ (21)
Cj
where αj = Cj /aj . Substituting Eq.(21) into Eq.(17) gives
ds = Ap dep − B p dz p + D p dθ (22)
where
n
 n
 l
 
(Cj )
Ap = Cj (θ), Bp = αj Q(j) , Dp = Q(j) (23)
j=1 j=1 j=1
Cj
With the derived ep and dep , the stress in the austenite can be obtained as
sa = 2Ga (θ)eae = 2Ga (θ)(e − ep ) (24)
a
where G (θ) is the elastic shear modulus of the austenite. The differential form of Eq.(24) is

(Ga ) a
dsa = 2Ga (θ)(de − dep ) + s dθ (25)
Ga
The bulk stress-strain response of the martensite can be expressed as
m
σkk = 3K m (θ)εme m mθ
kk = 3K (θ)(εkk − εkk ) (26)
m
where K (θ) is the bulk modulus of the martensite. The differential form of Eq.(24) is

m
  (K m ) m
dσkk = 3K m(θ) dεkk − dεmθkk + σ dθ (27)
K m kk
Similarly, the bulk stress-strain response of the austenite and its differential form can be expressed
respectively as  
a
σkk = 3K a (θ)εae a
kk = 3K (θ) εkk − εkk

(28)
and
a
  K a a
dσkk = 3K a (θ) dεkk − dεaθkk + σ dθ (29)
K a kk
In Eq.(6), ξ can be determined with the following modified Tanaka’s transformation rule[24] :

exp [−Aa θ − [As + B a σe + g (ε̄pmax )]] M → A
ξ= (30)
1 − exp [−Am Ms + B m σe − θ] A→M

where σ e = (3/2)s : s is the equivalent stress, Aa , Am , B a and B m are material constants related
to forward and inverse transformation, and ε̄pmax = max(ε̄p ), in which ε̄p = (2/3)ep : ep . It can be
seen that Eq.(30) takes into account the effect of plastic deformation on the inverse transformation
by introducing an additional term g(ε̄pmax ). If we choose the simplest linear relation g(ε̄pmax ) = hε̄pmax ,
(h > 0), and define As = As + hε̄pmax , we can see that As increases with the increase of ε̄pmax . On the
other hand, fixing the temperature θ, the σe required for the inverse transformation may decrease with
the increase of g(ε̄pmax ) or ε̄pmax , as observed by many researchers in experiments.
In order to describe the temperature-induced inverse transformation (e.g., SME), we define a macro-
scopic average reorientation strain ēmr = ξemr and assume demr /dξ = β  emr /ξ = β  ēmr /ξ 2 . Making
use of Eq.(30), we have
ēmr
dēmr = emr dξ + ξdemr = β dξ = −βAa ēmr < dθ > (31)
ξ
where β = 1 + β  , and

dθ if dθ > 0 and As ≤ θ ≤ Af
< dθ >= (32)
0 otherwise
Suppose ēmr∗ is the macroscopic average residual reorientation strain at the onset of heating, the
recovery of the ēmr∗ will take place as θ > As . Assuming ēmr = 0.001ēmr∗ at θ = Af , it can be obtained
from Eqs.(31) and (32) that βAa = 6.9078/(Af − As ), where As and Af are the austenitic transformation
start and finish temperatures, respectively, taking into account the effect of plastic deformation.
Vol. 25, No. 3 Xianghe Peng et al.: A Constitutive Model of Shape Memory Alloys · 291 ·

III. INCREMENTAL FORM OF THE MODEL


It has been shown that a direct application of the differential form of the constitutive model may
involve remarkable errors[40]. In order to make the numerical process more accurate, the incremental
form of Eqs.(12) and (13) can be obtained by following the approach suggested by Peng and Fan[40] as
 
m(j) mr m mr m m(j) mr
(Djm ) m(j)
P k+1 = kj Dj (θ) ek+1 − λj P k zk+1 + P k θk+1 (33)
Djm
l
 m(j) mr mr
sm
k+1 = P k+1 =Āmr mr mr
k ek+1 − B̄ k zk+1 + D̄ k θk+1 (34)
j=1

with mr
1 − ez̄j (Djm )
kjmr = , z̄jmr = λm mr
j zk+1 + θk+1 (35)
z̄jmr Djm
l l l  m 
 mr
 m(j) mr
 Dj m(j)
Āmr
k = kjmr Djm , B̄ k = kjmr λm
j Pk , D̄ k = mr
kj m Pk (36)
j=1 j=1 j=1
D j

in which the quantities with subscript k denote those of or after the kth increment of loading.
Similarly, the incremental form of Eqs.(21) and (22) can be derived as

(j) p p(j) (j) p (Cj ) (j)


Qk+1 = kj Cj (θ) ek+1 − αj Qk zk+1 + Qk θk+1 (37)
Cj
l
 (j) p p
sk+1 = Qk+1 =Āpk epk+1 − B̄ k zk+1
p
+ D̄ k θk+1 (38)
j=1

with p
1 − e−z̄j (Dj )
kjp = , z̄jp = αj zk+1
p
+ θk+1 (39)
z̄jp Dj
n
 n
 l
 
p (j) p (Cj ) (j)
Āpk = kjp Cj (θ), B̄ k = kjp αj Qk , D̄ k = kjp Qk (40)
j=1 j=1 j=1
Cj
mr emr mr
k+1 :ek+1
Making use of zk+1 = (fkmr )2 zk+1
mr obtained from Eq.(11), it can be solved from Eq.(34) that
−1 −1 mr
emr m
k+1 = T k : sk+1 − T k : D̄ k θk+1 (41)
where
mr emrk+1
T k = Āmr
k I 4 − B̄ k ⊗ (42)
(fkmr )2 zk+1
mr

and I 4 is the identity tensor of rank four.


p ep :ep
Similarly, making use of zk+1 = (f k+1
p 2 p
k+1
) zk+1
obtained from Eq.(20), it can be solved from Eq.(38)
k
that
p
epk+1 = R−1 −1
k : sk+1 − Rk : D̄ k θk+1 (43)
where
p epk+1
Rk = Āpk I 4 − B̄ k ⊗ p 2 p (44)
(fk ) zk+1
Equation (16) can be rewritten as

p Gm  m
sm m mr
k+1 = 2G (θ)[ ek+1 − a ek+1 − ek+1 ] + s θk+1 (45)
Gm k
with 
1, if emr mr
eq ≤ e0
a= (46)
0, if eeq > emr
mr
0
· 292 · ACTA MECHANICA SOLIDA SINICA 2012

By substituting Eqs.(41) and (43) into Eq.(45), one obtains the following relationship for the marten-
site:

m −1 m m −1 m −1 mr m −1 p (Gm 
k ) m
sk+1 = Ak : [2Gk ek+1 − 2Gk Rk sk+1 + 2aGk T k : Dk + 2Gk Rk : Dk + sk
Gm
k
· θk+1 ] (47)

with 
−1
Ak = I 4 + 2aGm
k Tk (48)
On the other hand, one can also obtain the relationship for the austenite as follows by substituting
Eq.(43) into Eq.(25):

p (Gak ) a
sak+1 = 2Gak ek+1 − 2Gak R−1 s k+1 + 2Ga −1
R
k k : D + s θk+1 (49)
k k
Gak k
The incremental form of Eq.(6) at the (k+1)th increment of loading can be expressed as

sk+1 = ξ sm a m a
k+1 + (1 − ξ) sk+1 + ξk+1 (sk − sk ) (50)

Substituting Eqs.(47) and (49) into Eq.(50) yields


−1 −1 a −1 −1

sk+1 = I 4 + 2ξGm k Ak : Rk + 2(1 − ξ)Gk Rk


p (Gm 
k ) m
: ξA−1k : 2G m
k e k+1 + 2aGm −1
k T k : D mr
k + 2G m −1
k R k : D k + sk θ k+1
Gm
k


a a −1 p (Gak ) a m a
+(1 − ξ) 2Gk ek+1 + 2Gk Rk : D k + s θk+1 + ξk+1 (sk − sk ) (51)
Gak k
The incremental volumetric responses of the two phases can be obtained from Eqs.(4), (26) and (28)
as
 
m (Kkm ) m m 
( σii )k+1 = 3Kkm ( εii )k+1 + (σ )
ii k − 9K m
k (θ k )[αm
k + (α k ) (θ k − θ 0 )] θk+1 (52)
Kkm
 a  
a a (Kk ) a a a a 
( σii )k+1 = 3Kk ( εii )k+1 + (σii )k − 9Kk (θk )[αk + (αk ) (θk − θ0 )] θk+1 (53)
Kka
Given ε and θ, the deviatoric stress increment in the material can be determined with Eq.(51), then
the deviatoric stress increment in each phase can be determined with Eqs.(47) and (49), respectively.
The volumetric stress increment in each phase can be determined with Eqs.(52) and (53), and the
volumetric stress increment in the material can be determined with the incremental form of the second
equation in Eq.(6).

IV. APPLICATIONS AND VERIFICATION


4.1. Identification of Material Constants
The material parameters related to the thermal-mechanical loading induced transformation, reori-
entation and plastic behavior of an SMA can be separated into five groups: the parameters of the
elastic behavior of the austenite, those of the elastic and reorientation behavior of the martensite, the
parameters related to the plastic behavior of the SMA, the parameters related to forward and inverse
martensitic transformation, and the dependence of material constants on temperature. In principle,
these material constants can be identified by the following procedures:
(1) At θ > Af , before stress-induced martensitic transformation occurs, the SMA to be tested is
composed of the pure austenite, the following parameters can be identified with a simple elastic tension
(or torsion) test: the Young’s modulus E a (or shear modulus Ga ) and the Poisson’s ratio ν a of the
austenite.
(2) At θ < Mf , the SMA to be tested is composed of pure martensite, and deformation can not
induce further macroscopic martensitic transformation. If the deformation of the martensite can be
Vol. 25, No. 3 Xianghe Peng et al.: A Constitutive Model of Shape Memory Alloys · 293 ·

approximately separated into elastic and reorientation stages, and the overlap and interaction between
different kinds of deformation can be neglected, the constants related to the elastic and reorientation
properties of the martensite can be identified with different parts of the stress-strain curve of a simple
tension (or simple torsion) test.
(3) The parameters related to the plastic property of the SMA can be identified with the σ-εp relation,
which can be easily abstracted from the experimental result involving plastic deformation. It should
be noted that plastic deformation can hardly take place simultaneously in both the phases, depending
mainly on the working temperature, but can be inherited between different phases. Thus, taking the
plastic deformation as the behavior of the overall material and considering its temperature-dependence
may, on one hand, distinguish different resistances against the plastic deformation on the austenite
or the martensite, and on the other hand, provide an easier way for both modeling and application.
The temperature-dependence of plastic deformation can also be determined with the experimental σ-εp
relations at different temperatures.
(4) Mf , Ms , As , Af are usually provided by material suppliers, and the rest parameters related to
forward and inverse martensitic transformation can be identified by using pseudoelastic effects: (a) Am
and B m in a loading process involving stress induced martensitic transformation, and (b) Aa and B a in
the subsequent unloading process involving stress induced inverse transformation; the function g(ε̄pmax )
and the corresponding parameters related to the effect of plastic deformation on As can be determined
by measuring a set of As after different levels of plastic deformation.
(5) If the constitutive behavior of the SMA tested exhibits distinct dependence on temperature, the
variation of the corresponding material constants can be determined with the results tested at different
temperatures.

4.2. Analysis of Constitutive Behavior of SMA Au-47.5 at.% Cd


The ferroelasticity (FE) and pseudoelasticity (PE) of SMA Au-47.5 at.% Cd subjected to ten-
sion/compression were investigated experimentally by Nakanishi et al.[41] , which are described with
the proposed model. Since the experimental results involve mainly elastic, transformation and reori-
entation behaviors, plastic deformation is insignificant. When making use of the proposed model, the
plastic deformation can be ignored by simply letting n = 1, and α1 and C1 /α1 be sufficiently large.
It was found from the experimental results[41] that the effect of temperature on material constants
is less significant. The martensitic transformation finish and start temperatures, Mf and Ms , and the
austenitic transformation start and finish temperatures, As and Af were given as 308 K, 331 K, 345 K
and 358 K, respectively[41] . For simplicity, we choose l = 2, f mr (z mr ) = 1, and neglect the temperature
dependence of the material constants, the other material constants are identified as follows:
Elastic constants: Gm = 2.0 GPa, Ga = 2.5 GPa, ν m = ν a = 0.28,
Reorientation constants: D1m = 9.5 GPa, D2m = 0.05 GPa, λm m mr
1 = 500, λ2 = 5, e0 = 0.036,
Transformation constants: A , B , A , B = 0.45 K , 0.32 MPa K, 0.25 K , 0.47 MPa−1 K.
a a m m −1 −1 −1

Fig. 3. σ-ε curves of Au-47.5 at.% Cd under ten- Fig. 4. σ-ε curves of Au-47.5 at.% Cd under ten-
sion/compression at 300 K. sion/compression at 357K.
· 294 · ACTA MECHANICA SOLIDA SINICA 2012

Figure 3 shows the description of the response of the SMA subjected to tensile/compressive strain
cycles with different amplitudes at 300 K. Since θ < Mf , the stress could not induce further macroscopic
transformation, the SMA should be martensitic during the deformation and the ferroelastic behavior
should be attributed to the reorientation or detwinning of the martensite variants. The deformation is
initially elastic, and turns to be inelastic when reorientation takes place. The reorientation proceeds until
the equivalent macroscopic reorientation strain reaches a critical value, i.e., emr mr
eq = e0 (Eq.(15)), then
the response turns to the elastic response of the martensite after the reorientation. It should be noted
that for nonproportional loading, reorientation of new martensite variants may be activated when the
loading path changes its direction, which can be included in the proposed model by simply considering
the critical value emr
0 as the function of nonproportionality.
Figure 4 shows the tensile/compressive curve of the Au-47.5 at.% Cd at 357 K. It is known that the
material is almost composed of austenite at the temperature (θ ≈ Af = 358 K) before deformation, with
the increase of strain, Ms +B m σe increases, and martensitic transformation takes place as Ms +B m σe > θ.
During unloading, As +Ba σe decreases and the inverse transformation occurs as As+B a σe < θ. Compared
with experimental results[41] , it can be seen that the pseudoelastic behavior can be well described with
the proposed model.
The tensile/compressive curves of the Au-47.5 at%Cd at different temperatures are shown in Fig.5.
At 330 K the constitutive behavior is typically ferroelastic. At 366 K (θ > Af = 358 K) the material
is mainly composed of austenite before deformation and the constitutive behavior at this temperature
should be pseudoelastic during deformation, similar to that shown in Fig.4, but the critical stress
should increase due to the increase of temperature, because the martensitic transformation at a higher
temperature may need a larger driving force. At θ = 385 K, the critical stress further increases due
to that larger stress is needed for the martensitic transformation, and the hysteretic loop becomes
much smaller, indicating that the contribution of the martensite becomes less significant because of the
smaller martensite volume fraction. It can be conjectured that if the temperature is sufficiently high
so that no stress-induced martensitic transformation can take place, the material would be composed
of the austenite and the behavior of the material would be determined by the austenite.
Figure 6 shows the prediction of the shape memory effect of the Au-47.5 at%Cd. The material is firstly
elongated at 300 K to ε = 0.0342 and then unloading. In this stage, because θ < Mf , the deformation
of the material can be attributed to the elastic deformation of the martensite and the reorientation of
martensite variants. It can be seen that the residual strain after unloading is about ēmr∗ = 0.0294. Then
the material is heated and the variation of strain against temperature is also plotted in Fig.6, where we
assume αm = αa = 10−5 K−1 . It can be seen that the strain increases a little before the temperature
reaches As (345 K) due to thermal expansion, but it recovers quickly once the temperature exceeds this
threshold when inverse martensitic transformation takes place. The residual strain tends to vanish as
the temperature is sufficiently high when the material transforms completely back to its parent phase.
Since we ignore the macroscopic volumetric change during transformation, the strain after heating may
consist of two parts: the strain of the residual martensite (we assume ēmr = 0.001ēmr∗ at θ = Af ,

Fig. 5. σ-ε curves of Au-47.5 at.% Cd under ten- Fig. 6. SME of Au-47.5 at.% Cd subjected to tension-
sion/compression at different θ. unloading at 300 K followed by heating.
Vol. 25, No. 3 Xianghe Peng et al.: A Constitutive Model of Shape Memory Alloys · 295 ·

as mentioned previously) and thermal strain. If the temperature then falls to below Mf , martensitic
transformation would occur and the material would return to its original state. This process can be
quite satisfactorily described with the proposed model.

4.3. Analysis of Constitutive Behavior of an NiTi SMA


Wang et al.[4] investigated in details the pseudoelastic behavior of a kind of NiTi SMA, involving
plastic deformation and its effect. The corresponding results are simulated with the proposed model.
For the NiTi SMA, Mf , Ms , As and Af have been provided as 164 K, 244 K, 232 K and 270 K
respectively in Ref.[4]. For simplicity, we choose l = 1, n = 2, and f mr = f p = 1, and assume g(ε̄pmax )
to have the following simplest linear form:

g(ε̄pmax ) = g0 + hε̄pmax (54)

the corresponding material constants can be identified as Gm = 12 GPa, Ga = 24 GPa, ν m = ν a = 0.28,


αm = 10−5 K−1 , αa = 2.2 × 10−5 K−1 , D1m = 1530 GPa, λm mr
1 = 4000, e0 = 0.059, g0 = −30 K, h = 800
K, A , B , A , B = 0.15 K , 0.32 MPa K, 0.25 K , 0.26 MPa K, C1p = 1437 GPa, C2p = 3 GPa,
a a m m −1 −1 −1 −1

αp1 = 3000, αp2 = 30.


The experiments were performed at room temperature of θ = 296 K (θ > Af = 270 K) in an open
testing environment[4] , so that the SMA should be austenitic before deformation. Figure 7 shows the
σ-ε curve of the NiTi SMA subjected to pure tension until ε ≈ 0.11 followed by unloading. The response
should be of austenitic elasticity at the beginning of deformation. Soon after the onset of martensitic
transformation, the stress drops slightly and then almost keeps constant until the completion of the
forward transformation and the reorientation of the martensite variants. Then the response turns to the
elastic response of the detwinned martensite, i.e., the stress increases elastically with the development
of strain. Plastic deformation speeds up as the stress exceeds some critical value. During the unloading,
the response is initially elastic until the inverse transformation occurs when the stress falls to a critical
value. The reorientation strain recovers as the detwinned martensite transforms back to its parent
phase. When the inverse transformation finishes, the material is composed of austenite and the response
turns to austenitic elasticity. However, it can be seen that although the detwinned martensite has been
inversely transformed back to the parent phase, the plastic deformation occurring in the material remains
irrecoverable. The comparison with the experimental results shows that the main characteristics of the
constitutive behavior can be reasonably described.
The responses of the NiTi SMA subjected to tensile straining to different maximum strains, εmax ,
and unloading at 296 K are analyzed and compared with the experimental results, as shown in Fig.8. For
εmax = 6%, the response of the SMA exhibits typically pseudoelastic: a slight stress drop appears soon
after the start of the martensitic transformation, then the stress almost keeps constant until ε = εmax ,
followed by unloading. It can be seen that the strain recovers completely after unloading. For εmax = 8%,
the reorientation of the martensite variants finishes at ε ≈ 6.4%, then the response becomes almost
elastic until about σ = 600 MPa when the plastic deformation develops distinctly. During unloading, the

Fig. 7. σ-ε curve of NiTi SMA subjected to tension- Fig. 8. σ-ε curves of NiTi SMA subjected to tension un-
unloading at 296 K. loading at 296 K, with various strain amplitudes.
· 296 · ACTA MECHANICA SOLIDA SINICA 2012

reorientation strain recovers but the plastic strain remains irrecoverable. It can be seen by comparing
the results corresponding to εmax = 10% and εmax = 12% with those corresponding to εmax = 8%
that with the increase of εmax , the residual strain increases and the unloading parts of the σ-ε curves
descend distinctly. The comparison between the computed and experimental results indicates that the
proposed model can describe the main characteristics of the SMA, including the forward transformation
and the reorientation, the inverse transformation and the recovery of reorientation strain, and plastic
deformation and its effects.
It is known that plastic deformation dissipates the free energy stored in the microstress fields in
the material and, therefore, reduces the internal driving force for the inverse transformation. In other
words, more energy should be input for the inverse transformation, which accounts for the increase in
the temperature or the decrease in the stress for the inverse transformation. This mechanism can be
well described with the proposed model. It can be seen in Fig.8 that the larger the plastic strain, the
more the unloading parts of the curves descend, indicating that more energy is needed for the inverse
transformation. The dashed curves in Fig.9 correspond to the description of the stress-strain responses
without considering the effect of plastic deformation (which can be realized by setting h = 0 (Eq.(54))
on the inverse transformation, where the stresses at the start of the inverse martensitic transformation
of the material subjected to different plastic strain histories are almost the same as 296 MPa. In fact,
when εmax = 8%, 10% and 12%, hepmax ≈ 11.8 K, 30 K and 48.2 K, respectively, which can also be
regarded that the inverse transformation temperature increases.
Figure 10 shows the prediction of the SME of the NiTi SMA. The material is firstly elongated at
200 K to ε ≈ 0.12 and then unloading. According to the given temperature, we assume that the SMA is
mainly composed of the martensite. During loading, the deformation of the SMA contains four stages:
elastic deformation of both the phases incorporating quick transformation of the residual austenite,
reorientation of the martensite variants, elastic deformation of the detwinned martensite, and plastic
deformation. After unloading, the residual strain is about 0.0984. Then the material is heated and the
variation of strain against temperature is also plotted in Fig. 9. It can be seen that, without considering
the effect of plastic deformation (or h = 0), the strain almost keeps unchanged before the temperature
reaches As (232 K), but it recovers quickly once the temperature exceeds this critical value when inverse
transformation takes place (see the dashed curve). The residual strain tends to 0.051 if the temperature
is sufficiently high when the material transforms back to its parent phase. If the effect of the plastic
deformation is taken into account, the austenitic transformation start and finish temperatures will
increase about 40 K, respectively, and the variation of ε against θ is plotted with the curve in black.
Since we ignore the macroscopic volumetric change due to transformation, the residual strain after
heating may consist of three parts: the residual plastic strain, the residual reorientation strain (we
assume ēmr = 0.001ēmr∗ at θ = Af ) and the thermal strain. It can be seen that plastic deformation
induced increase in the inverse transformation temperature can be well described.

Fig. 9. σ-ε curves with and without considering the effect Fig. 10. SME of NiTi SMA elongated and unloaded at 200
of plastic deformation. K followed by heating.
Vol. 25, No. 3 Xianghe Peng et al.: A Constitutive Model of Shape Memory Alloys · 297 ·

V. SUMMARY
An SMA is a two-phase mixture composed of martensite and austenite, the volume fraction of each
phase is transformable, and the complicated behavior of an SMA is considered as the combination
of the relatively simple behavior of its two phases. The strain in the martensite consists of thermal
strain, elastic strain, reorientation strain and plastic strain; while the strain in the austenite consists of
thermal strain, elastic strain and plastic strain. Making use of thermomechanically consistent mechanical
models and the Voigt estimation, the constitutive modeling is formulated for SMAs. The corresponding
numerical algorithm is developed. The responses of Au-47.5 at.% Cd, including FE, PE and SME,
and that of an NiTi SMA, including PE, SME, plastic deformation and its effects, are analyzed. The
numerical processes are stable and convergent. The comparison between the computed results and the
experimental results shows satisfactory agreement, demonstrating the validity of the proposed model.
The proposed model can replicate the main features of SMAs and can also be easily applied in the
analysis for practical engineering problems.
It should be noted that the Voigt estimation is used for the evaluation of the effective mechanical
property of two-phase SMAs. The consideration of this simplification is the application of this model in
the analysis of practical engineering problems, and the extension of this model to ternary SMAs, such
as NiTiNb SMAs, where the Nb particles are embedded in the NiTi matrix as spherical inclusions. The
authors understand that the Eshelby’s inclusion theory based approach, such as Mori-Tanaka scheme
and self-consistent scheme, etc., may provide better estimation. However, for the problems involving
transformable volume fractions, reorientation and plastic deformation as well as the induced anisotropy,
there are still some problems that remain to be solved in our further research.

References
[1] Oyamada,O., Amano,K., Enomoto,K., Shigenaka,N., Matsumoto,J. and Asada,Y., Effect of environment
on static tensile and fatigue properties of Ni-Ti-Nb shape memory alloy. International Journal - Series A:
Solid Mechanics and Material Engineering, 1999, 42(2): 243-248.
[2] Kusagawa,M. and Asada,T., Fundamental deformation and recovery behavior of Ni-Ti-Nb shape memory
alloy. International Journal - Series A: Solid Mechanics and Material Engineering, 2001, 44(1): 57-63.
[3] Piao,M., Otsuka,K., Miyazaki,S. and Horikawa,H., Mechanism of the as temperature increase by pre-
deformation in thermoelastic alloys. Materials Transactions, JIM, 1993, 34(10): 919-929.
[4] Wang,X., Xu,B. and Yue,Z., Phase transformation behavior of pseudoelastic NiTi shape memory alloys
under large strain. Journal of Alloys and Compounds, 2008, 463: 417-422.
[5] Strnade,B., Ohashi,S., Ohtsuka,H., Miyazaki,S. and Ishihara,T., Effect of mechanical cycling on the pseu-
doelasticity characteristics of Ti-Ni and Ti-Ni-Cu alloys. Materials Science and Engineering: A, 1995, 203(1-
2): 187–196.
[6] Sehitoglu,H., Anderson,R., Karaman,I., Gall,K. and Chumlyakov,Y., Cyclic deformation behavior of single
crystal NiTi. Materials Science and Engineering: A, 2001, 314: 67-74.
[7] Gall,K. and Maier,H.J., Cyclic deformation mechanisms in precipitated NiTi shape memory alloys. Acta
Materialia, 2002, 50: 4643-4657.
[8] Brinson,L.C., 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, 1993, 4(2): 229-242.
[9] Miller,D.A. and Lagoudas,D.C., Thermo-mechanical characterization of NiTiCu and NiTi SMA actuators:
influence of plastic strains. Smart Materials and Structures, 2000, 5: 640-652.
[10] McKelvey,A.L. and Ritchie,R.O., Fatigue crack propagation in Nitinol, a shape-memory and superelastic
endovascular stent material. Journal of Biomedical Materials Research, 1999, 47: 301-308.
[11] Sehitoglu,H., Karaman,I., Anderson,R., Zhang,X., Gall,K., Maier,H.J. and Chumlyakov,Y., Compressive
response of NiTi single crystals. Acta Materialia, 2000, 48: 3311-332.
[12] Karaman,I., Yapici,G.G., Chumlyakov,Y.I. and Kireev,I.V., Deformation twinning in difficult-to-work al-
loys during severe plastic deformation. Materials Science and Engineering: A, 2005, 410-411: 243-247.
[13] Sun,Q.P. and Hwang,K.C., Micromechanics modeling for the constitutive behavior of polycrystalline shape
memory alloys – I. Derivation of general relations. Journal of the Mechanics and Physics of Solids, 1993a.,
41(1): 1-17.
[14] Sun,Q.P. and Hwang,K.C., Micromechanics modeling for the constitutive behavior of polycrystalline shape
memory alloys – II. Study of the individual phenomena. Journal of the Mechanics and Physics of Solids,
1993b, 41(1): 19-33.
· 298 · ACTA MECHANICA SOLIDA SINICA 2012

[15] Lexcellent,C., Goo,B.C., Sun,Q.P. and Bernardini,J., Characterization thermomechanical behaviour and
micromechanical-based constitutive model of shape-memory Cu-Zn-Al single crystals. Acta Materialia,
1996, 44(9): 3773-3780.
[16] Siredey,N., Patoor,E., Berveiller,M. and Eberhardt,A., Constitutive equations for polycrystalline thermoe-
lastic shape memory alloys. Part I. Intragranular interactions and behavior of the grain. International
Journal of Solids and Structures, 1999, 36: 3178-3204.
[17] Gao,X.J., Huang,M. and Brinson,L.C., A multivariant micromechanical model for SMAs Part 1. Crystal-
lographic issues for single crystal model. International Journal of Plasticity, 2000, 16: 1345-1369.
[18] Gao,X.J., Huang,M. and Brinson,L.C., A multivariant micromechanical model for SMAs Part 2. Polycrystal
model. International Journal of Plasticity, 2000, 16: 1371-1390.
[19] Blanc,P. and Lexcellent,C., Micromechanical modelling of a CuAlNi shape memory alloy behaviour. Ma-
terials Science and Engineering, 2004, A 378: 465-469.
[20] Patoor,E., Lagoudas,D.C., Entchev,P.B., Brinson,L.C. and Gao,X.J., Shape memory alloys, Part I: General
properties and modeling of single crystals. Mechanics of Materials, 2006, 38: 391-429.
[21] Boyd,J.G. and Lagoudas,D.C., A thermodynamic constitutive model for the shape memory materials, the
monolithic shape memory alloys. International Journal of Plasticity, 1996, 12(6): 805-842
[22] Auricchio,F., Taylor,R.L. and Lubliner,J., Shape-memory alloys: macromodelling and numerical simula-
tions of the superelastic behavior. Computer Methods in Applied Mechanics and Engineering, 1997, 146:
281-312.
[23] Raniecki,B. and Lexcellent,C., Thermodynamics of isotropic pseudoelasticity in shape memory alloys. Eu-
ropean Journal of Mechanics A-Solids, 1998, 17, 185-205.
[24] Tanaka,K., Kobayashi,S. and Sato,Y., Thermomechanics of transformation pseudoelasticity and shape
memory effect in alloys. International Journal of Plasticity, 1986, 2: 59-72.
[25] Liang,C. and Rogers,C.A., One-dimensional thermomechanical constitutive relations for shape memory
materials. Journal of Intelligent Material Systems and Structures, 1990, 1(2): 207-234.
[26] Brinson,L.C., Schmidt,I. and Lammering,R., Stress-induced transformation behavior of a polycrystalline
NiTi shape memory alloy: micro and macromechanical investigations via in situ optical microscopy. Journal
of the Mechanics and Physics of Solids, 2004, 52: 1549-1571.
[27] Peng,X., Yang,Y. and Huang,S., A comprehensive description for shape memory alloys with a two-phase
constitutive model. International Journal of Solids and Structures, 2001, 38: 6925-6940.
[28] Peng,X., Li,H. and Chen,W., A comprehensive description for SMAs with a two-phase mixture model
incorporating conventional theory of plasticity. Smart materials and Structures, 2005, 14: 425-433.
[29] Peng,X., Pi,W. and Fan,J., A microstructure-based constitutive model or the pseudoelastic behavior of
NiTi SMAs. International Journal of Plasticity, 2008, 24: 966-990.
[30] Leclercq,S. and Lexcellent,C., A general macroscopic description of the thermomechanical behavior of shape
memory alloys. Journal of the Mechanics and Physics of Solids, 1996, 44(6): 953-957.
[31] Popov,P. and Lagoudas,D.C., A 3-D constitutive model for shape memory alloys incorporating pseudoe-
lasticity and detwinning of self-accommodated martensite. International Journal of Plasticity, 2007, 23:
1679-1720.
[32] Bo,Z.H. and Lagoudas,D.C., Thermomechanical modeling of polycrystalline SMAs under cyclic loading.
International Journal of Engineering Science, 1999, 37: 1089-1249.
[33] Panico,M. and Brinson,L.C., A three-dimensional phenomenological model for martensite reorientation in
shape memory alloys. Journal of the Mechanics and Physics of Solids, 2007, 55: 2491-2511.
[34] Dong,L. and Sun,Q., Stress hysteresis and domain evolution in thermoelastic tension strips. Acta Mechanica
Solida Sinica, 2009, 22(5): 399-406.
[35] He,Y. and Sun,Q., Non-local modeling on macroscopic domain patterns in phase transformation of NiTi
tubes. Acta Mechanica Solida Sinica, 2009, 22(5): 407-417.
[36] Souza,A.C., Mamiya,E.N. and Zouain,N., Three-dimensional model for solids undergoing stress-induced
phase transformations. European Journal of Mechanics A-Solids, 1998, 17: 789-806.
[37] Kan,Q., Kang,G. and Qian,L., Super-elastic constitutive model considering plasticity and its finite element
implementation. Acta Mechanica Solida Sinica, 2010, 23(2): 95-105.
[38] Peng,C., Wang,X. and Huo,Y., Characteristics of stress-induced transformation and microstructure evo-
lution in Cu-based SMA. Acta Mechanica Solida Sinica, 2008, 21(1): 1-8.
[39] Fan,J. and Peng,X., A physically based constitutive description for nonproportional cyclic plasticity. Journal
of Engnging Materials and Technology, 1991, 113(2): 254-262
[40] Peng,X. and Fan,J., A Numerical approach for nonclassical plasticity. Computers & Structures, 1993, 47(2):
313-320.
[41] Nakanishi,N., Mori,T., Miura,S., Murakami,Y. and Kachi,S., Pseuelasticity in Au-Cd thermoelastic marten-
site. Philosophical Magazine, 1973, 28: 277-292.

You might also like