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

ARTICLE IN PRESS

Journal of the Mechanics and Physics of Solids


53 (2005) 2320–2346
www.elsevier.com/locate/jmps

Experimental characterization and


micromechanical modeling of superelastic
response of a porous NiTi shape-memory alloy
Sia Nemat-Nassera,, Yu Sua, Wei-Guo Guob, Jon Isaacsa
a
Center of Excellence for Advanced Materials, Department of Mechanical and Aerospace Engineering,
University of California, San Diego, 9500 Gilman Drive, La Jolla, CA 92093-0416, USA
b
Department of Aircraft Engineering (118#), Northwestern Polytechnical University, Xi’An City,
Shaanxi Province, 710072, PR China
Received 19 May 2004; received in revised form 26 March 2005; accepted 29 March 2005

Abstract

Porous shape-memory alloys are usually brittle due to the presence of various
nickel–titanium intermetallic compounds that are produced in the course of most commonly
used synthesizing techniques. We consider here a porous NiTi shape-memory alloy (SMA),
synthesized by spark-plasma sintering, that is ductile and displays full shape-memory effects
over the entire appropriate range of strains. The porosity however is only 12% but the basic
synthesizing technique has potential for producing shape-memory alloys with greater porosity
that still are expected to display superelasticity and shape-memory effects. The current
material has been characterized experimentally using quasi-static and dynamic tests at various
initial temperatures, mostly within the superelastic strain range, but also into the plastic
deformation regime of the stress-induced martensite phase. To obtain a relatively constant
strain rate in the high strain-rate tests, a novel pulse-shaping technique is introduced. The
results of the quasi-static experiments are compared with the predictions by a model that can
be used to calculate the stress–strain response of porous NiTi shape-memory alloys during the
austenite-to-martensite and reverse phase transformations in uniaxial quasi-static loading and
unloading at constant temperatures. In the austenite-to-martensite transformation, the porous
shape-memory alloy is modeled as a three-phase composite with the parent phase (austenite)

Corresponding author. Tel.: +1 858 534 4772; fax: +1 858 534 2727.
E-mail address: sia@ucsd.edu (S. Nemat-Nasser).

0022-5096/$ - see front matter r 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jmps.2005.03.009
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2321

as the matrix and the product phase (martensite) and the voids as the embedded inclusions,
reversing the roles of austenite and martensite during the reverse transformation from fully
martensite to fully austenite phase. The criterion of the stress-induced martensitic
transformation and its reversal is based on equilibrium thermodynamics, balancing the
thermodynamic driving force for the phase transformation, associated with the reduction of
Gibbs’ free energy, with the resistive force corresponding to the required energy to create new
interface surfaces and to overcome the energy barriers posed by various microstructural
obstacles. The change in Gibbs’ free energy that produces the driving thermodynamic force for
phase transformation is assumed to be due to the reduction of mechanical potential energy
corresponding to the applied stress, and the reduction of the chemical energy corresponding to
the imposed temperature. The energy required to overcome the resistance imposed by various
nano- and subnano-scale defects and like barriers, is modeled empirically, involving three
constitutive constants that are then fixed based on the experimental data. Reasonably good
correlation is obtained between the experimental and model predictions.
r 2005 Elsevier Ltd. All rights reserved.

Keywords: Porous shape-memory alloy; NiTi; Experimental characterization; Thermodynamics;


Micromechanics; Free energy

1. Introduction

In the absence of an applied stress, a solid shape-memory alloy (SMA) undergoes


phase transformation from the austenitic to the martensitic phase when cooled below
the so called martensitic start temperature, M s . This process evolves continually until
the martensitic finish temperature, M f , is reached, completing the martensitic
transformation. The austenitic phase re-forms when the temperature is raised above
the austenitic start temperature, As , and this process is complete when the austenitic
finish temperature, Af , is attained. Usually As and Af are not correspondingly equal
to M s and M f . In the presence of an applied stress, the transformation from the
austenitic to the martensitic phase may start at a temperature higher than M s .
Crystallographically, the austenite-to-martensite phase transformation consists of
the Bain distortion and a lattice-invariant shear. The Bain distortion is a lattice
deformation in the parent phase that produces a new crystal structure. The invariant
shear involves lattice twinning within the new crystal. During the transformation,
martensites are formed in the shape of thin plates. The surface of such plates
represents the habit plane on which a material line element vector remains
unchanged in length and direction. Such a geometrical theory was first introduced by
Wechsler et al. (1953) and Bowles and Mackenzie (1954) to characterize martensitic
transformations. The determination of such an invariant plane from the crystal
structure of the parent and product phases was later more generally studied by
Khachaturyan (1983). Therefore, the normal and shear components of the phase
transformation strain could be calculated based on the theory of the invariant plane
strain.
Currently, there is considerable interest in producing light-weight ductile SMAs by
introducing suitable porosity in the material. Such porous SMAs may have
ARTICLE IN PRESS
2322 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

advantages over the traditional SMAs in energy absorbing applications. There also
are some applications of porous SMAs in the biomedical field; see, e.g.,
Shabalovskaya et al. (1994) and Ayers et al. (1999). Thus the stress-induced
response of porous SMAs is of current research interest.
Various methods have been used to synthesize porous SMAs. Both pressureless
sintering and self-propagating high-temperature synthesis (SHS) methods have been
used by Li and coworkers to produce porous NiTi SMAs; see Li et al. (1998a, b,
2000). Unfortunately, both techniques produce various Ni–Ti intermetallics (e.g.,
Ni3Ti, Ti2Ni, and Ni4Ti3) that are brittle, limiting the superelastic and other
mechanical attributes of the resulting porous material. A somewhat similar outcome
results with hot-isostatic pressuring (HIP) approach, where similar intermetallics
seem to emerge, limiting the shape-memory attributes; see, Lagoudas and Vandygriff
(2002).
Starting with uncontaminated powders of 49.1Ni–50.9Ti (atom%), porous SMAs
have been synthesized using spark-plasma sintering (SPS) at suitable temperatures
and pressures; Taya (2004). The porosity may be controlled in a limited way by
adjusting the particle size distribution. However, greater porosity may be achieved
by using hollow shape-memory powders; Berkowitz (2004). In the present work, we
examine experimentally and theoretically a pilot porous SMA which has been
produced using solid powders, resulting in about 12–13% porosity.
To characterize the response of the material, we have performed compression
experiments at high and low (quasi-static) strain rates. Since, for SMAs, it is not
possible to obtain a constant strain rate in conventional Hopkinson bar systems
(Nemat-Nasser et al., 2005a), a novel pulse-shaping technique is introduced that
does create suitable stress pulses that in turn can produce relatively constant strain
rates. The results show that this porous SMA is superelastic and displays full shape-
memory effects. It has been shown by Nemat-Nasser and Choi (2005) that the
deformation of SMAs generally is strongly strain-rate dependent, with the transition
stress and the yield stress of the parent austenite both increasing with the increasing
strain rate, and eventually attaining a common value at a critical strain rate within
the superelastic temperature range. At strain rates below the critical level, stress-
induced martensites are formed, whereas the dislocation-induced slip deformation
occurs at strain rates above this level. The yield stress of the material follows the
yielding behavior of the resulting martensite at strain rates below the critical level,
and the yield stress of the parent austenite for strain rates exceeding this level. The
present porous SMA seems to possess similar characteristics. We are not aware of
any viable model that addresses the kinetics of such complex deformation
mechanisms. Therefore, here, we seek to model only the constant-temperature,
quasi-static response of the material, using standard micromechanics and
phenomenological thermodynamics.
Most existing models of (quasi-static) phase transformation in SMAs exploit an
energy approach to construct a free energy function in terms of suitable state
variables, and then use this to evaluate the threshold and evolution of martensite
transformation; see, e.g., Lexcellent and Licht (1991), Raniecki et al. (1992),
Raniecki and Lexcellent (1998) and Abeyaratne and Knowles (1993). Recently,
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2323

based on the equilibrium thermodynamic principles and domain transformation


morphology, a micromechanical theory has been developed to predict the
stress–strain response of single crystal and polycrystal shape-memory alloys during
their phase transformation; see Lu and Weng (1997, 1998). It is recognized that the
stress-induced phase transformation of SMAs is governed by a reduction in Gibbs’
free energy, consisting of a change of the mechanical potential energy associated with
the applied stress and a change of the chemical free energy associated with the
imposed temperature. This method has turned out to be one of the successful
approaches to study phase transformation and domain reorientations of solid
materials. It has also been used for the constitutive modeling of the spontaneous
polarization and hysteresis behavior of ferroelectric crystals; see Li and Weng
(2001a), and Li and Weng (2001b). The method has so far been applied to model the
phase transformation of SMAs, assuming that the local elastic moduli of the
martensite and austenite phases are the same. This assumption, however, may not be
appropriate for most shape-memory alloys.
Lagoudas et al. (2000, 2001), Qidwai et al. (2001), and Entchev and Lagoudas
(2002), have integrated the principles of thermodynamics into the framework of
micromechanics to model porous SMAs. In the present paper, we seek to use the
micromechanical model developed for SMAs to calculate the quasi-static stress–
strain response of our porous shape-memory alloy during its stress- and
temperature-induced phase transformation, and compare the model predictions
with our quasi-static experimental results. The porous material is modeled as a three-
phase composite system with the austenitic phase as the matrix, the product phase
(martensite) as the inclusion, and the pores with a constant volume fraction also as
an inclusion phase. The thermodynamic formulation is based on balancing the
thermodynamic driving force with that due to the energy required to overcome the
resistance to transformation posed by various microstructural obstacles. The
morphology and the volume fraction of the product phase are considered as the
characteristics of the microstructure that are involved in the course of phase
transformation. The results of one quasi-static experiment are used to fix the
constitutive parameters associated with the model and then the model predictions
are checked against experimental data obtained at other (constant) temperatures,
arriving at reasonably good agreements.

2. Material, experimental procedure, and results

2.1. Material and sample preparation

A Porous NiTi SMA disc is fabricated from high-purity powders of 49.1Ni–50.9Ti


(atom%) by 5 min spark plasma sintering (SPS) at 800 1C temperature and 25 MPa
pressure; Taya (2004). The microstructure of the resulting porous material is shown
in Fig. 1. It has an average grain size of about 165 mm. Cylindrical samples (5 mm
in nominal diameter and 3.8 mm in height) are then cut from a disc of this material
by electric discharge machine (EDM). Finally these porous NiTi alloy samples are
ARTICLE IN PRESS
2324 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

Fig. 1. Microstructure of porous NiTi shape-memory alloy.

heat-treated for 30 min at 320 1C temperature, and then quenched in ice water. The
samples have 12% porosity, and 5.7 g/cc mass density. The resulting austenite start
and finish temperatures, As and Af , are about 1.3 1C and 23.8 1C, respectively.

2.2. Quasi-static experiments

Uniaxial compression tests are performed on the specimens using an INSTRON


hydraulic testing machine. The strain rate is kept at 103 =s. The compression tests
are performed at several constant temperatures ranging from 253 to 363 K. To
reduce the end friction, the ends of the specimens are slightly polished using
waterproof silicon carbide paper, 4000 grit, and then they are greased for room-
temperature tests. A molybdenum-powder lubricant is used for the low- and high-
temperature experiments. Elevated temperatures are attained with a high-intensity
quartz lamp, in a radiant-heating furnace. The temperature is measured using a
thermocouple arrangement, and is maintained constant to within a deviation of
2  C. The deformation of the specimen is measured by an LVDT that is mounted in
the testing machine. Before each test, the LVDT is calibrated. The low temperature is
obtained by immersing the specimen with the testing fixture (Al2O3 ceramic bars) in
a bath of acetone and adding dry ice.
Two examples of the resulting stress–strain curves of this porous NiTi SMA are
depicted in Fig. 2. The closed loop (solid triangles) corresponds to the longitudinal
loading and unloading at a 296 K temperature, within the superelastic limit which
exceeds 6% for the present porous SMA. The curve marked by solid circles
illustrates the behavior of the porous SMA beyond the superelastic range to failure.
As is seen, the material can sustain strains up to and exceeding 15% before failing. In
Fig. 3, the stress–strain response of various samples, each at the indicated
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2325

1600

Porous NiTi, T=296K, 0.001/s, f=12%

1200
Compression, σ (MPa)

800

Experiment No.1
Experiment No.2
400

0
0 5 10 15 20 25
Longitudinal strain, ε (%)

Fig. 2. Compressive stress–strain curve of porous NiTi shape-memory alloy at the indicated initial
temperature; forward A–M and the reversed M–A transformation within superelastic range (solid
triangles), and to failure (solid circles).

1200

Porous NiTi, 0.001/s, f =12%


1000
Compression, σ (MPa)

800

600

Exp 363K
400 Exp 343K
Exp 323K
Exp 296K
200 Exp 273K
Exp 253K

0
0 2 4 6 8 10
Longitudinal strain, ε (%)

Fig. 3. Measured (initial) temperature dependence of martensite transformation for porous NiTi shape-
memory alloy; only the loading regimes are shown.

temperature, is displayed in the order of, from bottom to top, 253, 273, 296, 323, 343
and 363 K test temperatures; a fresh sample is used for each test. Fig. 4 shows the
results of the experiment at 323 K temperature, to more than a 6% strain. Upon
ARTICLE IN PRESS
2326 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

1200

Porous NiTi, 323K, 0.001/s, f =12%


1000
Compression, σ (MPa)

800

600

400

200

0
0 2 4 6 8
Longitudinal strain, ε (%)

Fig. 4. Compressive stress–strain curve of porous NiTi shape-memory alloy 323 K initial temperature;
unloading leaves about 1% residual strain, which is fully recovered once the sample is removed from the
testing machine and slightly heated.

unloading, the sample retains about 1% residual strain, but this is fully recovered
once the sample is removed from the testing machine and is slightly heated.

2.3. Dynamic experiments

We are not aware of any high strain-rate experiments on porous SMAs, although
Nemat-Nasser and coworkers have recently published the results of a series of high
and very high strain-rate tests of solid SMAs; see, Nemat-Nasser and Choi (2005)
and Nemat-Nasser et al. (2005a, b). Here we report the results of some new high
strain-rate Hopkinson compression experiments using our new porous SMA, where
we have also devised a new pulse-shaping technique to produce relatively constant
high strain rates. These dynamic compression tests are performed at room
temperature, using UCSD’s recovery Hopkinson technique;1 see Nemat-Nasser
et al. (1991). In the Hopkinson tests, the sample strain rate depends on the incident
elastic pulse as well as on the pulse transmitted through the sample into the
transmission bar. The transmitted and reflected pulses represent the sample response.
The reflected pulse, representing the difference between the incident and the
1
Even though the experiments are performed at room temperature, deformation at high strain rates is
basically adiabatic and hence the sample temperature changes with loading and unloading, since, in
addition to the deformation-induced heating, the forward and reverse martensitic transformations are
exothermic and endothermic, respectively.
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2327

transmitted pulses, essentially defines the sample strain rate. To achieve a constant
strain rate, therefore, the time-variation of the incident pulse must be similar to that
of the stress transmitted by the sample into the transmission bar. For most metals, it
is often easy to produce an incident pulse of a suitable time-profile to render the
resulting sample strain rate essentially constant, since the time-variation of the
incident pulse can be controlled by adjusting the geometry and properties of a pulse
shaper (that is placed between the striker and incident bars) according to the length
and velocity of the striker bar; see Nemat-Nasser et al. (1991) who analyze this
problem in detail and provide examples.
The stress–strain relation for SMAs within the superelastic range is partly concave
down and partly concave up, requiring special pulse shaping in order to ensure a
constant strain rate. In recent articles, Nemat-Nasser et al. (2005a, b) and Nemat-
Nasser and Choi (2005) have examined the suitability of several simple pulse shapers
for high strain-rate tests of solid SMAs. Here we consider a new composite pulse
shaper, consisting of a central SMA cylinder encased in an aluminum ring, as shown
in Fig. 5 [left]. The dimensions of the components of this pulse shaper can be
optimized to produce the desired effect. Fig. 5 [right] shows a typical time-variation
of the resulting strain. The corresponding stress and strain-rate curves, obtained
using this pulse shaper, are shown in Fig. 6 as functions of the strain. As is seen, a
relatively constant strain rate is attained in this test. We have observed that the best
result is obtained when the central part of the composite pulse shaper is made of the
same SMA that is being tested. Fig. 7 shows the effect of different compositions for
the pulse shaper on the resulting stress–strain relation.
The results of several different tests are shown in Fig. 8, including three quasi-
static test results and one experiment without a pulse shaper (the upper black curve,
with failure at about 15% strain). As is seen, the response of this porous SMA is in
full correspondence with that of any solid ductile SMA. As has been discussed by
Nemat-Nasser and Choi (2005), the deformation of the shape-memory alloy is

Fig. 5. [left] Composite pulse shaper; and [right] A typical time-variation of the resulting strain.
ARTICLE IN PRESS
2328 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

Fig. 6. A typical stress and strain rate as function of the corresponding strain obtained using a composite
pulse shaper.

Fig. 7. Effect of pulse shaper on the response of the porous SMA.

strongly strain-rate dependent. These authors show that the transition stress and the
yield stress of the parent austenite both increase with increasing strain rate and
eventually attain a common value at a critical strain rate within the superelastic
temperature range. At strain rates below the critical level, stress-induced martensites
are formed, whereas the dislocation-induced slip deformation occurs at strain rates
above this level. The yield stress of this material follows the yielding behavior of the
resulting martensite at strain rates below the critical level, and the yield stress of
the parent austenite for strain rates exceeding this level. As is seen in Fig. 8, the
transition stress also increases with increasing strain rate for our porous SMA, as
does the yield stress of the resulting stress-induced martensite phase.
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2329

Fig. 8. Stress–strain curves of the porous SMA at indicated strain rates; the sample corresponding to the
upper black curve failed at a strain somewhat greater than 15%.

3. Constitutive modeling

While the kinetics and kinematics of the elastic–plastic deformation of crystalline


metals of a number of crystal structures have been fully understood and modeled
based on dislocation-induced crystallographic slip and micromechanical averaging
techniques (see, e.g., Nemat-Nasser, 2004, Chapters 6 and 8, and references cited
therein), we are not aware of similar developments for superelasticity and plasticity
of SMAs. Most basic developments in this area have focused on the kinematics of
phase transformation and phenomenological thermodynamics of quasi-static
deformations. Therefore, in what follows, we will only focus on modeling the
quasi-static response of our porous SMA, using standard micromechanical models
(Nemat-Nasser and Hori, 1999) and phenomenological thermodynamics, leaving the
modeling of the strain-rate effects to a future effort.

3.1. Mori– Tanaka method for three-phase composite

The micromechanical model follows the classical inclusion approach (Mura, 1987;
Nemat-Nasser and Hori, 1993), based on the Eshelby (1957) and Mori and Tanaka
(1973) models, with a uniform phase transformation strain, eT , assigned to each
aligned ellipsoidal inclusion (martensite plates). In the forward (reverse) transforma-
tion, the martensite (austenite) product phase and voids are modeled as inclusions
embedded in an infinitely extended austenite (martensite) medium. The effect of the
interaction between the martensite (austenite) plates and voids is ignored. A
schematic diagram for the modeling purposes is depicted in Fig. 9. The elasticity
tensors of the austenite and martensite phases are denoted by CA and CM ,
ARTICLE IN PRESS
2330 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

Fig. 9. Schematics of a standard three-phase model of porous NiTi shape-memory alloy.

respectively2. Under an applied far-field uniform stress, r̄, the stress–strain relation
of a homogenous comparison material with the austenite elasticity, is given by

s̄ij ¼ C A 0
ijkl kl , (3.1)

where quantities with subscripts represent components referred to a fixed rectangular


Cartesian coordinate system, repeated indices are summed, and e0 is the
corresponding uniform strain. When the three-phase system is subjected to the
same far-field stress, r̄, the presence of the inclusions alters the mean stress in the
~ that corresponds to a mean strain, e~ .
matrix from r̄ by a mean interaction stress, r,
The average (uniform) stress and strain of the austenitic phase are given by
0 
sA ~ ij ¼ C A
ij ¼ s̄ij þ s ~kl ;
ijkl kl þ 
(3.2)
A 0
ij ¼ ij þ ~ij ;

so that the interaction stress and strain are related by

s~ ij ¼ C A ~ kl .
ijkl  (3.3)

The stress and strain fields within the martensite inclusion, rM and eM , are also taken
to be uniform. They consist of the corresponding terms for the austenite phase plus a
2
The superscripts ‘‘A’’, ‘‘M’’ and ‘‘v’’ are used to represent the uniform quantities (unless explicitly stated
otherwise) corresponding to the austenite, martensite, and the voids, respectively, without super bars, for
notational simplicity.
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2331

perturbation term due to the presence of inhomogeneities within the composite.


Thus, rM and eM are written in component form as
 
sM ~ ij þ spt
ij ¼ s̄ij þ s ij
M
¼ CM 0 ~kl þ pt
ijkl kl þ  kl
M
 Tkl ;
(3.4)
M 0
ij ¼ ij þ ~ij þ pt
ij
M
;

where rpt M and ept M are the corresponding (uniform) perturbation fields in the
martensite inclusions, and eT is the inelastic strain due to the martensitic phase
transformation. In accordance with the Eshelby model, an inhomogeneity can be
viewed equivalently as an inclusion of the same elasticity as that of the matrix but
with an equivalent eigenstrain; see, e.g., Mura (1987) and Nemat-Nasser and Hori
(1993). The equivalent (uniform) eigenstrain is denoted by enM for the martensite
plates, and by env for the voids. With the additional phase transformation strain, eT ,
the total eigenstrain in the product phase is enM þ eT . The average perturbed strain,
ept M , in (3.4) relates to the total eigenstrain, enM þ eT , by
 nM 
pt
ij
M
¼ SM T
ijkl kl þ kl , (3.5)
where S M ijkl is Eshelby’s S-tensor for the inhomogeneity (martensite plate or voids).
The stress and strain in the martensite phase are now expressed as
M
0  nM  
sMij ¼ C ijkl kl þ  ~kl þ SM T T
klmn mn þ mn  kl
 nM  (3.6)
M 0
ij ¼ ij þ  ~ij þ S M T
ijkl kl þ kl :

On the other hand, based on the equivalent inclusion method, the stress in an
inclusion can be equivalently expressed as
 
sM A 0
~kl þ pt
ij ¼ C ijkl kl þ  kl
M
 Tkl  kl
nM

h   i
ð4sÞ  T
¼ CA 0
~kl þ S M
ijkl kl þ  klmn  1klmn
nM
mn þ mn , ð3:7Þ

where 1ð4sÞ is the fourth-order symmetric identity tensor with components,

1ð4sÞ 1
ijkl ¼ 2ðdik djl þ dil djk Þ, (3.8)

with dij being the Kronecker delta. The consistency of (3.6) and (3.7) now requires
that
0  nM  
CM ijkl kl þ  ~ kl þ S M T
klmn mn þ mn  kl
T
h   i
ð4sÞ  T
¼ CA 
ijkl kl
0
þ ~
 kl þ S M
klmn  1 klmn  mn þ  nM
mn , ð3:9Þ

yielding the total eigenstrain in the martensite phase,


h  i1
ijnM þ Tij ¼ C M A M A
ijmn  C ijmn S mnkl þ C ijkl
h    i
  CM klpq  C A
klpq C 1A
pqgh s̄gh þ ~
 pq þ C M T

klpq pq . ð3:10Þ
ARTICLE IN PRESS
2332 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

Likewise, the average strain and stress for the voids are expressed by
h   i
ð4sÞ
svij ¼ C A 
ijkl kl
0
þ ~
 kl þ S v
klmn  1 nv
klmn mn ¼ 0,

vij ¼ 0ij þ ~ij þ pt


ij
v
¼ 0ij þ ~ij þ S vijkl kl
nv
, ð3:11Þ
where the corresponding stress, strain, and the S-tensor are marked with superscript
‘‘v’’ to represent the values associated with the voids. From (3.11) the eigenstrain for
the voids becomes
 1  
ijnv ¼  S vijkl  1ð4sÞ
ijkl C 1A ~kl .
klmn s̄mn þ  (3.12)

Taking the volume average of the stress field, it follows that


s̄ij ¼ f A sA M
ij þ f M sij , (3.13)
where f A and f M denote the volume fraction of the austenite and martensite,
respectively. Substituting from (3.2) and (3.6) into (3.13) and using (3.10), we obtain
the (common, uniform) interaction strain, e~, to be
h   M i1
~ ij ¼ 1  f v 1ð4sÞ
ijkl  f M U ijmn C mnkl  C A
mnkl
nh   i o
ð4sÞ
 f M U klgh C M ghpq  C A
ghpq þ f 1
v klpq C 1A

pqrs rs  f U C M
 T
M klgh ghpq pq ,

ð3:14Þ
where f v is the volume fraction of the voids, and the tensor U is defined by
 h  i1
ð4sÞ
U ijkl ¼ S Mijmn  1 ijmn C M
mnpq  C A
mnpq S M
pqkl þ C A
mnkl . (3.15)

Taking the volume average of the strain field, we have


¯ ij ¼ f A A M v
ij þ f M ij þ f v ij , (3.16)
and upon substitution from (3.2), (3.6), and (3.11), we have
 
nM T nv
¯ ij ¼ C 1A
ijkl s̄ kl þ f 
M ij þ ij þ f v ij . (3.17)

Eq. (3.17) in conjunction with (3.10), (3.12), and (3.14) can be used to estimate the
stress–strain response of a porous SMA, once the volume fraction of martensite, f M ,
is known.

3.2. Multi-phase composite model (MPCM)

Compared to the modified Mori–Tanaka method introduced in Section 3.1, there


is a more general method, called the double-inclusion method, introduced by Nemat-
Nasser and Hori (1993, 1999), where the interaction between inclusion and matrix,
and inclusion and inclusion is included in the analysis to a certain extent. It has been
proved by Hori and Nemat-Nasser (1993) and Hu and Weng (2000) that the double-
inclusion method includes, as special cases, some of the often-used homogenization
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2333

Fig. 10. A standard multi-phase composite model for porous NiTi shape-memory alloy.

methods, including the self-consistent model (Hill, 1965), the Mori–Tanaka


model (Mori and Tanaka, 1973), the Kuster and Toksoz model (Kuster and
Toksoz, 1974), and the Ponte Castaneda and Willis model (Ponte Castaneda and
Willis, 1995).
Based on the double-inclusion method, Nemat-Nasser and Hori (1993, 1999)
introduce a multi-phase composite model (MPCM), where the interaction among
adjacent inclusions is estimated by considering the interaction of a composite system
with its surrounding homogenized infinite domain. Using this approach, Nemat-
Nasser and Hori suggest that, in a composite consisting of inclusions of different
materials, the interaction among various constituents may be accounted for, in an
approximate manner, by choosing a suitable arrangement of different inclusions in a
multi-phase composite model (Fig. 10). We briefly describe this multi-phase
composite model in Appendix A.

4. Thermodynamics and transformation criterion

The decrease in the Gibbs’ free energy is assumed to provide the driving force for
the formation of martensites during phase transformation. With (uniform)
temperature and stress as the two external loading parameters, the change in the
Gibbs’ free energy is assumed to be dominated by two components. The first is the
change in the mechanical potential energy, and the second is the difference between
ARTICLE IN PRESS
2334 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

the chemical energy of the new and parent phases,


DGðs̄ij ; T; f M ; f v Þ ¼ DGme ðs̄ij ; f M ; f v Þ þ DG ch ðT; f M Þ, (4.1)
where G is the Gibbs’ free energy, and T is the absolute temperature. The subscripts
‘‘me’’ and ‘‘ch’’ stand for the mechanical potential energy and chemical free energy.
The mechanical potential energy, G me , is given by the stress work and the work of the
applied surface tractions,
Z Z
1
G me ðs̄ij ; f M ; f v Þ ¼ sij ij dV  F i ui dS
2 V S
Z
1
¼ ðs̄ij þ s~ ij þ spt 0
~ij þ pt
ij Þðij þ 
T
ij  ij Þ dV
2 V
Z
 F i ðu0i þ u~ i þ upti Þ dS, ð4:2Þ
S

where quantities denoted by the tilde and ‘‘pt’’ are now field variables, corresponding
to variable stress, strain, and displacement. In (4.2), the surface tractions, F, relate to
the uniform overall stress, r̄, by F j ¼ ni s̄ij , where n with components ni , is the
outward unit vector normal to the external surface of the material.
The mechanical potential energy at the reference state, with no martensite phase, is
written as
Z   
1
G 0me ðs̄ij ; f v Þ ¼ s̄ij þ s~ ij þ spt
ij  0
ij þ ~
 ij þ  pt
ij dV
2 V
Z
 
 F i u0i þ u~ i þ upt i dS, ð4:3Þ
S

where all the field quantities now correspond to a two-phase system consisting of the
austenite matrix and voids. The change of the mechanical potential energy thus
becomes
DG me ðs̄ij ; f M ; f v Þ ¼ G me ðs̄ij ; f M ; f v Þ  G 0me ðs̄ij ; f v Þ. (4.4)
Upon substitution from (4.2) and (4.3) into (4.4), and using the model developed in
Section 3, we obtain
1 h   i
DG me ðs̄ij ; f M ; f v Þ ¼  f M s̄ij ijnM þ Tij þ sM 
ij ij
T
2
1    1
1
 f v s̄ij ijnv þ 1  f v Svijkl  1ð4sÞ
ijkl C 1 A
klmn mn . ð4:5Þ

2
The change in the free chemical energy from the reference to the current state can
be expressed as
DG ch ðT; f M Þ ¼ f M ðT  T 0 ÞDS A!M , (4.6)
where DSA!M is the change of entropy from austenite to martensite, T 0 is the
equilibrium reference temperature, and T is the current temperature. Substituting
Eqs. (4.6) and (4.5) into (4.1), we arrive at an expression for the total change of
Gibbs’ free energy, DG.
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2335

During the phase transformation from austenite to martensite, the reduction of


Gibbs’ free energy gives rise to a thermodynamic force, f drv , that drives the process.
This thermodynamic driving force is the negative of the gradient of the change in the
Gibbs’ free energy with respect to the volume fraction of martensites,
qDGðs̄ij ; T; f M ; f v Þ qDG me ðs̄ij ; f M ; f v Þ
f drv ¼  ¼ þ ðT  T 0 ÞDS A!M . (4.7)
qf M qf M
The creation of martensitic plates requires surface (interface) energy, Gs . In
addition, the phase transformation is a dynamic process, requiring activation energy
and involving energy dissipation. Here, however, we consider only equilibrium states
where the driving thermodynamic force is balanced by an effective resistive force,
associated with an energy term, say, G d , giving rise to a conjugate thermodynamic
resistance force.
The surface energy is related to the total surface area of the martensite plates. For
a unit volume of the three-phase system, the change of surface energy can be
presented as
DG s ¼ gs AM ¼ f M gs ðAM =V M Þ ¼ 2f M gs =t, (4.8)
where gs is the interface energy density, V M and AM are the total volume and surface
area of the product domain, and the average thickness of the martensite plates is
denoted by t. The resistive energy term, Gd , is associated with the domain wall
movement. The sources of this resistance during phase transformations may be
related to the interactions of moving interfaces with various defects, e.g., point defects
(solute and impurity atoms, and vacancies), dislocations, grain, subgrain and twin
boundaries, and the lattice resistance or the Peierls barrier, as has been pointed out by
Levitas and Shvedov (2002). Such an energy term is often used as a criterion or
threshold for the phase transformation; see e.g., Buisson et al. (1991) and Berveiller
et al. (1991), as well as Ghosh and Olson (1994), Levitas (1997) and Levitas (1998).
Experimental data suggest that the growth of a new martensitic phase first
proceeds slowly when the temperature drops below M s , then it becomes rapid, and
finally slow again before it is completed. To define the energy barrier, G d , we note
that it should be an increasing function of the martensite volume fraction, and
consider the following empirical relation:
DG d ¼ Hðeb0 f M þ eb1 f M  2Þ, (4.9)
where H, b0 , and b1 are constants to be determined, but only b0 and b1 are
independent model constants. A similar expression has been used by Li and Weng
(2001b) to characterize the energy barrier during the domain transformation of
ferroelectric materials. This function is convex and it reduces to zero when f M is
equal to zero. As is shown below, this empirical relation actually works well to
reproduce the experimental stress–strain relation in quasi-static loading. We note
however that, at high deformation rates, it would be necessary to consider rate
equations based on the kinetics of phase transformation, including drag and other
time-dependent dissipative mechanisms. Recent experimental results of Nemat-
Nasser and coworkers show a strong rate-dependence of martensitic phase
ARTICLE IN PRESS
2336 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

transformation that requires fundamental kinetic modeling, currently unavailable;


see Nemat-Nasser and Choi (2005) and Nemat-Nasser et al. (2005a, b).
The total resistance force thus is
qðDG s þ DG d Þ
f resis ¼ ¼ gs ðAM V M Þ þ Hðb0 eb0 f M  b1 eb1 f M Þ. (4.10)
qf M
Setting the thermodynamic driving force in Eq. (4.7) equal to the resistance force in
Eq. (3.10), a balance equation that governs the martensitic transformation is
obtained,
qGme ðs̄ij ; f M ; f v Þ  
 þ ðT  T 0 ÞDS A!M ¼ 2gs =t þ H b0 eb0 f M  b1 eb1 f M .
qf M
(4.11)
This equation gives the equilibrium volume fraction of the product phase, f M , for
given external stress and temperature. Then the overall stress–strain behavior of the
porous SMA is obtained using Eq. (3.17). As is well-known (Hill, 1965), in the
Eshelby approach to micromechanical modeling of composites and polycrystals, we
do not distinguish inclusions from the matrix: in the basic homogenization
equations, they are symmetrically treated. In the present case also, a similar
situation exists: for the reverse transformation from fully martensite to austenite, the
austenites (and voids) are viewed as embedded inclusions in a martensite matrix.

5. Application of model and results

The three-phase model developed in Section 3.1 is used to simulate the overall
stress–strain curve of the porous NiTi SMA during the quasi-static stress-induced phase
transformation at various constant temperatures. The Young’s moduli and Poisson’s
ratios for solid austenite and martensite are taken as E A ¼ 70 GPa; E M ¼ 30 GPa; nA ¼
0:33 and nM ¼ 0:33. The phase transformation is accompanied by a change in the crystal
structure from the body-centered-cubic to the monoclinic state. The local normal and
shear components of the phase transformation strain have been calculated according to
Khachaturyan (1983) as normal ¼ 0:1174% and shear ¼ 6:643%. For our modeling
purposes, we consider a total of 24 potential plate orientations for the NiTi crystal during
phase transformation, which accords with the invariant plane theory. At a given stress,
the orientation that yields the greatest driving force in (4.7) is assumed to be the most
favorably oriented one. A list of the 24 potential orientations considered here can be
found in Lu and Weng (1998), and is not repeated here.
Eqs. (4.11) and (3.17) are used to calculate the stress–strain response for the
forward martensite transformation under given stresses and temperatures. It is
noticed that Eq. (4.11) could be further simplified when the test is under constant3
3
For high rate deformations, only the initial temperature can be controlled, since the response is
essentially adiabatic and both mechanical deformation and phase transformation affect the temperature of
the sample.
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2337

temperature T. The independent constants are set as b0 ¼ 1:00 and b1 ¼ 1:40. The
entropy change during martensite transformation is found by Lu and Weng (1998) to
be 0:7024 J=mm3 . The transformation stress, rs , at which the martensite
transformation starts, and the finish stress, rf , at which the transformation finishes,
are obtained from the compression test at 296 K. It is observed from Fig. 2 that
ss33
200 MPa and sf33
800 MPa. Thus we obtain the initial and finishing
conditions as
s̄ij ¼ ssij ; T ¼ 296 K; f M ¼ 0,
s̄ij ¼ sfij ; T ¼ 296 K; f M ¼ 1  f v. ð5:1Þ
By applying these two conditions in (5.1) to (4.11), we calculate the values of the
constants H and 2gs =t. We then use these constants for all other tests at various
temperatures. The relatively large number of experimental curves at various test
temperatures then verifies the appropriateness of the model assumptions. For the
backward transformation from martensite to austenite upon unloading, the
micromechanical and thermodynamic relations in Sections 3 and 4 remain
applicable, since, as pointed out before, the model treats both phases symmetrically;
see Hill (1965). Hence, simple modification of the basic equations, i.e., interchanging
the roles of the parent and product phases, will produce the necessary equations.
The computed result of the stress–strain response at a temperature of 296 K is
displayed in Fig. 11 along with the experimental data (same as in Fig. 2) which are
used to fix the model parameters; hence showing good correlation. The selected and
fixed coefficients, b0 and b1 , also enable the model to capture some of the essential

1600

Porous NiTi, T=296K, 0.001/s, f =12%

1200
Compression, σ (MPa)

800

Exp No.1
Exp No.2
400 Computation

0
0 5 10 15 20 25
Longitudinal strain, ε (%)

Fig. 11. Calculated and measured results of the stress–strain response for porous NiTi shape-memory
alloy at a constant temperature of 296 K.
ARTICLE IN PRESS
2338 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

1200

Porous NiTi, 0.001/s, f =12%


1000
Compression, σ (MPa)

800

T=323K
600

T=296K
400
T=273K
200 T=253K

0
0 2 4 6 8 10
Longitudinal strain, ε (%)

Fig. 12. Calculated and measured results of the temperature dependence of the martensite transformation
for porous NiTi shape-memory alloy.

characteristics of the phase transformation. The temperature dependence of the


martensite transformation is shown in Fig. 12. The theoretical and experimental
results are compared at temperatures of, from bottom to top in Fig. 12, 253, 273, 296
and 323 K. The solid lines are the corresponding theoretical, and the marked dashed
lines are the experimental results. From the results in Fig. 12, we note that the
transformation stress, rs , increases as we increase the imposed (constant)
temperature. Such temperature dependence of rs can be observed more clearly in
Fig. 13, where both the calculated and measured transformation stresses are plotted
as a function of the temperature. It is seen that the current model has good
temperature sensitivity for temperature induced martensite transformation. The
loading and unloading curves corresponding to Fig. 2 are shown in Fig. 14. Since the
plastic deformation may exist at temperatures higher than 296 K, the theoretical
result for 323 K is not given at this time. Figs. 15 and 16 provide addition verification
of the model predictions.
Among all the theoretical and experimental results shown above, the results at
constant temperatures 273 and 296 K are two important representatives, because
these temperatures are very close to As and Af that are reported in Section 5.1. This
means that, without considering the plastic deformation, all the results above 296 K
fall into the category of superelasticity. The strain due to the martensite phase
transformation will be fully recovered upon unloading. On the other hand, the rest of
the results fall into temperatures ranging from M s to As , where we expect M s to be
slightly lower than 253 K. The strain due to phase transformation cannot be fully
recovered upon unloading as shown in the results. Since the applied temperatures are
all above M f , which is expected to be less than 233 K, only phase transformation is
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2339

700

Porous NiTi, 0.001/s, f =12%


600

Transition stress, σs (MPa) 500 Experiment


Computation

400

300

200

100

0
250 275 300 325 350 375
Temperature, T (K)

Fig. 13. Temperature dependence of the transformation stress rs during the martensite.

1200

Porous NiTi, T=296K, 0.001/s, f =12%


1000
Compression, σ (MPa)

800

600

400

Experiment
200 Computation

0
0 2 4 6 8 10
Longitudinal strain, ε (%)

Fig. 14. Calculated and measured stress–strain response of porous NiTi shape-memory alloy at a constant
temperature of 296 K, during the A–M and M–A transformation.

considered. This means that there is no shape-memory effect due to the reorientation
of the twins above M f . Thus, it is seen from both the theoretical and experimental
results that SMAs with low porosity still display superelasticity and shape-memory
effect, similarly to the corresponding solid SMAs, described by Sun and Hwang
(1993).
ARTICLE IN PRESS
2340 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

1200

1000 Porous NiTi, T=273K, 0.001/s, f =12%


Compression, σ (MPa)

800

600

400

Experiment
200 Computation

0
0 2 4 6 8 10
Longitudinal strain, ε(%)

Fig. 15. Calculated and measured stress–strain response of porous NiTi shape-memory alloy at a constant
temperature of 273 K, during the A–M and M–A transformation.

1000

Porous NiTi, T=253K, 0.001/s, f =12%


800
Compression, σ (MPa)

600

400

200
Experiment
Computation

0
0 2 4 6 8 10
Longitudinal strain, ε (%)

Fig. 16. Calculated and measured stress–strain response of porous NiTi shape-memory alloy at a constant
temperature of 253 K, during the A–M and M–A transformation.

6. Concluding remarks

In a series of uniaxial, quasi-static and dynamic, compression tests of a new


porous SMA of 12% porosity, the superelastic response is experimentally
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2341

characterized at various temperatures. For the dynamic tests, a new pulse-shaping


technique is introduced to obtain relatively constant strain rates. These tests show
that, our porous SMA has strain-rate sensitivity similar to the solid NiTi SMAs. In
addition, a micromechanical model is presented to predict the (quasi-static and
constant-temperature) stress-induced phase transformation of a porous shape-
memory alloy with small porosity. This model is based on the modified equivalent
inclusion method for a three-phase system, and equilibrium thermodynamics. This
three-phase micromechanics model is a special case of the multi-phase composite
model (MPCM), and can be obtained as a special case of MPCM by assuming that
the martensite phase and the pores are ellipsoids with common aspect ratios and
orientations. An equation that balances the thermodynamic driving force with the
corresponding resistive force is set up to calculate the volume fraction of the
martensitic phase for given stress and temperature. The overall stress–strain
behavior of the porous SMA and its temperature dependence are then obtained
using the martensite volume fraction. The theoretical results are found to be in
general agreement with our quasi-static experimental data.

Acknowledgements

This work was supported by ONR (MURI) Grant N000140210666 to the


University of California, San Diego, with Dr. Roshdy G. Barsoum as Program
Manager. The authors wish to express their appreciation to Professor M. Taya of the
University of Washington for supplying the samples of the porous NiTi SMA.

Appendix A. Mori–Tanaka method and MPCM

The multi-phase composite model (MPCM) is based on the averaging procedure


of the double-inclusion model proposed by Nemat-Nasser and Hori (1993). To find
the relation between the multi-phase composite model and the modified Mori–
Tanaka method used in Section 3.1, we now consider a three-phase composite
without the strain due to phase transformation. We denote the austenite as phase 0;
the martensite and pores as phases 1 and 2, respectively.
The consistency conditions for the modified Mori–Tanaka method are
h i h   i
ð4sÞ
C 1ijkl 1
klmn þ S 1
 n1
klpq pqmn ¼ C 0
1
ijkl klmn þ S 1
klpq  1  n1
klpq pqmn
h i h   i
ð4sÞ
C 2ijkl 1 2 n2 0 1 2 n2
klmn þ S klpq pqmn ¼ C ijkl klmn þ S klpq  1klpq pqmn ¼ 0. ðA:1Þ

This leads to the following expressions for the eigenstrains:


 1 1  
ijn1 ¼ C 0ijmn  C 1ijmn C 0mnkl  S 1ijkl 1
kl ¼ A 1
ijkl  S 1 1
ijkl kl
 1
ijn2 ¼ 1ð4sÞ
ijkl  S 2
ijkl 1
kl . ðA:2Þ
ARTICLE IN PRESS
2342 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

The overall average stress and strain are then given by


h   i
ð4sÞ
s̄ij ¼ f 0 C 0ijkl 1
kl þ f C 0

1 ijkl kl
1
þ S 1
klmn  1 n1
klmn mn
  1
 
¼ C 0ijkl 1  f 2 1ð4sÞklmn þ f 1 S 1
klpq  1 ð4sÞ
klpq A1
pqmn  S 1
pqmn 1
mn ðA:3Þ

and
   
1 n1 2 n2
¯ ij ¼ f 0 1 1 1
ij þ f 1 ij þ S ijkl kl þ f 2 ij þ S ijkl kl
 1
ð4sÞ 1
 1 1
1 2 ð4sÞ 2
¼ 1ijkl þ f 1 S ijmn Amnkl  S mnkl þ f 2 Sijmn 1mnkl  S mnkl 1
kl . ðA:4Þ

The corresponding effective elasticity tensor then is


  1
MMT 0
  ð4sÞ 1 ð4sÞ 1 1
C̄ ijkl ¼ C ijmn 1  f 2 1mnpq þ f 1 S mngh  1mngh Aghpq  S ghpq
 1 1
 1 1
 1ð4sÞ
pqkl þ f S 1
1 pqrs A rskl  S 1
rskl þ f S 2
2 pqrs 1 ð4sÞ
rskl  S 2
rskl .

ðA:5Þ
On the other hand, we can obtain the effective elasticity tensor using the
multi-phase composite model. For a three-phase composite system, illustrated in
Fig. 9, we may consider an ellipsoid V containing two ellipsoidal heterogeneities,
Oa ; ða ¼ 1; 2Þ; see Fig. 10. The remaining part of V is denoted by G. The elasticities of
Oa and G are Ca and C0 , respectively. This multi-phase composite is then embedded
in an infinite domain B with uniform elasticity C. The consistency conditions for the
MPCM are
   
C 0ijkl 1 d 0
kl þ hkl i ¼ C ijkl 1 d 0
kl þ hkl i  
n0
¯ kl ,
1
1 d 1
 1 d 1 n1

C ijkl kl þ hkl i ¼ C ijkl kl þ hkl i  ¯ kl ,
   
C 2ijkl 1 d 2
kl þ hkl i ¼ C ijkl 1 d 2
kl þ hkl i  
n2
¯ kl , ðA:6Þ
or simply written as
1   
CK d K
ijkl kl þ hkl i ¼ C ijkl 1 d K
kl þ hkl i  
nK
¯ kl ,
K ¼ 0; 1; 2 (not summed), ðA:7Þ
nK d K
where ē is the average eigenstrain, and he i is the volume average of strains,
produced by the eigenstrain field, taken over the corresponding region. To estimate
hed ia for a ¼ P 1; 2, we assume that ē na is uniformly distributed in Oa , and the average
2 nb a
eigenstrain, baa ðf b =1  f a Þē , is uniformly distributed in V  O . Then the
a
average strain over O becomes
 1  V  
hdij i1 ¼ S 1ijkl ¯ kl
n1
þ 1 f1 Sijkl  S1ijkl f 0 ¯ kln0 n2
þ f 2 ¯ kl ,
    
1
hdij i2 ¼ S 2ijkl ¯ kl
n2
þ 1 f2 SV 2
ijkl  S ijkl
n0
f 0 ¯ kl n1
þ f 1 ¯ kl , ðA:8Þ
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2343

or simply written as
( )
  X2
fb nb
hdij ia ¼ S aijkl ¯akl þ SV a
ijkl  S ijkl ¯ ; a ¼ 1; 2 (not summed),
baa
1  f a kl
(A.9)
a V a
where S and S are Eshelby’s S-tensors for O and V , respectively. For the average
strain in G, from hed iV ¼ SV : hē n iV and (A.8), it follows that
  1  V  
hdij i0 ¼ S V 
¯ n0
ijkl kl þ f f
1 0 1  f 1 S ijkl  S 1
ijkl
n1
¯ kl  ¯ nkl
  1  V  
þ f2 f0 1f2 S ijkl  S 2ijkl ¯ kl
n2
 ¯ nkl . ðA:10Þ

The average eigenstrain over V is given by


¯ nij ¼ f 0 ¯ ijn0 þ f 1 ¯ ijn1 þ f 2 ¯ ijn2 . (A.11)
Now, assume that the elasticity tensor of phase 0 is the same as that of the infinite
matrix B, i.e., C ijkl ¼ C 0ijkl , and that all these phases are coaxial ellipsoids with a
common aspect ratio. The Eshelby’s S-tensors are all the same, i.e.,
Sijkl ¼ S V 1 2
ijkl ¼ S ijkl ¼ S ijkl , (A.12)
and we have
hdij i0 ¼ S ijkl ¯ kl
n0
,
hdij i1 ¼ S ijkl ¯ kl
n1
,
hdij i2 ¼ S ijkl ¯ kl
n2
. ðA:13Þ
By substituting these expressions back into the consistency equations, one obtains
the corresponding eigenstrains as
¯ ijn0 ¼ 0,
 1 1  1
¯ ijn1 ¼ C 0ijmn  C 1ijmn C 0mnkl  Sijkl 1 1
kl ¼ Aijkl  S ijkl 1
kl
 1
¯ ijn2 ¼ 1ð4sÞ
ijkl  S ijkl 1
kl . ðA:14Þ

The overall eigenstrain defined by (A.11) then is


¯ nij ¼ f 0 ¯ ijn0 þ f 1 ¯ ijn1 þ f 2 ¯ ijn2
 1  1
¼ f 1 A1ijkl  S ijkl þ f 2 1ð4sÞ
ijkl  S ijkl 1 1
kl ¼ Āijkl kl , ðA:15Þ

which also defines Āijkl . Then the overall stress and strain are given by
h   i
ð4sÞ
s̄ij ¼ C 0ijkl 1
kl þ S klpq  1klpq  ¯ npq
h   i
¼ C 0ijkl 1ð4sÞ
klmn þ S klpq  1 ð4sÞ
klpq Ā pqmn 1
mn ðA:16Þ
ARTICLE IN PRESS
2344 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

and
 
ð4sÞ
¯ ij ¼ 1 ¯ nmn ¼ 1ijkl þ Sijmn Āmnkl 1
ij þ S ijmn  kl . (A.17)

Thus the overall effective elasticity tensor of the composite system can be written as
h   ih i1
MPCM
C̄ ijkl ¼ C 0ijgh 1ð4sÞ
ghmn þ S ghrs  1 ð4sÞ
ghrs Ā rsmn 1 ð4sÞ
mnkl þ S Ā
mnpq pqkl
n   1
ð4sÞ ð4sÞ
¼ C 0ijgh 1ghmn þ f 1 Sghrs  1ghrs A1rsmn  S rsmn
  1 o
þf 2 S ghrs  1ð4sÞ
ghrs 1 ð4sÞ
rsmn  S rsmn
 1  1 1
ð4sÞ 1 ð4sÞ
 1mnkl þ f 1 Smnpq Apqkl  S pqkl þ f 2 Smnpq 1pqkl  S pqkl .

ðA:18Þ
After rearrangement of (A.18), the overall effective elasticity tensor based the
MPCM is given by
h   1 i
MPCM ð4sÞ  1
C̄ ijkl ¼ C 0ijgh ð1  f 2 Þ1ð4sÞ
ghmn þ f 1 S ghrs  1ghrs Arsmn  S rsmn
 1   1
 1ð4sÞmnkl þ f S
1 mnpq A1
pqkl  S pqkl þ f S
2 mnpq 1 ð4sÞ
pqkl  S 1
pqkl .

ðA:19Þ
It can be seen that this final expression is the same as the result of the modified
Mori–Tanaka method given by (A.5). This shows that the three-phase model
introduced in Section 3.1 is a special case of the MPCM, which is based on the
double-inclusion method.

References

Abeyaratne, R., Knowles, J.K., 1993. A continuum model of a thermoelastic solid capable of undergoing
phase transformations. J. Mech. Phys. Solids 41 (3), 541–571.
Ayers, R.A., Simske, S.J., Bateman, T.A., Petkus, A., Sachdeva, R.L.C., Gyunter, V.E., 1999. Effects of
nitinol implant porosity on cranial bone ingrowth and apposition after 6 weeks. J. Biomed. Mater. Res.
45 (1), 42–47.
Berkowitz, A., 2004. Private communication.
Berveiller, M., Patoor, E., Buisson, M., 1991. Thermomechanical constitutive equations for shape memory
alloys. J. Phys. IV 1 (C4), 387–396.
Bowles, J.S., Mackenzie, J.K., 1954. The crystallography of martensitic transformations. Acta Metall. 2,
129–147.
Buisson, M., Patoor, E., Berveiller, M., 1991. Interfacial motion in shape memory alloys. J. Phys. IV 1
(C4), 463–466.
Entchev, P.B., Lagoudas, D.C., 2002. Modeling porous shape memory alloys using micromechanical
averaging techniques. Mech. Mater. 34, 1–24.
Eshelby, J.D., 1957. The determination of the elastic field of an ellipsoidal inclusion, and related problems.
Proc. R. Soc. Lond. A 241, 376–396.
ARTICLE IN PRESS
S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346 2345

Ghosh, G., Olson, G.B., 1994. Kinetics of F.C.C.-B.C.C. heterogeneous martensitic nucleation—I. The
critical driving force for athermal nucleation. Acta Metall. Mater. 42, 3361–3370.
Hill, R., 1965. A self-consistent mechanics of composite materials. J. Mech. Phys. Solids 13, 213–222.
Hori, M., Nemat-Nasser, S., 1993. Double-inclusion model and overall moduli of multiphase composites.
Mech. Mater. 14, 189–206.
Hu, G.K., Weng, G.J., 2000. The connections between the double-inclusion model and the Ponte
Castaneda–Willis, Mori–Tanaka, and Kuster–Toksoz model. Mech. Mater. 32, 495–503.
Khachaturyan, A.G., 1983. Theory of Structural Transformation in Solids. Wiley, New York.
Kuster, G.T., Toksoz, M.N., 1974. Velocity and attenuation of seismic waves in two-phase media: I
theoretical formulation. Geophysics 39, 587–606.
Lagoudas, D.C., Vandygriff, E.L., 2002. Processing and characterization of NiTi porous SMA by elevated
pressure sintering. J. Intell. Mater. Systems and Structures 13 (12), 837–850.
Lagoudas, D.C., Entchev, P.B., Vandygriff, E.L., Qidwai, M.A., DeGiorgi, V.G., 2000. Modeling of
thermomechanical response of porous shape memory alloys. Proc. SPIE-Int. Soc. Opt. Eng. 496–508.
Lagoudas, D.C., Entchev, P.B., Vandygriff, E.L., 2001. Fabrication, modeling and characterization of
porous shape memory alloys. Proc. SPIE-Int. Soc. Opt. Eng.
Levitas, V.I., 1997. Phase transformations in elastoplastic materials: continuum thermomechanical theory
and examples of control—Part I. J. Mech. Phys. Solids 45 (6), 923–947.
Levitas, V.I., 1998. Thermomechanical theory of martensitic phase transformations in inelastic materials.
Int. J. Solids Struct. 35, 889–940.
Levitas, V.I, Shvedov, L.K., 2002. Low-pressure phase transformation from rhombohedral to cubic BN:
experiment and theory. Phys. Rev. B 65, 104109.
Lexcellent, C., Licht, C., 1991. Some remarks on the modeling of the thermomechanical behaviour of
shape memory alloys. J. Phys. IV 1 (C4), 35–39.
Li, J., Weng, G.J., 2001a. A micromechanics-based hysteresis model for ferroelectric ceramics. J. Intell.
Mater. Systems and Struct. 12, 79–91.
Li, W.F., Weng, G.J., 2001b. Micromechanics simulation of spontaneous polarization in ferroelectric
crystals. J. Appl. Phys. 90 (5), 2484–2491.
Li, B.-Y., Rong, L.-J., Li, Y.-Y., 1998a. Porous NiTi alloy prepared from elemental powder sintering.
J. Mater. Res. 13, 2847–2851.
Li, B.-Y., Rong, L.-J., Luo, X.-H., Li, Y.-Y., 1998b. Transformation behavior of sintered porous NiTi
alloys. Metall. Mater. Trans. A 30A, 2753–2756.
Li, B.-Y., Rong, L.-J., Li, Y.-Y., Gjunter, V.E., 2000. Synthesis of porous Ni-Ti SMA by SHS: reaction
mechanism and anisotropy in pore structure. Acta Mater. 48 (15), 3895–3904.
Lu, Z.K., Weng, G.J., 1997. Martensitic transformation and stress–strain relations of shape-memory
alloys. J. Mech. Phys. Solids 45 (11/12), 1905–1928.
Lu, Z.K., Weng, G.J., 1998. A self-consistent model for the stress–strain behavior of shape-memory alloy
polycrystals. Acta Mater. 46 (15), 5423–5433.
Mori, T., Tanaka, K., 1973. Average stress in matrix and average elastic energy of materials with misfitting
inclusions. Acta Metall. 21, 571–574.
Mura, T., 1987. Micromechanics of Defects in Solids, 2nd ed. Martinus Nijhoff Publishers, Dordrecht.
Nemat-Nasser, S., 2004. Plasticity: A Treatise on the Finite Deformation of Heterogeneous Inelastic
Materials. Cambridge University Press, Cambridge.
Nemat-Nasser, S., Choi, Y.C., 2005. Strain rate dependence of deformation mechanisms in a Ni–Ti–Cr
shape-memory alloy. Acta Mater. 53, 449–454.
Nemat-Nasser, S., Hori, M., 1993. Micromechanics: Overall Properties of Heterogeneous Materials, first
ed. Elsevier, Amsterdam.
Nemat-Nasser, S., Hori, M., 1999. Micromechanics: Overall Properties of Heterogeneous Materials,
second revised ed. Elsevier, Amsterdam.
Nemat-Nasser, S., Isaacs, J.B., Starrett, J.E., 1991. Hopkinson techniques for dynamic recovery
experiments. Proc. Roy. Soc. Lond. 435A, 371–391.
Nemat-Nasser, S., Choi, J.Y., Guo, W.-G., Isaacs, J.B., 2005a. Very high strain-rate response of a NiTi
shape-memory alloy. Mech. Mater. 37, 287–298.
ARTICLE IN PRESS
2346 S. Nemat-Nasser et al. / J. Mech. Phys. Solids 53 (2005) 2320–2346

Nemat-Nasser, S., Choi, J.Y., Guo, W.-G., Isaacs, J.B., Taya, M., 2005b. High strain-rate, small strain
response of a NiTi shape-memory alloy. J. Eng. Mater. Tech. 127, 1–7.
Ponte Castaneda, P., Willis, J.R., 1995. The effect of spatial distribution on the effective behavior of
composite materials and cracked media. J. Mech. Phys. Solids 43, 1919–1951.
Qidwai, M.A., Entchev, P.B., Lagoudas, D.C., DeGiorgi, V.G., 2001. Modeling of the thermomechanical
behavior of porous shape memory alloys. Int. J. Solids Struct. 38, 8653–8671.
Raniecki, B., Lexcellent, C., 1998. Thermodynamics of isotropic pseudoelasticity in shape memory alloys.
Eur. J. Mech. A/Solids 17, 185–205.
Raniecki, B., Lexcellent, C., Tanaka, K., 1992. Thermodynamic models of pseudoelastic behaviour of
shape memory alloys. Arch. Mech. 44 (3), 261–284.
Shabalovskaya, S.A., Itin, V.I., Gyunter, V.E., 1994. Porous Ni–Ti—a new material for implants and
prostheses. Proceedings of the First International Conference on Shape Memory and Superelastic
Technologies, pp. 7–12.
Sun, Q.P., Hwang, K.C., 1993. Micromechanics modeling for the constitutive behavior of polycrystalline
shape memory alloys—I. Derivation of general relations. J. Mech. Phys. Solids 41 (1), 1–17.
Taya, M., 2004. Private communication.
Wechsler, M.S., Lieberman, D.S., Read, T.A., 1953. On the theory of formation of martensite. Trans.
TMS - AIME 197, 1503–1515.

You might also like