Sunteți pe pagina 1din 61

Accepted Manuscript

A micromechanical constitutive model for grain size dependent thermo-mechanically


coupled inelastic deformation of super-elastic NiTi shape memory alloy

Chao Yu, Guozheng Kang, Qianhua Kan

PII: S0749-6419(17)30606-X
DOI: 10.1016/j.ijplas.2018.02.005
Reference: INTPLA 2301

To appear in: International Journal of Plasticity

Received Date: 25 October 2017


Revised Date: 23 January 2018
Accepted Date: 7 February 2018

Please cite this article as: Yu, C., Kang, G., Kan, Q., , A micromechanical constitutive model for grain
size dependent thermo-mechanically coupled inelastic deformation of super-elastic NiTi shape memory
alloy, International Journal of Plasticity (2018), doi: 10.1016/j.ijplas.2018.02.005.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

A micromechanical constitutive model for grain size dependent


thermo-mechanically coupled inelastic deformation of super-elastic
NiTi shape memory alloy
Chao Yua,b, Guozheng Kanga,b*, Qianhua Kana,b

PT
a
State Key Laboratory of Traction Power, Southwest Jiaotong University, Chengdu, Sichuan, China
b
Applied Mechanics and Structure Safety Key Laboratory of Sichuan Province, School of Mechanics

RI
and Engineering, Southwest Jiaotong University, Chengdu, Sichuan, China
*Corresponding author: Dr. Prof. G.Z. Kang, Tel: +86-28-87634671; Fax: +86-28-87600797

SC
E-mail address: guozhengkang@home.swjtu.edu.cn or guozhengkang@126.com

Abstract

U
A micromechanical constitutive model is constructed to describe the grain size dependent
AN
thermo-mechanically coupled inelastic deformation of polycrystalline super-elastic NiTi shape
memory alloys (SMAs). In the proposed model, the polycrystalline aggregate is regarded as a
M

composite material, i.e., each grain-interior (GI) phase is assumed to be a spherical inhomogeneity
D

embedded in a matrix of grain-boundary (GB). The constitutive relationship of GI phase is deduced


in the framework of irreversible thermodynamics, and the driving force of martensite transformation
TE

and the internal heat production caused by transformation latent heat and mechanical dissipation are
obtained by the Clausius’s dissipative inequality and energy-balance equation, respectively. But, the
EP

stress-strain relationship of GB phase is assumed to be linearly elastic without energy dissipation. To


describe the interaction between the GI and GB phases, a new incremental non-isothermal
C

Mori-Tanaka’s homogenization method is developed and the Eshelby’s tensor of a spherical


AC

inclusion embedded in a finite spherical domain is introduced. Moreover, the continuous tangent
moduli of GI and GB phases are deduced and the computational algorithms of the proposed
micromechanical model are developed from fully implicit backward Euler’s integration method.
Finally, the proposed micromechanical model is verified by comparing the simulated results with the
corresponding experimental ones. It is shown that the grain size dependent thermo-mechanically
coupled inelastic deformation of NiTi SMAs can be well captured by the proposed model.
Key words: Polycrystalline NiTi SMAs; grain size-dependence; micromechanical constitutive model;
ACCEPTED MANUSCRIPT

irreversible thermodynamics; Mori-Tanaka’s homogenization method.

1. Introduction
NiTi shape memory alloys (SMAs) are a kind of smart materials which exhibit many excellent
properties such as super-elasticity, shape memory effect, high damping capacity and good
biocompatibility. These excellent properties are determined mainly by a unique thermo-elastic

PT
martensite transformation in the NiTi SMAs. Recently, polycrystalline super-elastic NiTi SMAs have
been increasingly used in many engineering fields, such as aerospace, automotive industry, civil

RI
engineering, micro-electro-mechanical systems and biomedicine (Duerig et al., 1999; Choudhary and
Kaur, 2016; Kang and Kan, 2017; Leng et al., 2017).

SC
It is well-known that the martensite transformation in NiTi SMAs is a typical first-order phase
transition and can lead to a strong thermo-mechanically coupled effect. On the one hand, during a

U
mechanical loading, transformation latent heat and mechanical dissipation will lead to an internal
AN
heat production and then cause a temperature variation. The extent of temperature variation depends
greatly on the competition of internal heat production against the heat transfer/convection to the
M

environment, which is basically determined by the loading rate. On the other hand, existing
experimental results show that the transformation stresses of NiTi SMAs depend strongly on the test
D

temperature (Shaw and Kyriakides, 1995; Yu et al., 2014a). Such a thermo-mechanically coupled
TE

effect results in a low stability of NiTi SMAs, that is, the stress-strain responses are very sensitive to
the loading rate. For instance, the transformation hardening modulus increases with the increase of
EP

loading rate but the dissipation energy changes non-monotonically with a varied loading rate, as
shown in Fig. 1 (Grabe and Bruhns, 2008; He and Sun, 2010, 2011; Morin et al., 2011a, 2011b; Yin
C

et al., 2014; Kan et al., 2016; Zhang et al., 2017).


AC

In the last few decades, much effort had been done to improve the stability of polycrystalline NiTi
SMAs. So far, nano-crystallization has been recognized as a promising way since the grain size is
one of the most important parameters in the microscopic scale and determines the macroscopic
thermo-mechanical response of polycrystalline NiTi SMAs. Adopting the nano-crystallization, the
effect of grain size on the inelastic deformation of polycrystalline super-elastic NiTi SMAs had been
investigated by many experimental observations. For examples, Kim et al. (2006), Cho et al. (2006),
and Ahadi and Sun (2013; 2015) reported that the inelastic deformation of NiTi SMAs depended
ACCEPTED MANUSCRIPT

strongly on the grain size, i.e., when the grain size was larger than 80 nm, the stress-strain response
exhibited a transformation plateau and large hysteresis loop; however, with the decrease of grain size,
the stress-strain response changed from the “plateau-type” one to “hardening-type” one gradually,
and the dissipation energy decreased monotonically, as shown in Fig. 2. As discussed by Ahadi and
Sun (2013; 2015), there are two types of nanoscale interfaces in polycrystalline NiTi SMAs, i.e., the

PT
grain-boundary (GB) (which serves as non-transformable region and separates the grains with
different crystallographic orientations) and austenite-martensite (A-M) interface (which serves as a

RI
transition layer with a nanometer thickness and separates the lattices of austenite and martensite
phases). For the coarse-grained NiTi SMAs, the volume fractions of the two nanoscale interfaces are

SC
very small, and their deformation is dominated by the martensite transformation of grain-interior (GI)
phase. With the decrease of grain size, the volume fractions of the two nanoscale interfaces become

U
larger and larger. When the grain size is reduced to a nanoscale, the mechanical constraints of the GB
AN
to the martensite transformation in the GI phase become very significant, and the deformation of
nano-grained NiTi SMAs is dominated by the GB and A-M interfaces.
M

It should be noted that the aforementioned studies (Kim et al., 2006; Cho et al., 2006; Ahadi and
Sun 2013; 2015) focused only on the grain size-dependent inelastic deformation of NiTi SMAs under
D

an isothermal loading condition. Recently, Ahadi and Sun (2014) further investigated the grain size
TE

dependent thermo-mechanically coupled inelastic deformation of polycrystalline NiTi SMA, and the
results showed that with the reduction of grain size, the stress-strain response of the NiTi SMA
EP

exhibited much weaker rate-dependence, which meant that its stability was significantly improved.
Such a decreased rate-dependence is caused by small temperature variations during the martensite
transformation and its reverse (due to the reduction of internal heat source) and a significantly
C

decreased temperature-dependence of transformation stresses with the reduction of grain size.


AC

In the last three decades, many constitutive models were established to describe the
thermo-mechanically coupled inelastic deformation of SMAs, as reviewed by Lagoudas (2008),
Cisse et al. (2016a; 2016b) and Kang and Kan (2017). The existing models can be classified as three
groups, i.e., the macroscopic phenomenological ones, the crystal plasticity based ones and the
micromechanical ones. The first ones regarded the martensite transformation as a macroscopic
phenomenological process, and the constitutive models were constructed directly from the
macroscopic experimental observations. The macroscopic phenomenological models were firstly
ACCEPTED MANUSCRIPT

constructed to describe the isothermal deformation of SMAs in one loading-unloading cycle under
proportional loading conditions (Brinson, 1993; Boyd and Lagoudas 1996; Leclercq and Lexcellent,
1996; Auricchio et al., 1997; Zaki and Moumni 2007a), and further extended to describe some
complex phenomena, including thermo-mechanically coupled inelastic deformation (Morin et al.
2011a, 2011b; Yu et al., 2015a; Wang et al., 2017a; 2017b; Gu et al., 2017); super-elasticity

PT
degeneration (Lagoudas and Entchev 2004; Auricchio, 2007; Zaki and Moumni, 2007b; Yu et al.,
2014a; 2015a; Zhang et al., 2014; 2016; Wang et al., 2017b); large deformation (Müller and Bruhns

RI
2006; Reese and Christ, 2008; Arghavani et al., 2011; Xiao, 2013; Wang et al., 2017a; 2017b);
non-proportional deformation (Bouvet et al., 2004; Panico and Brinson, 2007; Arghavani et al., 2010;

SC
Chemisky et al., 2011; Lagoudas et al., 2012; Zaki, 2012; Chatziathanasiou et al., 2016), and so on.
The phenomenological models are good candidates for predicting the thermo-mechanically coupled

U
inelastic deformation of SMA components and devices since they do not consider the complicated
AN
microstructures of the material and can be easily integrated into a finite element method. The crystal
plasticity based models was constructed first in the single crystal scale with considering various
M

martensite variants, and then extended to the polycrystalline version by a finite element method or
explicit scale-transition rule (Gall et al., 2000; Thamburaja and Anand, 2001; Anand and Gurtin,
D

2003; Thamburaja, 2005; Patoor et al. 2006; Manchiraju and Anderson, 2010; Yu et al., 2013, 2014b,
TE

2015b; Long et al., 2017). As commented by Yu et al. (2015b), the crystal plasticity based models are
very effective to predict the thermo-mechanical responses of NiTi SMAs under complex loading
EP

conditions (for instance, the non-proportional multiaxial stress-controlled cyclic loading) since the
microstructures of NiTi SMAs are considered in a certain extent here. The micromechanical models
are developed from the Eshelby’s inclusion theory and consider the martensite phases as the
C

inhomogeneities embedded in the matrix of austenite ones. The macroscopic overall stress-strain
AC

response of NiTi SMAs is then obtained by certain homogenization methods such as the
Mori-Tanaka’s scheme (Sun and Hwang 1993a, 1993b; Yu et al., 2015c) and self-consistent one
(Peng et al., 2008). Interactions between the austenite and martensite phases can be reflected by such
micromechanical models reasonably.
It should be noted that all the aforementioned constitutive models belong to local-type ones (that is,
the gradient terms of strain or martensite volume fraction are not included in the constitutive
equations) and aim to predict the thermo-mechanical response of coarse-grained SMAs. By
ACCEPTED MANUSCRIPT

introducing the gradient terms of strain or martensite volume fraction, some nonlocal models were
constructed. For example, by employing a non-convex and non-local 1D chain element model, Sun
and He (2008) established a one-dimensional (1D) nonlocal continuum theory and derived an
analytical scaling law for the grain size-dependent dissipation energy. It was shown that such a
theory could well capture some key physical natures of martensite transformation, e.g., the sequential

PT
nucleation and growth of martensite domain during a loading and the vanishing during an unloading.
The 1D theory was further verified by a two-dimensional (2D) phase field model (Sun and He, 2008),

RI
and the results predicted by both 1D and 2D models agreed well with the available experimental data
of polycrystalline NiTi SMAs quantitatively. It is noted that in Sun and He (2008), the GB was

SC
regarded as a non-transformable sharp interface with zero thickness and prohibiting the martensite
transformation in the GI phase. Considering the thickness and volume fraction of the GB, a new 1D

U
nonlocal continuum theory was established by Li and Sun (2017). The total free energy of the
AN
polycrystalline aggregate was considered as the combination of the convex energy of GB phase,
non-convex energy of GI phase and the gradient energy of the A-M interfaces. It was shown that
M

with the decrease of grain size, the energy of the GB that was ignorable in the coarse-grained
situation gradually became as one dominant factor over other energy terms, which eventually
D

brought fundamental changes in the martensite transformation by making the total free energy of the
TE

system convex, i.e., the coexistence breakdown of austenite and martensite phases and the vanishing
of hysteresis loop. By extending the strain gradient plasticity theory (Gurtin and Anand 2005; Anand
EP

et al., 2005), Qiao et al. (2011) proposed a 1D non-local constitutive model by introducing a gradient
energy and considering two internal length scales, i.e., an energetic length scale and a dissipative one,
simultaneously. Numerical results showed that the external size (i.e., the geometric dimension of the
C

specimen, rather than the internal grain size) dependent deformation of SMAs could be reasonably
AC

described by such a model. The model was further extended to a 3D version and used to predict the
grain size dependent deformation of NiTi SMAs (Qiao et al., 2016).
As mentioned above, for a nonlocal model, the gradient terms of strain or martensite volume
fraction are involved, which will lead to a high computational complexity. Recently, adopting some
simplifications and assumptions, Yu et al., (2017) derived an analytical solution for the 1D nonlocal
model proposed by Qiao et al. (2011) and obtained a scaling law of transformation hardening
modulus with respect to grain size. Adopting the proposed energy-equivalent rule, the existing
ACCEPTED MANUSCRIPT

nonlocal model (Qiao et al., 2011) was converted to an equivalent local one. It was shown that the
equivalent local model proposed by Yu et al. (2017) could well describe the grain size dependent
stress-strain responses of NiTi SMAs and could ensure the high calculation efficiency and easy
application availability.
It should be noted that the existing models focus only on the grain size-dependent inelastic

PT
deformation of SMAs under isothermal loading condition (Sun and He, 2008; Qiao et al., 2011, 2016;
Li and Sun, 2017; Yu et al., 2017). The thermo-mechanically coupled inelastic deformation of SMAs

RI
has not been considered in the constitutive models yet. Moreover, most of the existing models
belonged to the 1D ones (Sun and He, 2008; Qiao et al., 2011; Li and Sun, 2017; Yu et al., 2017),

SC
which limited the wide applications of them. Therefore, a three-dimensional (3D) local constitutive
model which reasonably describes the grain size-dependent thermo-mechanically coupled inelastic

U
deformation of super-elastic NiTi SMAs is extremely necessary.
AN
The aim of this work is to establish a 3D micromechanical constitutive model which accounts for
the grain size-dependent and thermo-mechanically coupled inelastic deformation, simultaneously. In
M

the proposed model, the polycrystalline aggregate is regarded as a composite material, i.e., each GI
phase is assumed to be a spherical inhomogeneity embedded in a matrix of GB. For the GI phase, a
D

1D scaling law of transformation hardening modulus with respect to grain size (Yu et al., 2017) is
TE

extended to a 3D one and the transformation hardening energy is introduced into the Helmholtz free
energy to reflect the constraint of GB on martensite transformation in GI phase. The constitutive
EP

relationship of GI phase is deduced in the framework of irreversible thermodynamics. The driving


force of martensite transformation and the internal heat production caused by transformation latent
heat and mechanical dissipation are obtained by the Clausius’s dissipative inequality and
C

energy-balance equation, respectively. For the GB phase, the stress-strain relationship is assumed to
AC

be linearly elastic without energy dissipation since its yielding stress is very high. A new incremental
non-isothermal Mori-Tanaka’s homogenization method is developed and the Eshelby’s tensor in a
finite spherical domain (Li et al., 2007a; 2007b) is introduced. The interaction between the GI and
GB phases is described by the new proposed homogenization method. Moreover, the continuous
tangent moduli of GI and GB phases are deduced and the computational algorithms of the proposed
micromechanical model are developed from fully implicit backward Euler’s integration method.
Finally, the proposed model is verified by comparing the simulated results with the corresponding
ACCEPTED MANUSCRIPT

experimental ones (Ahadi and Sun, 2014). It is shown that the proposed model can well describe the
grain size-dependent thermo-mechanically coupled inelastic deformation of polycrystalline NiTi
SMAs.

2. Constitutive models for GI and GB phases

PT
Figs. 3a and 3b show the macroscopic specimen and microstructure of the polycrystalline
aggregate, respectively. In the coarse-grained polycrystalline materials, the GB phase can be treated

RI
as a sharp interface without volume fraction. However, when the grain size decreases to a nanoscale,
the volume fraction of GB phase increases significantly and influences the thermo-mechanical

SC
response of the polycrystalline aggregate strongly. To model the effect of GB phase on the overall
deformation of nano-crystalline materials, a two-phase composite model has been adopted widely

U
(Capolungo et al., 2005; 2007; Li and Weng, 2007; Lan et al., 2008). The basic idea of such a model
AN
is that the polycrystalline aggregate is assumed to be a composite material consisting of two distinct
phases (i.e., the GI and GB ones), and the GI phase is assumed to be a spherical inhomogeneity
M

embedded in a matrix of GB phase, as shown in Fig. 3c. The interaction between the inhomogeneity
and matrix phases is described by certain micromechanical homogenization methods, such as
D

Mori-Tanaka’s (Li and Weng, 2007) and self-consistent ones (Capolungo et al., 2005; 2007; Lan et
TE

al., 2008).
By the two-phase composite model, the volume fractions of GI and GB phases can be calculated
EP

as
3
 l 
fGI = g  (1-a)
l +t 
C

 g b
AC

3
 lg 
fGB = 1 − fGI = 1− 
 l + t 
(1-b)
 g b
where, l g and tb are the grain size of GI phase and the thickness of GB phase, respectively, as

shown in Fig. 3c. f GI and fGB are the volume fractions of GI and GB phases in the composite

model, respectively.
2.1 Constitutive model for GI phase
Based on the hypothesis of small deformation, the total strain tensor ε of a material point in GI
ACCEPTED MANUSCRIPT

phase can be decomposed into two parts, i.e., the elastic strain ε e and transformation strain εtr .

That is,

ε = εe + εtr (2)

Referring to Lagoudas and Entchev (2004), the transformation strain εtr and the martensite

PT
volume fraction ξ can be linked by the following flow rule, i.e.,

εɺ tr = g tr Nξɺ (3-a)

RI
 3 s
 ξɺ ≥ 0

SC
 2 s
N= (3-b)
 3 ε t −r
ξɺ < 0
 2 ε t −r

U
where, gtr is the magnitude of transformation strain generated in a full forward transformation, N is
AN
the direction tensor of martensite transformation. s is the deviator of stress tensor σ , ε t − r denotes
M

the transformation strain at the point with the maximum volume fraction of martensite phase during

the forward transformation. Therefore, ε t − r is a constant tensor during the reverse transformation
D

from the induced martensite to austenite phase.


TE

The thermodynamic states of the material can be defined by the observable external variables and
some introduced internal variables. The Helmholtz free energy Ψ of a material point in GI phase
EP

can be given as
  T 
Ψ ( ε, εtr , T , ξ ) = ( ε − εtr ) : EGI : ( ε − εtr ) + γ (T − T0 ) ξ + c (T − T0 ) − T ln   
1
C

2   T0   (4)
1 ξ
+ H (ξ ) + ∫ α (T − T0 ) dξ
AC

2 0

where, EGI is the fourth-order elastic tensor of GI phase, and in this work, the GI phase is assumed

EGIν GI E
to be elastically isotropic for simplicity, i.e., E GI = 1 ⊗ 1 + GI I , I and I is
(1 − 2ν GI )(1 + ν GI ) 1 + ν GI

the fourth- and second-order unit tensors, and EGI and ν GI are the elastic modulus and Poisson’s

ratio of GI phase, respectively; T is the absolute temperature, T0 is the balance temperature, c is


ACCEPTED MANUSCRIPT

the specific heat at a constant volume, γ and α are two material parameters, H is the

transformation hardening modulus. Generally speaking, H depends on the grain size l g and is a

function of σ , N and T . The transformation hardening modulus will be discussed in details in


Section 3.2. The first term on the right side of Eq. (4) is the elastic energy; the second and third terms

are the thermal energy by referring to Anand and Gurtin (2003) and Qiao et al. (2011); γ (T − T0 ) ξ

PT
  T 
is the part of thermal energy coming from transformation latent heat and c T − T0 − T ln    is

RI
  T0  
the one coming from atomic thermal vibration. It should be noted that the part of thermal energy

SC
coming from thermal expansion is neglected in this work, since the temperature variation caused by
the internal heat production is not large enough and the resulting thermal expansion is very small

U
(since the coefficient of thermal expansion of NiTi SMA is about 20 × 10-6/K). The fourth term in Eq.
AN
(4) is the transformation hardening energy which originates from the gradient energy in the
microscopic scale (Yu et al., 2017); and the fifth term is a new introduced one, and will be explained
M

in detail after the constitutive model of the GI phase has been established.
The well-known Clausius’s dissipative inequality (i.e., the second law of thermodynamics) can be
D

written as
ɺ − ηTɺ − q ⋅ ∇T ≥ 0
TE

Γ = σ : εɺ − Ψ (5)
T
where, Γ is the total dissipation, q is the heat flux, η is the entropy.
EP

Substituting Eq. (4) into Eq. (5), and considering the Fourier’s law of heat flux, i.e.,

q = − k ⋅ ∇T (6)
C

it yields
AC

 ∂Ψ   ∂Ψ  ɺ  ∂Ψ tr ∂Ψ ɺ  k : ( ∇T ⊗ ∇T )
Γ = σ −  : εɺ − η +  T −  tr : εɺ + ξ + ≥0
 ∂ε   ∂T   ∂ε ∂ξ  
T  (7)
 Heat flux dissipation
Intrinsic dissipation

where, k is the second-order thermal conductivity tensor. The third and fourth terms on the right
side of Eq. (7) are the intrinsic dissipation caused by the inelastic deformation and the heat flux
dissipation, respectively.

Since Eq. (7) should be always satisfied with arbitrary values of εɺ and Tɺ , it can give
ACCEPTED MANUSCRIPT

∂Ψ
σ= = EGI : ( ε − ε tr ) (8-a)
∂ε

∂Ψ T 
η=− = γξ + c ln   (8-b)
∂T  T0 
Eqs. (8-a) and (8-b) are stress-elastic strain and entropy-temperature relationships, respectively.

PT
It is very easy to prove that the dissipation of heat flux is non-negative since thermal conductivity
tensor k is positive, i.e.,

RI
k : ( ∇T ⊗ ∇T )
Γ flux = ≥0 (9)
T

SC
Then, the intrinsic dissipation inequality can be written as

 ∂Ψ ∂Ψ ɺ 
Γint = −  tr : εɺ tr + ξ ≥0
∂ξ 
(10)

U
 ∂ε
AN
Substituting Eq. (4) into Eq. (10), and considering the relationship between εɺ tr and ξɺ (Eq. 3-a),

it yields
M

Γint = ( g tr σ : N − β ( T − T0 ) − H ξ ) ξɺ ≥ 0 (11)

where, β = γ + α .
D
TE

By Eq. (11), the thermodynamic driving force of martensite transformation can be defined as

π tr = g tr σ : N − β ( T − T0 ) − H ξ (12)
EP

It is seen that the thermodynamic driving force contains three terms, i.e., the one from stress

g tr σ : N , the one from temperature − β (T − T0 ) and the one overcoming the transformation
C

hardening − H ξ . For stress-induced martensite transformation, the term − β (T − T0 ) also reflects


AC

the dependence of transformation stress on temperature, as observed by Shaw and Kyriakides (1995).
By Eq. (11), it is seen that if no martensite transformation occurs (e.g., during the elastic

loading-unloading), i.e., ξɺ = 0 , the dissipative inequality can be satisfied automatically and there is

no constraint for the thermodynamic force π tr . If the forward transformation (from the austenite to

martensite phase) occurs, i.e., ξɺ > 0 , the dissipative inequality can be satisfied when π tr is

positive. Similarly, If the reverse transformation (from the induced martensite to austenite phase)
ACCEPTED MANUSCRIPT

occurs, i.e., ξɺ < 0 , the dissipative inequality can be satisfied when π tr is negative. Thus, the

thermodynamic constraints on the martensite transformation can be written as

Forward transformation: ξɺ > 0 , π tr = Y (13-a)

Reverse transformation: ξɺ < 0 , π tr = −Y (13-b)

PT
Elastic loading-unloading: ξɺ = 0 no constraint on π tr (13-c)

where, Y is a positive constant, which controls the width of the stress-strain hysteresis loop.

RI
Considering the thermodynamic constraints (Eq. (13)), the forward and reverse transformation
surfaces can be defined as

SC
Φ for = π tr − Y = g tr σ : N − β ( T − T0 ) − H ξ − Y ≤ 0 (14-a)

Φ rev = −π tr − Y = − g tr σ : N + β ( T − T0 ) + H ξ − Y ≤ 0

U
(14-b)
AN
Then, the Kuhn–Tucker conditions of forward and reverse transformations can be written as

ξɺ ≥ 0 Φ for ≤ 0 ξɺΦ for = 0 (15)


M

ξɺ ≤ 0 Φ rev ≤ 0 ξɺΦ rev = 0 (16)


D

The forward and reverse transformation surfaces and the Kuhn–Tucker conditions are used to
calculate the evolution of martensite volume fraction during the forward/reverse transformation.
TE

The well-known energy-balance equation (i.e., the first law of thermodynamics) can be written as

uɺ = σ : εɺ − ∇ ⋅ q (17)
EP

where, u is the internal energy. Considering the relationship between the internal energy u and

Helmholtz free energy Ψ , i.e., Ψ = u − Tη , the energy-balance equation can be rewritten as an


C
AC

equivalent form

Tηɺ + ∇⋅ q = σ : εɺ − Ψ
ɺ − ηTɺ (18)

Substituting Eqs. (4), (6), (8-b) and (12) into Eq. (18), it yields

cTɺ − ∇ ⋅ ( k ⋅ ∇ T ) = γ T ξɺ + π trξɺ (19)

Eq. (19) is the thermodynamic equilibrium equation. The first and second terms on the left side of
Eq. (19) represent the variation of temperature and heat transfer, respectively; while, the first and
second terms on the right side of Eq. (19) represent the internal heat production caused by
ACCEPTED MANUSCRIPT

transformation latent heat and mechanical dissipation, respectively.


From Eqs. (12) and (19), it is seen that two fundamental properties, i.e., the temperature
dependence of martensite transformation stress and the magnitude of transformation latent heat, are

controlled by two parameters, i.e., β (noted that β = γ + α ) and γ , respectively. It should be

noted that in the existing thermo-mechanically coupled constitutive models (Morin et al. 2011a,

PT
2011b; Yu et al., 2015a; Wang et al., 2017a; 2017b; Gu et al., 2017), the additional term
ξ
∫ α (T − T ) d ξ
0
in Helmholtz free energy was not considered. Thus, only a single parameter γ

RI
0

was used to characterize such two fundamental properties simultaneously. However, recent

SC
experimental results (Ahadi and Sun, 2014) showed that with the reduction of grain size, the
temperature-dependence of transformation stress decreases more rapidly than the magnitude of

transformation latent heat, which cannot be fully characterized only by one parameter γ . For

U
AN
instance, when the grain size decreases to 10 nm, the temperature-dependence of transformation
stress almost vanished, while a large amount of transformation latent heat was still detected (Ahadi
M

and Sun, 2014). Therefore, two different parameters (i.e., β and γ ) are used by introducing the
ξ
additional term ∫ α (T − T ) d ξ into the Helmholtz free energy, in this work, as shown in Eq. (4).
D

0 0
TE

2.2 Constitutive model for GB phase


As mentioned above, the GB phase in the polycrystalline NiTi SMAs can be regarded as a
EP

non-transformable region (Ahadi and Sun, 2013). Therefore, the stress-strain relationship of the GB
phase is assumed to be pure elastic since its plastic yielding stress is very high (Delville et al., 2011;
C

Ahadi and Sun, 2013; 2014). In this case, the plastic deformation of the GB phase is neglected for
AC

simplicity and the constitutive relationship of a material point can be given as

σ = EGB : ε (20)

where, EGB is the fourth-order elastic tensor of the GB phase assumed to be elastically isotropic,

EGBν GB E
i.e., EGB = 1 ⊗ 1 + GB I , EGB and ν GB are the elastic modulus and Poisson’s
(1 − 2ν GB )(1 + ν GB ) 1 + ν GB

ratio of GB phase, respectively


ACCEPTED MANUSCRIPT

Since the stress-strain relationship of GB phase is assumed to be linearly elastic without any
martensite transformation and energy dissipation, the thermodynamic equilibrium equation of a
material point in GB phase can be written as:

cTɺ − ∇ ⋅ ( k ⋅ ∇T ) = 0 (21)

PT
2.3 Continuous tangent moduli of GI and GB phases
From Eq (8-a), the rate form of the stress-strain relationship of GI phase can be written as

RI
σɺ = EGI : ( εɺ − εɺ tr ) (22)

SC
Substituting Eq. (3-a) into Eq. (22), it yields

(
σɺ = EGI : εɺ − g tr Nξɺ ) (23)

U
During the forward transformation, ξɺ > 0 , the Kuhn–Tucker condition requires that Φ for = 0 . In
AN
this case, Φ
ɺ for = 0 , i.e., (noted that H is a function of σ , N and T ).

∂Φ for ∂Φ for ɺ ∂Φ for ɺ ∂Φ for ɺ


ξ
M

Φ =
ɺ for
: σɺ + :N+ T+
∂σ ∂N ∂T ∂ξ
(24)
ɺ − β Tɺ − H ξɺ − ξ  ∂H : σɺ + ∂H : Ν
= g σɺ : N + g σ : N
tr tr ɺ + ∂H Tɺ  = 0
 
D

 ∂σ ∂N ∂T 
TE

During the revere transformation, ξɺ < 0 , the Kuhn–Tucker condition requires that Φ rev = 0 . In

this case, Φ
ɺ rev = 0 , i.e.,
EP

ɺ rev = ∂Φ : σɺ + ∂Φ : N ɺ + ∂Φ Tɺ + ∂Φ ξɺ
rev rev rev rev
Φ
∂σ ∂N ∂T ∂ξ
(25)
C

ɺ − β Tɺ − H ξɺ − ξ  ∂H : σɺ + ∂H : Ν
= g σɺ : N + g σ : N
tr tr ɺ + ∂H Tɺ  = 0
 
 ∂σ ∂N ∂T 
AC

From Eqs. (24) and (25), the rates of martensite volume fraction during the forward and reverse
transformations can be obtained as

ξɺ =
1
H
( X σ : σɺ + X N : Nɺ − X T Tɺ ) (26)

where,
∂H
Xσ = g tr N − ξ (27-a)
∂σ
ACCEPTED MANUSCRIPT
∂H
X N = g tr σ − ξ (27-b)
∂N
∂H
XT = β + ξ (27-c)
∂T
Substituting Eq. (26) into Eq. (23), it yields

 g tr N ⊗ Xσ g tr N ⊗ X N ɺ g tr NX T ɺ 
ɺσ = EGI :  εɺ − :σ−
ɺ :N+ T (28)
 H H H 

PT
Recalling the definition of N (i.e., (Eq. 3-b)), its rate form can be written as
  2 

RI
 3  I − N ⊗ N  : sɺ
  3  ξɺ ≥ 0
N=
ɺ (29)
2 s

SC
0 ξɺ < 0
Substituting Eq. (29) into Eq. (28), after a series of derivation, it yields

U
  dev 2  
 g tr  I − N ⊗N 
g N ⊗ XN 
( )
tr
3  Θ ξɺ  : σɺ
AN
3
I + EGI : N ⊗ Xσ + EGI : :
 H 2 H s 
  (30)
M

g tr NX T
= EGI : εɺ + EGI : Tɺ
H
1
where, I dev is the fourth-order unit deviatoric tensor, i.e., I dev = I − (1 ⊗ 1) , Θ is the step
D

3
function, when x > 0 , Θ ( x ) = 1 ; when x ≤ 0 , Θ ( x ) = 0 . It is noted that Eq. (30) gives an explicit
TE

relationship among σɺ , εɺ and Tɺ during the forward and reverse transformations. In the case of
EP

pure elastic deformation (i.e., ξɺ = 0 ), Eq. (23) is reduced to

σɺ = EGI : εɺ
C

(31)
AC

By Eqs. (30) and (31), the relationship among σɺ , εɺ and Tɺ at a material point in the GI phase

can be written as the following unified formula:

σɺ = DGI
ε : ε + DT T
ɺ GI ɺ
(32)

where, D GI
ε and DTGI are called as the stress-strain and stress-temperature tangent moduli of GI

phase, respectively, i.e.,


ACCEPTED MANUSCRIPT

M : EGI ξɺ ≠ 0
DεGI =  (33-a)
EGI ξɺ = 0

 g tr M : EGI : NX T ɺ
 ξ ≠0
DTGI =  H (33-b)
0 ξɺ = 0

−1
  dev 2  

PT
 g tr g tr N ⊗ X N I − N ⊗N 
M = I+

 H
EGI : N ⊗ Xσ +
3
2
EGI :
H
:
3
s
 Θ ξɺ 

( ) (33-c)
 

RI
For a material point in the GB phase, the continuous tangent modulus DGB
ε is equal to the elastic

SC
modulus EGB since the stress-strain relationship is considered as pure elastic, i.e.,

U
σɺ = E GB : εɺ = D GB ɺ
ε :ε (34)
AN
2.4 Overall thermo-mechanical properties of GI and GB phases
M

In sub-sections 2.1, 2.2 and 2.3, the constitutive models for a material point in the GI and GB
phases are established, respectively. However, the local and overall constitutive equations of GI and
D

GB phases are not identical. Therefore, in this section, the overall thermo-mechanical responses of
TE

GI and GB phases are derived. In general, the stress, strain and temperature fields in the whole GI
and GB phases are not uniform. Therefore, all the external and internal variables should be the
EP

functions of space coordinates. Recalling the constitutive models constructed in sub-sections 2.1, 2.2
and 2.3, the main equations of constitutive relationships at a spatial point x can be given as
C

follows:
when x ∈ GI
AC

ε ( x ) = ε e ( x ) + ε tr ( x ) (35-a)

σɺ ( x ) = EGI : εɺ e ( x ) (35-b)

εɺ tr ( x ) = g tr N ( σ ( x ) ) ξɺ ( x ) (35-c)

Φ for ( x ) = g tr σ ( x ) : N ( σ ( x ) ) − β (T ( x ) − T0 ) − H ( σ ( x ) , T ( x ) ) ξ ( x ) − Y ≤ 0 (35-d)

Φ rev ( x ) = − g tr σ ( x ) : N ( σ ( x ) ) + β (T ( x ) − T0 ) + H ( σ ( x ) , T ( x ) ) ξ ( x ) − Y ≤ 0 (35-e)
ACCEPTED MANUSCRIPT

ε ( σ ( x ) , T ( x ) , ξ ( x ) ) : ε ( x ) + DT ( σ ( x ) , T ( x ) , ξ ( x ) ) T ( x )
σɺ ( x ) = DGI ɺ GI ɺ (35-f)

when x ∈ GB

σɺ ( x ) = EGB : εɺ ( x ) (36)

Now, let’s rewrite Eq. (35-f) in an abstract functional form:

σɺ ( x ) = f σ ( x ) , T ( x ) , ξ ( x ) , εɺ ( x ) , Tɺ ( x )  x ∈ GI (37-a)

PT
f σ ( x ) , T ( x ) , ξ ( x ) , εɺ ( x ) , Tɺ ( x ) 
(37-b)
ε ( σ ( x ) , T ( x ) , ξ ( x ) ) : ε ( x ) + DT ( σ ( x ) , T ( x ) , ξ ( x ) ) T ( x )

RI
= DGI ɺ GI ɺ

The average values of the stress, temperature, strain rate, martensite volume fraction (i.e., σ ( x ) ,

SC
T ( x ) , εɺ ( x ) and ξ ( x ) ) in the GI phase are defined as

U
1
σ GI
= ∫ σ ( x ) dV (38-a)
VGI VGI
AN
1
T GI
= ∫ T ( x ) dV (38-b)
VGI VGI
M

1
εɺ GI
=
VGI ∫ εɺ ( x ) dV
VGI
(38-c)
D

1
ξ ξ ( x ) dV
TE

GI
=
VGI ∫
VGI
(38-d)

where, VGI is the volume of the GI phase. In this work, the volume averages in the GI and GB
EP

phases are denoted by the operators i GI


and i GB
, respectively. For instance, A GI
and A GB
C

represent the average values of A ( x ) in the GI and GB phases, respectively.


AC

From Eq. (37-a), it is seen that σɺ ( x ) in the GI phase is a function of σ ( x ) , T ( x ) , ξ ( x ) , εɺ ( x ) and

Tɺ ( x ) . Adopting the Taylor expansion of σɺ ( x ) around the point ( σ GI


, T GI
, ξ GI
, εɺ GI
, Tɺ )
GI

and neglecting the second-order small quantities, it yields


ACCEPTED MANUSCRIPT

σɺ ( x ) = f σ ( x ) , T ( x ) , ξ ( x ) , εɺ ( x ) , Tɺ ( x ) 

= f (σ GI
, T GI
, ξ GI
, εɺ GI
, Tɺ
GI
) + ∂σ∂(f x ) : σ ( x ) − σ GI


∂f ∂f x ∈ GI (39)
+ : T ( x ) − T  + : ξ ( x ) − ξ 
∂T ( x )  GI
∂ξ ( x )  GI

∂f ∂f
+ : εɺ ( x ) − εɺ  + : Tɺ ( x ) − Tɺ 
∂εɺ ( x )  GI
∂Tɺ ( x )  GI 

PT
By Eqs. (39), the average value of σɺ ( x ) in the GI phase can be obtained as (noted that

RI
1
VGI ∫
VGI  A ( x ) − A GI  dV = A GI
− A GI
= 0)

SC
σɺ GI
=
1
VGI ∫VGI
σɺ ( x ) dV =
1
VGI ∫VGI
f (σ GI
, T GI
, ξ GI
, εɺ GI
, Tɺ
GI
) dV
(σ )
U
= f GI
, T GI
, ξ GI
, εɺ GI
, Tɺ (40)
GI

= DεGI ( σ ): + DTGI ( σ ) Tɺ
AN
GI
, T GI
, ξ GI
εɺ GI GI
, T GI
, ξ GI GI

From Eq. (40), it is seen that under the first-order approximation of Taylor expansion, the
M

functional relationship among σɺ GI


, σ GI
, T GI
, ξ GI
, εɺ GI
and Tɺ keeps the same as
GI

that among σɺ ( x ) , σ ( x ) , T ( x ) , ξ ( x ) , εɺ ( x ) and Tɺ ( x ) .


D
TE

Adopting the deduction procedure identical to Eqs. (39) and (40), the overall constitutive
relationships of GI and GB phases can be obtained as follows:
EP

Overall constitutive relationship of GI phase:

ε GI
= εe + εtr (41-a)
GI GI
C

σɺ GI
= EGI : εɺ e (41-b)
GI
AC

εɺ tr = g tr N ( σ GI ) ξɺ (41-c)
GI GI

Φ for = g tr σ GI
:N( σ GI )−β ( T GI
− T0 ) − H ( σ GI
, T GI )ξ GI
−Y ≤ 0 (41-d)
GI

Φ rev = − g tr σ GI
:N( σ GI )+ β ( T GI
− T0 ) + H ( σ GI
, T GI )ξ GI
−Y ≤ 0 (41-e)
GI

σɺ GI ε ( σ
= DGI GI
, T GI
, ξ GI ): εɺ GI
+ DTGI ( σ GI
, T GI
, ξ GI ) Tɺ GI
(41-f)

Overall constitutive relationship of GB phase:


ACCEPTED MANUSCRIPT

σɺ GB
= EGB : εɺ GB
(42)

Therefore, it is proved that under the first-order approximation of Taylor expansion, the local and
overall constitutive relationships of each phase are identical.

3. Homogenization method and grain size dependent transformation

PT
hardening modulus

RI
3.1 Average temperature of the specimen
Fig. 4a shows the heat transfer in a polycrystalline NiTi SMA specimen subjected to a uniaxial

SC
mechanical loading, as tested by Ahadi and Sun (2014). Referring to Yin et al. (2014), the heat flow
through the two grips can be modeled as heat conduction through two boundary cross-sections of the

U
specimen gauge-length for simplification, as shown in Fig. 4a. Thus, the total specimen surface S is
AN
1 2
decomposed into two parts, i.e., the lateral surface (Slat) and the cross-section ones ( Scro and Scro ).

Then, the initial and boundary conditions of thermodynamic equilibrium equation can be written as
M

(as shown in Fig. 4b)

T (x, t = 0) = Tr in Ω initial condition (43-a)


D

q ⋅ n lat = h (T − Tr ) in Slat boundary condition (43-b)


TE

q ⋅ n cro = λ (T − Tr ) in Scro boundary condition (43-c)


EP

where, Tr is the ambient temperature, Ω is the domain occupied by the specimen with the volume

of VΩ ( VΩ = VGI + VGB ), nlat and ncro are the surface normal vectors of the lateral surface (Slat) and
C

1 2
AC

cross-section ones ( Scro and Scro ), respectively, h is the heat exchange coefficient of ambient media,

λ is a constant which reflects the heat conduction through the specimen cross-section surface,

Scro = Scro
1
+ Scro
2
.

Recalling Eqs. (19) and (21), the thermodynamic equilibrium equation of a material point in the
specimen can be written as

cTɺ ( x ) − ∇ ⋅ ( k ⋅∇T ( x ) ) = 
(
 g σ ( x ) , T ( x ) , ξ ( x ) , ξɺ ( x ) ) x ∈ GI
(44)
0 x ∈ GB
ACCEPTED MANUSCRIPT

where,

( )
g σ ( x ) , T ( x ) , ξ ( x ) , ξɺ ( x ) = γ T ( x ) ξɺ ( x ) + π tr σ ( x ) , T ( x ) , ξ ( x )  ξɺ ( x ) (45)

Adopting the Taylor expansion of function g around the point (σ GI


, T GI
, ξ GI
, ξɺ
GI
) and

neglecting the second-order small quantities, it yields

( )
g σ ( x ) , T ( x ) , ξ ( x ) , ξɺ ( x ) = g σ ( , ξ , ξɺ ) + ∂σ∂(gx ) : σ ( x ) − σ 

PT
GI
, T GI GI GI
GI

∂g ∂g ∂g  ɺ (46)
+ : T ( x ) − T  + ξ ( x ) − ξ  + ξ ( x ) − ξɺ 
∂T ( x )  ∂ξ ( x )  ∂ξɺ ( x )  

RI
GI GI GI

Integrating Eq. (44) in the whole domain Ω , it yields

SC
∫VΩ VΩ VGI
(
cTɺ ( x )dV − ∫ ∇ ⋅ ( k ⋅ ∇T ( x ) )dV = ∫ g σ ( x ) , T ( x ) , ξ ( x ) , ξɺ ( x ) dV ) (47)

U
By using the Gauss’s theory, the volume integrals can be converted to the surface integrals, then, the
AN
second term on the left side of Eq. (47) can be rewritten as

∫ ∇ ⋅ ( k ⋅∇T ( x ) )dV = ∫ n ⋅ ( k ⋅∇T ( x ) ) dS = −∫ n ⋅ qdS


VΩ Slat + Scro Slat + Scro
(48)
M

= Slat h (Tr − Tcol ) + Scro λ (Tr − Tcro )


D

where, Tlat and Tcro are the average temperatures in Slat and Scro, respectively.
TE

1
Tlat = ∫ T ( x ) dS (49-a)
Slat Slat
EP

1
Tcro = ∫ T ( x ) dS (49-b)
S cro Scro

Substituting Eqs. (43-b), (43-c), (46) and (48) into Eq. (47), it yields
C

VΩ cTɺ − Slat h (Tr − Tlat ) − Scro λ (Tr − Tcro )


AC

(50)
= VGI γ T GI
ξɺ + VGI π tr  σ GI
, T GI
, ξ  ξɺ
GI 
GI GI

where, T is the average temperature of the specimen, i.e.,

1
T= ∫ T ( x ) dV (51)
VGI + VGB VGI +VGB

In this work, the volume average in the whole specimen is denoted by the operator i . For instance,
ACCEPTED MANUSCRIPT

A represents the average value of A ( x ) in the whole specimen.

It is noted that the volume fractions of GI and GB phases are defined as

VGI
fGI = (52-a)
VΩ

VGB
f GB = (52-b)

PT
VΩ

Then, Eq. (50) can be rewritten in an equivalent form:

RI
S h (T − T ) S λ (Tr − Tcro )
cTɺ − lat r lat − cro
VΩ VΩ (53)

SC
= fGI γ T GI ξɺ + fGI π tr  σ GI , T GI
, ξ GI
 ξɺ
GI GI

Furthermore, in this work, the heterogeneity of temperature field in the specimen is neglected for

U
simplicity. Adopting this assumption of uniform temperature field, it yields
AN
Tlat = Tcro = T GI
= T GB
=T (54)

In this case, the evolution equation of average temperature (i.e., (Eq. 53)) can be simplified as
M

Shɶ (Tr − T )
cTɺ − = fGI γ T ξɺ + fGI π tr  σ GI
,T , ξ GI
 ξɺ (55)
VΩ GI GI
D

where, hɶ is the effective heat exchange coefficient, i.e.,


TE

S h + S cro λ
hɶ = lat (56)
S
EP

The assumption of uniform temperature field has been widely adopted and its effectiveness has been
verified (Yin et al., 2014; Ahadi and Sun, 2014b; Yu et al., 2014b; Wang et al., 2017).
C

3.2 Grain size dependent transformation hardening modulus


AC

As mentioned in the section of Introduction, the transformation hardening modulus of


polycrystalline NiTi SMAs strongly depends on the grain size, as shown in Fig. 2. Recently, based on
the existing 1D nonlocal model (Qiao et al., 2011) and a new proposed energy-equivalent rule, a
scaling law of transformation hardening modulus with respect to the grain size (i.e., the explicit
functional relationship between the transformation hardening modulus and grain size) was derived by
Yu et al. (2017). It should be noted that in Yu et al. (2017), the temperature field is assumed to be
uniform. Thus, in this work, the grain size dependent transformation hardening modulus is discussed
ACCEPTED MANUSCRIPT

here since the assumption of uniform temperature has been introduced (in Section 3.1). The 1D
scaling law is given as (Yu et al., 2017)
12 S0le2 8S0le2
 l2 0≤F ≤
lg2
 g
 2
(
H ( lg ) =  S0le 4lg lɶ − 6lg lɶ + 3lg lɶ
3 2 2 3
) F>
8S0le2 (57-a)
 2
lg2
 6  − lɶ + lg l + 2 S0le h 
3 ɶ2 2

PT
  3 2 F 

 

σ g tr − β (T − T0 ) − Y ξɺ > 0

RI
F =  tr (57-b)
σ g − β (T − T0 ) + Y ξɺ < 0

SC
where, S0 is a positive constant with a dimension of stress and le is an internal energetic length
1 2 l
l g − 8 S 0le2 F and lɶ = g − h . σ is the uniaxial stress (noted that only the uniaxial

U
scale. h =
2 2
AN
stress is considered in Yu et al. (2017) since such a scaling law is derived from a 1D nonlocal model).

From Eq. (17-a), it is seen that when the applied stress is not very high (e.g., F ≤ 8 S 0le2 l g2 ), H is
M

inversely proportional to l g2 . However, when the applied stress reaches to a critical value (e.g.,

F > 8 S 0le2 l g2 ), the functional relationship between H and lg becomes much complex. Here, only
D
TE

the final expressions of the scaling law are given, and some important details can be found in the
Appendix.
EP

The scaling law proposed by Yu et al. (2017) can capture the main physical nature of the grain size
dependent transformation hardening modulus, i.e., the GB are considered as obstacles to the
martensite transformation of GI phase, and the macroscopic transformation hardening originates
C

from the gradient energy in the microscopic scale. Thus, in this work, the scaling law proposed by Yu
AC

et al. (2017) is adopted. Surely, the scaling law should be extended to a 3D version since the aim of
this work is to establish a 3D constitutive model. However, it is a very challenging work to obtain an
analytical solution of a 3D nonlocal model. For this reason, here, we don’t solve a real 3D nonlocal

problem but adopt a simple method, i.e., replacing the term σ g tr in Eq. (57-b) (which represents

the stress driving force in the 1D constitutive model) by g tr σ : N (which represents the stress

driving force in the 3D constitutive model, as shown in Eq. (12)), i.e.,


ACCEPTED MANUSCRIPT

 g tr σ : N − β (T − T0 ) − Y ξɺ > 0
F =  tr (58)
 g σ : N − β (T − T0 ) + Y ξɺ < 0

Other variables keep unchanged since they are scalar quantities, and there is no difference in 1D
and 3D cases.

PT
3.3 New incremental non-isothermal Mori-Tanaka’s homogenization method
As mentioned in Section 2, in this work, the polycrystalline aggregate is modeled as a two-phase

RI
composite, i.e., the GI phase is assumed to be an inhomogeneity embedded in a matrix of GB phase.

SC
To describe the interaction between the inhomogeneity and matrix phases reasonably, an appropriate
homogenization method is needed. In 1973, Mori and Tanaka (1973) proposed a rational approach to

U
predict the overall elastic behavior of composite materials, which is the well-known Mori-Tanaka’s
homogenization method. After several decades, original Mori-Tanaka’s method had been extended
AN
and successfully used to describe the thermal expansion (Hu et al., 2002; Lu et al., 2013),
viscoelasticity (Berbenni et al., 2015; Lavergne et al., 2016), elastic-plasticity (Doghri and Ouaar,
M

2003; Pierard et al., 2007; Guo et al., 2011; Peng et al., 2016) and elastic-viscoplasticity (Pierard and
D

Doghri, 2006; Mercier and Molinari, 2009; Doghri et al., 2010; Guo et al., 2013) of composite
materials. It is noted that in the original and extended Mori-Tanaka’s homogenization methods, the
TE

Eshelby’s tensor is obtained from the assumption that an inclusion is embedded in an infinite matrix.
This is a good approximation only if the size effect of the inclusion is negligible, i.e., the size of the
EP

inclusion is small enough compared to the size of the representative volume element (RVE).
However, such an assumption is not applicable any more for modelling the deformation of
C

nano-crystalline materials over a wide range of grain size since the volume fraction of the inclusion
AC

(GI phase) is close to 100% when the grain size is large. To overcome such shortcomings, Li et al.
(2007a; 2007b) derived an analytical solution of Eshelby’s tensor for a spherical inclusion in a finite
spherical domain and further extended the Mori-Tanaka’s method. However, only the isothermal
elastic deformation of the composite was considered in the work of Li et al. (2007a; 2007b). Thus,
such an extended Mori-Tanaka’s method cannot be directly used to predict the grain size-dependent
thermo-mechanically coupled inelastic deformation of NiTi SMAs since the martensite
transformation is an inelastic and non-isothermal process. Therefore, in this subsection, a new
ACCEPTED MANUSCRIPT

incremental non-isothermal Mori-Tanaka’s homogenization method is developed, based on the


analytical solution of Eshelby’s tensor in a finite spherical domain (Li et al., 2007a; 2007b).
Referring to Li et al. (2007b), the average strain rate in each phase can be decomposed into two
parts, i.e., the background strain and the average disturbance one, i.e.,

εɺ GI
= εɺ b + εɺ d (59-a)
GI

PT
εɺ GB
= εɺ b + εɺ d (59-b)
GB

where, εɺ b is the background strain rate, εɺ d and εɺ d are the average disturbance strain rates

RI
GI GB

in the inhomogeneity (GI) and matrix (GB) phases, respectively.

SC
According to the Eshelby’s equivalent inclusion principle for inhomogeneity problem, the total
strain field of this composite under external loads can be equivalently represented by an artificially

U
introduced eigenstrain strain field plus the homogeneous strain field when the inhomogeneity is
AN
assumed to be absent. Then, the inhomogeneity problem is converted into an equivalently
homogeneous inclusion problem. That is, the overall constitutive relationship of GI phase can be
M

rewritten in the following equivalent form:

σɺ GI
= DGI ɺ
ε : ε GI
+ DTGI Tɺ = EGB : εɺ ( GI
− εɺ * ) (60)
D

where, εɺ * is the introduced eigenstrain rate, which is an unknown variable.


TE

By Eq. (60), the relationship among εɺ GI


, εɺ * and Tɺ can be obtained as

= Aε : εɺ* + AT Tɺ
EP

εɺ GI
(61-a)

A ε = ( EGB − DεGI ) : E GB
−1
(61-b)
C

A T = ( E GB − DGI
ε ) : DT
GI −1
(61-c)
AC

The average disturbance strain rates and the introduced eigenstrain rate can be linked by two
Eshelby’s tensors, i.e.,

εɺ d =Π
ɶ GI : εɺ * (62-a)
GI

εɺ d =Π
ɶ GB : εɺ * (62-b)
GB

ɶ GI and Π
where, Π ɶ GB are named as the interior and exterior Eshelby’s tensors (Li et al., 2007a;

2007b). It should be noted that when the RVE is infinite, the exterior Eshelby’s tensor is equal to
ACCEPTED MANUSCRIPT
ɶ GB = 0 . Thus, in the traditional Mori-Tanaka’s homogenization method, the term
zero, i.e., Π εɺ d
GB

does not occur.

ɶ GI and Π
In the case of spherical inclusion, there are analytical solutions for Π ɶ GB (Li et al.,

2007a; 2007b), i.e.,


ɶ GI = s I I vol + s I I dev
Π (63-a)

PT
1 2

ɶ GB = s E I vol + s E I dev
Π (63-b)
1 2

RI
(1 +ν GB )(1 − fGI ) 2 ( 4 − 5ν GB )(1 − f GI ) 21 fGI (1 − f GI2/3 )
2

sI
= s = I
− (63-c)
3 (1 −ν GB ) 15 (1 −ν GB ) 10 (1 −ν GB )( 7 − 10ν GB )
1 2

SC
(1 +ν GB ) fGI 2 ( 4 − 5ν GB ) f GI 21 f GI2 (1 − f GI2/3 )
2

sI
=− s =−
I
+ (63-d)
3 (1 −ν GB ) 15 (1 −ν GB ) 10 (1 −ν GB )( 7 − 10ν GB )(1 − f GI )
1 2

U
1
where, I vol is the fourth-order unit spherical tensor, i.e., I vol = (1 ⊗ 1 ) .
AN
3
Substituting Eqs. (61-a) and (62-a) into Eq. (59-a), it yields

(
εɺ * = A ε − Π
ɶ GI )
−1
(
: εɺ b − AT Tɺ ) (64)
M

The macroscopic strain rate can be written as


D

εɺ = f GI εɺ GI
+ (1 − fGI ) εɺ GB
(65)
TE

Substituting Eqs. (59-a), (59-b), (62-a), (62-b) and (64) into Eq. (65), the background strain rate
can be expressed in terms of the macroscopic strain rate and temperature rate:
EP

εɺ b = B ε : εɺ + B T Tɺ (66-a)

( ) ( )
−1
ɶ GI :  A − (1 − f ) Π
Bε = A ε − Π ɶ GB 
ɶ GI − Π (66-b)
 ε 
C

GI

( )
−1
BT =  A ε − (1 − f GI ) Π ɶ GI + (1 − f ) Π
ɶ GB  :  fΠ
ɶ GI − Π ɶ GB  : A
AC

(66-c)
  GI  T
Substituting Eqs. (62-a), (64) and (66-a) into Eq. (59-a), the average strain rate in the GI phase can
be rewritten as
GI ɺ
εɺ GI
= CGI
ε : ε + CT T
ɺ (67-a)

( )
−1
 :B
ε = I + Π
CGI ɶ GI : A − Π
ɶ GI (67-b)
 ε
 ε

( ) ( )
−1 −1
CTGI = I + Π
ɶ GI : A − Π
ɶ GI :B −Π
ɶ GI : A − Π
ɶ GI : AT (67-c)
 ε  T ε
ACCEPTED MANUSCRIPT

And substituting Eqs. (62-b), (64) and (66-a) into Eq. (59-b), the average strain rate in the GB phase
can be rewritten as
GB ɺ
εɺ GB
= CGB
ε : ε + CT T
ɺ (68-a)

( )
−1
CGB = I + Π
ɶ GB : A − Π
ɶ GI :B (68-b)
ε  ε  ε

( ) ( )
−1 −1
CTGB = I + Π
ɶ GB : A − Π
ɶ GI :B −Π
ɶ GB : A − Π
ɶ GI : AT (68-c)

PT
 ε
 T ε

Like the macroscopic strain rate, the macroscopic stress rate can be written as

( )

RI
σɺ = f GI σɺ GI
+ (1 − f GI ) σɺ GB
= f DGI
ε : ε
ɺ GI
+ DTGI Tɺ + (1 − f GI ) E GB : εɺ GB (69)

Substituting Eqs. (67-a) and (68-a) into Eq. (69), it yields

SC
σɺ = Dε : εɺ + DT Tɺ (70)

U
where, Dε and DT are the macroscopic stress-strain and stress-temperature tangent moduli,
AN
respectively, i.e.,

Dε = f GI DεGI : CεGI + (1 − f GI ) EGB : CεGB (71-a)


M

DT = fGI ( DGI
ε : CT + DT ) + (1 − f GI ) EGB : CT
GI GI GB
(71-b)
D

It should be noted that the elastic and inelastic deformations and the internal stress caused by
temperature variation are comprehensively considered in the proposed homogenization method. So,
TE

such a homogenization method can be further used to predict the grain size dependent deformation of
other metallic materials and the thermo-mechanical properties of composite materials.
EP

So far, the micromechanical constitutive model for the grain size dependent thermo-mechanically
coupled inelastic deformation of super-elastic NiTi SMAs has been established. The main equations
C

of the proposed model are summarized in Table 1.


AC

4. Numerical implementation
From Section 3, it is seen that the proposed homogenization method and the constitutive
relationship of the GI phase are highly nonlinear. To solve such a nonlinear problem accurately and
efficiently, an appropriate computational algorithm is needed. In 2003, Doghri and Ouaar (2003)
proposed a successive substitution algorithm to solve the nonlinear elastic-plastic problem involved
in the Mori-Tanaka’s homogenization method. Meanwhile, the classic return-mapping algorithm has
ACCEPTED MANUSCRIPT

been widely adopted to solve the nonlinear elastic-plastic problem in the constitutive equations of a
material point (Simon and Taylor, 1985; 1986). However, the existing successive substitution and
return-mapping algorithms are limited to the isothermal loading condition. In this Section, such two
algorithms are further extended to solve the problem of nonlinear phase transition faced in the NiTi
SMAs under a non-isothermal loading condition.

PT
4.1 Computational algorithm for the proposed homogenization method

RI
Consider the interval from the step n to n+1 with the time increment ∆ t . By adopting the fully
implicit backward Euler’s integration method, main equations of the proposed Mori-Tanaka’s

SC
homogenization model can be discretized as follows:

∆ε GI
= C GI
ε , n +1 : ∆ ε + C T , n +1 ∆ T
GI
(72-a)

∆ε = C GB
U
ε , n +1 : ∆ ε + C T , n +1∆ T
GB
(72-b)
AN
GB

∆ σ = D ε , n +1 : ∆ ε + D T , n +1∆ T (72-c)

Dε ,n +1 = fGI DεGI,n +1 : CεGI,n +1 + (1 − fGI ) EGB : CεGB,n +1


M

(72-d)

DT , n+1 = fGI ( DGI


ε , n +1 : CT , n +1 + DT , n +1 ) + (1 − f GI ) EGB : CT , n +1
GI GI GB
(72-e)
D

( )
−1
 :B
ε , n +1 =  I + Π ε , n +1 − Π
CGI ɶ GI : A ɶ GI (72-f)
TE

  ε ,n +1

( ) ( )
−1 −1
CTGI,n +1 =  I + Π
ɶ GI : A
ε , n +1 − Π
ɶ GI :B −Π
ɶ GI : A
ε , n +1 − Π
ɶ GI : AT ,n +1 (72-g)
  T ,n +1
EP

( )
−1
 :B
ε , n +1 =  I + Π ε , n +1 − Π
CGB ɶ GB : A ɶ GI (72-h)
  ε ,n +1

( ) ( )
−1 −1
CTGB,n +1 = I + Π
ɶ GB : A
ε , n +1 − Π
ɶ GI :B −Π
ɶ GB : A
ε , n +1 − Π
ɶ GI : AT ,n +1 (72-i)
  T ,n +1
C

( ) ( )
−1

 ε , n +1 − (1 − f GI ) Π − Π
ɶ GI :  A
B ε , n +1 = A ε , n +1 − Π ɶ GI ɶ GB 
AC

 (72-j)

( )
−1
B T , n +1 =  A ε , n +1 − (1 − f GI ) Π  
ɶ GI + (1 − f ) Π
ɶ GB  :  fΠ
ɶ GI − Π
GI
ɶ GB  : A
 T , n +1 (72-k)

A ε , n +1 = ( E GB − DεGI,n +1 ) : E GB
−1
(72-l)

AT ,n +1 = ( EGB − DGI
ε , n +1 ) : DT , n +1
GI −1
(72-m)

It is assumed that the all the external and internal variables are known at the step n , including the

macroscopic stress, strain and temperature, i.e., σ n , εn and Tn , the average stress and strain
ACCEPTED MANUSCRIPT

tensors in the GI and GB phases, i.e., σ n GI


, εn GI
, σn GB
, εn GB
and the average volume

fraction of martensite phase in the GI phase ξn GI


. For a given increment of macroscopic strain

tensor ∆ ε , the macroscopic strain in the step n+1 can be obtained, i.e., εn+1 = εn + ∆ε . The problem

is to update other variables in the step n+1, i.e., σ n+1 , Tn +1 , σ n+1 GI


, ε n +1 GI
, σ n +1 GB
, ε n +1 GB

PT
and ξ n +1 , according to the adopted model. Here, a successive substitution algorithm is proposed
GI

to solve such a highly nonlinear problem, which is developed from the work of Doghri and Ouaar

RI
(2003). That is:

SC
Initialization: ∆ε GI
← ∆ε

 Iteration

U
∆ ε − fGI ∆ε
1) Calculate the average strain in the GB phase: ∆ε = GI

(1 − fGI )
AN
GB

2) Substitute ∆ε GI
into the constitutive relationship of GI phase, calculate ∆σ GI
, ∆T , ∆ξ GI
,
M

DGI GI
ε , n +1 and D T , n +1 (The details will be discussed in subsection 4.2).

∆ε ∆σ
D

3) Substitute GB
into the constitutive relationship of GB phase, calculate GB
(The details
TE

will be discussed in subsection 4.2).

4) Calculate A ε ,n +1 , AT ,n +1 , B ε , n +1 , BT ,n +1 , CGI GI GB GB
ε , n +1 , CT , n +1 , Cε , n +1 and C T , n +1 .
EP

5) Check the compatibility of average strain in the inhomogeneity (i.e., GI phase) by computing the
residual:
C

R = C GI
ε , n +1 : ∆ ε + C T , n +1∆ T − ∆ ε
GI
GI
.
AC

6) If R ≤ TOL, then exit the iteration loop.

7) Else, new iteration (go to step 1) with a new ∆ε GI


: ∆ε GI
← ∆ε GI
+R.

 After convergence, calculate the macroscopic stress-strain and stress-temperature tangent moduli

Dε ,n +1 and DT , n +1 , and the increment of macroscopic stress tensor: ∆ σ = D ε , n +1 : ∆ ε + D T , n +1∆ T .

4.2 Computational algorithms for the constitutive models of GI and GB phases


ACCEPTED MANUSCRIPT

Considering the interval from the step n to n+1 with a time increment ∆ t . The main equations of
the constitutive model of GI phase are discretized by the fully implicit backward Euler’s method as

∆ε I
= ∆ε e + ∆ε tr (73-a)
GI GI

∆σ GI
= EGI : ∆ε ( GI
− ∆ε tr
GI ) (73-b)

∆εtr = g tr N n +1 ∆ξ (73-c)

PT
GI GI

 3 s n +1

GI
∆ξ GI
≥0

RI
 2 s n +1 GI
N n +1 =  t −r (73-d)
 3 ε n +1 GI
 ∆ξ <0

SC
ε tn−+r1
GI
2
 GI

Φ nfor+1 = g tr σ n +1 GI
: N n +1 − β (Tn +1 − T0 ) − H n +1 ξ n +1 GI
−Y
GI

U
(73-e)
− β (Tn +1 − T0 ) − H n +1 ξ n +1
3 tr
= g s n +1 GI GI
−Y ≤ 0
2
AN
Φ rev
n +1 = − g tr σ n +1 GI
: N n +1 + β (Tn +1 − T0 ) + H n +1 ξ n +1 GI
−Y
GI
(73-f)
= − g tr s n +1 : N n +1 + β (Tn +1 − T0 ) + H n +1 ξ n +1 −Y ≤ 0
M

GI GI

M n +1 : EGI ∆ξ ≠0
DεGI, n +1 = 
GI
(73-g)
EGI ∆ξ =0
D

GI

 g tr M n +1 : EGI : N n +1 X T ,n +1
TE

 ∆ξ ≠0
DTGI,n +1 =  H GI
(73-h)
0 ∆ξ =0
 GI
EP

−1
  dev 2 
I − N n +1 ⊗ N n +1  
 g tr 3 g tr N n +1 ⊗ X N ,n +1  3 
M n +1 =  I + EGI : N n +1 ⊗ Xσ ,n +1 + EGI : : (73-i)
 H 2 H s n +1 
C

 
AC

Similar, the evolution equation of average temperature can be discretized as:

Shɶ (Tr − Tn +1 ) ∆t
c∆T − = fGI γ Tn+1 ∆ξ GI
VΩ (74)
+ fGI  g tr
σ n +1 GI
: Nn +1 − β (Tn+1 − T0 ) − H n +1 ξ n+1 GI  ∆ξ
 GI

It is assumed that all the external and internal variables of GI phase are known at the step n ,

including the average stress and strain σ n GI


and ε n GI
, the macroscopic temperature Tn , and the
ACCEPTED MANUSCRIPT

average volume fraction of martensite phase ξn GI


. For a given increment of average strain tensor

∆ε GI
, the average strain in the step n+1 can be obtained, i.e., ε n +1 GI
= εn GI
+ ∆ε GI
. The

problem is also to calculate ∆σ GI


, ∆T and ∆ξ GI
. Here, the return-mapping algorithm with an

elastic predictor and inelastic corrector is adopted.

PT
The elastic predictor is taken to be an elastic tentative stress:

σ*n +1
GI
= EGI : ε n +1 ( GI
− ε trn
GI )= σ n GI + EGI : ∆ε GI (75)

RI
Substituting Eq. (73-b) into Eq. (75) gives

σ n +1 = σ*n +1 − EGI : ∆ε tr (76)

SC
GI GI GI

Taking the deviatoric part of Eq. (76), it yields (noted that EGI : ∆ε trn +1 = 2 µGI ∆ε trn +1 since the
I I

U
elastic isotropy and martensite transformation incompressibility, and µGI is the shear modulus of GI
AN
EGI
phase, i.e., µGI = ):
2 (1 + ν GI )
M

s n +1 = s*n +1 − 2µGI g tr ∆ξ GI
N n +1 (77)
GI GI
D

The tentative direction of transformation N*n +1 can be written as


TE

 3 s*n +1
 GI
∆ξ ≥0

GI
2 s *
 n +1 GI
N*n +1 =  (78)
EP

 3 ε tn− r

GI
∆ξ <0
2 t −r GI
 ε n GI
C

The forward and reverse transformation conditions are then checked according to
AC

g s n +1 − β (Tn − T0 ) − H n ξ n
3 tr *
Φ nfor+1* = GI
−Y (79-a)
GI 2

Φ rev
n +1
,*
= − g tr s*n +1 : N *n +1 + β (Tn − T0 ) + H n ξ n GI
−Y (79-b)
GI GI

(1) If Φ nfor+1* > 0 and ξn < 1 , it implies that the forward transformation occurs. Noted that
GI GI

2 2 *
s n +1 GI
= s n +1 GI
N n +1 and s*n +1 = s n +1 N*n +1 , then, Eq. (77) can be rewritten as
3 GI 3 GI
ACCEPTED MANUSCRIPT

 2  2 *
 s n +1 GI
+ 2 µGI g tr ∆ξ GI  N n +1 = s n +1 N*n +1 (80)
 3 3 GI

By Eq. (80), it can be obtained that

s n +1 = s*n +1 − 6 µGI g tr ∆ξ GI (81-a)


I I

N n +1 = N*n +1 (81-b)

PT
Substituting Eq. (81-a) into the forward transformation condition (i.e., Eq. (73-e)), it yields

− 3 ( g tr ) µGI ∆ξ
3 tr *

RI
2
Φ nfor+1 = g s n +1 GI
GI 2 GI
(82)
− β (Tn +1 − T0 ) − H n +1 ξ n +1 −Y = 0

SC
GI

Eqs. (74) and (82) are two nonlinear equations of ∆T and ∆ξ GI


, and can be solved by the

U
Newton-Raphson’s method. Then, the increment of transformation strain tensor ∆ε tr can be
GI
AN
obtained as

∆ε tr = g tr N n +1 ∆ξ GI
= g tr N*n +1 ∆ξ GI
(83)
GI
M

(2) If Φ rev
n +1
*
> 0 and ξ n < 0 , it implies that the reverse transformation occurs. Noted that
GI GI
D

N n+1 keeps as a constant tensor during the reverse transformation, and is a known quantity. So,
TE

multiplying both sides of Eq. (77) by N n+1 , it gives

s n +1 : N n +1 = s*n +1 : N n +1 − 3µGI g tr ∆ξ GI
(84)
GI GI
EP

Substituting Eq. (84) into Eq. (73-f), the reverse transformation condition can be rewritten as

: N n +1 + 3 ( g tr ) µ GI ∆ξ
2
Φ rev = − g tr s*n +1
C

n +1 GI GI GI
(85)
+ β (Tn +1 − T0 ) + H n +1 ξ n +1 −Y = 0
AC

GI

Eqs. (74) and (85) are two nonlinear equations of ∆T and ∆ξ GI


, and can be solved by the

Newton-Raphson’s method. Then, ∆ε tr can be obtained by Eq. (73-c).


GI

(3) If the aforementioned two conditions (i.e., Φ nfor+1*


GI
>0 , ξn GI
< 1 or Φ rev
n +1
*
GI
>0 ,

ξn GI
< 0 ) cannot be satisfied, it implies that a pure elastic deformation takes place. So,
ACCEPTED MANUSCRIPT

∆ξ GI
=0, ∆ε tr = 0 . In this case, ∆T can be obtained by solving Eq. (74), i.e.,
GI

Shɶ (Tn − Tr ) ∆t
∆T = (86)
cV + Shɶ∆t

After ∆ε tr is obtained, ∆σ GI
can be calculated by Eq. (73-b). Finally, the tangent moduli
GI

D GI GI
ε , n +1 and D ε , n +1 can be easily obtained by Eqs. (73-g), (73-h) and (73-i).

PT
For the GB phase, ∆σ GB
can be directly calculated by the following equation once the

RI
increment of strain tensor ∆ε GB
is given, i.e.,

SC
∆σ GB
= EGB : ∆ε GB
(87)

U
5. Model verification
AN
In this section, the proposed micromechanical model is verified by simulating the grain size
dependent thermo-mechanically coupled inelastic deformation of polycrystalline NiTi SMAs (as
M

observed by Ahadi and Sun 2014). In the experiment of Ahadi and Sun (2014), dog-bone shaped
tensile specimens with a gauge length (L) of 17 mm, width of 1.3 mm, and thickness of 1 mm were
D

used, and were cut from the polycrystalline NiTi SMA sheets with five different mean grain sizes, i.e.,
TE

90, 68, 42, 27 and 10 nm. The initial phases of all the specimens with different grain sizes were
austenite ones at room temperature (i.e., 26ºC) and the materials exhibited a super-elasticity. All the
EP

tests were performed under a displacement-controlled tension-unloading condition and at room


temperature, as shown in Figs. 4a and 4b. Then, the mechanical boundary condition can be given as
C

u x = 0, u y = 0, u z = 0 in Scro
1
(88-a)
AC

umax
ux = (1 − cos ωt ) , u y = 0, uz = 0 2
in Scro (88-b)
2
where, ux , u y and uz are the displacements in x, y and z directions, respectively; umax is the

maximum end-displacement and ω is the angular frequency. By such a boundary condition, the
applied strain rate in the axial direction can be defined as
umax
ε (t ) = (1 − cos ωt ) (89-a)
2L
ACCEPTED MANUSCRIPT
umaxω
εɺ (t ) = sin ω t (89-b)
2L
From Eq. (89-b), it is seen that the applied strain rate is not a constant during the loading and
unloading. The average strain rate can be defined as:
umaxω
εɺa = (90)

It should be noted that in Ahadi and Sun (2014), the average strain rate εɺa is adopted to

PT
characterize the loading-unloading speed. In this case, the angular frequency can be calculated by Eq.

RI
(90).

SC
5.1 Determination of parameters

In the existing constitutive models, the GB thickness tb was often set as a constant ranged from 1

U
to 3 nm (Jiang and Weng, 2004; Li and Weng, 2007; Li and Sun, 2017). However, recent
AN
experimental observations (Ahadi and Sun, 2013) and molecular dynamic simulations (Ko et al.,
2017) showed that the GB thickness increased rapidly when the grain size decreased to a critical
M

value. Therefore, in this work, tb is set as a function of l g , i.e.,

tb = tb0 1 + ( d1 lg ) 
m1
D

(91)
 
TE

where, tb0 represents the GB thickness in the coarse-grain case. d1 and m1 are the two material

parameters. tb0 , d1 and m1 are set as 1 nm, 12 nm and 5, respectively. Adopting such parameters,
EP

the relationship between tb and l g are drawn in Fig. 5. It is seen that the thickness of GB phase
C

increases significantly when the grain size falls below 20 nm, while for the grain sizes of over 20 nm
it remains almost constant (e.g., 1 nm).
AC

Referring to Yin et al. (2014), the specific heat c and the effective heat exchange coefficient hɶ

are set to be 3.225 MJm-3K-1 and 100 Wm-2K-1, respectively; and the parameter γ which controls

the transformation latent heat is set as 0.39 MPa K-1.


In order to quantitatively characterize the temperature dependence of the start stress of martensite

transformation, Ahadi and Sun (2014) introduced a variable dσ sf dT (where, σ sf is the start
ACCEPTED MANUSCRIPT

stress of forward transformation) and found that dσ sf dT decreased rapidly with the decrease of

grain size. For instance, the values of dσ sf dT for the specimens with the grain sizes of 90, 68, 42,

27 and 10 nm are measured as 6.43, 6.3, 5.21, 4.37 and 0.3 MPa/K, respectively. Therefore, in the

proposed model, β (which controls the dependence of transformation stress on temperature, as

PT
shown in Eq. (12)) is set as a function of l g , i.e.,

β = g tr β 0 1 − ( d 2 l g ) 
m
2
(92)
 

RI
where, β 0 , d 2 and m2 are three parameters, which can be determined by fitting the experimental

SC
data of dσ sf dT and using the least square method.

The Poisson’s ratios of GI and GB phases are set to be the same for simplicity, i.e., ν GI =ν GB =0.3.

U
The other material parameters cannot be directly extracted from the experimental data due to the
AN
interaction between the GI and GB phases. However, when the grain size is large, the volume
fraction of GB phase is very small. In this case, the macroscopic thermo-mechanical property of the
M

composite can be approximately regarded as only that of the GI phase. Therefore, in this work, the
D

parameters EGI , g tr , T0 , Y are all determined from the experimental stress-strain responses of
TE

the polycrystalline aggregate with a large grain size (e.g., 90 nm) under an isothermal condition, as
shown in Fig. 6a.
EP

EGI can be directly determined from the elastic part of stress-strain response. The parameter g tr

can be extracted by fitting the maximum transformation strain.


C

The balance temperature T0 is chosen as


AC

Y − X0
T0 = Tr + (93)
β

where, X 0 is a new introduced parameter and reflects the critical energy of martensite nucleation.

From Eq. (93), it is seen that T0 is not an independent parameter and can be obtained after three

basic parameters X 0 , Y and β are determined.

Substituting Eq. (93) into Eqs. (15-a) and (15-b), the forward and reverse transformation surfaces
ACCEPTED MANUSCRIPT

can be rewritten as

Φ for = g tr σ : N − β (T − Tr ) − H ξ − X 0 ≤ 0 (94-a)

Φ rev = − g tr σ : N + β (T − Tr ) + H ξ + X 0 − 2Y ≤ 0 (94-b)

At room temperature, with an applied uniaxial stress, the transformation conditions at the start
point of forward transformation and the finish point of reverse transformation can be written as

PT
Φ for = g trσ sf,90 − X 0 = 0 (95-a)

RI
Φ rev = − g trσ rf ,90 + X 0 − 2Y ≤ 0 (95-b)

where, σ sf,90 and σ rf ,90 are the experimental measured start stress of forward transformation and

SC
finish stress of reverse transformation for the polycrystalline aggregate with a grain size of 90 nm, as

U
shown in Fig. 6a.
By Eqs. (95-a) and (95-b), it yields
AN
X 0 = g trσ sf,90 (96-a)

g tr (σ sf,90 − σ rf ,90 )
M

Y= (96-b)
2
D

It is seen that the parameters X 0 and Y control the start stress of forward transformation and
TE

the width of stress-strain hysteresis loop, respectively, and can be determined by Eqs. (96-a) and
(96-b), respectively.
EP

Finally, from Eq. (57-a), it is seen that the grain size-dependent transformation hardening under an

isothermal loading condition is controlled by the parameter S0le2 . Thus, it can be extracted by fitting
C

the transformation hardening modulus of the polycrystalline aggregate with a grain size of 27 nm at a
AC

very low strain rate (e.g., 4×10-5/s), as shown in Fig. 9a. After EGI is determined, the elastic

modulus of GB phase ( EGB ) can be extracted by fitting the elastic modulus of the polycrystalline

aggregate with a grain size of 10 nm, as shown in Fig. 10a.


All the material parameters are listed in Table 2.

5.2 Verification and discussion


ACCEPTED MANUSCRIPT

Figs. 6a to 6g show the stress-strain responses of polycrystalline NiTi SMA with a grain size of 90
nm at seven different strain rates (from 4×10-5/s to 1×10-1/s), and Fig. 6h shows the variation of
average temperature during loading and unloading. It is seen that the stress-strain responses strongly
depend on the strain rate, i.e., the transformation hardening modulus increases with the increase of
strain rate but the dissipation energy changes non-monotonically with the varied strain rate. The main

PT
physical reason for such a rate-dependence of stress-strain response is the thermo-mechanically
coupled nature of martensite transformation in NiTi SMAs, i.e., the competition between the

RI
release/absorption of the latent heat and the heat transfer with the environment, and the
temperature-sensitive stress-strain responses. When the applied strain rate is low (4×10-5/s), the heat

SC
transfer via the convection and conduction is much faster than the heat production. In this case, the
deformation of NiTi SMAs can be approximately regarded as an isothermal process and the

U
thermo-mechanically coupling effect is quite weak. At the intermediate strain rates (from 5×10-4/s to
AN
1×10-2/s), the effects of heat production and heat transfer on the stress-strain responses of NiTi
SMAs become significant due to the characteristic time of heat production is comparable to that of
M

heat transfer. Let’s take the deformation at the strain rate of 5×10-3/s as an example (as shown in Figs.
6d and 6h): at the beginning of loading (o-a), the average temperate of the specimen keeps the same
D

as ambient temperature since only elastic deformation occurs without internal heat production;
TE

during the forward transformation (a-c), the average temperature firstly increases with the increase of
accumulated strain (a-b) due to the release of latent heat, however, further loading (b-c) leads to a
EP

decrease of average temperature. In fact, as mentioned above, a cosine wave loading-unloading mode
is adopted in the experiment of Ahadi and Sun (2014), and the strain rate is not a constant during the
C

loading-unloading. As shown in Eqs. (89-a) and (89-b), when the applied strain approximates to the
maximum controlled strain (i.e., ωt → π ), the applied strain rate tends to be zero. In this case, the
AC

internal heat production in a time unit is close to zero, and the average temperature will decrease in
the subsequent loading process, as shown in Fig. 6h. At the beginning of unloading (c-d), elastic
deformation occurs. The average temperature decreases with the increase of accumulated strain,
which is caused by the heat transfer with the environment. During the reverse transformation (d-e),
the average temperature further decreases due to the absorption of latent heat, as shown in Fig. 6h.
After the reverse transformation is completed, further loading leads to an elastic deformation (e-o).
From Figs. 6b-6e, it is seen that the temperature oscillation can further lead to an additional
ACCEPTED MANUSCRIPT

transformation hardening since the transformation stress increases with the increasing temperature.
When the applied strain rate is high (5×10-2/s and 1×10-1/s), the heat production is much faster than
the heat transfer. In this case, the material responses approach approximately an adiabatic condition,
so the temperature oscillation and additional transformation hardening become more and more
obvious, as shown in Figs. 6f, 6g and 6h.

PT
Figs. 7, 8, 9 and 10 show the thermo-mechanically coupled deformation of polycrystalline NiTi
SMAs with the grain sizes of 68, 42, 27 and 10 nm, respectively. The experimental results show that

RI
the stress-strain responses strongly depend on the grain size, i.e., at a given strain rate, the
transformation hardening modulus and dissipation energy monotonically decrease with the decrease

SC
of grain size. The significant grain size effect is primarily caused by the constraint of GB on
martensite transformation (Ahadi and Sun, 2013; 2015). Moreover, it is seen that with the reduction

U
of grain size, the stress-strain responses gradually change from a strongly rate-dependent type to a
AN
rate-insensitive one. Meanwhile, the amplitude of temperature oscillation decreases with the
reduction of grain size. As discussed by Ahadi and Sun (2014), such decreased thermo-mechanical
M

coupled effect is caused by the combination of small temperature variations occurred during the
martensite transformation and its reverse (due to the reduction of internal heat production) and the
D

significant decrease in the temperature-dependence of the transformation stress, with the reduction of
TE

grain size.
Fig. 11 shows the curves of maximum stress vs. strain rate for the polycrystalline NiTi SMAs with
EP

various grain sizes. It is seen that with the reduction of grain size, the dependence of the maximum
stress on the loading rate decreases, and such dependence completely disappears when the grain size
C

decreases to 10 nm. It means that the stability of NiTi SMAs can be significantly improved by the
nano-crystallization of grain size.
AC

Comparing the simulated results with the experimental ones, it can be concluded that the grain size
dependent thermo-mechanically coupled inelastic deformation of polycrystalline NiTi SMAs can be
reasonably described by the proposed model, since the internal heat production, heat transfer,
interaction between the GI and GB phases, and grain size-dependent thermo-mechanical response
have been reasonably considered. However, some slight discrepancies between the experimental and
simulated results are still existed. For instance, the amplitude of temperature oscillation for the
polycrystalline aggregate with a grain size of 10 nm is underestimated by the proposed model. In fact,
ACCEPTED MANUSCRIPT

the parameter γ which controls the transformation latent heat of GI phase is set as a constant for

simplicity in the proposed model, but the dependence of γ on the grain size is neglected. Surely, it

is believed that the simulated results can be further improved by introducing a functional relationship

between γ and l g , but more material parameters will be involved.

PT
6. Conclusions

RI
(1) A micromechanical constitutive model is established to describe the grain size dependent
thermo-mechanically coupled inelastic deformation of polycrystalline NiTi SMAs. The

SC
polycrystalline aggregate is regarded as a two-phase composite material, i.e., each GI phase is
assumed to be a spherical inhomogeneity embedded in a matrix of GB phase.

U
(2) The constitutive relationship of GI phase is deduced in the framework of irreversible
AN
thermodynamics. The 1D scaling law of transformation hardening modulus with respect to the grain
size proposed by Yu et al. (2017) is extended to a 3D version. The driving force of martensite
M

transformation and the internal heat production caused by the transformation latent heat and
mechanical dissipation are obtained by the Clausius’s dissipative inequality and energy-balance
D

equation, respectively. The GB phase is considered as a non-transformable region without energy


TE

dissipation and its constitutive relationship is assumed to be linearly elastic.


(3) A new incremental non-isothermal Mori-Tanaka’s homogenization method is developed. To
EP

describe the interaction between the GI and GB phases accurately, the Eshelby’s tensor of a spherical
inclusion embedded in a finite spherical domain is introduced.
C

(4) It is proved that under the first-order approximation of Taylor expansion, the local and overall
AC

constitutive relationships of each phase are identical.


(5) A new numerical algorithm is proposed to solve the nonlinear problems faced in the
micromechanical model, which is based on the successive substitution algorithm proposed by Doghri
and Ouaar (2003) and the classical return-mapping algorithm, but further extended to the
non-isothermal case with considering internal heat conduction.
(6) Given the comparison between experimental and simulated results of the proposed model
considering full thermo-mechanical coupling, interaction between GI and GB phases, and grain size
ACCEPTED MANUSCRIPT

effect, it is shown that the proposed micromechanical model is able to describe thermo-mechanical
coupled inelastic grain size dependent polycrystalline NiTi SMAs response in a reasonable manner.

Acknowledgements
Financial supports by the National Natural Science Foundation of China (11532010, 11602203),

PT
Young Elite Scientist Sponsorship Program by CAST (No. 2016QNRC001) are appreciated.

RI
Appendix. The derivation of 1D scaling law
It should be noted that it is a very challenging work to obtain an analytical solution of 3D nonlocal

SC
problem. To overcome this shortcoming, Sun and He (2008) introduced a simplified 1D chain
element model to characterize the polycrystalline aggregates, i.e., the grains are regarded as the

U
transformable lathy sub-chains with a length of l g and separated by a non-transformable GB. In the
AN
1D chain element model, the stress field is assumed to be uniform in the whole chain and equal to the
applied stress σ , as shown in Fig. A1(a).
M

Let’s focus on an individual grain ( x ∈  0, l g  , as shown in Fig. A1(b)). Referring to the 1D


D

nonlocal model proposed by Qiao et al. (2011), for an arbitrary material point in this GI phase, the
TE

evolution of the volume fraction of martensite phase can be given as (more details can be found in
the original literature, i.e., Yu et al. (2017))
 
EP

Y ζɺ ∂  Yld2ζɺ, x 
σ g − β (T − T0 ) + S l ζ , xx =
tr 2
−  (A1)
∂x 2 
( ) ( ) ( ) ( )
0 e 2 2 2
ζɺ + l ζɺ, x
2
 ζɺ + ld2 ζɺ, x 
d

C

where, σ is the uniaxial stress, S0 is a positive constant with a dimension of stress, le and ld
AC

are the internal energetic and dissipative length scales, respectively, ζ is the volume fraction of

martensite phase in this material point. It should be noted that with the introduction of ld , numerical

results by the nonlocal model show that the dissipation energy changes non-monotonically, i.e., the
dissipation energy first increases and then decreases with the decreasing grain size. However,
existing experimental observations demonstrate that the dissipation energy monotonically decreases
with the decreasing grain size in the polycrystalline NiTi SMAs, as shown in Fig. 2 in the main
ACCEPTED MANUSCRIPT

content. So, in Yu et al. (2017) the role of dissipative length scale is deleted and the ld is set as zero

for simplicity, i.e., ld = 0 . In this case, Eq. (A1) can be simplified as:

σ g tr − β (T − T0 ) + S0le2ζ , xx = Y (A2)

As mentioned above, in the polycrystalline aggregates, the grains are separated by the

ζ ( x = 0 ) = ζ ( x = lg ) = 0 is adopted to

PT
non-transformable GB phase. Thus, the boundary condition

reflect that two ends of each grain are obstacles to the martensite transformation. Then, considering

RI
the constraint of the volume fraction of martensite phase, i.e, 0 ≤ ζ ( x ) ≤ 1 and adopting the

SC
assumption of uniform temperature field, the analytical solution of Eq. (A2) was derived in Yu et al.
(2017) as

U
F 2 Flg
ζ ( x) = − x + x x ∈[0, lg ] when 0≤F ≤
8 S 0 l e2
(A3-a)
2S0le2 2S0le2 l g2
AN
 F Flg lg lg
− x2 + x x ∈ [0, − h] ∪ [ + h, lg ]
 2 S0le
2 2
2 S 0le 2 2 8 S 0 le2
ζ ( x) =  when F > (A3-b)
M

 lg lg l g2
1 x ∈ ( − h, + h )
2 2
D

By Eqs. (A3-a) and (A3-b), the average volume fraction in this individual grain ( x ∈  0, l g  ) can
TE

be calculated as:

 F 8S0le2
 l2 when F ≤
EP

2 g
1 lg 12 S l lg2
ξ = ∫ ζ ( x )dx = 
0 e
(A4)
lg 0 ɶ3 ɶ2 8S l 2
− Fl + Fl + 2h when F > 02 e
 3S0le2lg 2S0le2 lg lg
C


AC

In the 1D nonlocal model proposed by Qiao et al. (2011), a gradient energy is introduced into the
Helmholtz free energy to reflect the grain size dependent transformation hardening of SMAs, that is,

ϕ grad = S0le2 (ζ , x ) . To obtain the transformation hardening modulus, Yu et al. (2017) proposed an
1 2

2
energy-equivalent rule, i.e.,

S0le (ζ , x ) dx = H (ξ )
1 lg 1 2 2 1

2
(A5)
lg 2
0 2

Eq. (A5) implies that the grain size dependent transformation hardening in the macroscopic scale
ACCEPTED MANUSCRIPT

originates from the gradient energy (or internal length scale) in the microscopic scale. Substituting
Eqs. (A3) and (A4) into Eq. (A5), the scaling law (i.e., the functional relationship between H and

l g ) can be obtained, as shown in Eq. (57-a).

References

PT
Ahadi, A., Sun, Q., 2013. Stress hysteresis and temperature dependence of phase transition stress in
nanostructured NiTi-effects of grain size. Appl. Phys. Lett. 103, 021902.

RI
Ahadi, A., Sun, Q., 2014. Effects of grain size on the rate-dependent thermomechanical responses of

SC
nanostructured superelastic NiTi. Acta Mater. 76, 186-197.
Ahadi, A., Sun, Q., 2015. Stress-induced nanoscale phase transition in superelastic NiTi by in situ

U
X-ray diffraction. Acta Mater. 90, 272-281.
Anand, L., Gurtin, M.E., 2003. Thermal effects in the superelasticity of crystalline shape-memory
AN
materials. J. Mech. Phys. Solids 51, 1015-1058.
Anand, L., Gurtin, M. E., Lele, S. P., Gething, C., 2005. A one-dimensional theory of strain-gradient
M

plasticity: formulation, analysis, numerical results. J. Mech. Phys. Solids 53, 1789-1826.
D

Arghavani, J., Auricchio, F., Naghdabadi, R., Reali, A., Sohrabpour, S., 2010. A 3-D
phenomenological constitutive model for shape memory alloys under multiaxial loadings. Int. J.
TE

Plast. 26, 976-991.


Arghavani, J., Auricchio, F., Naghdabadi, R., 2011. A finite strain kinematic hardening constitutive
EP

model based on Hencky strain: General framework, solution algorithm and application to shape
memory alloys. Int. J. Plast. 27, 940-961.
C

Auricchio, F., Taylor, R. L., Lubliner, J., 1997. Shape-memory alloys: macromodelling and numerical
AC

simulations of the superelastic behavior. Comput. Method Appl. M. 146, 281-312.


Auricchio, F., Reali, A., Stefanelli, U., 2007. A three-dimensional model describing stress-induced
solid phase transformation with permanent inelasticity. Int. J. Plast. 23, 207-226.
Berbenni, S., Dinzart, F., Sabar, H., 2015. A new internal variables homogenization scheme for linear
viscoelastic materials based on an exact Eshelby interaction law. Mech. Mater. 81, 110-124.
Bouvet, C., Calloch, S., Lexcellent, C., 2004. A phenomenological model for pseudoelasticity of
shape memory alloys under multiaxial proportional and nonproportional loadings. Eur. J. Mech.
ACCEPTED MANUSCRIPT

A-Solid. 23, 37-61.


Boyd, J. G., Lagoudas, D. C., 1996. A thermodynamical constitutive model for shape memory
materials. Part I. The monolithic shape memory alloy. Int. J. Plast. 12, 805-842.
Brinson, L. C., 1993. One-dimensional constitutive behavior of shape memory alloys:
thermomechanical derivation with non-constant material functions and redefined martensite

PT
internal variable. J. Intel. Mat. Syst. Str. 4, 229-242.
Capolungo, L., Jochum, C., Cherkaoui, M., Qu, J., 2005. Homogenization method for strength and

RI
inelastic behavior of nanocrystalline materials. Int. J. Plast. 21, 67-82.
Capolungo, L., Cherkaoui, M., Qu, J., 2007. On the elastic–viscoplastic behavior of nanocrystalline

SC
materials. Int. J. Plast. 23, 561-591.
Chatziathanasiou, D., Chemisky, Y., Chatzigeorgiou, G., Meraghni, F., 2016. Modeling of coupled

U
phase transformation and reorientation in shape memory alloys under non-proportional
AN
thermomechanical loading. Int. J. Plast. 82, 192-224.
Chemisky, Y., Duval, A., Patoor, E., Ben Zineb, T., 2011. Constitutive model for shape memory
M

alloys including phase transformation, martensitic reorientation and twins accommodation.


Mech. Mater. 43, 361-376.
D

Cisse, C., Zaki, W., Zineb, T. B., 2016a. A review of constitutive models and modeling techniques for
TE

shape memory alloys. Int. J. Plast. 76, 244-284.


Cisse, C., Zaki, W., Zineb, T. B., 2016b. A review of modeling techniques for advanced effects in
EP

shape memory alloy behavior. Smart Mater. Struct. 25, 103001.


Cho, G. B., Kim, Y. H., Hur, S. G., Yu, C. A., Nam, T. H., 2006. Transformation behavior and
C

mechanical properties of a nanostructured Ti-50.0 Ni (at.%) alloy. Met. Mater. Int. 12, 181-187.
Choudhary, N., Kaur, D., 2016. Shape memory alloy thin films and heterostructures for MEMS
AC

applications: a review. Sensor. Actuat. A-Phys. 242, 162-181.


Delville, R., Malard, B., Pilch, J., Sittner, P., Schryvers, D., 2011. Transmission electron microscopy
investigation of dislocation slip during superelastic cycling of Ni-Ti wires. Int. J. Plast. 27,
282-297.
Doghri, I., Ouaar, A., 2003. Homogenization of two-phase elasto-plastic composite materials and
structures: study of tangent operators, cyclic plasticity and numerical algorithms. Int. J. Solids
Struct. 40, 1681-1712.
ACCEPTED MANUSCRIPT

Doghri, I., Adam, L., Bilger, N., 2010. Mean-field homogenization of elasto-viscoplastic composites
based on a general incrementally affine linearization method. Int. J. Plast. 26, 219-238.
Duerig, T. W., Pelton, A., Stöckel, D., 1999. An overview of nitinol medical applications. Mater. Sci.
Eng. A-Struct. 273, 149-160.
Gall, K., Lim, T. J., McDowell, D. L., Sehitoglu, H., Chumlyakov, Y. I., 2000. The role of

PT
intergranular constraint on the stress-induced martensitic transformation in textured
polycrystalline NiTi. Int. J. Plast. 16, 1189-1214.

RI
Grabe, C., Bruhns, O.T., 2008. On the viscous and strain rate dependent behavior of polycrystalline
NiTi. Int. J. Solids Struct. 45, 1876-1895.

SC
Gu, X., Zhang, W., Zaki, W., Moumni, Z., 2017. An extended thermomechanically coupled 3D
rate-dependent model for pseudoelastic SMAs under cyclic loading. Smart Mater. Struct. 26,

U
095047.
AN
Guo, S., Kang, G., Zhang, J., 2011. Meso-mechanical constitutive model for ratchetting of
particle-reinforced metal matrix composites. Int. J. Plast. 27, 1896-1915.
M

Guo, S., Kang, G., Zhang, J., 2013. A cyclic visco-plastic constitutive model for time-dependent
ratchetting of particle-reinforced metal matrix composites. Int. J. Plast. 40, 101-125.
D

Gurtin, M. E., Anand, L., 2005. A theory of strain-gradient plasticity for isotropic, plastically
TE

irrotational materials. Part I: Small deformations. J. Mech. Phys. Solids 53, 1624-1649.
He, Y.J., Sun, Q.P., 2010. Frequency-dependent temperature evolution in NiTi shape memory alloy
EP

under cyclic loading. Smart Mater. Struct. 19, 115014.


He, Y.J., Sun, Q.P., 2011. On non-monotonic rate dependence of stress hysteresis of superelastic
C

shape memory alloy bars. Int. J. Solids Struct. 48, 1688-1695


Hu, G. K., Sun, Q. P., 2002. Thermal expansion of composites with shape memory alloy inclusions
AC

and elastic matrix. Compos. Part A-Appl. S. 33, 717-724.


Jiang, B., Weng, G. J., 2004. A generalized self-consistent polycrystal model for the yield strength of
nanocrystalline materials. J. Mech. Phys. Solids 52, 1125-1149.
Kan, Q., Yu, C., Kang, G., Li, J., Yan, W., 2016. Experimental observations on rate-dependent cyclic
deformation of super-elastic NiTi shape memory alloy. Mech. Mater. 97, 48-58.
Kang, G., Kan, Q., 2017. Cyclic Plasticity of Engineering Materials: Experiments and Models. John
Wiley & Sons.
ACCEPTED MANUSCRIPT

Kim, Y. H., Cho, G. B., Hur, S. G., Jeong, S. S., Nam, T. H., 2006. Nanocrystallization of a Ti-50.0
Ni (at.%) alloy by cold working and stress/strain behavior. Mater. Sci. Eng. A-Struct. 438,
531-535.
Ko, W. S., Maisel, S. B., Grabowski, B., Jeon, J. B., Neugebauer, J., 2017. Atomic scale processes of
phase transformations in nanocrystalline NiTi shape-memory alloys. Acta Mater. 123, 90-101.

PT
Lagoudas, D.C., Entchev, P.B., 2004. Modeling of transformation-induced plasticity and its effect on
the behavior of porous shape memory alloys. Part I: constitutive model for fully dense SMAs.

RI
Mech. Mater. 36, 865-892.
Lagoudas, D., Hartl, D., Chemisky, Y., Machado, L., Popov, P., 2012. Constitutive model for the

SC
numerical analysis of phase transformation in polycrystalline shape memory alloys. Int. J. Plast.
32, 155-183.

U
Lagoudas, D. C., 2008. Shape memory alloys: modeling and engineering applications. Springer
AN
Science & Business Media.
Lan, J., Hong, Y., 2008. Micromechanics modeling of strength for nanocrystalline copper. Arch. Appl.
M

Mech. 78, 465-476.


Lavergne, F., Sab, K., Sanahuja, J., Bornert, M., Toulemonde, C., 2016. Homogenization schemes for
D

aging linear viscoelastic matrix-inclusion composite materials with elongated inclusions. Int. J.
TE

Solids Struct. 80, 545-560.


Leclercq, S., Lexcellent, C., 1996. A general macroscopic description of the thermomechanical
EP

behavior of shape memory alloys. J. Mech. Phys. Solids, 44, 953-980.


Leng, J., Yan, X., Zhang, X., Qi, M., Liu, Z., Huang, D., 2017. A novel bending fatigue test device
C

based on self-excited vibration principle and its application to superelastic Nitinol microwire
study. Smart Mater. Struct. 26, 105020.
AC

Li, J., Weng, G. J., 2007. A secant-viscosity composite model for the strain-rate sensitivity of
nanocrystalline materials. Int. J. Plast. 23, 2115-2133.
Li, S., Sauer, R. A., Wang, G., 2007a. The Eshelby tensors in a finite spherical domai-part I:
theoretical formulations. J. Appl. Mech. 74, 770-783.
Li, S., Wang, G., Sauer, R. A., 2007b. The Eshelby tensors in a finite spherical domain-Part II:
applications to homogenization. J. Appl. Mech. 74, 784-797.
Li, M. P., Sun, Q. P., 2017. Nanoscale phase transition behavior of shape memory alloys-Closed form
ACCEPTED MANUSCRIPT

solution of 1D effective modelling. J. Mech. Phys. Solids 110, 21-37.


Long, X., Peng, X., Fu, T., Tang, S., Hu, N., 2017. A micro-macro description for pseudoelasticity of
NiTi SMAs subjected to nonproportional deformations. Int. J. Plast. 90, 44-65.
Lu, P., 2013. Further studies on Mori-Tanaka models for thermal expansion coefficients of
composites. Polymer, 54, 1691-1699.

PT
Manchiraju, S., Anderson, P.M., 2010. Coupling between martensitic phase transformations and
plasticity: a microstructure-based finite element model. Int. J. Plast. 26, 1508-1526.

RI
Mercier, S., Molinari, A., 2009. Homogenization of elastic-viscoplastic heterogeneous materials:
Self-consistent and Mori-Tanaka schemes. Int. J. Plast. 25, 1024-1048.

SC
Mori, T., Tanaka, K., 1973. Average stress in matrix and average elastic energy of materials with
misfitting inclusions. Acta Metall. 21, 571-574.

U
Morin, C., Moumni, Z., Zaki, W., 2011a. A constitutive model for shape memory alloys accounting
AN
for thermomechanical coupling. Int. J. Plast. 27, 748-767.
Morin, C., Moumni, Z., Zaki, W., 2011b. Thermomechanical coupling in shape memory alloys under
M

cyclic loadings: experimental analysis and constitutive modeling. Int. J. Plast. 27, 1959-1980.
Müller, C., Bruhns, O.T., 2006. A thermodynamic finite-strain model for pseudoelastic shape
D

memory alloys. Int. J. Plast. 22, 1658-1682.


TE

Panico, M., Brinson, L.C., 2007. A three-dimensional phenomenological model for martensite
reorientation in shape memory alloys. J. Mech. Phys. Solids 55, 2491-2511.
EP

Patoor, E., Lagoudas, D.C., Entchev, P.B., Brinson, L.C., Gao, X., 2006. Shape memory alloys, Part I:
general properties and modeling of single crystals. Mech. Mater. 38, 391-429.
C

Peng, X., Pi, W., Fan, J., 2008. A microstructure-based constitutive model for the pseudoelastic
behavior of NiTi SMAs. Int. J. Plast., 24; 966-990.
AC

Peng, X., Tang, S., Hu, N., Han, J., 2016. Determination of the Eshelby tensor in mean-field schemes
for evaluation of mechanical properties of elastoplastic composites. Int. J. Plast. 76, 147-165.
Pierard, O., Doghri, I., 2006. An enhanced affine formulation and the corresponding numerical
algorithms for the mean-field homogenization of elasto-viscoplastic composites. Int. J. Plast. 22,
131-157.
Pierard, O., LLorca, J., Segurado, J., Doghri, I., 2007. Micromechanics of particle-reinforced
elasto-viscoplastic composites: finite element simulations versus affine homogenization. Int. J.
ACCEPTED MANUSCRIPT

Plast. 23, 1041-1060.


Qiao, L., Rimoli, J. J., Chen, Y., Schuh, C. A., Radovitzky, R., 2011. Nonlocal superelastic model of
size-dependent hardening and dissipation in single crystal Cu-Al-Ni shape memory alloys. Phys.
Rev. Lett. 106, 085504.
Qiao, L., Radovitzky, R., 2016. Computational modeling of size-dependent superelasticity of shape

PT
memory alloys. J. Mech. Phys. Solids 93, 93-117.
Reese, S., Christ. D., 2008. Finite deformation pseudo-elasticity of shape memory alloys-constitutive

RI
modelling and finite element implementation. Int. J. Plast., 24, 455-482
Shaw, J.A., Kyriakides, S., 1995. Thermomechanical aspects of NiTi. J. Mech. Phys. Solids 43,

SC
1243-1281.
Simo, J. C., Taylor, R. L., 1985. Consistent tangent operators for rate-independent elastoplasticity.

U
Comput. Method. Appl. M. 48, 101-118.
AN
Simo, J. C., Taylor, R. L., 1986. A return mapping algorithm for plane stress elastoplasticity. Int. J.
Numer. Meth. Eng. 22, 649-670.
M

Sun, Q.P., Hwang, K.C., 1993a. Micromechanics modelling for the constitutive behavior of
polycrystalline shape memory alloys-I. Derivation of general relations. J. Mech. Phys. Solids 41,
D

1-17.
TE

Sun, Q.P., Hwang, K.C., 1993b. Micromechanics modelling for the constitutive behavior of
polycrystalline shape memory alloys-II. Study of the individual phenomena. J. Mech. Phys.
EP

Solids 41, 19-33.


Sun, Q. P., He, Y. J., 2008. A multiscale continuum model of the grain-size dependence of the stress
C

hysteresis in shape memory alloy polycrystals. Int. J. Solids Struct. 45, 3868-3896.
Thamburaja, P., Anand, L., 2001. Polycrystalline shape-memory materials: effect of crystallographic
AC

texture. J. Mech. Phys. Solids 49, 709-737.


Thamburaja, P., 2005. Constitutive equations for martensitic reorientation and detwinning in
shape-memory alloys. J. Mech. Phys. Solids 53, 825-856.
Wang, J., Moumni, Z., Zhang, W., 2017a. A thermomechanically coupled finite-strain constitutive
model for cyclic pseudoelasticity of polycrystalline shape memory alloys. Int. J. Plast. 97,
194-221.
Wang, J., Moumni, Z., Zhang, W., Zaki, W., 2017b. A thermomechanically coupled finite
ACCEPTED MANUSCRIPT

deformation constitutive model for shape memory alloys based on Hencky strain. Int. J. Eng.
Sci. 117, 51-77.
Xiao, H., 2013. Pseudoelastic hysteresis out of recoverable finite elastoplastic flows. Int. J. Plast., 41,
82-96.
Yin, H., He, Y., Sun, Q., 2014. Effect of deformation frequency on temperature and stress oscillations

PT
in cyclic phase transition of NiTi shape memory alloy. J. Mech. Phys. Solids 67, 100-128
Yu, C., Kang, G., Kan, Q., Song, D., 2013. A micromechanical constitutive model based on crystal

RI
plasticity for thermo-mechanical cyclic deformation of NiTi shape memory alloys. Int. J. Plast.
44, 161-191.

SC
Yu, C., Kang, G., Kan, Q., 2014a. A physical mechanism based constitutive model for
temperature-dependent transformation ratchetting of NiTi shape memory alloy:

U
one-dimensional model. Mech. Mater. 78, 1-10.
AN
Yu, C., Kang, G., Kan, Q., 2014b. Study on the rate-dependent cyclic deformation of super-elastic
NiTi shape memory alloy based on a new crystal plasticity constitutive model. Int. J. Solids
M

Struct. 51, 4386-4405.


Yu, C., Kang, G., Kan, Q., Zhu, Y., 2015a. Rate-dependent cyclic deformation of super-elastic NiTi
D

shape memory alloy: thermo-mechanical coupled and physical mechanism-based constitutive


TE

model. Int. J. Plast. 72, 60-90.


Yu, C., Kang, G., Song, D., Kan, Q., 2015b. Effect of martensite reorientation and
EP

reorientation-induced plasticity on multiaxial transformation ratchetting of super-elastic NiTi


shape memory alloy: new consideration in constitutive model. Int. J. Plast. 67, 69-101.
C

Yu, C., Kang, G., Kan, Q., 2015c. A micromechanical constitutive model for anisotropic cyclic
deformation of super-elastic NiTi shape memory alloy single crystals. J. Mech. Phys. Solids 82,
AC

97-136.
Yu, C., Kang, G., Kan, Q., 2017. An equivalent local constitutive model for grain size dependent
deformation of NiTi polycrystalline shape memory alloys. Int. J. Mech. Sci. Submitted for
publication.
Zaki, W., Moumni, Z., 2007a. A three-dimensional model of the thermomechanical behavior of shape
memory alloys. J. Mech. Phys. Solids 55, 2455-2490.
Zaki, W., Moumni, Z., 2007b. A 3D model of the cyclic thermomechanical behavior of shape
ACCEPTED MANUSCRIPT

memory alloys. J. Mech. Phys. Solids 55, 2427-2454.


Zaki, W., 2012. An efficient implementation for a model of martensite reorientation in martensitic
shape memory alloys under multiaxial nonproportional loading. Int. J. Plast. 37, 72-94.
Zhang, X., Yan, X., Xie, H., Sun, R., 2014. Modeling evolutions of plastic strain, maximum
transformation strain and transformation temperatures in SMA under superelastic cycling.

PT
Comp. Mater. Sci. 81, 113-122.
Zhang, X., Huang, D., Yan, X., Zhou, X. 2016. Modeling functional fatigue of SMA using a more

RI
accurate subdivision of martensite volume fractions. Mech. Mater. 96, 12-29.
Zhang, Y., You, Y., Moumni, Z., Anlas, G., Zhu, J., Zhang, W., 2017. Experimental and theoretical

SC
investigation of the frequency effect on low cycle fatigue of shape memory alloys. Int. J. Plast.
90, 1-30.

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

Table 1 Summary of the proposed model

Overall constitutive relationship of GI phase


Strain decomposition:

ε GI
= εe + εtr
GI GI

Stress-elastic strain relationship:

PT
σɺ GI
= EGI : εɺ e
GI

Flow rule:

RI
 3 s

GI
ξɺ ≥0
2 s GI

SC
 GI
εɺ tr = g tr N ξɺ N=
GI GI
 3 εt − r
 t −r
GI
ξɺ <0
2 ε GI


U
GI

Transformation surfaces:
AN
Φ for = g tr σ GI
: N − β (T − T0 ) − H ξ GI
−Y ≤ 0
GI

: N + β (T − T0 ) + H ξ
M

Φ rev = − g tr σ GI GI
−Y ≤ 0
GI

Transformation hardening modulus:


D

12S0le2 8S0le2
 2 0≤ F ≤
TE

 lg l g2
 2 3
(
H =  S0le 4lg lɶ − 6lg lɶ + 3l g lɶ
2 2 3
) F>
8S0le2
 ɶ 2 2S l 2 h  2 lg2
  lɶ lg l
EP

 6 − 3 + 2 + F 
0 e

  

: N − β (T − T0 ) − Y
C

 g tr σ ξɺ >0
 GI
F =
GI

: N − β (T − T0 ) + Y ξɺ
AC

tr
 g σ GI
<0
GI

Continuous tangent moduli:

σɺ GI
= DGI ɺ
ε : ε GI
+ DTGI Tɺ

Overall constitutive relationship of GB phase

σɺ GB
= EGB : εɺ GB

Evolution of average temperature


ACCEPTED MANUSCRIPT

Shɶ (Tr − T )
ɺ
cT − = fGI γ T ξɺ + fGI  g tr σ GI
: N − β (T − T0 ) − H ξ GI
 ξɺ

VΩ GI GI

Homogenization method
Strain concentration relationships:
GI ɺ
εɺ GI
= CGI
ε : ε + CT T
ɺ

PT
GB ɺ
εɺ GB
= CGB
ε : ε + CT T
ɺ

Overall continuous tangent modulus of the composite:

RI
σɺ = Dε : εɺ + DT Tɺ

SC
Dε = f GI DεGI : CεGI + (1 − f GI ) EGB : CGB
ε DT = fGI ( DGI
ε : CT + DT ) + (1 − f GI ) EGB : CT
GI GI GB

Intermediate variables:

( )
U
−1
 :B
ε = I + Π
CGI ɶ GI : A − Π
ɶ GI
  ε
AN
ε

( ) ( )
−1 −1
CTGI = I + Π
ɶ GI : A − Π
ɶ GI :B −Π
ɶ GI : A − Π
ɶ GI : AT
 ε
 T ε

( )
−1
= I + Π :B
M

CGB ɶ GB : A − Π
ɶ GI
ε
 ε
 ε

( ) ( )
−1 −1
CTGB = I + Π
ɶ GB : A − Π
ɶ GI :B −Π
ɶ GB : A − Π
ɶ GI : AT
  T
D

ε ε

( ) ( )
−1
ɶ GI :  A − (1 − f ) Π
Bε = A ε − Π ɶ GB 
ɶ GI − Π
TE

 ε GI 

( )
−1
BT =  A ε − (1 − f GI ) Π   GI
ɶ GI + (1 − f ) Π
ɶ GB  :  f Π
ɶ GI − Π
GI
ɶ GB  : A
 T
EP

Aε = ( EGB − DεGI ) : EGB AT = ( EGB − DGI


ε ) : DT
−1 GI −1
C
AC
ACCEPTED MANUSCRIPT

Table 2 Material parameters used in the proposed model


Parameters related to GB thickness:

tb0 =1nm; d1 =12nm; m1 =5;

Elastic constants:

EGI =13GPa; EGB =43GPa; ν GI =ν GI =0.3;

PT
Parameters related to martensite transformation

β 0 =7.46MPaK-1; d 2 =9.1nm; m2 =1.34; Y =3.5MPa; X 0 =11MPa; S0le2 =0.8 × 10-9Pam2;

RI
Parameters related to heat transfer:

SC
c =3.225MJ m-3K-1; hɶ =100Wm-2K-1; γ =0.39MPa K-1.

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT
600

500

400

Stress (MPa) 300


-4
3.3×10 /s
-4
200 6.6×10 /s

PT
-3
1.0×10 /s
-3
3.3×10 /s
100 -2
1.0×10 /s
-2
3.3×10 /s

RI
0
0 2 4 6 8 10 12

Strain (%)

SC
Fig. 1 Rate-dependent stress-strain responses of polycrystalline NiTi SMA (cited from Kan et al.,
2016)

U
AN
1500
Grain size 1.5µm
M

Grain size 80nm


Grain size 64nm
1200 Grain size 42nm
Grain size 27nm
Stress (MPa)

Grain size 18nm


900 Grain size 10nm
TE

Ahadi and Sun (2013)

600
EP

300

0
C

0 1 2 3 4 5
Strain (%)
AC

Fig. 2 Grain-size dependent stress-strain responses of polycrystalline NiTi SMA (cited from Ahadi
and Sun, 2013)
ACCEPTED MANUSCRIPT

PT
Fig. 3 (a) macroscopic specimen; (b) GI and GB phases in polycrystalline (cited from Ko et al.,

RI
2017); (c) the two-phase composite model.

U SC
x Mechanical loading
AN
(a) (b) 1

Heat conduction
Heat conduction 2
Scro
Grip
M

Heat conduction
Heat convection
Gauge length L

Simplification
Slat
Heat convection
Heat convection
D
TE

z
1

Heat conduction
Model 1
Scro
y
EP

Experiment Heat conduction

Fig. 4 (a) Mechanical and heat transfer in the experiment and the model; (b) thermal and mechanical
C

boundary conditions.
AC
ACCEPTED MANUSCRIPT

Thickness of GB (nm)
4

PT
2

RI
0
10 20 30 40 50 60 70 80 90 100

SC
Grain size (nm)

Fig. 5 Relationship between the thickness of GB phase and grain size.

U
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

600 600
(a) Grian size 90nm (b) Grian size 90nm
-5 -4
Strain rate 4×10 /s Strain rate 5×10 /s
500 500
Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014)
Simulation Simulation
Stress (MPa)

Stress (MPa)
400 400
f
σs90
300 300

200 200

100 100

PT
r
σf90

0 0
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)

RI
600 600
(c) Grian size 90nm (d) Grian size 90nm
-3 -3
Strain rate 2×10 /s Strain rate 5×10 /s
500 500
Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014) b c
Simulation Simulation
Stress (MPa)

400

Stress (MPa)
400

SC
a
300 300 d

200 200

U
100 100
o e
0 0
AN
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
600 600
(e) Grian size 90nm (f) Grian size 90nm
-2
-2
Strain rate 1×10 /s Strain rate 5×10 /s
M

500 500
Experiment (Ahadi and Sun, 2014)
Simulation
Stress (MPa)

400
Stress (MPa)

400

300 300
D

200 200
TE

100 100 Experiment (Ahadi and Sun, 2014)


Simulation
0 0
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
EP

600 60
(g) Grian size 90nm (h) Grian size 90nm
-1
Strain rate 1×10 /s Experiment Simulation
500 -3
1×10 /s -3
1×10 /s
50 -3
Temperature (°C)

5×10 /s -3
5×10 /s
-1
b
1×10 /s
Stress (MPa)

-1
400
C

c 1×10 /s
40
d
300
AC

30
200
o a
Experiment (Ahadi and Sun, 2014) 20
100
Simulation
Loading Unloading
e
0 10
0 2 4 6 8 6 5
0 1 2 3 4 5 6 7 8 9 10 3 12
4 11 2 13 0
1 14
Strain (%) Strain (%)

Fig. 6 Thermo-mechanically coupled deformation of polycrystalline NiTi SMA with a grain size
of 90 nm: (a)-(g) stress-strain responses at different average strain rates (4×10-5/s, 5×10-4/s, 2×10-3/s,
5×10-3/s, 1×10-2/s, 5×10-2/s, 1×10-1/s); (h) evolutions of average temperature at various strain rates
(1×10-3, 5×10-3 and 1×10-1/s (the experimental results are cited from Ahadi and Sun, 2014).
ACCEPTED MANUSCRIPT

800 800
(a) Grian size 68nm (b) Grian size 68nm
-5 -4
Strain rate 4×10 /s Strain rate 5×10 /s

600 Experiment (Ahadi and Sun, 2014) 600 Experiment (Ahadi and Sun, 2014)
Simulation Simulation
Stress (MPa)

Stress (MPa)
400 400

200 200

PT
0 0
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)

RI
800 800
(c) Grian size 68nm (d) Grian size 68nm
-3 -3
Strain rate 2×10 /s Strain rate 5×10 /s
Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014)
600 600
Simulation Simulation
Stress (MPa)

Stress (MPa)

SC
400 400

200 200

U 0
AN
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
800 800
(e) Grian size 68nm (f) Grian size 68nm
-2 -2
Strain rate 1×10 /s Strain rate 5×10 /s
M

600 Experiment (Ahadi and Sun, 2014) 600 Experiment (Ahadi and Sun, 2014)
Simulation Simulation
Stress (MPa)

Stress (MPa)

400 400
D

200 200
TE

0 0
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
EP

800 60
(g) Grian size 68nm (h) Grian size 68nm
-1
Strain rate 1×10 /s Experiment Simulation
-3 -3
Experiment (Ahadi and Sun, 2014)
50 1×10 /s 1×10 /s
Temperature (°C)

600 -3
5×10 /s
-3
5×10 /s
Simulation
Stress (MPa)

-1 -1
1×10 /s 1×10 /s
C

40
400
AC

30

200
20 loading unloading

0 10
0 2 4 6 8 8 59 10
0 1 2 3 4 5 6 7 6 3 12
4 11 2 13 0
1 14
Strain (%) Strain (%)

Fig. 7 Thermo-mechanically coupled deformation of polycrystalline NiTi SMA with a grain size
of 68 nm: (a)-(g) stress-strain responses at different average strain rates (4×10-5/s, 5×10-4/s, 2×10-3/s,
5×10-3/s, 1×10-2/s, 5×10-2/s, 1×10-1/s); (h) evolutions of average temperature at various strain rates
(1×10-3, 5×10-3 and 1×10-1/s (the experimental results are cited from Ahadi and Sun, 2014).
ACCEPTED MANUSCRIPT

800 800
(a) Grian size 42nm (b) Grian size 42nm
-5 -4
Strain rate 4×10 /s Strain rate 5×10 /s
Experiment (Ahadi and Sun, 2014)
600 Experiment (Ahadi and Sun, 2014) 600
Simulation
Simulation
Stress (MPa)

Stress (MPa)
400 400

200 200

PT
0 0
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
800 800

RI
(c) Grian size 42nm (d) Grian size 42nm
-3 -3
Strain rate 2×10 /s Strain rate 5×10 /s
Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014)
600 600 Simulation
Simulation
Stress (MPa)

Stress (MPa)

SC
400 400

200 200

U 0
AN
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
800 800
(e) Grian size 42nm (f) Grian size 42nm
-2
-2
Strain rate 1×10 /s Strain rate 5×10 /s
M

Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014)
600 600 Simulation
Simulation
Stress (MPa)

Stress (MPa)

400 400
D

200 200
TE

0 0
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
EP

800 60
(g) Grian size 42nm (h) Grian size 42nm
-1
Strain rate 1×10 /s Experiment Simulation
-3
50 -3
1×10 /s 1×10 /s
Experiment (Ahadi and Sun, 2014)
600
Temperature (°C)

-3
Simulation
-3
5×10 /s 5×10 /s
-1
Stress (MPa)

-1
1×10 /s 1×10 /s
C

40
400
AC

30

200
20
loading unloading

0 10
0 2 4 6 8 5 10
6 9
0 1 2 3 4 5 6 7 8 3 12
4 11 22 13 00
1 14
Strain (%) Strain (%)
0

Fig. 8 Thermo-mechanically coupled deformation of polycrystalline NiTi SMA with a grain size
of 42 nm: (a)-(g) stress-strain responses at different average strain rates (4×10-5/s, 5×10-4/s, 2×10-3/s,
5×10-3/s, 1×10-2/s, 5×10-2/s, 1×10-1/s); (h) evolutions of average temperature at various strain rates
(1×10-3, 5×10-3 and 1×10-1/s (the experimental results are cited from Ahadi and Sun, 2014).
ACCEPTED MANUSCRIPT

800 800
(a) Grian size 27nm (b) Grian size 27nm
-4
-5
Strain rate 4×10 /s Strain rate 5×10 /s

600 600

Stress (MPa)
Stress (MPa)

400 400

200 200
Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014)

PT
Simulation Simulation
0 0
0 2 4 6 8 0 2 4 6 8

Strain (%) Strain (%)


800

RI
800
(c) Grian size 27nm (d) Grian size 27nm
-3
-3
Strain rate 2×10 /s Strain rate 5×10 /s

600 600
Stress (MPa)

Stress (MPa)

SC
400 400

200 200

U
Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014)
Simulation Simulation
0 0
AN
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
800 800
(e) Grian size 27nm (f) Grian size 27nm
-2
-2
Strain rate 1×10 /s Strain rate 5×10 /s
M

600 600
Stress (MPa)

Stress (MPa)

400 400
D

200 200
TE

Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014)
Simulation Simulation
0 0
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
EP

800 60
(g) Grian size 27nm (h) Grian size 27nm
-1
Strain rate 1×10 /s
Experiment Simulation
50 -3
1×10 /s -3
1×10 /s
Temperature (°C)

600 -3
5×10 /s -3
5×10 /s
Stress (MPa)

-1
C

1×10 /s -1
1×10 /s
40
400
AC

30

200
20
Experiment (Ahadi and Sun, 2014) loading unloading
Simulation
0 10
0 2 4 6 8 6 5
0 1 2 3 4 5 6 7 8 9 10 3 12
4 11 22 13 00
1 14
Strain (%) Strain (%)

Fig. 9 Thermo-mechanically coupled deformation of polycrystalline NiTi SMA with a grain size
of 27 nm: (a)-(g) stress-strain responses at different average strain rates (4×10-5/s, 5×10-4/s, 2×10-3/s,
5×10-3/s, 1×10-2/s, 5×10-2/s, 1×10-1/s); (h) evolutions of average temperature at various strain rates
(1×10-3, 5×10-3 and 1×10-1/s (the experimental results are cited from Ahadi and Sun, 2014).
ACCEPTED MANUSCRIPT

1800 1800
(a) Grian size 10nm (b) Grian size 10nm
1600 -5
Strain rate 4×10 /s
1600 -4
Strain rate 5×10 /s
1400 1400
Stress (MPa)

Stress (MPa)
1200 1200
1000 1000
800 800
600 600
400 400
Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014)

PT
200 Simulation 200 Simulation
0 0
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
1800 1800

RI
(c) Grian size 10nm (d) Grian size 10nm
1600 -3
Strain rate 2×10 /s
1600 -3
Strain rate 5×10 /s
1400 1400
Stress (MPa)

Stress (MPa)
1200 1200

SC
1000 1000
800 800
600 600
400 400

U
Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014)
200 Simulation 200 Compression
0 0
AN
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
1800 1800
(e) Grian size 10nm (f) Grian size 10nm
1600 -2 1600 -2
Strain rate 5×10 /s
Strain rate 1×10 /s
M

1400 1400
Stress (MPa)

1200
Stress (MPa)

1200
1000 1000
D

800 800
600 600
400 400
TE

Experiment (Ahadi and Sun, 2014) Experiment (Ahadi and Sun, 2014)
200 Simulation 200 Simulation
0 0
0 2 4 6 8 0 2 4 6 8
Strain (%) Strain (%)
EP

1800 60
(g) Grian size 10nm (h) Grian size 10nm
1600 -1
Strain rate 1×10 /s Experiment Simulation
1400 50 -3
1×10 /s -3
1×10 /s
Temperature (°C)

-3
5×10 /s -3
5×10 /s
Stress (MPa)

1200 -1
40 1×10 /s -1
1×10 /s
1000
800
AC

30
600
400 20 loading unloading
Experiment (Ahadi and Sun, 2014)
200 Simulation
0 10
0 2 4 6 8 0 1 2 3 4 5 6
4 3
7 8
2 9
1 0
10
Strain (%) Strain (%)

Fig. 10 Thermo-mechanically coupled deformation of polycrystalline NiTi SMA with a grain size
of 10 nm: (a)-(g) stress-strain responses at different average strain rates (4×10-5/s, 5×10-4/s, 2×10-3/s,
5×10-3/s, 1×10-2/s, 5×10-2/s, 1×10-1/s); (h) evolutions of average temperature at various strain rates
(1×10-3, 5×10-3 and 1×10-1/s (the experimental results are cited from Ahadi and Sun, 2014).
ACCEPTED MANUSCRIPT

1800

1500

Experiment Simulation

Stress (MPa)
1200 90nm
68nm
42nm
900 27nm
10nm

PT
600

300

RI
0
1E-5 1E-4 1E-3 0.01 0.1 1

SC
-1
Strain rate (s )

Fig. 11 Curves of maximum stress vs. strain rate for the polycrystalline NiTi SMAs with various

U
grain sizes (the experimental results are cited from Ahadi and Sun, 2014).
AN
M

σ σ
D
TE

lg
σ σ
EP

o x
C

Fig. A1 1D chain element model.


AC
ACCEPTED MANUSCRIPT

Research Highlights for the paper

(1) A micromechanical constitutive model is established.


(2) The polycrystalline NiTi SMA is regarded as a two-phase composite of GI and GB phases
(3) A new incremental non-isothermal Mori-Tanaka’s homogenization method is developed.

PT
(4) A 3D scaling law of transformation hardening modulus w.r.t. the grain size is proposed.
(5) Grain size dependent thermo-mechanically coupled deformation of NiTi SMAs is described

RI
well by the proposed model.

U SC
AN
M
D
TE
C EP
AC

S-ar putea să vă placă și