28.01.2013 Views

A Simplified Multivariant SMA Model Based on Invariant Plane ...

A Simplified Multivariant SMA Model Based on Invariant Plane ...

A Simplified Multivariant SMA Model Based on Invariant Plane ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

A <str<strong>on</strong>g>Simplified</str<strong>on</strong>g> <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>SMA</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> <str<strong>on</strong>g>Based</str<strong>on</strong>g> <strong>on</strong> <strong>Invariant</strong> <strong>Plane</strong><br />

Nature of Martensitic Transformati<strong>on</strong><br />

Xiujie Gao and L. Catherine Brins<strong>on</strong><br />

Department of Mechanical Engineering, Northwestern University<br />

2145 Sheridan Road, Evanst<strong>on</strong> IL 60208<br />

ABSTRACT<br />

The <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model describing shape memory alloy c<strong>on</strong>stitutive behavior is further<br />

developed in this paper for improved predicti<strong>on</strong> of thermomechanical resp<strong>on</strong>se. In the<br />

original formulati<strong>on</strong> of the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model, an interacti<strong>on</strong> energy, representative of<br />

the incompatibility of an inclusi<strong>on</strong> to the matrix, was calculated by a micromechanics<br />

model in which every variant group (inclusi<strong>on</strong>) was embedded in the austenite phase with<br />

numerous other inclusi<strong>on</strong>s. However, experiments show that martensite variants tend to<br />

form large plates most of which have an invariant plane interface with the austenite and<br />

reach the grain boundary. Hence, to better simulate material behavior, in the revised<br />

model the micromechanical interacti<strong>on</strong> energy is replaced by a small, c<strong>on</strong>stant term. The<br />

predicti<strong>on</strong>s by the new model for different uniaxial tensi<strong>on</strong> directi<strong>on</strong>s <strong>on</strong> a single CuAlNi<br />

crystal have excellent agreement with the experimental results. Furthermore, the counter-<br />

intuitive results for a polycrystal CuZnAl under triaxial loading are also well captured by<br />

the new model. As the revised model removes an iterative procedure for interacti<strong>on</strong><br />

energy, calculati<strong>on</strong>s are simplified, making the multivariant model more suitable for<br />

larger scale computati<strong>on</strong>s.<br />

1. INTRODUCTION<br />

Due to the nature of martensitic transformati<strong>on</strong>s with stress, research and applicati<strong>on</strong>s of<br />

shape memory alloys have often been <strong>on</strong>e-dimensi<strong>on</strong>al in nature in order to obtain<br />

maximal recoverable deformati<strong>on</strong>. However, certain classes of applicati<strong>on</strong>s, such as <str<strong>on</strong>g>SMA</str<strong>on</strong>g><br />

hybrid (self-healing) materials and many shape c<strong>on</strong>trol or frequency c<strong>on</strong>trol applicati<strong>on</strong>s,<br />

1


will subject the material to multi-dimensi<strong>on</strong>al stresses, partial loading and unloading,<br />

reorientati<strong>on</strong> of loading, as well as n<strong>on</strong>-uniform temperature c<strong>on</strong>diti<strong>on</strong>s. Thus an<br />

understanding of <str<strong>on</strong>g>SMA</str<strong>on</strong>g> resp<strong>on</strong>se to these n<strong>on</strong>-simple loading c<strong>on</strong>diti<strong>on</strong>s is needed. In<br />

additi<strong>on</strong>, better understanding of the directi<strong>on</strong>ality of <str<strong>on</strong>g>SMA</str<strong>on</strong>g> resp<strong>on</strong>se in general 3-D stress<br />

states will open the doors to a new set of potential uses.<br />

There are several 3-D <str<strong>on</strong>g>SMA</str<strong>on</strong>g> models currently in the literature (Auricchio and Taylor,<br />

1997; Boyd and Lagoudas, 1994; Boyd and Lagoudas, 1996; Graesser and Cozzarelli,<br />

1994, Patoor, 1994 #26; Lagoudas et al., 1996; Lu and Weng, 1997; Sun and Hwang,<br />

1993a; Sun and Hwang, 1993b). Predominantly, the existing c<strong>on</strong>tinuum level models<br />

approach <str<strong>on</strong>g>SMA</str<strong>on</strong>g> c<strong>on</strong>stitutive resp<strong>on</strong>se by making use of macroscale plasticity theory<br />

(often complete with backstresses, flow rules and yield functi<strong>on</strong>s). While it is appealing<br />

to utilize the vast plasticity model literature and methodology, this approach must be<br />

exercised with cauti<strong>on</strong> since the <str<strong>on</strong>g>SMA</str<strong>on</strong>g> deformati<strong>on</strong>/stress resp<strong>on</strong>se is based <strong>on</strong> different<br />

underlying mechanisms than irreversible plastic resp<strong>on</strong>se. To most accurately represent<br />

<str<strong>on</strong>g>SMA</str<strong>on</strong>g> material behavior, the problem is best approached from the fundamental material<br />

resp<strong>on</strong>se up, accounting appropriately for the mechanisms which impart the unique<br />

reversible, c<strong>on</strong>trollable nature to <str<strong>on</strong>g>SMA</str<strong>on</strong>g>s. Specifically, macroscopic resp<strong>on</strong>se of <str<strong>on</strong>g>SMA</str<strong>on</strong>g> can<br />

be built up from deformati<strong>on</strong> at the variant level for single crystal behavior and<br />

polycrystalline resp<strong>on</strong>se can be based <strong>on</strong> individual grain (single crystal) resp<strong>on</strong>se.<br />

In recent years, several researchers have proposed <str<strong>on</strong>g>SMA</str<strong>on</strong>g> models based <strong>on</strong> a<br />

micromechanics approach (Goo and Lexcellent, 1997; Huang and Brins<strong>on</strong>, 1998; Lu and<br />

Weng, 1997; Patoor et al., 1994). These models use thermodynamics laws to describe the<br />

transformati<strong>on</strong> while using micromechanics to estimate the interacti<strong>on</strong> energy due to the<br />

phase transformati<strong>on</strong>. All are based up<strong>on</strong> transformati<strong>on</strong> strains of individual variants,<br />

utilized in varying fashi<strong>on</strong>s in the micromechanics formulati<strong>on</strong>. Most of the models<br />

handle <strong>on</strong>e variant <strong>on</strong>ly and will therefore not allow for loading involving reorientati<strong>on</strong> of<br />

martensitic variants. One excepti<strong>on</strong> is the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model by Huang and Brins<strong>on</strong><br />

(Huang and Brins<strong>on</strong>, 1998) and a further developed versi<strong>on</strong> by Gao et al. (Gao et al.,<br />

2000), which not <strong>on</strong>ly accounts for all variants separately, but also includes the self-<br />

2


accommodated grouping structure for the habit plane variants. Therefore, re-orientati<strong>on</strong>,<br />

i.e. transformati<strong>on</strong> from <strong>on</strong>e martensite variant to another martensite variant, is accounted<br />

for. Although a wide array of thermomechanical resp<strong>on</strong>se, including multiaxial loading,<br />

can be simulated with the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> and the results are in good<br />

qualitative agreement with the experimental data, a few difficulties in the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g><br />

model such as high transformati<strong>on</strong> stress and l<strong>on</strong>g calculati<strong>on</strong> time prevent it from being<br />

practically used.<br />

Observati<strong>on</strong> of both stress and temperature induced martensitic transformati<strong>on</strong> reveal that<br />

martensite variants tend to form large plates, most of which have invariant plane<br />

interfaces with the austenite and reach the grain boundary. Hence the original formulati<strong>on</strong><br />

of the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model in which every variant group (inclusi<strong>on</strong>) is embedded in the<br />

austenite phase with numerous other inclusi<strong>on</strong>s is not physically representative of<br />

material behavior and therefore predicts an interacti<strong>on</strong> barrier to transformati<strong>on</strong> that is too<br />

large. In this paper, a simplified <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model is developed in which the energy<br />

c<strong>on</strong>tributed by the incompatibility of inclusi<strong>on</strong> to the matrix is taken to be a small<br />

c<strong>on</strong>stant.<br />

Similar simplified models (with some variati<strong>on</strong>) can be found in literature. However, no<br />

systematic study or comparis<strong>on</strong> to experiments with a model such as described here has<br />

been performed. Patoor et al. (Patoor et al., 1993a) neglected the interacti<strong>on</strong> between<br />

martensite plates and parent phase in the single crystal simulati<strong>on</strong>. However, <strong>on</strong>ly<br />

forward transformati<strong>on</strong>s of orientati<strong>on</strong> dependence etc. were simulated. In the polycrystal<br />

case, no intergranular effect was taken into account, and thus stair-like curves were<br />

obtained by averaging. Subsequent papers by Patoor et al. (Patoor et al., 1993b; Patoor et<br />

al., 1994) use an interacti<strong>on</strong> matrix to account for the interacti<strong>on</strong> between plates and<br />

austenite phase in single crystal case, and a self-c<strong>on</strong>sistent modulus predicting scheme to<br />

obtain better averaging am<strong>on</strong>g grains in the polycrystal case. Other micromechanics<br />

models (Lexcellent et al., 1996; Lu and Weng, 1997) use different methods to account for<br />

the interacti<strong>on</strong> energy, and these will be discussed in more detail in a later secti<strong>on</strong>. In<br />

many cases, quantitative correlati<strong>on</strong> with experimental results has been difficult to obtain.<br />

3


We propose that a major obstacle in many of these models has been an overly large<br />

estimati<strong>on</strong> of interacti<strong>on</strong> energy provided by the micromechanics methods.<br />

In next secti<strong>on</strong> a summary of the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model is given, then difficulties<br />

for the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model are reviewed. The simplified <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model is then<br />

derived, motivated by reas<strong>on</strong>able interacti<strong>on</strong> energies based <strong>on</strong> experimental observati<strong>on</strong>s<br />

and implementati<strong>on</strong>s in previous micromechanical models. Finally, results predicted by<br />

both the original and revised <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> models are compared to experimental results.<br />

2. SUMMARY OF THE ORIGINAL MULTIVARIANT MODEL<br />

In this secti<strong>on</strong>, the previously developed <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> will be briefly summarized<br />

for both the single crystal and polycrystal forms. Although this introducti<strong>on</strong> is intended to<br />

be short, some details are still given so that readers can understand the basis of the<br />

simplified model development. The readers are also referred to previous work <strong>on</strong> the<br />

<str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> (Gao et al., 2000; Huang et al., 2000; Huang and Brins<strong>on</strong>, 1998).<br />

2.1. Single Crystal <str<strong>on</strong>g>Model</str<strong>on</strong>g> <str<strong>on</strong>g>Based</str<strong>on</strong>g> <strong>on</strong> Micromechanics<br />

The <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model begins with the complementary free energy of the material<br />

n<br />

ΨΣ ( , T, f ) = − [ ∆G + W + W − Σ E ]<br />

(1)<br />

ij<br />

ch mech sur ij ij<br />

the rate of change of which is set equal to the rate of change of the dissipati<strong>on</strong> energy<br />

dΨ | Σ ij ,T = dW d ≥ 0 (2)<br />

For the chemical free energy, ∆G ch , a standard expressi<strong>on</strong> is used, linearly proporti<strong>on</strong>al<br />

to temperature change; the surface energy, W sur , is small in comparis<strong>on</strong> to other terms and<br />

is neglected. The mechanical energy, W mech , is approximated by a micromechanics<br />

technique in which the formati<strong>on</strong> of martensitic variants in a shape memory alloy under<br />

loading is viewed as a superpositi<strong>on</strong> of the external loading acting <strong>on</strong> a homogeneous<br />

material and a set of transforming inclusi<strong>on</strong>s (see figure 1). In the simplified model, this<br />

4


last assumpti<strong>on</strong> will be revisited. With the micromechanics approach it can be shown that<br />

the mechanical energy reduces to a stored elastic energy term plus an interacti<strong>on</strong> energy<br />

term, Eint ,<br />

W mech = 1<br />

∫<br />

2V σ e<br />

ijεijdV V<br />

= 1<br />

2 Σ −1<br />

ijCijklΣkl − 1<br />

5<br />

∫<br />

II tr<br />

εij dV<br />

2V σij � � � V�<br />

� � �<br />

where Σ ij is the macroscopic applied stress, σ ij , ε ij are the local stress and strain<br />

(superscripts e and tr <strong>on</strong> the strain represent elastic and transformati<strong>on</strong> strains<br />

respectively), and C ijkl is the modulus of the material. In order to use the closed form<br />

expressi<strong>on</strong> of the Eshelby tensor in the subsequent Eshelby inclusi<strong>on</strong> analysis (Mura,<br />

1987), C ijkl is taken to be isotropic and identical for austenite and martensite. In the new<br />

model, since the Eshelby Tensor is no l<strong>on</strong>ger needed, the anisotropic effects of the<br />

material’s crystal lattice can be easily captured in the single crystal case.<br />

To calculate the original form of the interacti<strong>on</strong> energy, the formati<strong>on</strong> of martensitic<br />

variants in self-accommodated groups of compatible variants was recognized, which<br />

minimized the interacti<strong>on</strong> energy compared to previous similar approaches. Using this<br />

key c<strong>on</strong>cept, the interacti<strong>on</strong> energy is approximated as a sum over G groups of M self-<br />

accommodating variants each<br />

where σ ij<br />

g<br />

Eint =− 1<br />

2V σ ∫ ij<br />

V<br />

II tr<br />

εij dV<br />

E int<br />

(3)<br />

=− 1<br />

G<br />

g<br />

g<br />

∑ σij ε ij f<br />

2<br />

g (4)<br />

g g<br />

,ε ij , f are the average stress in an inclusi<strong>on</strong> of group g, average<br />

transformati<strong>on</strong> strain of group g and the volume fracti<strong>on</strong> of group g, respectively. The<br />

n<br />

transformati<strong>on</strong> strain of each variant, εij (n=1,2, ... , GxM), is calculated from the basic<br />

crystallographic informati<strong>on</strong> <strong>on</strong> the habit plane normal, n, invariant plane shear directi<strong>on</strong>,<br />

m, and magnitude of the invariant plane shear, g. For a given material:<br />

tr 1<br />

ε = gnm ( + nm)<br />

(5)<br />

ij<br />

i j j i<br />

2<br />

Equati<strong>on</strong> (4) can be shown to reduce to<br />

g=1


where<br />

g<br />

and Sijkl can be written as<br />

L<br />

NM<br />

G<br />

G<br />

1 g g m m g<br />

Eint =− ε � σ − f � σ f<br />

(6)<br />

∑<br />

ij ij<br />

2 g=<br />

1<br />

m=<br />

1<br />

6<br />

∑<br />

ij<br />

O<br />

QP<br />

g<br />

g g g<br />

� σ = C ( S ε − ε )<br />

(7)<br />

ij<br />

ijkl klmn<br />

mn<br />

is the Eshelby tensor of group g. The final system of equati<strong>on</strong>s obtained from (2)<br />

n<br />

F<br />

d n<br />

tr E<br />

F f B T T ij ij<br />

F if f<br />

f<br />

n<br />

Fext<br />

F F<br />

C n<br />

n n<br />

n<br />

int<br />

n<br />

n<br />

wall<br />

fric<br />

� int<br />

=− ( − ) + − , �<br />

∂<br />

��������� 0<br />

+ − ><br />

0 Σ ε λ λ<br />

< 0<br />

∂<br />

��� ���<br />

∓<br />

where superscript n represents the n th of the N=GxM variants. In the external driving<br />

n<br />

force Fext, B is a linearized factor of difference in the chemical free energies per unit<br />

n<br />

volume of the two phases near the thermodynamic equilibrium temperature T0. Fwall represents a boundary force applied <strong>on</strong> the system to keep the martensite fracti<strong>on</strong> within<br />

physical bounds ( 0 ≤ f n ≤ 1), and it <strong>on</strong>ly appears for a variant when the volume fracti<strong>on</strong><br />

n C<br />

of that variant is very close to 0 or 1. Ffric is a c<strong>on</strong>stant F<br />

kl<br />

6<br />

(8)<br />

if variant n is actively<br />

transforming forward, and -F C<br />

when transforming backward. The system of equati<strong>on</strong>s (8)<br />

is solved numerically for a single c<strong>on</strong>stitutive point; a short descripti<strong>on</strong> of the algorithm<br />

can be found in the appendix.<br />

Note that the formulati<strong>on</strong> of the interacti<strong>on</strong> energy imposes that an “inclusi<strong>on</strong>” is a group<br />

of variants; at the same time, however, the volume fracti<strong>on</strong>s for each individual variant<br />

are tracked so that a final product of a single variant is possible.<br />

2.2. Polycrystal <str<strong>on</strong>g>Model</str<strong>on</strong>g> <str<strong>on</strong>g>Based</str<strong>on</strong>g> <strong>on</strong> Micromechanics<br />

The behavior of single and polycrystalline <str<strong>on</strong>g>SMA</str<strong>on</strong>g>s are quite different. Single crystal <str<strong>on</strong>g>SMA</str<strong>on</strong>g>s<br />

are highly anisotropic and this fact has serious implicati<strong>on</strong>s in both the modeling and<br />

experimental resp<strong>on</strong>se. An untextured polycrystalline material, however, c<strong>on</strong>sists of


many single crystal grains which when randomly oriented lead to an overall isotropic<br />

resp<strong>on</strong>se.<br />

In the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g>, simulati<strong>on</strong> of a polycrystalline specimen is c<strong>on</strong>ducted by<br />

assuming that a number of randomly oriented single crystal grains are in the poly-<br />

crystalline specimen. The single crystal model developed is used for each single crystal<br />

grain and a self-c<strong>on</strong>sistent method (Budiansky and Wu, 1962; Kröner, 1961) is used to<br />

account for the interacti<strong>on</strong> between grains. The main difference between the original<br />

<str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model and the simplified model proposed in this paper is at the single crystal<br />

level (described in the next secti<strong>on</strong>). Therefore the method described here for the<br />

polycrystalline formulati<strong>on</strong> is applicable for both the original and the improved<br />

<str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model.<br />

We assume a homogeneous stress field at infinity, and that every grain is in turn under<br />

the same external homogeneous stress field. Prior to any transformati<strong>on</strong>, all grains see the<br />

same overall stress field. However, since the grains are differently oriented some grains<br />

will begin transformati<strong>on</strong> at a lower external stress level. Therefore the transformati<strong>on</strong><br />

strains in each grain will differ. Due to the incompatibility of these transformati<strong>on</strong> strains<br />

am<strong>on</strong>g the interacting grains, the internal stresses due to phase transformati<strong>on</strong> also differ<br />

and are calculated via an Eshelby inclusi<strong>on</strong> technique.<br />

γ<br />

The actual transformati<strong>on</strong> strain in grain γ is εij , which is<br />

ε ij<br />

N var<br />

n<br />

∑ εij (9)<br />

γ = f n<br />

n=1<br />

where Nvar is the number of variants (same as N in single crystal case). If all grains have<br />

the same transformati<strong>on</strong> strain, the stress due to the incompatibility of transformati<strong>on</strong><br />

strain is zero. To utilize the Eshelby's soluti<strong>on</strong>, we assume that the difference between the<br />

actual transformati<strong>on</strong> strain of each grain and the average transformati<strong>on</strong> strain is the<br />

“effective transformati<strong>on</strong> strain” or eigenstrain. Thus the effective transformati<strong>on</strong> strain<br />

in grain γ is �ε γ<br />

ij :<br />

7


grain<br />

γ γ 1<br />

n<br />

�εij = εij − ∑εij<br />

N<br />

grain n=<br />

1<br />

where Ngrain is the number of grains. The stress in grain γ due to incompatibility with<br />

other grains can be calculated by using Eshelby's soluti<strong>on</strong>.<br />

8<br />

N<br />

(10)<br />

γ γ γ<br />

� σ = C ( S � ε − � ε )<br />

(11)<br />

ij ijkl klmn mn kl<br />

where Sklmn is the Eshelby tensor of spherical inclusi<strong>on</strong>. The average stress in grain γ is<br />

the summati<strong>on</strong> of the external stress and the stress due to incompatibility, that is<br />

γ γ<br />

Σ = Σ + � σ<br />

(12)<br />

ij ij ij<br />

Note that for the polycrystalline material, each grain is assumed to be isotropic to<br />

simplify calculati<strong>on</strong>s, as the closed form soluti<strong>on</strong>s for the Eshelby tensor require isotropy.<br />

Ongoing work is pursuing modificati<strong>on</strong>s to allow anisotropy in the polycrystalline model.<br />

In all calculati<strong>on</strong>s shown here, a 10-grain specimen will be used to simulate a<br />

polycrystalline material. These grains are randomly oriented to avoid texture effect in the<br />

material. Earlier results show the difference between 10 and 100 grains simulati<strong>on</strong>s to be<br />

minimal for random orientati<strong>on</strong>s (Huang et al., 2000). The <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> polycrystalline<br />

resp<strong>on</strong>se for a single c<strong>on</strong>stitutive point is calculated numerically, modifying the stress<br />

state and temperature according to loading history and iterating to soluti<strong>on</strong> at each step<br />

(see also appendix).<br />

3. MODIFIED MULTIVARIANT MODEL<br />

Temperature induced transformati<strong>on</strong>s, shape memory effect, pseudoelasticity and<br />

ferroelasticity are all accounted for properly by the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model. In additi<strong>on</strong><br />

multiaxial loading cases were also simulated and the results are in good qualitative<br />

agreement with the experimental data. The reader is referred to previous papers showing<br />

these results (Gao et al., 2000; Huang et al., 2000; Huang and Brins<strong>on</strong>, 1998). However,<br />

close examinati<strong>on</strong> of results reveals a few difficulties in the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model:


1.Experimental data indicate that stress induced martensite in a single crystal is<br />

formed by a single habit plane variant spanning the specimen, in c<strong>on</strong>trast to the<br />

Eshelby inclusi<strong>on</strong> approach which assumes many inclusi<strong>on</strong>s in a matrix.<br />

2.The transformati<strong>on</strong> stresses remain a factor of two higher than experiments.<br />

3.Anisotropy of austenite and martensite phases are not accounted for.<br />

4.For polycrystalline materials, the iterati<strong>on</strong>s involved in solving the system of<br />

equati<strong>on</strong>s for each grain become computati<strong>on</strong>ally expensive.<br />

5.The self-accommodated grouping structure currently used is based <strong>on</strong> an 18R<br />

structure and habit plane variants; it is oversimplified for B19′ (NiTi), 3R and 2H<br />

(CuAlNi).<br />

Am<strong>on</strong>g these issues, item 1 presents the fundamental insight leading to the simplified<br />

<str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model. Items 2 and 4 are easily resolved by the simplified <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g><br />

model as shown in the next secti<strong>on</strong>. Item 3 can be easily accommodated in the single<br />

crystal case for the new model, but difficulty remains in the polycrystal case. Although<br />

item 5 has not yet been addressed in either model, it will be easier to resolve in the<br />

simplified model.<br />

Returning to item 1 c<strong>on</strong>sider figure 2a which shows several thermally induced martensitic<br />

habit plane variants in a CuAlNi single crystal. It is obvious that the martensite plates are<br />

not <strong>on</strong>ly large in size but also often span the specimen. This phenomen<strong>on</strong> is also comm<strong>on</strong><br />

for stress induced martensite plates and figures 2b-2d show typical examples. Since the<br />

habit plane is an invariant plane, the stress and strain fields of the austenitic phase around<br />

the martensite plate are not disturbed by the transformati<strong>on</strong> strain of the martensite plate.<br />

Thus the original model formulati<strong>on</strong> in which an interacti<strong>on</strong> energy is approximated by<br />

c<strong>on</strong>sidering each martensite variant group as an inclusi<strong>on</strong> in a many inclusi<strong>on</strong> problem<br />

(see figure 1) is physically inappropriate. In fact in many cases, such as formati<strong>on</strong> of<br />

CuAlNi γ 1 ′ (2H) phase from austenite, the undistorted habit plane indicates very small<br />

interacti<strong>on</strong> energy since the incompatibilities of the transformati<strong>on</strong> strain are<br />

accommodated by the finely twinned microstructure. Thus to modify the multivariant<br />

model, we have taken the simplest possible approach and removed the micromechanics<br />

9


calculati<strong>on</strong> of the interacti<strong>on</strong> energy completely and replaced it by a small, c<strong>on</strong>stant<br />

c<strong>on</strong>tributi<strong>on</strong> to F C , the transformati<strong>on</strong> resistance force. Now, equati<strong>on</strong> (8) reduces to<br />

d n<br />

tr<br />

F f B T T F if f<br />

n<br />

Fext<br />

F F<br />

C n<br />

n<br />

n<br />

n<br />

��������� wall<br />

fric ���<br />

� 0<br />

=− ( − ) + Σ ε + λ − λ ∓ , � > < 0<br />

0<br />

ij ij<br />

This change also allows us to incorporate the anisotropic effect into the single crystal<br />

simulati<strong>on</strong>s as Eshelby tensor calculati<strong>on</strong>s are avoided. The numerical procedure for the<br />

<str<strong>on</strong>g>Simplified</str<strong>on</strong>g> model is exactly same as that for the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model.<br />

A careful look at previous micromechanical model calculati<strong>on</strong>s also reveals<br />

accommodati<strong>on</strong> of the different models to the physically very small magnitude of an<br />

interacti<strong>on</strong> energy when calculati<strong>on</strong>s and comparis<strong>on</strong>s to experimental data are made. A<br />

few relevant models are reviewed here since this issue is not explicitly menti<strong>on</strong>ed in<br />

some of the original papers. In Table 1, several micromechanics models for <str<strong>on</strong>g>SMA</str<strong>on</strong>g><br />

c<strong>on</strong>stitutive resp<strong>on</strong>se are listed, with references and expressi<strong>on</strong>s for the interacti<strong>on</strong> energy<br />

and interacti<strong>on</strong> force used in the model. The last column lists pertinent features of each<br />

model. Note that all of these models originally start with an integral expressi<strong>on</strong> for<br />

interacti<strong>on</strong> energy that is identical to that shown in equati<strong>on</strong> (4) of this paper. From that<br />

comm<strong>on</strong> starting point, different assumpti<strong>on</strong>s are made to enable realistic calculati<strong>on</strong>s.<br />

The interacti<strong>on</strong> force, Fint, the gradient of the interacti<strong>on</strong> energy, is ultimately what is<br />

used in calculating the transformati<strong>on</strong> kinetics (as in equati<strong>on</strong> (8) here).<br />

Due to complexity in the evaluati<strong>on</strong> of interacti<strong>on</strong> energy, Patoor et al. (Patoor et al.,<br />

1994; Patoor et al., 1996) replaced the original integral expressi<strong>on</strong> for Eint with a c<strong>on</strong>stant<br />

interacti<strong>on</strong> matrix H nm , representing the resistance to transformati<strong>on</strong> (see table 1). Two<br />

kinds of interacti<strong>on</strong> are assumed: a weak H 1 corresp<strong>on</strong>ding to self-accommodated<br />

variants, and str<strong>on</strong>g H 2 for n<strong>on</strong>-self-accommodated variants, with typical values provided<br />

in Table 1. Note that due to the use of the interacti<strong>on</strong> matrix, for n<strong>on</strong>-zero values of H 1<br />

and H 2 a strain-hardening effect is seen even in predicti<strong>on</strong>s for single crystal stress-strain<br />

resp<strong>on</strong>se (see figure 9 in (Patoor et al., 1996)). Unfortunately, experimental results <strong>on</strong><br />

single crystal (and polycrystal) materials at varying strain rates (Brins<strong>on</strong> et al., 2002;<br />

10<br />

n<br />

6<br />

(13)


Shaw and Kyriakides, 1995; Shaw and Kyriakides, 1997) have revealed that strain<br />

hardening effects are due predominantly to temporary temperature changes of the<br />

specimen with release of latent heat and therefore depend <strong>on</strong> specimen geometry,<br />

surrounding fluid, and strain rate. The baseline predicti<strong>on</strong> from a c<strong>on</strong>stitutive model for a<br />

material point therefore should not include these strain hardening effects by default.<br />

Thus, <strong>on</strong>ly for very small values of H i will the material predicti<strong>on</strong> be appropriate.<br />

Referring to Table 1, the model by Lexcellent and co-workers (Goo and Lexcellent, 1997;<br />

Raniecki et al., 1992; Vivet and Lexcellent, 1998) and the model by Lu and Weng (Lu<br />

and Weng, 1997) start with the same form for the interacti<strong>on</strong> energy (the integral in<br />

equati<strong>on</strong> (4)) and both approximate the integral as a sum by c<strong>on</strong>sidering the martensitic<br />

plates to be transforming inclusi<strong>on</strong>s. The energy between martensitic plates and the<br />

austenitic matrix for the two models is identical. Lexcellent and co-workers have an<br />

additi<strong>on</strong>al term accounting for interacti<strong>on</strong> between martensitic variants in the general<br />

expressi<strong>on</strong>, but in applicati<strong>on</strong> they assume <strong>on</strong>ly a single active variant such that this<br />

sec<strong>on</strong>d term drops out. Lu and Weng also in applicati<strong>on</strong> of the model assume <strong>on</strong>ly a<br />

single active variant and then perform the calculati<strong>on</strong>s based <strong>on</strong> that variant al<strong>on</strong>e. The<br />

original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model has a very similar expressi<strong>on</strong> for the interacti<strong>on</strong> energy,<br />

being also based <strong>on</strong> the transforming inclusi<strong>on</strong> c<strong>on</strong>cept, but the exact form differs due to<br />

the grouping of variants into self-accommodated clusters in the calculati<strong>on</strong>.<br />

In the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model, this grouping of self-accommodated variants<br />

dramatically decreased the interacti<strong>on</strong> energy over what was calculated when c<strong>on</strong>sidering<br />

each inclusi<strong>on</strong> to be a single variant. Since both the Lexcellent and Weng models take the<br />

inclusi<strong>on</strong> to be a single variant, there remains yet another difference to account for the<br />

different magnitudes of the interacti<strong>on</strong> force, seen in the 4 th column of Table 1. The form<br />

and magnitudes of the Eshelby tensor, S, in each of these models is based <strong>on</strong> the aspect<br />

ratio of the penny-shaped inclusi<strong>on</strong>. In the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model, the aspect ratio of<br />

the inclusi<strong>on</strong>/variant was selected to be approximately 0.01, based <strong>on</strong> experimental<br />

observati<strong>on</strong>s (Saburi and Wayman, 1979). The images in Figure 2b-2d also support this<br />

ratio for width to length of the martensitic plates formed. Lu and Weng, however,<br />

11


explicitly take α→0 which enables a much smaller final interacti<strong>on</strong> force; while not<br />

stated in the paper, the aspect ratio for Vivet and Lexcellent can be calculated * to be<br />

approximately 0.001 in order to obtain the value for Ws cited given the other material<br />

c<strong>on</strong>stants provided.<br />

Taken together, a trend is obvious: starting from the basic c<strong>on</strong>cept of treating the<br />

martensitic variants as transforming inclusi<strong>on</strong>s in order to calculate an interacti<strong>on</strong> energy<br />

(a barrier to transformati<strong>on</strong>), each of the micromechanical models has developed a<br />

different “coping mechanism” to enable realistic calculati<strong>on</strong>s when compared with<br />

experimental data. Furthermore, when the physical nature of the martensitic plate<br />

formati<strong>on</strong>, as illustrated in Figure 2, is viewed, it is obvious that the c<strong>on</strong>cept of a many-<br />

inclusi<strong>on</strong> problem (see figure 1) where each inclusi<strong>on</strong> is surrounded by matrix material<br />

and embedded in an infinite medium of additi<strong>on</strong>al inclusi<strong>on</strong>s and matrix material, is<br />

incorrect. Since this latter criteri<strong>on</strong> is the key to the Eshelby micromechanics approach,<br />

we propose that discarding the interacti<strong>on</strong> energy calculati<strong>on</strong> by this micromechanics<br />

method, as we have d<strong>on</strong>e in equati<strong>on</strong> (13), is perhaps the most realistic strategy. It also<br />

has additi<strong>on</strong>al advantages of practically improving the computati<strong>on</strong> time required by<br />

removing a level of iterati<strong>on</strong> from the problem.<br />

Finally, it should be menti<strong>on</strong>ed that n<strong>on</strong>-invariant plane martensitic transformati<strong>on</strong> has<br />

been found by Sun et al. (Sun et al., 1999; Sun et al., 1997) when austenite transforms to<br />

β 1 ′ (pseudoelasticity) for a single CuAlNi crystal. In this case, there is a distorti<strong>on</strong> of the<br />

austenitic strain field near the A-M interface while the stress and strain fields away from<br />

the interface are uniform. These results indicate that a slightly larger, but still small<br />

interacti<strong>on</strong> energy would be suitable for the β1→ β 1 ′ transformati<strong>on</strong>. Appropriate means<br />

to calculate an interacti<strong>on</strong> energy in this n<strong>on</strong>-invariant plane case will be c<strong>on</strong>sidered in<br />

* p p p<br />

Using the parameters µ=12Gpa, ν=0.3, ε 11=<br />

- 0.1174%, ε 12<br />

=ε21 =6.646%, and zeros<br />

for other transformati<strong>on</strong> strain comp<strong>on</strong>ents, the authors calculated the value of<br />

p p<br />

LI ( − S)<br />

ε ε . It is 4.16 and 0.416 Mpa for aspect ratio of 0.01 and 0.001 respectively.<br />

12


the future. The same experiments also indicate that the A/M interface for the n<strong>on</strong>-<br />

invariant plane is macroscopically a straight line with an angle very close to the<br />

predicti<strong>on</strong> by phenomenological theory of martensitic transformati<strong>on</strong>. Therefore, in our<br />

calculati<strong>on</strong>s we will c<strong>on</strong>tinue to use the phenomenological theory of martensitic<br />

transformati<strong>on</strong> for the values of transformati<strong>on</strong> strain, equati<strong>on</strong> (5), and we will ignore<br />

the distorti<strong>on</strong> near the interface in the calculati<strong>on</strong> for a β to β 1 ′ (pseudoelasticity) for a<br />

single CuAlNi.<br />

4. RESULTS<br />

In this secti<strong>on</strong>, the shape memory resp<strong>on</strong>se under pure tensi<strong>on</strong> for the <str<strong>on</strong>g>Simplified</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g><br />

and the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model will be compared first. Then the predicted orientati<strong>on</strong><br />

dependence of both models will be compared to experimental results by Shield (Shield,<br />

1995) <strong>on</strong> a Cu-13.95Al-3.93Ni (wt%) single crystal. Finally, triaxial loading simulati<strong>on</strong><br />

results predicted by both models are compared to experimental results by Gall et al. (Gall<br />

et al., 1998) for a polycrystalline Cu71Zn25Al4 (wt%) material with a volume change of -<br />

0.3%.<br />

Since not all parameters necessary for the simulati<strong>on</strong> were readily available for the<br />

experimental materials, some parameters (e.g. lattice parameters necessary to calculate �n,<br />

�m and g for equati<strong>on</strong> (5)) were found in the literature for materials with very similar<br />

compositi<strong>on</strong>s. Table 2a c<strong>on</strong>tains parameters known for the materials used in the<br />

experiments and table 2b lists other necessary parameters for similar materials attained<br />

elsewhere in the literature. Although some parameters (e.g. B and F C obtained/used in<br />

equati<strong>on</strong> (13)) are tailored from the experimental materials, they are listed in table 2b<br />

because they are not part of the experimental results. In the calculati<strong>on</strong> of �n, �m and g for<br />

equati<strong>on</strong> (5) for Cu-14.0Al-4.2Ni (wt%), martensite phase lattice parameters were from<br />

Otsuka and Shimizu (Otsuka and Shimizu, 1981). Since no parent phase parameter is<br />

�<br />

available, a 0 = 5.836 A<br />

is used (Otsuka and Shimizu, 1974).<br />

13


First examining the different pseudoelasticity resp<strong>on</strong>se predicted by both models under<br />

uniaxial tensi<strong>on</strong>, Figure 3 shows the comparis<strong>on</strong> for single crystal Cu70.17Zn25.63Al4.2<br />

(wt.%). The transformati<strong>on</strong> stress predicted by the simplified <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model is much<br />

lower than that predicted by the original model. Note that experimental results <strong>on</strong> single<br />

crystal Cu70.17Zn25.63Al4.2 (wt%) alloy by Goo and Lexcellent (Goo and Lexcellent, 1997),<br />

and Lexcellent et al. (Lexcellent et al., 1996) are used to choose the parameters B and FC<br />

used in the simulati<strong>on</strong> and that the same values are used for both models; these<br />

parameters can be used to tune the model to the experiment, but even values of zero,<br />

which are not physically reas<strong>on</strong>able (and would result in loss of hysteresis and<br />

temperature effects) could not bring the stress predicti<strong>on</strong>s of the original multivariant<br />

model in line with experimental results. For example, setting F c to 0 <strong>on</strong>ly reduces the<br />

critical transformati<strong>on</strong> stress by approximately 30Mpa. Figure 4 shows the comparis<strong>on</strong><br />

for the polycrystal pseudoelasticity effect <strong>on</strong> the same material. Again, the transformati<strong>on</strong><br />

stress has dropped substantially to observed experimental values. In additi<strong>on</strong>, since the<br />

micromechanical interacti<strong>on</strong> energy calculati<strong>on</strong> is removed in the simplified model, the<br />

computati<strong>on</strong>al time required to solve the system of equati<strong>on</strong>s has dropped by nearly two<br />

orders of magnitude. For example, with n<strong>on</strong>-optimized code <strong>on</strong> a HP 712/80 workstati<strong>on</strong>,<br />

a single crystal case such as Figure 3 completes in less than a minute, while a polycrystal<br />

case such as in Figure 4 takes <strong>on</strong> the order of an hour. The decreased computati<strong>on</strong> time<br />

addresses c<strong>on</strong>cern 4 in the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model and renders the model suitable for further<br />

practical use.<br />

One major advantage to the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> has been its good qualitative agreement<br />

to more complicated experimental results. Thus, here we now examine orientati<strong>on</strong><br />

dependence in single crystals, then a multiaxial loading case for polycrystals. Figure 5<br />

shows results for uniaxial tensi<strong>on</strong> experiments performed <strong>on</strong> three single crystal<br />

specimens of Cu-13.95Al-3.93Ni (wt%) by Shield (Shield, 1995). Specimen T1 had a<br />

tensile axis of [2.43 1 0]; T2 was oriented 15 degree from [-1 -1 1] directi<strong>on</strong> and had a [-1<br />

-1 1.73] tensile axis directi<strong>on</strong>; T3 had the [-1 -1 1] directi<strong>on</strong> as its tensile axis.<br />

14


There are several notable features of these results. First, we see that the initial moduli of<br />

all the specimens differ, as do the transformati<strong>on</strong> stresses. Specimen T1 has the lowest<br />

initial modulus, T3 the largest <strong>on</strong>e, and T2 an intermediate modulus. The difference in<br />

the initial modulus is c<strong>on</strong>sistent with the behavior of a cubic crystal (Nye, 1985; Shield,<br />

1995) where the maximum and minimum Young’s moduli are in the and directi<strong>on</strong>s, respectively. The trend for magnitude of the transformati<strong>on</strong> stress is similar<br />

to the trend in the initial moduli (Shield, 1995) such that specimen T1 ([2.43 1 0]<br />

directi<strong>on</strong>) has the lowest transformati<strong>on</strong> stress and specimen T3 ([-1 -1 1] directi<strong>on</strong>) has<br />

the largest transformati<strong>on</strong> stress.<br />

In order to verify the applicability of the single crystal model, we calculated three tensile<br />

loading and unloading cases with identical orientati<strong>on</strong>s to those in Shield’s experiments<br />

for the CuAlNi material listed in table 2. In the simulati<strong>on</strong> the forward equilibrium<br />

temperature T0 is approximated by (MS+Af)/2 (≈14 °C) while for the reverse<br />

transformati<strong>on</strong> equilibrium temperature T0′ is taken as (Mf+AS)/2 (≈ - 5 °C). T0 is the<br />

metastable equilibrium temperature in transformati<strong>on</strong> from austenite to martensite in<br />

P→ M<br />

which the difference in chemical free energy of the two phases, ∆Gc , is zero<br />

(Kaufman and Cohen, 1958). During reverse transformati<strong>on</strong>, the n<strong>on</strong>-chemical free<br />

M→ P<br />

energy accumulated during the A→M transformati<strong>on</strong> ( ∆Gnc , strain energy, surface<br />

energy, etc) plays a role as a driving force for the transformati<strong>on</strong>. The inverse role of this<br />

n<strong>on</strong>-chemical energy shifts the metastable equilibrium temperature from T0 to T0′ (T<strong>on</strong>g<br />

M→P M→P and Wayman, 1974). Thus T0′ is where ∆G + ∆G<br />

= 0 and is always less than AS.<br />

c<br />

The result predicted by the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> is shown in Figure 6, and that by<br />

the simplified model is shown in Figure 7.<br />

It is observed that both models predict the proper sequence of transformati<strong>on</strong> stresses<br />

compared to the experiments with T3 [-1 -1 1] having the highest transformati<strong>on</strong> stress,<br />

and T1 [2.43 1 0] the lowest. Moreover, the experimental transformati<strong>on</strong> stresses have a<br />

maximum to minimum ratio of 4.9, while the model predicti<strong>on</strong>s provide 4.0 and 4.2 for<br />

15<br />

nc


the original model and simplified models respectively. These results show that both<br />

models are c<strong>on</strong>sistent with the main trends of stress-strain relati<strong>on</strong>ship in <str<strong>on</strong>g>SMA</str<strong>on</strong>g>s.<br />

However, several significant discrepancies can be seen between the original model<br />

predicti<strong>on</strong>s (Figure 6) and the experimental data (Figure 5). The magnitude of the<br />

transformati<strong>on</strong> stresses are nearly three times that of the experimental results; two of the<br />

orientati<strong>on</strong>s are unable to recover the strain pseudoelastically up<strong>on</strong> unloading; and the<br />

moduli differences are clearly missing due to the isotropy assumpti<strong>on</strong>. With a move to<br />

the simplified model (Figure 7), the results are strikingly improved and very close to the<br />

experimental results in terms of transformati<strong>on</strong> stresses, strains and moduli. Note that the<br />

parameters F c and B for the model were determined from the experimental result for T3<br />

and there are no other free parameters in the model. Using values from the literature for<br />

the anisotropic elastic c<strong>on</strong>stants (see Table 2b), the ratio of maximum to minimum<br />

moduli for the three orientati<strong>on</strong>s shown is predicted to be 6.9, while the experimental<br />

value is 5.9.<br />

It should be menti<strong>on</strong>ed that the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model often predicted appearance of<br />

multiple variants: for the simulati<strong>on</strong>s here the variants active in cases T1, T2 and T3 were<br />

(17 & 18), (6 & 15) and (6, 8, 13, 15, 18, & 19) respectively. Experimentally however, a<br />

final product of <strong>on</strong>e variant is expected in most cases, although in directi<strong>on</strong>s with special<br />

symmetry several variants are sometimes seen (Lexcellent et al., 1996). With the<br />

simplified <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model, the variants predicted in cases T1, T2 and T3 are (17 &<br />

18), (6) and (8) respectively. Since all the materials simulated have 18R structure, the<br />

variant order is same as Saburi and Wayman (Saburi et al., 1980) except here we identify<br />

1(+) as 1, 1(-) as 2, …, and 6'(-) as 24. Angles between the tensi<strong>on</strong> directi<strong>on</strong> and the<br />

intersecti<strong>on</strong> directi<strong>on</strong> of the habit plane and the sample surface are also measured in the<br />

experiment and compared to predicti<strong>on</strong> based <strong>on</strong> Type I or Type II twinning for a 2H<br />

system (Shield, 1995). Since the material is at a temperature 17° higher than the Af<br />

(23°C), the CuAlNi material should transform between the parent and a 18R martensite<br />

phase which does not have twinning substructure (Otsuka and Shimizu, 1981; Sun et al.,<br />

1999). Therefore using lattice parameters (table 2b), we calculated the habit plane<br />

16


normal, invariant plane shear directi<strong>on</strong>, magnitude of the shear and transformati<strong>on</strong> strains<br />

for the 18R system, shown in table 2b. The predicted angles between the tensi<strong>on</strong> directi<strong>on</strong><br />

and the intersecti<strong>on</strong> directi<strong>on</strong> of the habit plane and the sample surface are shown in table<br />

3. Agreement between predicti<strong>on</strong> and data is quite good with the excepti<strong>on</strong> of the highly<br />

symmetric directi<strong>on</strong> T3, where the transformati<strong>on</strong> strain levels of several potential<br />

variants are very close. The variants with largest transformati<strong>on</strong> strain will show up in the<br />

single crystal calculati<strong>on</strong> since we are using a maximum work criteri<strong>on</strong> and all variants<br />

are under same stress state. The model predicts variant 8, with a habit plane angle very<br />

different from the experimentally observed value, note, however, that a minor change in<br />

lattice parameters can cause selecti<strong>on</strong>s of variants 6,15,18 or19 (with angles very close to<br />

the experiments). For such a symmetric loading directi<strong>on</strong>, it is very likely that a small<br />

amount of misalignment of or imperfecti<strong>on</strong> in the specimen may change the observed<br />

variant from <strong>on</strong>e to another, or high resoluti<strong>on</strong> microscopy during loading would reveal<br />

presence of several, nearly equivalent variants.<br />

There are <strong>on</strong>ly a few triaxial loading experiments which have been performed to<br />

investigate the effect of three dimensi<strong>on</strong>al stress states <strong>on</strong> martensitic transformati<strong>on</strong> in<br />

<str<strong>on</strong>g>SMA</str<strong>on</strong>g>s. Jacobus et al. (Jacobus et al., 1996) investigated the effects of triaxial loading in a<br />

polycrystalline Ni-Ti <str<strong>on</strong>g>SMA</str<strong>on</strong>g>. Lim and McDowell (Lim and McDowell, 1999) investigated<br />

the resp<strong>on</strong>se of a Ni-Ti <str<strong>on</strong>g>SMA</str<strong>on</strong>g> specimen under axial-torsi<strong>on</strong>al proporti<strong>on</strong>al and n<strong>on</strong>-<br />

proporti<strong>on</strong>al loading. Gall et al. (Gall et al., 1998) investigated triaxial loading in a<br />

polycrystalline Cu71Zn25Al4 (wt%, ∆V/V ≈ -0.3%) <str<strong>on</strong>g>SMA</str<strong>on</strong>g> (shown in figures 8 and 9).<br />

Under uniaxial loading, it was found that the compressive stress level required to<br />

macroscopically trigger the transformati<strong>on</strong> was 34% larger than the required tensile<br />

stress. The triaxial tests produced effective stress-strain curves with critical effective<br />

transformati<strong>on</strong> stress levels in between the tensile and compressive results.<br />

The effective stress-strain plots at a temperature above Af under a loading rate of dε/dt =<br />

10 -4 s are shown in figure 8. Curves for the effective stress at the <strong>on</strong>set of transformati<strong>on</strong><br />

vs. the hydrostatic stress at the <strong>on</strong>set of transformati<strong>on</strong> are shown in figure 9. The<br />

effective stress and effective strain are calculated by:<br />

17


2<br />

2<br />

2<br />

σ eff = [( σ11 − σ 22)<br />

+ ( σ11 − σ 33)<br />

+ ( σ 22 −σ<br />

33)<br />

]<br />

2<br />

/<br />

2<br />

2<br />

2<br />

εeff = [( ε11 − ε22) + ( ε11 − ε33) + ( ε22 −ε33)<br />

]<br />

3<br />

/<br />

where σnn and εnn are the n th principal stress and strain respectively. For the tests in<br />

Figure 8 and 9 the applied stress state, the effective stress and the relative hydrostatic<br />

pressure are shown in table 4.<br />

We simulated the <strong>on</strong>e dimensi<strong>on</strong>al and three dimensi<strong>on</strong>al loading paths listed in table 4<br />

using both <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> models. The effective stress-strain curves similar to figure 8<br />

predicted by the simplified model are plotted in figure 10. The effective stress at the <strong>on</strong>set<br />

of transformati<strong>on</strong> vs. the hydrostatic stress at the <strong>on</strong>set of transformati<strong>on</strong> similar to figure<br />

9 predicted by both models are shown in figure 11. One sees that the new model captures<br />

the magnitude of transformati<strong>on</strong> stress, and the trend of the effective transformati<strong>on</strong><br />

stress with increasing of hydrostatic pressure. The counter-intuitive experimental results<br />

of increasing transformati<strong>on</strong> stress with increasing hydrostatic pressure for a material<br />

with negative volume change are also well captured by the simplified model. The<br />

compressive stress level required to macroscopically trigger the transformati<strong>on</strong> was 16%<br />

larger than the required tensile stress in the simulati<strong>on</strong>.<br />

4. DISCUSSION<br />

The <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model is <strong>on</strong>e of the few that has been tested against complex<br />

thermomechanical loading, and incorporates both variant level predicti<strong>on</strong>s as well as<br />

c<strong>on</strong>tinuum level resp<strong>on</strong>se. As illustrated in the previous secti<strong>on</strong>, the simplified and<br />

improved model provides excellent qualitative and quantitative agreement with<br />

experimental data for sophisticated loading. Of the original list of 5 c<strong>on</strong>cerns with respect<br />

to the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model, the improved model has addressed the first four. 1) By<br />

replacing the micromechanical interacti<strong>on</strong> energy calculati<strong>on</strong>, which assumed formati<strong>on</strong><br />

of many different martensitic plates (inclusi<strong>on</strong>s), with a small c<strong>on</strong>stant interacti<strong>on</strong> force<br />

included in F C , the model now more closely resembles experimental observati<strong>on</strong>s. 2) This<br />

18<br />

2 1 2<br />

2 1 2<br />

(14)<br />

(15)


same change then removed a significant barrier to transformati<strong>on</strong> and thereby lowered<br />

transformati<strong>on</strong> stresses to appropriate magnitudes. 3) Anisotropy can be fully accounted<br />

for in the single crystal model, although work is still <strong>on</strong>going to bring that aspect into the<br />

polycrystalline calculati<strong>on</strong>s; this change would be essential to properly modeling textured<br />

materials. 4) Removing the interacti<strong>on</strong> energy also eliminated <strong>on</strong>e level of iterati<strong>on</strong>s in<br />

the calculati<strong>on</strong>s and therefore significantly decreased computati<strong>on</strong> time.<br />

C<strong>on</strong>cern 5 was related to the self-accommodated grouping structure and the applicability<br />

of the model to habit plane variants <strong>on</strong>ly. While eliminating the interacti<strong>on</strong> energy<br />

removes much of the explicit grouping structure from the model, the transformati<strong>on</strong><br />

strains (equati<strong>on</strong> (5)) are still calculated based <strong>on</strong> crystallographic parameters for habit<br />

plane variants. This equati<strong>on</strong> is appropriate for materials such as CuZnAl with stacking<br />

faults acting as the invariant plane shear prior to de-faulting (Miyazaki et al., 1984). It is<br />

also appropriate for materials such as γ 1 ′ CuAlNi, and B19' NiTi with internally twinned<br />

structure before detwinning of corresp<strong>on</strong>dence variants occurs. The c<strong>on</strong>versi<strong>on</strong> process<br />

for a thermally induced martensitic structure of γ 1 ′ type is indicated after work by Saburi<br />

(Saburi and Nenno, 1981) in Figure 12. Note that for clarity figure 12 ignores some finer<br />

levels of twinning and c<strong>on</strong>versi<strong>on</strong> (intermediate steps) in the two middle steps. Similar<br />

studies in microscale have not been performed for the stress-induced c<strong>on</strong>versi<strong>on</strong> process<br />

when starting from austenite. However, it is now comm<strong>on</strong>ly accepted (Horikawa et al.,<br />

1988; Leclercq and Lexcellent, 1996) that for such a material under loading, a distinct<br />

habit plane forms first separating austenite from an internally twinned or faulted<br />

martensitic structure (the habit plane variant); detwinning or de-faulting of the habit plane<br />

variant then occurs after a higher critical load is reached resulting in a single<br />

corresp<strong>on</strong>dence variant product. With further increases in load, reorientati<strong>on</strong> or<br />

c<strong>on</strong>versi<strong>on</strong> am<strong>on</strong>g corresp<strong>on</strong>dence variants can happen (as shown in the reorienting step<br />

of figure 12 for the thermally induced case). Ongoing work to address these issues is<br />

examining the stress induced detwinning/de-faulting and c<strong>on</strong>versi<strong>on</strong> process both<br />

experimentally by in situ SEM and numerically by including a subgrouping scheme for<br />

corresp<strong>on</strong>dence variants and direct calculati<strong>on</strong> of their transformati<strong>on</strong> strains, in an<br />

19


approach similar to that pursued recently by Govindjee and Miehe (Govindjee and Miehe,<br />

2001). Results will be presented in future work (Gao and Brins<strong>on</strong>, 2002).<br />

Finally, in this paper we have simply removed the micromechanics calculati<strong>on</strong> of the<br />

interacti<strong>on</strong> energy entirely, lumping any energy barrier due to incompatibilities of the<br />

transformati<strong>on</strong> strain with respect to the surrounding material into a c<strong>on</strong>stant resistance<br />

force, F C . It is seen experimentally that in most cases, the stress-induced martensite forms<br />

with a sharp habit plane with no distorti<strong>on</strong> due to a finely twinned substructure in the<br />

habit plane variant. Hence the incompatibility of the stress-induced variant is indeed quite<br />

small and the assumpti<strong>on</strong> here is justified. However, as menti<strong>on</strong>ed earlier, some recent<br />

results have shown that a n<strong>on</strong>-invariant plane relative to the austenite (Sun et al., 1999;<br />

Sun et al., 1997) is possible, accompanied by a finite amount of distorti<strong>on</strong> of the material<br />

around the interface. In such cases, (for example, β to β 1 ′ in CuAlNi alloy) a different<br />

calculati<strong>on</strong> of the interacti<strong>on</strong> energy due to the distorted regi<strong>on</strong> around the interface<br />

would be appropriate. Future work should investigate this possibility and its influence <strong>on</strong><br />

<str<strong>on</strong>g>SMA</str<strong>on</strong>g> models.<br />

5. CONCLUSION<br />

Here we have presented a simplified <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model for predicti<strong>on</strong> of shape memory<br />

alloy resp<strong>on</strong>se to thermomechanical loading. While the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model<br />

performed very well providing qualitative agreement with complex experiments,<br />

problems with transformati<strong>on</strong> stress magnitude, material anisotropy and exact variant<br />

predicti<strong>on</strong> remained. In additi<strong>on</strong>, due to the hierarchical iterative numerical approach,<br />

calculati<strong>on</strong> times for polycrystalline specimen resp<strong>on</strong>se were prohibitive. <str<strong>on</strong>g>Based</str<strong>on</strong>g> <strong>on</strong><br />

experimental results showing that both thermally and stress induced martensitic variants<br />

tend to form in large plates nearly spanning the single crystal (or grain), the original<br />

calculati<strong>on</strong> of an interacti<strong>on</strong> energy, in which variants were represented as inclusi<strong>on</strong>s in a<br />

many inclusi<strong>on</strong> problem, was removed. It was also dem<strong>on</strong>strated here that similar<br />

micromechanics approaches to <str<strong>on</strong>g>SMA</str<strong>on</strong>g> modeling have ultimately had to make significant<br />

20


compromises in calculati<strong>on</strong>s of the interacti<strong>on</strong> energy in order to achieve a small enough<br />

value to provide realistic predicti<strong>on</strong>s. By removing the micromechanical calculati<strong>on</strong> of<br />

interacti<strong>on</strong> energy, we achieved a model that is more physically based yet retains the<br />

essential crystallographic and thermodynamic framework and still utilizes<br />

micromechanics where appropriate in the multi-grain interacti<strong>on</strong>s for polycrystalline<br />

samples. The simplified model performs extremely well for quantitative comparis<strong>on</strong>s to<br />

experimental data <strong>on</strong> n<strong>on</strong>-simple loadings. In additi<strong>on</strong>, the computati<strong>on</strong> time is rapid,<br />

making the model feasible for higher level studies.<br />

Some further improvements to the model are under investigati<strong>on</strong>, most notably work to<br />

incorporate corresp<strong>on</strong>dence variant subunits and an effort to include anisotropy into the<br />

polycrystalline calculati<strong>on</strong>s. With these additi<strong>on</strong>al improvements the <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model<br />

will be a powerful tool for both calculati<strong>on</strong> of <str<strong>on</strong>g>SMA</str<strong>on</strong>g> c<strong>on</strong>stitutive resp<strong>on</strong>se and material<br />

level resp<strong>on</strong>se and will provide a c<strong>on</strong>venient test bed for simulating and understanding<br />

<str<strong>on</strong>g>SMA</str<strong>on</strong>g> resp<strong>on</strong>se to complex thermomechanical loading.<br />

APPENDIX<br />

The <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model is solved numerically for a single c<strong>on</strong>stitutive point given some<br />

initial c<strong>on</strong>diti<strong>on</strong>s (the initial volume fracti<strong>on</strong>s of all variants, typically these start from<br />

zeros). A brief descripti<strong>on</strong> of the algorithm is given here and the reader is referred to<br />

(Huang, 1997; Huang et al., 2000; Huang and Brins<strong>on</strong>, 1998) for more details.<br />

Single Crystal Algorithm:<br />

For each step of loading history (Temperature and Stress state)<br />

i i i i i<br />

1. Find net force for each variant Fnet = Fext + Fint + Ffric + Fwall<br />

�<br />

�<br />

2. If | Fnet | ≠ 0 , assign an initial trial step size h proporti<strong>on</strong>al to 1/| Fnet |<br />

�<br />

3. While | Fnet | ≠ 0 do<br />

Calculate ∆f i (volume fracti<strong>on</strong> change) and Err i (estimated error) by taking a Cash-<br />

Karp Runge-Kutta step<br />

21


MaxErr = max( Err i )<br />

If MaxErr > Tolerance<br />

Else<br />

Decrease h<br />

Increase h<br />

Advance f i by ∆f i<br />

i i i i i<br />

Find new net force for each variant F = F + F + F + F + F<br />

End {If}<br />

End {While}<br />

Polycrystal Algorithm:<br />

Each grain is treated as a single crystal and the single crystal algorithm above is used. In<br />

each loading step, the program checks each grain for n<strong>on</strong>-balanced driving forces (i.e.,<br />

�<br />

| Fnet | ≠ 0 ); if it finds a n<strong>on</strong>-balanced grain, it calculates the new volume fracti<strong>on</strong>s of each<br />

variant in that grain by <strong>on</strong>e single or several Runge-Kutta steps. The key point is that<br />

before iterati<strong>on</strong> with other grains the total volume fracti<strong>on</strong> change inside a grain should<br />

not exceed certain amount (e.g., 10 -5 ). After all the grains are checked and the new<br />

martensite fracti<strong>on</strong>s for n<strong>on</strong>-balanced grains are obtained, the program uses the new<br />

martensite fracti<strong>on</strong> of each variant in each grain to calculate the stresses in each grain by<br />

using (9) -.(12) Then the program checks all grains, advances the n<strong>on</strong>-balanced grains for<br />

<strong>on</strong>e single or several Runge-Kutta steps, and then recalculates the stresses in each grain<br />

again, repeating until all grains are balanced.<br />

REFERENCES<br />

Auricchio, F. and R.L. Taylor. 1997. "Shape-memory alloys: <str<strong>on</strong>g>Model</str<strong>on</strong>g>ing and numerical<br />

simulati<strong>on</strong>s of the finite-strain superplastic behavior", Comput. Methods Appl. Mech.<br />

Eng., 143(1-2): 175-194.<br />

Boyd, J.G. and D.C. Lagoudas. 1994. "Thermomechanical resp<strong>on</strong>se of shape memory<br />

composites", J. Intell. Mater. Syst. Struct., 5: 333-346.<br />

22<br />

net<br />

ext<br />

int<br />

fric<br />

wall<br />

0<br />

wall


Boyd, J.G. and D.C. Lagoudas. 1996. "A thermodynamical c<strong>on</strong>stitutive model for shape<br />

memory materials. Part I. The m<strong>on</strong>olithic shape memory alloy", Int. J. Plast., 12(6):<br />

805-842.<br />

Brins<strong>on</strong>, L.C., I. Schmidt and R. Lammering. 2001. "Micro and macromechanical<br />

investigati<strong>on</strong>s of transformati<strong>on</strong> behavior of a polycrystalline NiTi shape memory<br />

alloy using in situ optical microscopy", Acta Metall. Mater., submitted.<br />

Brins<strong>on</strong>, L.C., I. Schmidt and R. Lammering. 2002. "Micro and macromechanical<br />

investigati<strong>on</strong>s of CuAlNi single crystal and CuAlMnZn polycrystalline shape memory<br />

alloys", J. Intell. Mater. Syst. Struct., to appear.<br />

Budiansky, B. and T.T. Wu. 1962. "Theoretical predicti<strong>on</strong> of plastic strains of<br />

polycrystals", Fourth U. S. Nati<strong>on</strong>al C<strong>on</strong>gress of Applied Mechanics, Berkeley, CA,<br />

Proceedings of the Fourth U. S. Nati<strong>on</strong>al C<strong>on</strong>gress of Applied Mechanics, R.M.<br />

Rosenberg, ed. New York: ASME, pp.1175-1185.<br />

Gall, K., H. Sehitoglu, H.J. Maier and K. Jacobus. 1998. "Stress-induced martensitic<br />

phase transformati<strong>on</strong>s in polycrystalline CuZnAl shape memory alloys under different<br />

stress states", Metall. Trans. A, 29A(Mar.): 765-773.<br />

Gao, X. and L.C. Brins<strong>on</strong>. 2002. "A 3-D two-tier multivariant <str<strong>on</strong>g>SMA</str<strong>on</strong>g> model based <strong>on</strong><br />

hierachical structural characteristics of martensitic transformati<strong>on</strong>",, manuscript in<br />

preparati<strong>on</strong>.<br />

Gao, X., M. Huang and L.C. Brins<strong>on</strong>. 2000. "A multivariant micromechanical model for<br />

<str<strong>on</strong>g>SMA</str<strong>on</strong>g>s Part 1: Crystallographic issues for single crystal model", Int. J. Plast., 16(10-<br />

11): 1345-1369.<br />

Goo, B.C. and C. Lexcellent. 1997. "Micromechanics-based modeling of two-way shape<br />

memory effect of a single crystalline shape memory alloy", Acta Metall. Mater., 45(2):<br />

727-737.<br />

Govindjee, S. and C. Miehe. 2001. "A multi-variant martensitic phase transformati<strong>on</strong><br />

model: formulati<strong>on</strong> and numerical implementati<strong>on</strong>", Comput. Methods Appl. Mech.<br />

Eng., 191: 215-38.<br />

Graesser, E.J. and F.A. Cozzarelli. 1994. "A proposed three-dimensi<strong>on</strong>al c<strong>on</strong>stitutive<br />

model for shape memory alloys", J. Intell. Mater. Syst. Struct., 5(Jan.): 78-89.<br />

Hellwege, K.H. and A.M. Hellwege. 1984. "Elastic, Piezoelectric, Pyrelectric,<br />

Peizooptic, Electrooptic C<strong>on</strong>stants, and N<strong>on</strong>linear Dielectric Susceptibilities of<br />

Crystals." Landolt-Bornsterin: Numerical Data and Functi<strong>on</strong>al Relati<strong>on</strong>ships in<br />

Science and Technology: New Series: Group III: Crystal and Solid State Physics, K.H.<br />

Hellwge and O. Magdelung, eds., New York, NY: Springer-Verlag, Vol:18: 5-5.<br />

Horikawa, H., S. Ichinose, K. Morii, S. Miyazaki and K. Otsuka. 1988. "Orientati<strong>on</strong><br />

dependence of β1 to β'1 stressed-induced martensitic transformati<strong>on</strong> in a Cu-Al-Ni<br />

alloy", Metall. Trans. A, 19A: 915-923.<br />

Huang, M. 1997. "A <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> for Shape Memory Alloys", Ph. D.,<br />

Northwestern University, Evanst<strong>on</strong>, IL.<br />

23


Huang, M., X. Gao and L.C. Brins<strong>on</strong>. 2000. "A multivariant micromechanical model for<br />

<str<strong>on</strong>g>SMA</str<strong>on</strong>g>s Part 2: Polycrystal model", Int. J. Plast., 16(10-11): 1371-1390.<br />

Huang, M.S. and L.C. Brins<strong>on</strong>. 1998. "A multivariant model for single crystal shape<br />

memory alloy behavior", J. Mech. Phys. Solids, 46(8): 1379-1409.<br />

Jacobus, K., H. Sehitoglu and M. Balzer. 1996. "Effect of stress state <strong>on</strong> the stressinduced<br />

martensitic phase transformati<strong>on</strong> in polycrystalline Ni-Ti alloy", Metall.<br />

Trans. A, 27A(Oct.): 3066-3073.<br />

Kaufman, L. and M. Cohen. 1958. "Thermodynamics and kinetics of martensitic<br />

transformati<strong>on</strong>s", Prog. Met. Phys., 7: 165-246.<br />

Kröner, E. 1961. "Zur plastischen verformung des vielkristalls", Acta Metall., 9: 155-161.<br />

Lagoudas, D.C., Z. Bo and M.A. Qidwai. 1996. "A unified thermodynamic c<strong>on</strong>stitutive<br />

model for <str<strong>on</strong>g>SMA</str<strong>on</strong>g> and finite element analysis of active metal matrix composites", Mech.<br />

Compos. Mater. Struct., 3: 153-179.<br />

Leclercq, S. and C. Lexcellent. 1996. "A General macroscopic descripti<strong>on</strong> of the<br />

thermomechanical behaviors of shape memory alloys", J. Mech. Phys. Solids, 44(6):<br />

953-980.<br />

Lexcellent, C., B.C. Goo, Q.P. Sun and J. Bernardini. 1996. "Characterizati<strong>on</strong>,<br />

thermomechanical behavior and micromechanical-based c<strong>on</strong>stitutive model of shapememory<br />

Cu-Zn-Al single crystals", Acta Metall. Mater., 44(9): 3773-3780.<br />

Lim, T.J. and D.L. McDowell. 1999. "Mechanical behavior of an Ni-Ti shape memory<br />

alloy under axial-torsi<strong>on</strong>al proporti<strong>on</strong>al and n<strong>on</strong>proporti<strong>on</strong>al loading", Trans. Metall.<br />

Soc. AIME, 121(1): 9-18.<br />

Lu, Z.K. and G.J. Weng. 1997. "Martensitic transformati<strong>on</strong> and stress-strain relati<strong>on</strong>s of<br />

shape-memory alloys.", J. Mech. Phys. Solids, 45(11/12): 1905-1928.<br />

Miyazaki, S., S. Kimura, K. Otsuka and Y. Suzuki. 1984. "The habit plane and<br />

transformati<strong>on</strong> strains associated with the martensitic transformati<strong>on</strong>s in Ti-Ni single<br />

cystals", Scr. Metall., 18: 883-888.<br />

Mura, T., 1987. Micromechanics of Defects in Solids, 2nd, rev. ed., Bost<strong>on</strong>, MA: Kluwer<br />

Academic Publishers.<br />

Nye, J.F., 1985. Physical Properties of Crystals: Their Representati<strong>on</strong> by Tensors and<br />

Matrices, Repr. with correcti<strong>on</strong>s and new material ed., Oxford, UK: Clarend<strong>on</strong> Press.<br />

Otsuka, K. and K. Shimizu. 1974. "Morphology and crystallography of thermoelastic Cu-<br />

Al-Ni martensite analyzed by the phenomenological theory", Trans. Jpn. Inst. Met.,<br />

15(2): 103-108.<br />

Otsuka, K. and K. Shimizu. 1981. "Stress-induced martensitic transformati<strong>on</strong>s and<br />

martensite-to-martensite transformati<strong>on</strong>s", An Internati<strong>on</strong>al C<strong>on</strong>ference <strong>on</strong> Solid to<br />

Solid Phase Transformati<strong>on</strong>s, Pittsburgh, PA, Proceedings of An Internati<strong>on</strong>al<br />

C<strong>on</strong>ference <strong>on</strong> Solid to Solid Phase Transformati<strong>on</strong>s, H.I. Aar<strong>on</strong>s<strong>on</strong>, D.E. Laughlin,<br />

R.F. Sekerka et al., eds., Warrendale, Pa. : Metallurgical Society of AIME, pp.1267-<br />

1286.<br />

24


Otsuka, K. and C.M. Wayman, 1998. Shape Memory Materials, New York, NY:<br />

Cambridge University Press.<br />

Patoor, E., M.O. Bensalah, A. Eberhardt and M. Berveiller. 1993a. "Thermomechanical<br />

behaviour determining of shape memory alloys using a thermodynamical potential<br />

optimizing", Rev. Metall., 90: 1587-1592.<br />

Patoor, E., A. Eberhardt and M. Berveiller. 1993b. "Micromechanical <str<strong>on</strong>g>Model</str<strong>on</strong>g>ing of<br />

Superelasticity in Shape Memory Alloys." Pitman Research Notes in Mathematics<br />

Series: Progress in Partial Differential Equati<strong>on</strong>s: the Metz Surveys 2, M. Chipot, ed.,<br />

New York, NY: J. Wiley, Vol:296: 38-54.<br />

Patoor, E., A. Eberhardt and M. Berveiller. 1994. "Micromechanical modeling of the<br />

shape memory behavior", Internati<strong>on</strong>al Mechanical Engineering C<strong>on</strong>gress and<br />

Expositi<strong>on</strong>, Chicago, IL, Mechanics of phase transformati<strong>on</strong>s and shape memory<br />

alloys: presented at 1994 Internati<strong>on</strong>al Mechanical Engineering C<strong>on</strong>gress and<br />

Expositi<strong>on</strong>, L.C. Brins<strong>on</strong> and B. Moran, eds., New York, N.Y. : ASME, AMD - vol.<br />

189; PVP - vol. 292, pp.23-37.<br />

Patoor, E., A. Eberhardt and M. Berveiller. 1996. "Micromechanical modeling of<br />

superelasticity in shape memory alloys", MECAMAT '95 Internati<strong>on</strong>al Seminar <strong>on</strong><br />

Mechanics and Mechanisms of Solid-Solid Phase Transformati<strong>on</strong>s, La Bresse, France,<br />

J. Phys. IV, Colloq. (France), C. Lexcellent, ed. Editi<strong>on</strong>s de Physique, vol.6, no.C1,<br />

pp.277-292.<br />

Raniecki, B., C. Lexcellent and K. Tanaka. 1992. "Thermodynamic models of<br />

pseudoelastic behaviour of shape memory alloys", Arch. Mech., 44(3): 261-284.<br />

Saburi, T. and S. Nenno. 1981. "The shape memory effect and related phenomena", An<br />

Internati<strong>on</strong>al C<strong>on</strong>ference <strong>on</strong> Solid to Solid Phase Transformati<strong>on</strong>s, Pittsburgh, PA,<br />

Proceedings of An Internati<strong>on</strong>al C<strong>on</strong>ference <strong>on</strong> Solid to Solid Phase Transformati<strong>on</strong>s,<br />

H.I. Aar<strong>on</strong>s<strong>on</strong>, D.E. Laughlin, R.F. Sekerka et al., eds., Warrendale, Pa. :<br />

Metallurgical Society of AIME, pp.1455-1479.<br />

Saburi, T. and C.M. Wayman. 1979. "Crystallographic similarities in shape memory<br />

martensites", Acta Metall., 27: 979-995.<br />

Saburi, T., C.M. Wayman, K. Takata and S. Nenno. 1980. "The shape memory<br />

mechanism in 18R martensitic alloys", Acta Metall., 28: 15-32.<br />

Shaw, J.A. and S.K. Kyriakides. 1995. "Thermomechanical aspects of NiTi", J. Mech.<br />

Phys. Solids, 43(8): 1243-1281.<br />

Shaw, J.A. and S.K. Kyriakides. 1997. "On the nucleati<strong>on</strong> and propagati<strong>on</strong> of phase<br />

transformati<strong>on</strong> fr<strong>on</strong>ts in a NiTi alloy", Acta Metall. Mater., 45(2): 683-700.<br />

Shield, T.W. 1995. "Orientati<strong>on</strong> dependence of the pseudoelastic behavior of single<br />

crystals of Cu-Al-Ni in tensi<strong>on</strong>", J. Mech. Phys. Solids, 43(6): 869-895.<br />

Sun, Q.P. and K.C. Hwang. 1993a. "Micromechanics modeling for the c<strong>on</strong>stitutive<br />

behavior of polycrystalline shape memory alloys-I. Derivati<strong>on</strong> of general relati<strong>on</strong>s", J.<br />

Mech. Phys. Solids, 41(1): 1-17.<br />

25


Sun, Q.P. and K.C. Hwang. 1993b. "Micromechanics modeling for the c<strong>on</strong>stitutive<br />

behavior of polycrystalline shape memory alloys-II. Study of the individual<br />

phenomena", J. Mech. Phys. Solids, 41(1): 19-33.<br />

Sun, Q.P., T.T. Xu and X. Zhang. 1999. "On deformati<strong>on</strong> of A-M interface in single<br />

crystal shape memory alloys and some related issues", Trans. ASME, J. Eng. Mater.<br />

Technol., 121(Jan.): 38-43.<br />

Sun, Q.P., X. Zhang and T.T. Xu. 1997. "Some recent advances in experimental study of<br />

shape memory alloys", IUTAM Symposium <strong>on</strong> Macro- and Micro- Aspects of<br />

Thermoplasticity, Bochum, Germany, Solid Mechanics and Its Applicati<strong>on</strong>, O.T.<br />

Bruhns and E. Stein, eds., Vol. 62, pp.407-416.<br />

T<strong>on</strong>g, H.G. and C.M. Wayman. 1974. "Characteristic temperatures and other properties<br />

of thermoelastic martensites", Acta Metall., 22(July): 887-896.<br />

Vivet, A. and C. Lexcellent. 1998. "Micromechanical modeling for tensi<strong>on</strong>-compressi<strong>on</strong><br />

pseudoelastic behavior of AuCd single crystals", Eur. Phys. J. Appl. Phys., 4(2): 125-<br />

132.<br />

Zhu, W., B. Jiang and Z. Xu. 1986. "Crystallography of martensitic transformati<strong>on</strong> in a<br />

Cu-26Zn-4Al shape memory alloy", Acta Metall. Sin. (Chin. Ed.), 22(3): 229-236.<br />

26


Table 1. Interacti<strong>on</strong> energy and force comparis<strong>on</strong> between different models<br />

chem surf int<br />

Helmholtz’s free energy: Ψ = ∆G<br />

+ W + W + �<br />

Author Interacti<strong>on</strong> energy<br />

W int Interacti<strong>on</strong><br />

=<br />

1 II p<br />

−<br />

2V z σ ij ε ijdv<br />

v<br />

force<br />

int<br />

int W<br />

Fi<br />

=−<br />

f i<br />

∂<br />

Magnitude<br />

Features<br />

∂<br />

(unit volume)<br />

int<br />

Fi (Mpa)<br />

Patoor et 1<br />

al. (Patoor ∑ Hnm fnfm 2 nm ,<br />

et al., 1994;<br />

Patoor et<br />

al., 1996)<br />

−∑ Hnm fm<br />

m<br />

H 1 µ<br />

= ≈40<br />

1000<br />

H 2 µ<br />

= ≈270<br />

150<br />

Lu and<br />

Weng<br />

(Lu and<br />

Weng,<br />

1997)<br />

Vivet and<br />

Lexcellent<br />

(Vivet and<br />

Lexcellent,<br />

1998)<br />

Huang and<br />

Brins<strong>on</strong><br />

(Huang et<br />

al., 2000;<br />

Huang and<br />

Brins<strong>on</strong>,<br />

1998)<br />

1<br />

c1( 1−<br />

c1) L<br />

2<br />

p p<br />

( I − S)<br />

ε ε<br />

(<strong>on</strong>ly 1 variant<br />

c<strong>on</strong>sidered)<br />

N<br />

∑<br />

s=<br />

1<br />

−<br />

N<br />

f ( 1−<br />

f ) W<br />

s s s<br />

N<br />

∑ ∑<br />

f f W<br />

s t st<br />

s=<br />

1 s= 1, t= 1,<br />

t≠s 1 p p<br />

Ws= ε L( I −S)<br />

ε<br />

2<br />

G<br />

1<br />

p<br />

− σ ε f =<br />

1<br />

∑<br />

2 g=<br />

1<br />

G<br />

∑<br />

2 g=<br />

1<br />

G<br />

∑<br />

m=<br />

1<br />

g g<br />

p<br />

p<br />

ε [ LI ( − S)<br />

ε<br />

g<br />

p<br />

− f L( I −S)<br />

ε ] f<br />

g<br />

m m<br />

g<br />

g<br />

1<br />

− −<br />

2<br />

−<br />

1 2 ( c1) L<br />

p p<br />

( I S)<br />

ε ε<br />

−( 1−2f ) W<br />

s s<br />

(simplified form<br />

when <strong>on</strong>ly <strong>on</strong>e<br />

variant is<br />

c<strong>on</strong>sidered; same<br />

as Lu and Weng<br />

above)<br />

− ∂W<br />

∂<br />

f ( gn , )<br />

Gao and No explicit form. Small c<strong>on</strong>stant,<br />

Brins<strong>on</strong><br />

(this paper)<br />

lumped into Fc<br />

int<br />

< 1<br />

W s ≈ 0.468<br />

(Interacti<strong>on</strong><br />

between<br />

austenite and<br />

martensite)<br />

–40 to 60<br />

when µ=40<br />

-12 to 18<br />

when µ=12<br />

~1<br />

note Fc=4<br />

typically<br />

27<br />

A multivariant model<br />

Interacti<strong>on</strong> matrix H as approximati<strong>on</strong>:<br />

H 1 for self-accommodating variants<br />

H 2 for n<strong>on</strong>-self-accommodating <strong>on</strong>es<br />

Resistance force of F int increases with<br />

increased volume fracti<strong>on</strong>, causing a<br />

c<strong>on</strong>tinuous strain hardening effect<br />

Only values of H an order of magnitude<br />

lower than listed will produce a more<br />

appropriate n<strong>on</strong>-strain-hardening resp<strong>on</strong>se.<br />

Good qualitative agreement to uniaxial<br />

experimental results.<br />

A single variant model<br />

Analytical soluti<strong>on</strong> just for <strong>on</strong>e variant<br />

Resistance force of F int decreases with<br />

increased volume fracti<strong>on</strong>, becoming<br />

actually a driving force for f >0.5<br />

Aspect ratio of thin oblate was taken to be 0,<br />

reducing the magnitude of F int<br />

N<strong>on</strong>-linear dependence <strong>on</strong> f of dissipati<strong>on</strong><br />

Good agreement to uniaxial experimental<br />

results for plates of aspect ratio 0<br />

A single variant model<br />

Analytical soluti<strong>on</strong> just for <strong>on</strong>e variant<br />

Resistance force of F int decreases with<br />

increased volume fracti<strong>on</strong>, becoming<br />

actually a driving force for f >0.5<br />

Very thin oblate (a3/a1≈0.001) → lower Ws<br />

Very good agreement to uniaxial<br />

experimental results<br />

A multivariant model<br />

Analytical soluti<strong>on</strong> without additi<strong>on</strong>al<br />

approximati<strong>on</strong><br />

Self-accommodating groups of variants<br />

without interacti<strong>on</strong> am<strong>on</strong>g themselves<br />

Plate aspect ratio: a3/a1=0.01<br />

L<strong>on</strong>g time of calculati<strong>on</strong><br />

Good qualitative agreement to uni- or multiaxial<br />

experimental results<br />

A multivariant model<br />

<str<strong>on</strong>g>Based</str<strong>on</strong>g> <strong>on</strong> invariant plane nature of<br />

martensitic transformati<strong>on</strong><br />

Accounts for anisotropy in single crystal case<br />

Very fast calculati<strong>on</strong><br />

Excellent agreement to uni- or multi- axial<br />

experimental results


Table 2. Not all parameters necessary for the simulati<strong>on</strong> were readily available for the experimental<br />

materials. Therefore some parameters were found in the literature for materials with very similar<br />

compositi<strong>on</strong>s. Table 2a c<strong>on</strong>tains parameters known for the materials used in the experiments and table 2b<br />

lists other necessary parameters for similar materials attained elsewhere in the literature. Original sources<br />

are cited in the tables.<br />

Experimental<br />

Materials (wt%)<br />

Cu-13.95Al-3.93Ni<br />

(Shield, 1995) in Fig.<br />

5<br />

Cu70.17Zn25.63Al4.2<br />

(Goo and Lexcellent,<br />

1997; Lexcellent et al.,<br />

1996) in Fig. 3<br />

Cu71Zn25Al4<br />

(Gall et al., 1998) in<br />

Fig. 8, 9<br />

�n<br />

�m<br />

g<br />

volume change<br />

Transformati<strong>on</strong><br />

Temperatures<br />

Mf= - 13 °C, MS=5 °C<br />

AS=3 °C, Af=23 °C<br />

Mf= 12.5 °C, MS=22.5 °C<br />

AS=23.25 °C, Af=30 °C<br />

Mf= - 32 °C, MS= - 10 °C<br />

AS= - 15 °C, Af=5 °C<br />

0.6575, -0.7376, -0.1536<br />

-0.7256, -0.6737, -0.1403<br />

0.1903<br />

- 0.79%<br />

(Calculated by the authors using<br />

lattice parameters for Cu-14.0Al-<br />

4.2Ni (wt.%, β1 ′ 18R) (Otsuka and<br />

Shimizu, 1974; Otsuka and Shimizu,<br />

1981))<br />

-0.1489 0.7223, 0.6754<br />

-0.1350, 0.6550, -0.7434<br />

0.19386<br />

- 0.17%<br />

(Recalculated by the authors using<br />

lattice parameters for Cu70Zn26Al4<br />

(wt.%, 18R) (Zhu et al., 1986))<br />

Table 2a<br />

Equilibrium Test Volume Properties<br />

Temperatures Temperature change used for<br />

T0=(MS+Af)/2=14 °C<br />

T0′=(Mf+AS)/2= - 5 °C<br />

40 °C N/A Fig. 6, 7<br />

T0=(MS+Af)/2=26 °C<br />

T0′=(Mf+AS)/2= 18 °C<br />

T0=(MS+Af)/2= - 3 °C<br />

T0′=(Mf+AS)/2= - 24 °C<br />

Table 2b<br />

Elastic c<strong>on</strong>stants<br />

C11 = 142.6 (Gpa)<br />

C12 = 127.4 (Gpa)<br />

C44 = 97 (Gpa)<br />

(Hellwege and Hellwege, 1984)<br />

or<br />

µ=35 (Gpa)<br />

ν=0.3<br />

(Otsuka and Wayman, 1998; Shield,<br />

1995))<br />

µ=26 (Gpa)<br />

ν=0.3<br />

(Goo and Lexcellent, 1997; Lexcellent<br />

et al., 1996)<br />

28<br />

62 °C N/A Fig. 3, 4<br />

25 °C - 0.3% Fig.<br />

10 & 11<br />

B (MPaK -1 )<br />

F C (MPa)<br />

0.30<br />

4.1<br />

(tailored for<br />

(Shield, 1995))<br />

0.21<br />

1.7<br />

(tailored for<br />

(Goo and<br />

Lexcellent, 1997;<br />

Lexcellent et al.,<br />

1996))<br />

Properties<br />

used for<br />

All CuAlNi<br />

calculati<strong>on</strong>s<br />

Fig. 6 & 7<br />

All CuZnAl<br />

calculati<strong>on</strong>s<br />

Fig. 3, 4, 10 &<br />

11


Table 3. Measured and predicted angle between the tensi<strong>on</strong> directi<strong>on</strong> and the intersecti<strong>on</strong> directi<strong>on</strong> of the<br />

habit plane and the sample surface for figure 5 and figure 7. See text for details.<br />

Specimen<br />

A1-T1b<br />

[2.43 1 0]<br />

A1-T2b<br />

[-1 -1 1.73]<br />

A1-T3b<br />

[-1 -1 1]<br />

Experimentally<br />

measured angle<br />

for initial<br />

interface<br />

The axial strains of possible active variants<br />

al<strong>on</strong>g tensile directi<strong>on</strong> from<br />

phenomenological calculati<strong>on</strong> for β1′ (18R)<br />

46.8° V17: 0.0954020<br />

V18: 0.0954020<br />

21.4° V6: 0.0599651<br />

V15: 0.0599475<br />

9.8°<br />

29<br />

Calculated angle<br />

for initial<br />

interface<br />

48.402386°<br />

same as above<br />

24.997863°<br />

same as above<br />

Variants picked<br />

up by model<br />

√<br />

√<br />

√<br />

V6: 0.0228394 9.686614°<br />

V8: 0.0228493 55.529479° √<br />

V13: 0.0227965 55.529479°<br />

V15: 0.0228120 9.686614°<br />

V18: 0.0228072 8.748870°<br />

V19: 0.0228325 8.748870°<br />

Table 4. Uniaxial (1 and 3) and 3D (2 and 4 through 7) Stress State. Applied to the Polycrystalline CuZnAl<br />

Specimens.<br />

Test Number and<br />

Descripti<strong>on</strong><br />

#1<br />

Pure tensi<strong>on</strong><br />

#2<br />

Zero hydrostatic<br />

stress<br />

#3<br />

Pure compressi<strong>on</strong><br />

#4<br />

Triaxial compressi<strong>on</strong><br />

#5<br />

Triaxial compressi<strong>on</strong><br />

#6<br />

Triaxial compressi<strong>on</strong><br />

#7<br />

Pure hydrostatic<br />

stress<br />

Applied Stress State Effective<br />

Stress σ eff<br />

L<br />

M<br />

O<br />

P<br />

σ<br />

σ ij =<br />

0 0<br />

M0<br />

0 0<br />

NM<br />

P<br />

0 0 0QP<br />

L2σ0<br />

0<br />

σ ij = M<br />

O<br />

M 0 −σ0P<br />

NM<br />

P<br />

0 0 −σQP<br />

σ<br />

σ ij =<br />

L−<br />

0 0<br />

M<br />

O<br />

M 0 0 0P<br />

NM<br />

P<br />

0 0 0QP<br />

σ<br />

σ ij σ<br />

σ<br />

=<br />

L−3<br />

0 0<br />

M<br />

O<br />

M 0 − 0P<br />

NM<br />

P<br />

0 0 − QP<br />

σ<br />

σ ij σ<br />

σ<br />

=<br />

L−2<br />

0 0<br />

M<br />

O<br />

M 0 − 0P<br />

NM<br />

P<br />

0 0 − QP<br />

σ<br />

σ ij σ<br />

σ<br />

=<br />

L−17<br />

. 0 0<br />

M<br />

O<br />

M 0 − 0P<br />

NM<br />

P<br />

0 0 − QP<br />

σ<br />

σ ij σ<br />

σ<br />

=<br />

L−<br />

0 0<br />

M<br />

O<br />

M 0 − 0 P<br />

0 0 −<br />

NM<br />

QP<br />

Relative<br />

Hydrostatic<br />

Pressure σ h<br />

. σ eff<br />

σ +033<br />

3σ −000 . σ eff<br />

σ −033 . σ eff<br />

2σ −083 . σ eff<br />

σ −133 . σ eff<br />

0.7σ −176 . σ eff<br />

0 −∞σ eff


Figure 1. Typical micromechanics assumpti<strong>on</strong> <strong>on</strong> martensitic variants: an inclusi<strong>on</strong> (most models include<br />

<strong>on</strong>e variant <strong>on</strong>ly while a few include a group of variants) is embedded in an infinite matrix together with<br />

numerous other inclusi<strong>on</strong>s.<br />

c<br />

b<br />

a<br />

3.5mm<br />

Figure 2. Temperature-induced martensites (a) and stress-induced martensites (b-d). a) In a single CuAlNi<br />

crystal: two martensite plates are seen, <strong>on</strong>e toward the end of the narrow secti<strong>on</strong> and the other near the hole.<br />

Each martensite plate is composed of two HPVs; b) single CuAlNi crystal under uniaxial tensi<strong>on</strong>, strain<br />

rate 10 -4 s -1 used: <strong>on</strong>e large martensite band seen at the right edge (specimen width 3mm); c) a secti<strong>on</strong> of<br />

same single crystal loaded at strain rate of 1 s -1 showing evenly spaced martensitic plates spanning the<br />

specimen; and d) NiTi polycrystal under uniaxial tensi<strong>on</strong>, where martensitic plates are seen to typically<br />

span the entire width of the grains.<br />

30<br />

β1<br />

10 −4 m<br />

d


Stress σ 11 (x10 6 Pa)<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0<br />

10<br />

Single crystal pseudoelasticity<br />

Simple c<strong>on</strong>stant interacti<strong>on</strong><br />

Micromechanics interacti<strong>on</strong><br />

Experimental result<br />

20<br />

30<br />

31<br />

40<br />

Strain ε 11 (x10 -3 Pa)<br />

Figure 3. Single crystal pseudoelastic resp<strong>on</strong>se predicted by the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model with<br />

Micromechanics interacti<strong>on</strong> and the simplified <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> with simple c<strong>on</strong>stant interacti<strong>on</strong> for a<br />

Cu70.17Zn25.63Al4.2 (wt%) alloy. The test temperature is at 62 °C, T0= 26 °C, and T0′= 18 °C. Angles between<br />

tensile axis and [100], [010], and [001] of the single crystal are 46.16°, 70.68°, and 50.13° respectively.<br />

Experimental result is from (Goo and Lexcellent, 1997; Lexcellent et al., 1996).<br />

Stress σ 11 (x10 6 Pa)<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0<br />

10<br />

20<br />

30<br />

50<br />

40<br />

60<br />

Polycrystal pseudoelasticity<br />

Simple c<strong>on</strong>stant interacti<strong>on</strong><br />

Micromechanics interacti<strong>on</strong><br />

Strain ε (x10 11 -3 Pa)<br />

Figure 4. Polycrystal pseudoelastic resp<strong>on</strong>se predicted by the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model with<br />

Micromechanics interacti<strong>on</strong> and the simplified <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> with simple c<strong>on</strong>stant interacti<strong>on</strong> for a<br />

Cu70.17Zn25.63Al4.2 (wt%) alloy. The test temperature is at 62 °C, T0= 26 °C, and T0′= 18 °C (Goo and<br />

Lexcellent, 1997; Lexcellent et al., 1996).Note that a volume fracti<strong>on</strong> of martensite when averaged over all<br />

grains is approximately 50% at the stress level of 380Mpa in the simplified model. <str<strong>on</strong>g>Based</str<strong>on</strong>g> <strong>on</strong> recent<br />

experimental observati<strong>on</strong>s, such a state is reas<strong>on</strong>able in polycrystalline materials at the <strong>on</strong>set of significant<br />

permanent plastic deformati<strong>on</strong> and the stop of c<strong>on</strong>tinued martensite formati<strong>on</strong> (Brins<strong>on</strong> et al., 2001).<br />

70<br />

50


Figure 5: The stress-strain curves for fully austenitic specimens A1-T1b, A1-T2b and A1-T3b at 40 °C of a<br />

Cu-13.95Al-3.93Ni (wt%) single crystal (after (Shield, 1995)).<br />

Stress σ 11 (x10 9 Pa)<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.00<br />

0.02<br />

T3<br />

T2<br />

0.04<br />

Isotropic PE with<br />

Micromechanics interacti<strong>on</strong><br />

[-1 -1 1] or T3<br />

[-1 -1 1.73] or T2<br />

[2.43 1 0] or T1<br />

T1<br />

Strain ε11 Figure 6: Results predicted by the original <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> with interacti<strong>on</strong> for a Cu-13.95Al-3.93Ni<br />

(wt%) single crystal under different uniaxial loading directi<strong>on</strong>s. The test temperature is at 40 °C,<br />

T0=(MS+Af)/2=14 °C, and T0′=(Mf+AS)/2= - 5 °C. See text for details.<br />

32<br />

0.06<br />

0.08<br />

0.10


Stress Σ 11 (x10 6 Pa)<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0.00<br />

T3<br />

0.02<br />

T2<br />

Anisotropic PE with<br />

small c<strong>on</strong>stant interacti<strong>on</strong><br />

[-1 -1 1] or T3<br />

[-1 -1 1.73] or T2<br />

[2.43 1 0] or T1<br />

0.04<br />

Strain ε11 Figure 7: Results predicted by the simplified <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> <str<strong>on</strong>g>Model</str<strong>on</strong>g> without interacti<strong>on</strong> for a Cu-13.95Al-<br />

3.93Ni (wt%) single crystal under different uniaxial loading directi<strong>on</strong>s. The test temperature is at 40 °C,<br />

T0=(MS+Af)/2=14 °C, and T0′=(Mf+AS)/2= - 5 °C. See text for details.<br />

Figure 8: Effective stress-strain plots for a polycrystalline Cu71Zn25Al4 (wt%) <str<strong>on</strong>g>SMA</str<strong>on</strong>g> with a volume change<br />

of - 0.3% at a temperature of 25 °C. Each number corresp<strong>on</strong>ds to a 1D or 3D loading shown in table 4<br />

(after Gall et al. (Gall et al., 1998)).<br />

33<br />

0.06<br />

T1<br />

0.08<br />

0.10


Figure 9: Effective stress at the <strong>on</strong>set of transformati<strong>on</strong> vs. the hydrostatic stress at the <strong>on</strong>set of<br />

transformati<strong>on</strong> for a polycrystalline Cu71Zn25Al4 (wt%) <str<strong>on</strong>g>SMA</str<strong>on</strong>g> with a volume change of - 0.3% at a<br />

temperature of 25 °C. Each number corresp<strong>on</strong>ds to a 1D or 3D loading in table 4 (after Gall et al. (Gall et<br />

al., 1998)).<br />

Effective Stress (x10 6 Pa)<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0<br />

5<br />

10 15 20<br />

Effective Strain (x10 -3<br />

)<br />

34<br />

Pure Tensi<strong>on</strong> (#1)<br />

Zero Hydrostatic Pressure (#2)<br />

Pure Compressi<strong>on</strong> (#3)<br />

Triaxial Compressi<strong>on</strong> (#4)<br />

Triaxial Compressi<strong>on</strong> (#5)<br />

Triaxial Compressi<strong>on</strong> (#6)<br />

Figure 10: <str<strong>on</strong>g>Simplified</str<strong>on</strong>g> <str<strong>on</strong>g>Multivariant</str<strong>on</strong>g> model predicti<strong>on</strong>s for effective stress-strain plots for a polycrystalline<br />

Cu71Zn25Al4 (wt.%) with a volume change of - 0.3%. The test temperature is at 25 °C with the equilibrium<br />

temperature T0= - 3 °C, and T0′= - 24 °C. Each <strong>on</strong>e of the curves corresp<strong>on</strong>ds to a 1D or 3D loading shown<br />

in table 4. See text for details.<br />

25<br />

30


Effective Trans. Stress (x10 6 Pa)<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

σ eff Y offset<br />

0.100%<br />

0.050%<br />

0.025%<br />

With Micromechanics Interacti<strong>on</strong><br />

Pure Tensi<strong>on</strong> (#1) Zero Hydrostatic (#2)<br />

Pure Compressi<strong>on</strong> (#3) Triaxial Compressi<strong>on</strong> (#4)<br />

Triaxial Compressi<strong>on</strong> (#5) Triaxial Compressi<strong>on</strong> (#6)<br />

With Simple C<strong>on</strong>stant Interacti<strong>on</strong><br />

-300 -200 -100 0 100<br />

Hydrostatic Pressure (x10 6<br />

Pa)<br />

Figure 11: <str<strong>on</strong>g>Model</str<strong>on</strong>g> predicted effective stress at the <strong>on</strong>set of transformati<strong>on</strong> vs. the hydrostatic stress at the<br />

<strong>on</strong>set of transformati<strong>on</strong> for a polycrystalline Cu71Zn25Al4 (wt%) <str<strong>on</strong>g>SMA</str<strong>on</strong>g> with a volume change of - 0.3%. The<br />

test temperature is at 25 °C with the equilibrium temperature T0= - 3 °C, and T0′= - 24 °C. Each number<br />

corresp<strong>on</strong>ds to a 1D or 3D loading shown in table 4. See text for details.<br />

Figure 12: C<strong>on</strong>versi<strong>on</strong> process during loading of a thermally induced 2H (γ1′) <str<strong>on</strong>g>SMA</str<strong>on</strong>g> material at low<br />

temperature. Note that the final product after loading is not a single habit plane variant, but a<br />

corresp<strong>on</strong>dence variant (a detwinned habit plane variant). After work by Saburi and Nenno (Saburi and<br />

Nenno, 1981) where some informati<strong>on</strong> about the intermediate products are known, however the details of<br />

the loading and specimen orientati<strong>on</strong> are not given.<br />

35

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!