a unified thermodynamic constitutive model for sma

43
1 A Unified Thermodynamic Constitutive Model for SMA and Finite Element Analysis of Active Metal Matrix Composites Dimitris C. Lagoudas, Zhonghe Bo and Muhammad A. Qidwai Center for Mechanics of Composites Aerospace Engineering Department Texas A&M University College Station, Texas 77843-3141 ABSTRACT A unified thermodynamic constitutive model for Shape Memory Alloy (SMA) materials, derived based on the thermodynamic frame proposed by Boyd and Lagoudas (1995a), is presented in this paper. The specific selections for the form of Gibbs free energy associated with the martensitic volume fraction are identified for several earlier models. The thermal energy released or absorbed during the forward or reverse transformation predicted by the different models is compared with the experimental data from calorimetric measurements. The constitutive model is implemented in a finite element analysis scheme using a return mapping integration technique for the incremental formulation of the model. Finally, the constitutive model is utilized to analyze the thermomechanical response of an active metal matrix composite with embedded SMA fibers. Both tetragonal and hexagonal periodic arrangements of SMA fibers are considered in the calculation and the results are compared. 1. INTRODUCTION Shape memory alloy (SMA) materials (Buehler and Wiley, 1965, Wayman, 1983) are being used as active phases in active or "smart" composite structures ( Taya et al., 1993, Boyd and Lagoudas, 1994, Lagoudas et al., 1994a, Lagoudas et. al., 1994b). Due to the interactions of active

Upload: abdul-basit

Post on 30-Mar-2015

77 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: A Unified Thermodynamic Constitutive Model for SMA

1

A Unified Thermodynamic Constitutive Model for SMA and Finite Element Analysis of

Active Metal Matrix Composites

Dimitris C. Lagoudas, Zhonghe Bo and Muhammad A. Qidwai

Center for Mechanics of Composites

Aerospace Engineering Department

Texas A&M University

College Station, Texas 77843-3141

ABSTRACT

A unified thermodynamic constitutive model for Shape Memory Alloy (SMA) materials,

derived based on the thermodynamic frame proposed by Boyd and Lagoudas (1995a), is presented

in this paper. The specific selections for the form of Gibbs free energy associated with the

martensitic volume fraction are identified for several earlier models. The thermal energy released

or absorbed during the forward or reverse transformation predicted by the different models is

compared with the experimental data from calorimetric measurements. The constitutive model is

implemented in a finite element analysis scheme using a return mapping integration technique for

the incremental formulation of the model. Finally, the constitutive model is utilized to analyze the

thermomechanical response of an active metal matrix composite with embedded SMA fibers. Both

tetragonal and hexagonal periodic arrangements of SMA fibers are considered in the calculation and

the results are compared.

1. INTRODUCTION

Shape memory alloy (SMA) materials (Buehler and Wiley, 1965, Wayman, 1983) are being

used as active phases in active or "smart" composite structures ( Taya et al., 1993, Boyd and

Lagoudas, 1994, Lagoudas et al., 1994a, Lagoudas et. al., 1994b). Due to the interactions of active

Page 2: A Unified Thermodynamic Constitutive Model for SMA

2

SMA and non-active phases and the interactions among SMA particles themselves for high volume

fractions of SMA, complex 3-dimensional stress states result even when uniaxial loading is applied

to the composite. Therefore, accurate modeling of the thermomechanical response of SMA under

complex loading paths is necessary.

Extensive work has been done on the thermomechanical constitutive modeling of SMA

materials. In general, there have been two approaches: (1) the direct approach: the evolution laws

are obtained by either considering transformation micro-mechanisms (Tanaka, 1986, Sato and

Tanaka, 1988), or by directly matching experimental results (Liang and Rogers, 1990, 1991,

Graesser and Cozzarelli, 1991, Barrett, 1994); and (2) the thermodynamic approach: it starts with

constructing a free energy and then, by utilizing a dissipation potential in conjunction with the

second law of thermodynamics, the evolution laws for the internal state variables, i.e., the volume

fraction of the various forms of martensite, are derived (Berveiller et al., 1991, Ortin and Planes,

1989, 1991, Raniecki et al., 1994, Sun and Hwang, 1993a, 1993b, and Boyd and Lagoudas,

1995a,1995b). In the thermodynamic approach, the second law of thermodynamics applies

constraints to material constitution, but these constraints are usually weak. Hence, if a constitutive

model does not violate the thermodynamic constraints, usually it can also be derived by the

thermodynamic approach. It will be shown in this paper that several earlier models which were

derived based on different approaches are all related to each other under the thermodynamic

formulation.

Metal Matrix Composites (MMC) with SMA fibers are referred to as Active MMC

(AMMC) due to the incorporation of the active SMA phase. When SMA fibers are embedded in

elastoplastic metal matrix, they enhance the overall yield and hardening characteristics of the

composite at elevated temperatures, if they are properly prestrained at low temperature (Furuya et.

al., 1993, Taya et. al., 1993, Lagoudas et. al., 1994b). The thermomechanical response of AMMC

is influenced by the martensitic phase transformation in the SMA, which can be activated by applied

stresses and it occurs within a small temperature range. In this paper the effective thermomechanical

response of AMMC will be studied using the finite element method on a proper unit cell for a

periodic active composite. The purpose of this work is to investigate the interaction effects between

transforming SMA fibers and the elastoplastic matrix and study their impact on the effective

thermomechanical properties of the composite.

Page 3: A Unified Thermodynamic Constitutive Model for SMA

G( ij,T, , tij) G A( ij,T) G M( ij,T) G A( ij,T) G mix( ij,T, , t

ij)

G ( ij,T) 12

1 Sijkl ij kl1

ij ij(T T0) c (T T0) Tln( TTo

) so T uo

G mix( ij,T, , tij) f ( , t

ij)

G

G M A M A

G mix

3

(1)

(2)

(3)

In Section 2, a general form of a thermodynamic model, which is derived based on the

formalism of Boyd and Lagoudas (1995a) is given. The unification of two other models (Tanaka,

1986, Liang and Rogers, 1990) under the framework of the present thermodynamic model and

comparisons of these models with the experimental data is briefly discussed. In Section 3, a

numerical procedure for implementing the thermodynamic constitutive model is described in detail.

Finally, in Section 4 the development of a unit cell for periodic fibrous composites with tetragonal

and hexagonal periodic arrangements and the appropriate boundary conditions for the unit cell are

described, together with some selected results for the overall thermomechanical response and shape

memory characteristics of AMMC.

2. A UNIFIED THERMODYNAMIC CONSTITUTIVE MODEL FOR SMA

2.1 Derivation of the thermodynamic constitutive model

Based on the formulation of Boyd and Lagoudas (1995a), the total specific Gibbs free

energy, , of a polycrystalline SMA is assumed to be equal to the mass weighted sum of the free

energy , or , (superscript " " refers to martensitic phase and superscript " " refers to

the austenitic phase, respectively) of each phase plus the free energy, , of mixing:

The free energy of each species is written as

and the free energy of mixing is assumed to be given in this work by

Page 4: A Unified Thermodynamic Constitutive Model for SMA

u( teij , s, , t

ij) G Ts 1ij

teij

T ( teij

G

ij

) ij (s GT

)T ( ijG

tij

) tij

G 0

, T, ij,tij

, Sijkl, ij, c , so , uo

f( , tij )

u

teij ij

tij ij s

4

(4)

(5)

where are martensitic volume fraction, temperature, stress tensor and transformation

strain tensor, respectively, and are mass density, elastic compliance tensor,

thermal expansion coefficient tensor, specific heat, specific entropy at a reference state, and specific

internal energy at the reference state of the " " phase, respectively. The generic function

physically represents the elastic strain energy due to the interaction between martensitic variants and

the surrounding parent phase, and among the martensitic variants themselves. Reorientation

(detwinning) effects have been omitted for simplicity (Boyd and Lagoudas, 1995a).

Note that the present formulation does not account for permanent changes in microstructure

of SMA materials, i.e., after a complete thermomechanical loading cycle the material returns to its

original state without increase of any local entropy (Ortin and Planes, 1989). For modeling of cyclic

responses of SMA with changing of internal microstructures, the work by Bo and Lagoudas (1995)

can be referenced.

The Gibbs free energy defined above is related to the internal energy by the following

equation

where is the thermoelastic strain tensor, is the total strain tensor, is the entropy

per unit mass, and is the mass density of the SMA, assumed to be independent of the martensitic

volume fraction. Infinitesimal strains will be assumed throughout the paper, which is a reasonable

assumption for polycrystalline SMA materials undergoing constraint deformations as part of the

microstructure of the AMMC. By performing standard calculations (Malvern, 1969), the strong

form of the second law of thermodynamics (Truesdell and Noll, 1965) can be written in the

following form for the local internal dissipation rate

Page 5: A Unified Thermodynamic Constitutive Model for SMA

teij

G

ij

Sijkl kl ij(T T0)

s GT

1ij ij cln( T

T o) so

Sijkl S Aijkl (S M

ijkl S Aijkl ) , c c A (c M c A)

ijAij ( M

ijAij) , so s A

o (s Mo s A

o )

T effij

tij

G 0

tij ij

ij T G ij T

effij ij

ftij

tij

ij

5

(6)

(7)

(8)

(9)

where is the local dissipation rate (it does not include the entropy production rate due to heat

conduction) and the “dot” above is a symbol that indicates the increment of the corresponding

quantity. Since and are independent state variables, and is independent of and , the

following set of constitutive relations can be obtained by identically satisfying inequality (5)

where

The local internal dissipation rate is then given by

where the effective stress acts as a thermodynamic force conjugate to .

To simplify the formulation given above for the current case in which only transformation,

without reorientation of martensitic variants, is considered, the following assumption can be

introduced to relate the evolution of the transformation strain to the evolution of the martensitic

volume fraction (Bondaryev and Wayman, 1988, Boyd and Lagoudas, 1995a):

The transformation tensor is assumed to have the following form

Page 6: A Unified Thermodynamic Constitutive Model for SMA

i j

32

H ¯eff 1 effij , > 0

H ¯t 1 ti j , < 0

T effij ij

G 0

( ; ij,T, , tij)

H t max ¯eff 3

2

effij

effij

12

effij

effij

1

3

effkk ij

¯t 2

3

tij

tij

12

effij ij

G

> 0

< 0

( ; ij,T, , tij)

6

(10)

(11)

(12)

where is the maximum uniaxial transformation strain, ,

, and .

Using equation (9), the local dissipation rate given by equation (8) can be rewritten as

where is the thermodynamic force conjugate to . Inequality (11) is the

thermodynamic constraint to the SMA material system. For the forward phase transformation (when

) the driving force must be greater than zero, and for the reverse phase transformation

(when ) must be less than zero.

To obtain the evolution equation for the internal state variable , an additional assumption

must be introduced, either utilizing Edelen's formalism (Edelen, 1974) of dissipation potential theory

or directly assuming that satisfies a certain criterion during the phase transformation (Fischer et

al., 1994). The latter approach is a special case of Edelen's formulation. Following Edelen's

formalism, a dissipation potential, , can be introduced such that the evolution of the

internal state variable, , for the rate independent case, is given by (Boyd and Lagoudas, 1995a)

Page 7: A Unified Thermodynamic Constitutive Model for SMA

0 , Y , ( Y) 0

12

2

effij ij

1

2Sijkl ij kl ij ij T c T Tln T

T0soT

f( ) uo ±Y

Sijkl S Mijkl S A

ijkl , ijMij

Aij , c c M c A

so s Mo s A

o , uo u Mo u A

o , T T T0

Y 2Y

Y > 0

Y 0

> 0

Y

7

(13)

(14)

(15)

(16)

where satisfies the Kuhn-Tucker conditions

and is a material parameter. During phase transformation ( ), can be evaluated by using

the consistency condition, i.e., . If the potential is assumed to be the convex quadratic

function

then by Kuhn-Tucker condition given in equation (13) for the case of , we get

where

To satisfy the thermodynamic constraint given in equation (11), the plus sign in equation (15) should

be used for the forward phase transformation, while the minus sign should be used for the reverse

phase transformation. Equation (15) can be alternatively used to obtain the evolution equation for

the internal state variable , without utilizing evolution equation (12), since in the present case the

internal state variable is only a scalar.

Note that the material constant can be interpreted as the threshold value of the

thermodynamic force for the onset of the phase transformation. It is also a measure of the

internal dissipation due to microstructural changes during the phase transformation. Using equations

(11) and (15), the total internal dissipation during a complete forward and reverse isothermal phase

transformation cycle is given by

Page 8: A Unified Thermodynamic Constitutive Model for SMA

T Td d1

0

Y d0

1

( Y )d 2Y

f( , tij )

f

f f tij

f( )

f( )

f(0) 0 f( )

f( )

f( )

f( )

8

(17)

2.2. Unification of several SMA constitutive models

Different constitutive models are distinguished intrinsically by different selections of internal

state variables and their evolution equations. Since the interaction elastic strain energy, represented

by , describes transformation induced strain hardening in the SMA material, if the internal

state variables are selected to be the same for different SMA constitutive models, then they can

possibly be unified under the current thermodynamic framework. This can be done by selecting

different forms of the function to represent the transformation induced strain hardening as

predicted by these models.

In determination of the hardening function, , it is assumed that is independent of ,

which physically can be interpreted as the absence of kinematic transformation hardening. Then for

the following properties are assumed to be preserved: (1) parent austenitic phase is stress free

if no external mechanical loading is applied. This condition is satisfied by selecting to be zero

at the fully austenitic state, i.e., ; (2) the function must be non-negative since it

represents part of the elastic strain energy stored in the material. This requirement can be satisfied

by appropriately choosing material constants; and (3) the function must be continuous during

the phase transformation, including return points, for all possible loading paths (a return point

during the phase transformation is characterized by a change in the sign of ). Due to the formation

of different microstructures during the forward and reverse phase transformations, may have

different functional forms during the forward and reverse phase transformations, but these functions

must be continuously connected at return points.

Applying the above constraints, a form for the function can be selected as follows:

Page 9: A Unified Thermodynamic Constitutive Model for SMA

f( )f M( ) , > 0

f A( ) , < 0

f M( ) f M0( ) 11 R

f A( R) f M0( R)

f A( ) f A0( )R

f M( R) f A0( R)

f M0( )so

a Me

(1 )ln(1 ) (µe1 µe

2)

f A0( )so

a Ae

ln( ) 1 (µe1 µe

2)

f M0( )0

so

a Mc

cos 1(2 1) d (µc1 µc

2 )

f A0( )0

so

a Ac

cos 1(2 1) d (µc1 µc

2)

R

R 1

0 R f( )

f M0(0) f A0(0) 0 f M0( ) 0 f A0( ) 0 0 1 f M0(1) f A0(1)

f M0( ) f A0( )

9

(18)

(19)

(20)

(21)

where

and is the martensitic volume fraction at the return point. The range of for the forward phase

transformation is given by , while the range of for the reverse phase transformation is

given by . This selection of satisfies all of the properties from (1) to (3) given

above, provided , , for , and .

The functions and are selected differently for different models. For Tanaka's

exponential model (Sato and Tanaka, 1986), they can be selected as

while for Liang-Rogers' cosine model (Liang and Rogers, 1990), they take the form

Page 10: A Unified Thermodynamic Constitutive Model for SMA

effij ij

1

2Sijkl ij kl ij ij T c T Tln T

T0soT

s0

a Mc

cos 1(2 1) µc1 µc

2f A( R) f M0( R)

1 Ruo Y 0

12

cos a Mc (T M of)

a Mc

C MHij ij

12

Sijkl ij kl ij ij T 1

a Me , a A

e , a Mc , a A

c µ1 e,c µ2

f M0(1) f A0(1)

f

R 0

f A( R) f M0( R) 0

µc2

c 0 C Ms0

H

12

Sijkl ij kl

ij ij T

1H ij ij

10

(22)

(23)

where , and where are material constants, and the parameter can

be determined from the continuity condition .

The equivalence between the above selection of function and the corresponding original

models can be illustrated as follows by taking Liang-Rogers’ cosine model as an example:

Substituting equation (18) into equation (15), and using equations (19) and (21), for the forward

phase transformation the criterion given by equation (15) can be written explicitly as follows

For the case of a complete loading-unloading cycle, , then using equations (19) and (21) ,

. Equation (22) can then be rearranged to obtain explicitly in terms of

temperature and applied stress as follows

where Table 1 and the expression for from below (paragraph following equation (24)) are

utilized, and is assumed. The constant is called the martensite stress

influence coefficient in the original models. If one neglects the effect due to the terms

and , equation (23) is exactly the same as the one that has been proposed by Liang and

Rogers (1990), except for the term that is used here as the effective driving stress, instead

Page 11: A Unified Thermodynamic Constitutive Model for SMA

12

Sijkl ij kl ij ij T

1H

effij ij

¯eff

tij

M os, M of, A os A of

a Ae

ln(0.01)A os A of

a Ac

A of A osb A so A of A os

a Me

ln(0.01)M os M of

a Mc

M os M ofb M so M os M of

e 12

so M os 2A of A os c 12

so M os A of p 12

so M os A of

11

of only the applied stress in the original model. The corresponding expressions for Tanaka’s

exponential model as well as for the case of the reverse phase transformations for both models can

be obtained similarly.

The unification attempted in this work captures the main hardening features predicted by the

corresponding models, and it also gives a form of generalization to these models. The effect of the

change of the elastic compliances and thermal expansion coefficients is introduced naturally in the

current thermodynamic formalism by the terms, , and . Note also that in the

current model the quantity (generalized effective Von Mises stress) is utilized as the

driving force to the stress-induced phase transformation, instead of (effective Von Mises stress).

As a result, during the reverse phase transformation the stress applied to the opposite direction of

the accumulated transformation strain, , will help the reverse phase transformation. This

phenomenon has been observed in experiments (Graesser and Cozzarelli, 1991).

Table 1. Evaluation of material constants for given transformation temperatures.

Data given: and .

Exponential Cosine Polynomial

Page 12: A Unified Thermodynamic Constitutive Model for SMA

f M0( ) 1

2b M 2 (µp

1 µp2)

f A0( ) 1

2b A 2 (µp

1 µp2)

Ye12

so A os M os

so

2ln(0.01)M os M of A of A os

Yc12

so A of M os

14

so M os M of A of A os

Yp12

so A of M os

14

so M os M of A of A os

f( )

b M b A

b M b A

µp2 0 f( )

µp1

µ2 1

µ2 µe2

s0

21

a A

1a M

µc2

s0

41

a Mc

1

a Ac

µp2

14

b A b M

µ1

12

(24)

For a polynomial representation of the function including up to quadratic terms, the

different hardening effects during the forward and reverse phase transformations can be accounted

for by the following selection

where and are linear isotropic hardening moduli for the forward and reverse phase

transformations, respectively. If the hardening parameters and are selected to be the same,

then and the original form of reported by Boyd and Lagoudas (1995a) is recovered,

except for the additional linear term, , in equation (22).

The parameter is introduced to enforce the continuity condition at . The evaluation

of for each model is given by for the exponential model,

for the cosine model, and for the polynomial model,

respectively. The material constant is used to describe the accumulation of elastic strain energy

at the onset of the forward phase transformation. It is shown by Salzbrenner and Cohen (1979) that

Page 13: A Unified Thermodynamic Constitutive Model for SMA

f(0) 0 f( )

13

for polycrystalline SMA the forward and reverse phase transformation zones shift to lower

temperature, compared to their single crystal counterpart. This phenomenon indicates that

, and suggests that the function must contain the linear term.

Using the data given in Table 2, the uniaxial stress-strain curves predicted by the different

models are plotted in Fig. 1. The exponential law is plotted for two different ways of calibrating the

material constants (refer to Section 2.3). It is shown that the stress-strain curves of all three models

are similar, and they give the same amount of total hysteresis in a complete loading-unloading cycle,

except for the curve corresponding to the exponential model with input data given transformation

temperatures.

Table 2. Material constants for aluminum and SMA.

Material Constants for Aluminum at Room Temperature

E = 69x10 MPa3

= 0.33 α = 23.6x10 / C-6 o

Yield stress = 70 MPa

Material Constants for (Polynomial Hardening Law) SMA

Page 14: A Unified Thermodynamic Constitutive Model for SMA

u ij ij Q

Q cT T s

ijij ( T s ) g(T, ij)T hij(T, ij) ij

Q

14

(25)

(26)

E = 70.0 x 10 MPaA 3

E = 30.0 x 10 MPaM 3

= 0.33A

= 0.33M

α = 10.0x10 / CΑ -6 o

α = 10.0x10 / CΜ -6 o

H = 0.05 ρ∆c = 0.0 ρ∆s = -0.35 MPa/°Co

M = 3°Cos

M = 18°Cof

A = 22°Cos

A = 42°Cof

ρb = 7.0 MPa A

ρb = 5.25 MPaM

γ = -10.5 Mpa p

Y = 3.7625 MPa p*

To compare further the strain hardening behavior predicted by different models, the

experimental data from calorimetric measurements can be utilized. In order to perform the

comparison, the excess specific heat due to the phase transformation predicted by the models is

investigated first. The first law of thermodynamics reads

where is the heat input rate. By substituting equation (4) into (25) and using the constitutive

relations given by equation (6) the following equation can be derived

Table 3. Material constants for all three models.

Page 15: A Unified Thermodynamic Constitutive Model for SMA

E A, E M, A, M, A, M, H, c, so, Y

a Ae , a M

e , e a Ac , a M

c , c b A, b M, p

M sf, A sf, T AM Y

a Ae

ln(0.01)A sf

a Ac

A sfb A a 4 A fs

a Me

ln(0.01)M sf

a Mc

M sfb M a 4 M sf

e 12

so T AM A sf c 12

so T AM p 12

so T AM

M os 12

T AM Yso

12

1a M

1 ln(0.01)a A

M os 12

T AM Yso

14

M sf A sf

M os 12

T AM Yso

14

M sf A sf

A of 12

T AM Yso

12

1a M

1 ln(0.01)a A

A of 12

T AM Yso

14

M sf A sf

A of 12

T AM Yso

14

M sf A sf

15

Material constants in common for all three models

Material constants different

Exponential Cosine Polynomial

Table 4. Evaluation of material constants for given difference of transformation

temperatures, their average and the value of hysteresis.

Data given: and .

Exponential Cosine Polynomial

For the stress free case, equation (26) can be written in the following form

Page 16: A Unified Thermodynamic Constitutive Model for SMA

Q cT g(T)T

g(T) c ln( TT0

) soTY

T

g(T)

16

(27)

(28)

where, by utilizing equation (6b) and (15), the function predicted by the current model can be

written as

Page 17: A Unified Thermodynamic Constitutive Model for SMA

g(T)

17

The function can be

Page 18: A Unified Thermodynamic Constitutive Model for SMA

T g(T)

g(T)

E A, A E M, M

A M

E A,E M A, M A M

H

A os

c A c M

c

18

called the excess specific heat per unit mass due to the phase transformation.

The term on the left hand side of equation (28) can be measured by a calorimetric

measurement at constant temperature rate. Since is constant, the function can be obtained

using (27). Thus theoretical predictions can be compared directly with experimental results. The

curves of the function computed by using equation (28) for different models are plotted in Fig.

3, and an experimental curve from a calorimetric measurement is also plotted in the same figure.

It is shown that these different models give different predictions for the excess specific heat during

forward phase transformation (similar results are valid for the reverse phase transformation).

One observation from Figs. 2 and 3 is that, even though the curves of the excess specific heat

due to phase transformation are predicted to be remarkably different for the three models, the stress-

strain curves are similar, because the former curves are approximately proportional to the derivatives

of the latter ones.

2.3 Determination of material constants used in the SMA constitutive equations

For the polycrystalline SMA it can be reasonably assumed that the material is isotropic.

Then, the elastic compliance tensor for both phases can be represented by the Young’s modulus and

poission ratio and , respectively, and the thermal expansion coefficients tensor for

both phases can be reduced to two scalar constants and , respectively. Young’s moduli

, Poisson ratios , and thermal expansion coefficients and can be obtained by

performing standard tests at low temperature for martensitic constants and at high temperature for

austenitic constants, respectively.

The maximum uniaxial phase transformation strain can be obtained from a uniaxial test

at a temperature below austenitic start temperature by measuring the residual strain retained

after the specimen has been fully loaded to martensitic phase followed by complete unloading. The

specific heat of the two phases and can be determined from a calorimetric measurement test.

In the current model only the difference of the specific heat is used. Many calorimetry

Page 19: A Unified Thermodynamic Constitutive Model for SMA

ddT

so

H S T

c 0

so

ddT

so

H

Y

ijd ij

ijd ij 2Y Y

T eq

19

(29)

experiments show that the specific heat of both phases is almost equal, henceforth it is assumed that

. The selection of the remaining six material parameters is discussed next.

For the case of uniaxial tension, the stress-temperature transformation curves represented

by equation (15) for fixed martensitic volume fraction are plotted in Fig. 3. These curves can be

obtained experimentally (Miyazaki et al., 1981, Shaw and Kyriakides, 1995). The entropy difference

per unit volume between the phases can then be determined phenomenologically by the slope

of these curves at zero stress (Fig. 3). To analytically obtain the slopes, we first take the differential

of equation (15) for constant martensitic volume fraction , and after rearrangement of terms, the

slope of the curves can be obtained by the following equation:

Equation (29) is usually called the Clausius-Clapeyron equation (Ortin and Planes, 1989).

Substituting zero stress and neglecting the thermal term in the denominator in the above equation,

the slope of these curves is found out to be . This slope is predicted to be the same for

both forward and reverse phase transformations. This is consistent with experimental observations

(McCormick and Liu, 1994, Stalman et. al., 1992, Miyazaki, et. al., 1981).

The constant can be obtained from an isothermal uniaxial pseudo-elasticity test (stress

induced martensitic phase transformation fully reversible to the parent austenitic phase upon

unloading (Wayman, 1983)). The area enclosed by the stress-strain curve of a complete isothermal

pseudo-elasticity loading cycle can be calculated by integrating the stress over strain for a complete

cycle, i.e., . Utilizing equations (6) and (15) the integral can be explicitly integrated as

. Therefore, is directly correlated to the total hysteresis area enclosed by a stress-

strain curve which can be measured from an isothermal pseudo-elasticity test.

Following the arguments given by Salzbrenner and Cohen (1979), the phase equilibrium

temperature of single crystal SMA materials can be obtained by measuring the martensitic start

Page 20: A Unified Thermodynamic Constitutive Model for SMA

M os A of T eq 1

2(M os A of)

uo uo s0 T eq

uo

uo µ1 e,c,p

µ1 uo

M sf M os M of A sf A of A os

T AM 1

2(M os A of)

Y

M os, M of, A os A of

Y

20

and austenitic finish temperatures, and , and assuming . Then the

constant can be obtained as (Raniecki and Lexcellent, 1994). For

polycrystalline SMA materials can be obtained by performing tests on its single crystal

counterpart as described above. However, to evaluate the stress increment for given strain and

temperature increments, it is not necessary to determine and where

individually, but rather , since they both contribute to the linear part of the free

energy with respect to . Thus, there are three remaining material constants to be determined for

the three models, as shown on the bottom of Table 3.

After the above ten material constants are evaluated, the remaining three are determined

indirectly by specifying the total amount of hardening for forward and reverse phase

transformations, which is represented by , , and by

prescribing the average phase transformation temperature for the polycrystalline SMA,

. The pertinent evaluations of the remaining constants for the three models

are shown in Table 2. Note that since is utilized as input, the resulting transformation

temperatures for the three models are not identical, as it can be seen from Table 4.

An alternative way of calibrating these three remaining constants in the case that all four

transformation temperatures, i.e., and , are utilized as input is shown in Table

1. These temperatures are obtained from the stress-temperature transformation lines, shown in Fig.

3, by calculating the intersection points of these lines with the temperature axis. However, as Table

1 shows, the evaluation of for the exponential model is different from its evaluation for other

two models. This means that the total hysteresis area predicted by the exponential model is different

from the other two models.

Using the experimental data given by Jackson et. al. (1972), Howard (1995), Stalmans et.

al. (1992) and the way of determining the material constants discussed above, the material constants

needed for implementing the current model are determined as given in Table 2. The uniaxial stress-

Page 21: A Unified Thermodynamic Constitutive Model for SMA

Y , > 0

Y , < 0

ij Sijkl kl ijT Qij

> 0 0 1

teij

21

(30)

(31)

strain curves predicted by the different models are plotted in Fig. 1. The exponential law is plotted

for the two different ways of calibrating the material constants described above. It is shown that if

phase transformation start and finish stresses are specified to be the same, then the hysteresis loop

predicted by the exponential model will be different from the hysteresis loops predicted by the other

models. Conversely, if the hysteresis area is enforced to be the same for all models, then the

transformation start and finish stresses predicted by exponential model will be different from the

predictions of the other two models.

3. NUMERICAL IMPLEMENTATION OF THE CONSTITUTIVE MODEL

In typical finite element analysis, a displacement formulation is utilized, and for solving

nonlinear problems an incremental Newton-Raphson integration method is a usual choice

(Zienkiewicz, 1977). In implementing the SMA constitutive model the tangent stiffness tensor and

the stress tensor at each integration point of all elements should be updated in each iteration for given

increments of strain and temperature. In order to utilize established numerical schemes used in

plasticity theory, we define the transformation function as follows

The thermomechanical transformation zone described by Kuhn-Tucker condition, equation (13) can

equivalently be characterized by letting and . The transformation function takes

the similar role as the yield function in plasticity theory, but in the current case, an additional

constraint for must also be satisfied.

To derive the tangent stiffness tensor, using the definition of the thermal elastic strain tensor

, the stress-strain equations of state (6a) can be written in the following rate form

Page 22: A Unified Thermodynamic Constitutive Model for SMA

Qij Sijkl kl ij(T T0) ij

ijij T

T 0

ij Lijkl kl lijT

Lijkl Sijkl

Qijkl

1

, lij Lijkl Qkl kl

0

Lijkl lij

ij

±Qij

Lijkl Lklij

22

(32)

(33)

(34)

(35)

where

The consistency condition reads

Equation (31) and (33) can be used to eliminate and obtain the relationship between stress

increment and strain and temperature increments as

where the tangent stiffness tensor and tangent thermal moduli are defined by

Since , where the plus sign is used for the forward phase transformation and the minus

sign is for the reverse one, the tangent stiffness tensor is always symmetric, i.e., .

To calculate the increment of stress for given strain and temperature increments, a return

mapping integration algorithm proposed by Ortiz and Simo (1986) has been used. Because the

present formulation is time independent, equation (31) can be written in the following incremental

form

Page 23: A Unified Thermodynamic Constitutive Model for SMA

ij Sijkl1

kl kl T Qkl

1ij Sijkl

1kl kl T , 1

ij0ij

1ij

p 1ij Sijkl(

p) 1Q pkl

p 1

( ij , ) p( pij ,

p)p

ij

p 1ij

pp 1

p 1p

p

ij

sijkl1Q p

kl

p, t p 1

ijpij

p 1

0

ij 0 T 0 pth

p

T T p 1 0

0

t p 1

ij

p 1 t p 1

ij

23

(36)

(37)

(38)

(39)

(40)

The elastic predictor is calculated in the first step by letting , i.e.

An iterative scheme is then carried out to obtain the transformation corrector from equation (36) by

assuming , and , i.e., during the iteration

The transformation function, equation (30), is expanded into a Taylor series about the current value

of state variables, denoted by superscript " ", and is truncated at the linear part as shown below:

where temperature, , is fixed, and thus . By applying the Kuhn-Tucker condition given

by equation (13c) for the case of during phase transformation, and using equation (38), p+1

and can be obtained as follows:

The state variables and can then be updated as follows

Page 24: A Unified Thermodynamic Constitutive Model for SMA

p 1 p p 1 , t p 1

ijt p

ijt p 1

ij

p 1ij Sijkl(

p 1) 1kl kl(

p 1) T t p 1

kl

p 1

10 8

24

(41)

(42)

and the stress can then be obtained by using the constitutive equation (6a)

The iterative procedure ends if is less than a specified tolerance. If the convergence criterion

is not satisfied, calculations given by equations (38) through (42) are repeated until convergence is

achieved. For typical cases it takes only about three iterations to satisfy the tolerance of .

4. FINITE ELEMENT ANALYSIS FOR AMMC COMPOSITES

Due to the nature of fabrication techniques, one expects to have a certain periodic pattern of

SMA fiber placement in AMMC composites. In the present paper the tetragonal and hexagonal

arrangements, as shown in Fig. 4a and Fig. 4b, of SMA fibers will be investigated. The basic unit cells

for these two cases are tetragonal and hexagonal respectively (shown in the figure by dashed lines).

Assuming that the applied loads do not violate the geometric symmetry of the unit cells, the smallest

representative unit cells which can be used to predict the composite overall behavior, corresponding

to the tetragonal and hexagonal cases, are given in the Fig. 4c and Fig. 4d, respectively. In the

tetragonal case, Fig. 4c, the cross section of the representative cell is a square with length a, while

in the hexagonal case, Fig. 4d, the cell is a rectangle with length a, and width b, and b=0.866a. The

stress and strain in the unit cells will completely represent the whole

Page 25: A Unified Thermodynamic Constitutive Model for SMA

25

composi t e

response to the external applied loading because of the periodicity and symmetry conditions of the

composite geometry, material parameters, as well as the applied load.

To reduce the cost of FEM analysis, it is assumed that the temperature field within the

AMMC specimen is uniform. The three main reasons for this assumption are as follows: (1)

Consideration of heat conduction at the unit cell level, will destroy the periodicity of the

microstructure, thus rendering the whole analysis costly; (2) The temperature gradient in the specimen

will be small, if the thermomechanical loading processes are assumed to be slow (the meaning of slow

will be defined better at the end of this paragraph); (3) The main purpose of the current analysis is

to investigate the characteristics of the effective thermomechanical response of AMMC specimens,

and not their detailed response under temperature gradients. In heat transfer analysis that we

Page 26: A Unified Thermodynamic Constitutive Model for SMA

u1(0,x2,x3) u1(0) , u1(a,x2,x3) u1(a) ,u2(x1,0,x3) u2(0) , u2(x1,a,x3) u2(a) ,

12(o,x2,x3) 12(a,x2,x3) 12(x1,0,x3) 12(x1,a,x3) 0

u1(a2

,x2,x3) u1(a2

) , u1(a2

,x2,x3) u1(a2

) ,

u2(x1, b,x3) u2( b) ,

2mm

30%

26

(43)

(44)

(45)

performed using FEM on a cylindrical unit cell, which is close to that of the hexagonal case, with a

cylindrical SMA fiber of in diameter at the center of the unit cell and the volume fraction of the

SMA fiber of , we found that a temperature differential of ten degrees between the center of

the SMA fiber and the external adiabatic boundary of the unit cell, equilibrates to less than 0.1

degrees difference in temperature in less than 0.1 second. Although the geometry of the actual unit

cells for the tetragonal and hexagonal cases are different, it is expected that the time scale will remain

in the same order of the magnitude given above.

4.1 Boundary conditions for the unit cells

To derive the boundary conditions on the boundaries of the unit cells, let us first consider the

symmetry conditions on the surfaces represented by the lines AB, BC, CD, and AD in Fig. 4. On all

of these surfaces for the tetragonal case and the surfaces AB, BC, and AD for the hexagonal case,

the reflection symmetries require that the shear stresses vanish at all points and the normal

displacements are constant. These conditions can be mathematically expressed as follows for the

tetragonal case:

and for the hexagonal case:

Page 27: A Unified Thermodynamic Constitutive Model for SMA

12(a2

,x2,x3) 12(a2

,x2,x3) 12(x1 , b,x3) 0

u1( x1,0,x3) u1(x1,0,x3) , u2( x1,0,x3) u2(x1,0,x3)

12( x1,0,x3) 12(x1,0,x3) , 22( x1,0,x3) 22(x1,0,x3)

u3(x1,x2,0) u3(0) , u3(x1,x2,1) u3(1)

13(x1,x2,0) 23(x1,x2,0) 13(x1,x2,1) 23(x1,x2,1) 0

ui(xi) xi xi , i 1,2,3 i th xi

x3

x3

u1(0) u2(0) u3(0) 0 u1(a2

) u2( b) u3(0) 0

27

(46)

(47)

(48)

(49)

(50)

where is the displacement of the plane , in direction, and is a

constant. For the hexagonal case there is a 180 rotational symmetry about point O, which is at the

middle of line CD. The displacement boundary conditions on this surface are given by

the traction boundary conditions satisfy the following equations

In the longitudinal, , direction, a generalized plane strain state is assumed, and the

dimension of the unit cell in direction is taken to be unit length. The boundary condition on the

two surfaces can be expressed by the following

and

In implementing the above in a finite element analysis the boundary conditions stated in

equations (43) to (47) and (49) to (50) are standard. Neglecting rigid body motion, it can be set that

for the tetragonal case and for the

hexagonal case. The rest of the kinematic boundary conditions given in equations (43), (45), (47),

and (49) can be applied by using an equation type of constraints provided by most of commercially

available FEM software, such as ABAQUS. The boundary conditions given by equation (48),

however, are constraints for the components of the traction vector, and in general they can not be

explicitly enforced by existing FEM software. A detailed study and comparison with our in-house

Page 28: A Unified Thermodynamic Constitutive Model for SMA

28

numerical FEM code, has shown that these force constraints are implicitly enforced by applying the

kinematic boundary conditions given by equation (47), due to the way that the kinematic boundary

conditions are applied by operating on the global stiffness matrix.

4.2 Results and discussions

The results that follow the unit cell and boundary conditions discussed in Section 4.1 have

been used, while the constitutive response for the SMA fibers, presented in Section 2.2, have been

utilized, with material parameters given in Table 2. For the elastoplastic matrix a Von Mises yield

criterion with an associative flow rule and Prager-Ziegler kinematic hardening have been used with

material parameters, given in Table 2, corresponding to those of 6061 aluminum. The temperature

dependence of the material properties of aluminum has been considered for completeness (Military

Handbook, 1994). For the results presented here, the ABAQUS finite element program has been

utilized with an appropriate user supplied subroutine (UMAT) for the thermomechanical constitutive

response of the SMA fibers.

The temperature and mechanical loading history considered in the present work for finite

element analysis is schematically shown in Figs. 5 & 6, which is similar to the experimental processes

performed by Taya et al. (Taya et al. 1993, Furuya et al. 1993). Both axial and transverse loadings

are considered for tetragonal and hexagonal periodic arrangements. In the actual processing of the

composite residual stresses appear by annealing (austenite phase) at high temperature, i.e., 500 C,o

and then cooling to the point just before the forward transformation into martensite takes place, at

40 C. The effect of this process on the composite response, is also investigated in this study as can

be seen from loading step 1-2 in Fig. 6. Step 2-3 of the loading is to prestrain the composite. During

this step the SMA fibers fully transform into the martensitic phase. Residual stresses are expected

after unloading, at the end of step 3-4 due to the incompatible plastic deformation of aluminum matrix

and the transformation induced deformation (strains) of SMA fibers accumulated during loading step

2-3. Step 4-5 is introduced to simulate the elevated temperature service environment, at a

temperature above the austenitic finish temperature. Finally the composite is loaded mechanically at

elevated temperature in step 5-6.

Page 29: A Unified Thermodynamic Constitutive Model for SMA

29

In the results that follow, numbers on the graphs indicate the loading steps, while lower case

letters indicate the state of the SMA or aluminum matrix. The following letters are used: (a) for

aluminum yielding during loading, (b) for initiation of the forward stress induced transformation from

Page 30: A Unified Thermodynamic Constitutive Model for SMA

30

austenite to martensite, ( c) for the end of the forward transformation, (d) for aluminum yielding

during the composite unloading, (e) for initiation of the reverse transformation from martensite to

austenite, (f) for aluminum yielding at elevated temperature, and (g) for initiation of the forward stress

induced transformation from austenite to martensite at elevated temperatures.

The overall axial stress vs. strain curve for axial applied loading is shown in Fig. 7. Upon cooling

to 40 C from 500 C, the aluminum matrix develops residual tensile stresses whereas the SMA fibers

develop compressive stresses, due to the mismatch in the thermal expansion coefficients ( - =Al SMA

13.6x10 / C), as shown in Fig. 8 (points 2), for the average ( ) stress component for both-6 o33 zz

aluminum and SMA. This state of stress in SMA fibers causes stress induced phase transformation

at much higher temperature, about 39.7 C, than the stress-free martensitic start temperature, which

is about 18 C. The tensile residual stresses in aluminum matrix cause its early yielding at zero applied

axial stress due to residual tensile stresses (point a in Fig. 7). Stress induced martensitic phase

transformation in the SMA fibers starts at about 100 MPa overall applied stress (point b). The phase

transformation ends at about 170 MPa (point c), followed by elastic loading of the SMA fibers and

linear kinematic hardening of the aluminum matrix. Unloading of the composite takes place elastically

until aluminum yielding starts at about 160 MPa (point d). The reverse transformation back to

austenite starts at about 30 MPa (point e) but due to low temperature is only partially completed. The

temperature is then increased to 85 C and the SMA fibers fully transform into austenite at about 61

C, instead of the stress free finish temperature of 42 C. The recovery of the transformation straino o

in the SMA fibers induces axial compressive stress in the aluminum matrix, enhancing therefore the

required overall stress for composite yielding. When the composite is axially loaded, its overall yield

stress is increased to about 120 MPa (point f), which is much higher than the initial aluminum yield

stress value of 69 MPa (Table 2).

Contour plots of Von Mises stress and martensitic volume fraction, , for the applied axial loading

case are shown in Figs. 9 and 10, during step 2-3 at a point where the overall applied stress is about

140 MPa. Both the Von Mises stress and martensitic volume fraction in the SMA fibers are almost

uniform.

Page 31: A Unified Thermodynamic Constitutive Model for SMA

31

Page 32: A Unified Thermodynamic Constitutive Model for SMA

32

Fig. 11 is similar to Fig. 7 but for overall transverse loading applied to the composite. As seen

from Fig. 8, both the yield and the phase transformation points are not sharply defined in this case due

to highly nonuniform stress and inelastic strain fields in aluminum matrix. The contour plots of Von

Mises stress and martensitic volume fraction, , for the transverse loading case are shown in Figs. 12

and 13, during step 2-3 at a point where the overall applied stress is about 140 MPa. In the transverse

loading case it is observed that the Von Mises stress in the matrix is not uniform.

The comparison of the overall axial and transverse responses for both tetragonal and hexagonal

periodic arrangements in Figs. 14 and 15 reveal similar responses, however, tetragonal response is

stiffer compared to hexagonal response, for same fiber volume fraction. This can be explained by

observing the geometry of both arrangements. It can be seen that the placement of fibers in a

tetragonal periodic composite provides lines of stiffer resistance to applied force, unlike a hexagonal

periodic composite where the arrangement of fibers allows the softer matrix to contribute more to

the effective response, hence, a more compliant composite.

Page 33: A Unified Thermodynamic Constitutive Model for SMA

33

Finally, in Fig. 16 we compare the overall axial stress vs. overall axial strain for the loading case

described in Fig. 7, when the three SMA constitutive models discussed in Section 2 and whose

uniaxial response is plotted in Fig. 1 are utilized. One observes that the differences among the three

models during the forward and reverse transformations are less pronounced, with respect to the

monolithic SMA response shown in Fig. 1. This occurs mainly due to the fact that the aluminum

matrix has the same hardening response in all three cases, thus reducing the impact of the differences

among the SMA constitutive models to the overall composite response. Since the overall axial loading

is fiber dominated, the transverse loading case, which is matrix dominated in terms of deformations,

is expected to result to even less pronounced differences in the overall composite response for the

different SMA models.

The results obtained above agree qualitatively with the experimental data given by Taya et al.

(1993). The results acquired for the fiber volume fractions of 0.04 and 0.09, show a percentage

increase of about 52 % and 112 % respectively, in the effective yield stress of the composite relative

to the initial aluminum yield stress of 33 MPa, for the case of applied axial loading. The same trend

is observed in the present study, where for 0.3 fiber volume fraction, both tetragonal and hexagonal

periodic arrangements yield an increase of about 79 %, compared to the initial aluminum yield stress

of 70 MPa, for the case of applied axial loading.

5. CONCLUSIONS

A general thermodynamic constitutive model has been proposed based on the work done by

Boyd and Lagoudas (1994). Several earlier models can be identified by selecting different forms of

the function f( ) in the mixing free energy. The thermal energy output rates during forward (from

austenite to martensite) transformation predicted by these different models are compared with the

experimental data obtained from a calorimetric measurement. The results show that the different

models give substantially different predictions for the thermal energy output, even though the stress

strain curves predicted by the different models are similar to each other.

The effective thermomechanical response of Active Metal Matrix Composites (AMMC) with

SMA fibers has been modeled in this work. A 3-D constitutive model for SMA was employed in a

Page 34: A Unified Thermodynamic Constitutive Model for SMA

34

Page 35: A Unified Thermodynamic Constitutive Model for SMA

35

unit cell FEM analysis for a periodic AMMC to obtain the effective response of the elastoplastic

matrix / active fiber composite. A detailed procedure for implementing the constitutive model in the

finite element analysis scheme was given for both tetragonal and hexagonal periodic arrangements.

In addition to the overall thermomechanical response, the overall shape memory behavior of AMMC

was analyzed, and in particular its effect on the yield strength of the composite.

For all cases studied, it was demonstrated that both the axial and transverse prestraining have the

same strengthening effect for the final response of the composite, provided that the bonding between

matrix and SMA fibers is perfect. However, this condition may not be true for real composites

undergoing transverse loading due to the stress concentrations and large phase transformation strains.

Due to mechanical interactions, the phase transformation temperature range in AMMC has, finally,

been found to be much larger than the monolithic SMA.

Page 36: A Unified Thermodynamic Constitutive Model for SMA

36

Page 37: A Unified Thermodynamic Constitutive Model for SMA

37

6. ACKNOWLEDGMENTS

Page 38: A Unified Thermodynamic Constitutive Model for SMA

38

The authors acknowledge the financial support of the Army Research Office, contract No.

DAAL03-92-G-0123, monitored by Dr. G.L. Anderson.

7. REFERENCES

Barrett, D.J., 1994, " A One-Dimensional Constitutive Model for Shape Memory Alloys," to be

published.

Berveiller, M., Patoor, E., and Buisson, M., 1991, "Thermomechanical Constitutive Equations

for Shape Memory Alloys," Journal de Physique IV, Vol. 1, C.4,387, European Symposium

on Martensitic Transformation and Shape Memory Properties.

Bo, Z., and Lagoudas, D. C., 1994, "Comparison of Different Thermodynamic Models for Shape

Memory Alloys", Adaptive Structures and Composite Materials, Analysis and Applications,

E. Garcia, H. Cudney and A. Dasgupda, eds., AD-Vol. 45, MD-Vol. 54, pp. 9-19.

Bo, Z., and Lagoudas, D. C. 1995, “A Thermodynamic Constitutive Model for Cyclic Loading

of Shape Memory Alloy Materials with Application to Two Way Training,” Smart Structures

and Materials 1995, Proceedings of the 1995 SPIE North American Conference on Smart

Structures and Materials, San Diego, CA, SPIE Vol. 2441, pp. 118-128.

Bondaryev, E.N., and Wayman C.N., 1988, “Some Stress-Strain-Temperature Relationships for

Shape Memory Alloys,” Metallurgical Transactions A, Vol. 19A, pp. 2407-2413.

Boyd, J.G., and Lagoudas, D.C., 1994, "Thermomechanical Response of Shape Memory

Composites," Journal of Intelligent Materials and Structures, Vol. 5, pp. 333-346.

Boyd, J.G., and Lagoudas, D.C., 1995, "A Thermodynamic Constitutive Model for the Shape

Memory Materials. Part I. The Monolithic Shape Memory Alloys," to appear in International

Journal of Plasticity.

Boyd, J.G., and Lagoudas, D.C., 1995, "A Thermodynamic Constitutive Model for the Shape

Memory Materials. Part II. The SMA Composite Materials," to appear in International

Journal of Plasticity.

Buehler, W.J., and Wiley, R.C., 1965, "Nickel-Base Alloys," U.S. Patent 3,174,851. E d e l e n ,

D.G.B., 1974, "On the Characterization of Fluxes in Nonlinear Irreversible Thermodynamics,"

Page 39: A Unified Thermodynamic Constitutive Model for SMA

39

International Journal of Engineering Science, Vol. 12, pp. 397-411.

Fischer, F.D., Berveiller, M., Tanaka, K., Oberaigner, E.R., 1994, "Continuum Mechanical

Aspects of Phase Transformations in Solids," Archive of Applied Mechanics, Vol. 64, pp.54- 85.

Furuya, Y., Sasaki, A., and Taya, M., 1993, "Enhanced Mechanical Properties of TiNi Shape

Memory Fiber/Al Matrix Composite," Materials Transactions, The Japan Institute of Metals,

Vol. 34, No. 3, pp. 224-227.

Graesser, E.J., and Cozzarelli, F.A., 1991, "Shape-Memory Alloys as New Materials for Seismic

Isolation," Journal of Engineering Mechanics, Vol. 117, No. 11, pp. 2590-2608.

Howard, S. D., 1995, “The Thermomechanical Constitutive Experimentation of NiTi Shape

Memory Alloy Strips and Rods,” Master’s Thesis, Texas A&M University.

Jackson, C.M., Wagnar, H.J., and Wasilewski, R.J., 1972, “55-Nitinol-The Alloy with a

Memory: Its Physical Metallurgy, Properties, and Applications,” NASA-SP 5110.

Lagoudas, D.C., Boyd, J.G. and Bo, Z., 1994a, "Micromechanics of Active Composites with SMA

Fibers," ASME Journal of Material Science and Technology, Vol. 116, No. 3, pp. 337- 347.

Lagoudas, D.C., Bo, Z., and Qidwai, M.A., 1994b, “Micromechanics of Active Metal Matrix

Composites with Shape Memory Allot Fibers,” Inelasticity and Micromechanics of Metal Matrix

Composites, G.Z. Voyiadjis and J.W. Ju, eds., pp. 163-190.

Liang, C., and Rogers, C.A., 1990, "One-Dimensional Thermomechanical Constitutive Relations

for Shape Memory Materials," Journal of Intelligent Material Systems and Structures, Vol.

1, pp. 207-234.

Liang, C., and Rogers, C.A., 1992, "The Multi-Dimensional Constitutive Relations of Shape

Memory Alloys," Journal of Engineering Mathematics, Vol. 26, pp. 429-443.

McCormick, P.G., and Y. Liu, 1994, “Thermodynamic Analysis of the Martensitic

Transformation in TiNi-II. Effect of Transformation Cycling,” Acta Metallurgica el Materialia,

Vol. 42, No. 7, pp. 2407-2413.

Military Handbook, 1994, “Metallic Materials and Elements for Aerospace Vehicle Structures,”

Department of Defense, United States of America.

Miyazaki, S., Otsuka, K., and Suzuki, Y., 1981, “Transformation Pseudoelasticity and

Deformation Behavior in Ti-50.6at%Ni Alloy,” Scripta Metallurgica, Vol. 15, pp.287-292.

Page 40: A Unified Thermodynamic Constitutive Model for SMA

40

Ortin, J. and Planes, A., 1989, "Thermodynamics of Thermoelastic Martensitic Transformations,"

Acta Metall., Vol. 37, No. 5, pp. 1433-1441.

Ortin, J., and Planes, A., 1991, "Thermodynamics and Hysteresis Behavior of Thermoelastic

Martensitic Transformations," Journal de Physique IV, Vol. 1, Colloque C.4, 13, European

Symposium on Martensitic Transformation and Shape Memory Properties.

Ortiz, M., and Simo, J.C., 1986, "An Analysis of a New Class of Integration Algorithms for

Elastoplastic Constitutive Relations," International Journal for Numerical Methods in

Engineering, Vol. 23, pp. 353-366.

Raniecki, B., and Lexcellent, C., 1994, "R -Models of Pseudo elasticity and Their SpecificationL

for Shape Memory Solids," European Journal of Mechanics, A, Vol. 13, No.1, p. 21.

Salzbrenner R.L., and Morris Cohen, 1979, “On the Thermodynamics of Thermoelastic

Martensitic Transformations,” Acta Metallurgica, Vol. 27, pp. 739-748.

Sato, Y., and Tanaka, K., 1988, "Estimation of the Energy Dissipation in Alloys Due to

Stress-induced Martensitic Transformation," Res Mechanica, Vol. 23, pp. 381-393.

Shaw J. A., and Kyriakides, S., 1995, “Thermomechanical Aspects of NiTi,” Journal of

Mechanics and Physics of solids, Vol. 43, No. 8, pp. 1243-1282.

Stalmans, R., Humbeeck V. J., and Delaey, L., 1992, “The two way memory effect in Copper

based shape memory alloys--thermodynamics and mechanics”, Acta Metallurgica el

Materialia, Vol. 40, No. 11, pp. 2921-2931.

Sun, Q.P., and Hwang, K.C., 1993, "Micromechanics Modelling for the Constitutive Behavior

of Polycrystalline Shape Memory Alloys-I. Derivation of General Relations," Journal of

Mechanics and Physics of solids, Vol. 41, No. 1, pp.1-17.

Sun, Q.P., and Hwang, K.C., 1993, "Micromechanics Modelling for the Constitutive Behavior

of Polycrystalline Shape Memory Alloys-II. Study of the Individual Phenomena," Journal of

Mechanics and Physics of Solids, Vol. 41, No. 1, pp. 19-33.

Tanaka, K., 1986, "A Thermomechanical Sketch of Shape Memory Effect: One-Dimensional Tensile

Behavior," Res Mechanica, Vol. 18, pp. 251-263.

Taya, M. , Furuya, Y., Yamada, Y., Watanabe, R., Shibata, S., and Mori, T., 1993, "Strengthening

Mechanisms of TiNi Shape Memory Fiber/Al Matrix Composite," Proceedings of Smart

Page 41: A Unified Thermodynamic Constitutive Model for SMA

41

Materials, SPIE, eds., V.K. Varadan, Vol. 1916, pp. 373-383.

Truesdell, C., and Noll, W., “The Non-linear Field Theories of Mechanics,” in Encyclopedia of

Physics, ed. S. Flügge, Vol. 3/3, Berlin: Springer-Verlag, 1965.

Wayman, C.M., 1983, "Phase Transformations, Nondiffusive," in Physical Metallurgy," R.W.

Cahn and P. Haasen, eds., North-Holland Physics Publishing, NY, pp. 1031-1075.

Zienkiewicz, O.C., 1977, “The Finite Element Method,” 3rd., McGraw-Hill, London.

Page 42: A Unified Thermodynamic Constitutive Model for SMA

42

List of Figures

Fig. 1 Stress-temperature diagram for transformation start and finish lines.

Fig. 2 Uniaxial stress-strain behavior for different hardening laws.

Fig. 3 Heat energy output rate during austenite to martensite phase transformation.

Fig. 4 Tetragonal and hexagonal periodic arrangement with respective unit cells.

Fig. 5 Schematic representation of the loading history used in the analysis.

Fig. 6 Applied load and temperature history used in the analysis.

Fig. 7 Overall axial stress vs. axial strain for the case of applied axial loading.

Fig. 8 SMA and matrix axial average stress vs. overall axial strain for the case of applied axial

loading.

Fig. 9 Contour plot of Von-Mises stress for the case of applied axial loading at 140 MPa and 44 C.o

Fig. 10 Contour plot of martensitic volume fraction for the case of applied axial loading at 140 MPa

and 44 C.o

Fig. 11 Overall transverse stress vs. transverse strain for the case of applied transverse loading.

Fig. 12 Contour plot of Von-Mises stress for the case of applied transverse loading at 140 MPa and

44 C.o

Page 43: A Unified Thermodynamic Constitutive Model for SMA

43

Fig. 13 Contour plot of martensitic volume fraction for the case of applied transverse loading at 140

MPa and 44 C.o

Fig. 14 Comparison of overall axial stress vs. axial strain for the case of applied axial loading for

tetragonal and hexagonal pariodic arrangements.

Fig. 15 Comparison of overall transverse stress vs. transverse strain for the case of applied axial

loading for tetragonal and hexagonal pariodic arrangements.