Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/325530123

Dynamic Magnetic Shape Memory Alloys Responses: Eddy Current Effect and
Joule Heating

Article  in  Journal of Magnetism and Magnetic Materials · June 2018


DOI: 10.1016/j.jmmm.2018.05.107

CITATIONS READS

10 244

2 authors:

Krishnendu Haldar Dimitris C. Lagoudas


Indian Institute of Technology Bombay Texas A&M University
21 PUBLICATIONS   193 CITATIONS    268 PUBLICATIONS   7,526 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Active Origami Antennas View project

All content following this page was uploaded by Krishnendu Haldar on 08 June 2018.

The user has requested enhancement of the downloaded file.


Dynamic Magnetic Shape Memory Alloys
Responses: Eddy Current Effect and Joule
Heating

Krishnendu Haldar1∗and Dimitris C. Lagoudas2


1 Department of Aerospace Engineering,
Indian Institute of Technology Bombay, Mumbai 400076, India.
2 Department of Aerospace Engineering & Department of Materials Science and Engineering
Texas A&M University, College Station, TX 77843, USA.

Key Words: MSMA, variant reorientation, dynamic responses, eddy


current, Joule heating

ABSTRACT

Generating high actuation frequency (∼ 1.0kHz) is one of the potential appli-


cations of Magnetic Shape Memory Alloys (MSMAs). In this work, dynamic
responses of single crystal MSMAs due to variant reorientation are investi-
gated. Time dependent part of the Maxwell equations becomes significant
for a high frequency regime. Generation of an electric field and magnetic flux
linkage due to the motion of the material points during deformation create
a complex electro-magneto-mechanical coupling mechanism. We perform a
thermodynamically consistent study to capture the variation of electromag-
netic fields due to the deformation in the presence of fluctuating magnetic
field, mainly focusing on eddy current and Joule heating. A comparison
of MSMA responses with a typical ferromagnet/magnetostrictive material
responses is discussed.

1 Introduction
MSMAs are best known for their unique ability to produce Magnetic Field
Induced Strains (MFIS) up to 10% under a magnetic field [1–4]. Some of the
commonly used MSMA material systems are NiMnGa [5–10], FePd [11–15]

Corresponding author. Email: krishnendu@aero.iitb.ac.in

1
and NiMnX, where X = In, Sn, Sb [16–18].Among them, NiMnGa alloys are
widely investigated and will be the main focus of this study. Martensitic
transformations in Ni2 MnGa alloys were first reported by Webster et al. [19].
Zasimchuk et al. [20] and Martynov and Kokorin [21] performed detailed
studies on the crystal structure of martensite in the Ni2 MnGa alloys. Ullakko
et al. [22, 23] are credited with first suggesting the possibility of a magnetic
field-controlled shape memory effect in these materials. The unique magneto-
mechanical coupling makes MSMAs promising materials for multifunctional
structures, actuator and sensor applications [24–28].
The coupled MSMA behaviors can be modeled by considering the mate-
rial as an electromagnetic continuum. Extensive work on different electro-
magnetic formulations had been proposed in the literature [29–33] based on
different notion of breaking up long range and short range forces. A contin-
uum theory for deformable ferromagnetic materials and nonlinear magneto-
elasticity for magneto sensitive elastomers are discussed in [34–37]. On the
other hand, a study of electrostatic forces on large deformations of polariz-
able materials is discussed in[38, 39]. A theory for the equilibrium response of
magneto-elastic membranes is formulated by [40, 41] and a continuum theory
for the evolution of magnetization and temperature in a rigid magnetic body
for ferro/paramagnetic transition could be found in [42]. The variational
formulations for general magneto-mechanical materials have been proposed
by many authors [43–47].
The macroscopically observable MFIS in MSMAs is caused by the mi-
crostructural reorientation of martensitic variants [2, 48], field induced phase
transformation [17, 49–52] or a combination of the two mechanisms. In this
work, we will focus on variant reorientation. In the variant reorientation
mechanism, the variants have different preferred directions of magnetization
and the magnetic field is applied to select certain variants among others,
which results in the macroscopic shape change.
There are two major modeling approaches for variant reorientation mech-
anism. In microstructural based models, the resulting macroscopic strain
and magnetization response are predicted by minimizing a free energy func-
tional. Details on the microstructural based modeling approach can be found
in [1, 11, 53–56]. The second approach to study the material behavior is
through thermodynamics based phenomenological modeling [57–61]. Most
recent model development of variant reorientation in MSMAs and magneto-
mechanical loading analysis is reported in [62–64]. Moreover, a detailed
anisotropic consideration due to single crystal discrete symmetry is recently
presented in [65, 66]. A brief modeling approach for bulk polycrystalline ma-
terials is also mentioned in [65]. In this study, we systematically consider the
time dependent Maxwell equations and the rate forms of the mechanical and

2
magnetization constitutive equations that include single crystal anisotropy
[66].
Voltage generation (∼ 100 [V]) in MSMA due to mechanical impulse (∼
1.4 [m/s]) inside a biased magnetic induction (∼ 1.5 [T]) is reported in [67].
A similar study of stress induced variant reorientation under a fixed biased
magnetic field (1.6 [T]) is discussed in [28]. They found maximum voltage of
280 [mV] with the strain range of 4.9% at 10 [Hz]. Other relevant studies re-
garding voltage generation of MSMAs as sensors are reported in [68, 69]. Best
of our knowledge, effect of eddy current and corresponding Joule heating are
not so far reported for high-frequency actuation condition. Our main contri-
butions in this article are to predict the above mentioned quantities, based
on experimentally verified quasistatic constitutive equations, and to analyze
other field variable responses under dynamic conditions. We also compare
eddy current and Joule heating due to MSMA actuation with the 430-steel
(rigid ferromagnet) and Terfenol-D (magnetostrictive) material responses.

2 Electro-magneto-mechanical field equations


for dynamic system
We denote the reference configuration B0 , which is assumed to be free from
any externally applied stimuli, and the current configuration is denoted by
Bt . The body consists of material points X ∈ B0 . The spatial position in
the deformed configuration is denoted by x = ϕ(X, t), with t representing
time, so that the deformation gradient can be defined as F = ∇X ϕ with
J = det(F ) > 0.
In the deformed configuration Bt , we denote the magnetic induction by
b(x, t), the magnetic field by h(x, t), the magnetization vector by m(x, t),
and the electric displacement by d(x, t) with respect to a rest frame. We
consider that the body is only magnetizable and there are no polarization
and free charge. Magnetic and electric constitutive relations are written as

m = b/µ0 − h, d = ε0 e. (1)

Here µ0 is the magnetic permeability, and ε0 is the electric permittivity of


the free space. An electric field will be present due to the rate of change
of magnetic flux (Faraday’s law). Let a generic point P ∈ Ωt move with a
velocity ẋ during deformation. The motion of a magnetic body under applied
magnetic field will further produce motional electromotive force (emf), ẋ×b.
e = e + ẋ × b. At a moving point, the
The total electric field then becomes e

3
remaining electromagnetic variables {e
j f , h,
e e f} are written as
b, m

jf = jf ,
e b = b − c−2 (ẋ × e) ≈ b,
e m
f = m, e = h − ẋ × d
h

with respect to the rest frame variables {j f , h, b, m}, where j f is the free
current. The moving frame fields ej f and m
f are decoupled with the velocity
ẋ as there are no free charge qf and electrical polarization p, respectively.
The local form of the Maxwell equations are:
∗ ∗
∇x · d = 0, ∇x · b = 0, ∇x × h
e=e
j f + d, ∇x × e
e = −b, (2)

where, for a general vector α, convective time derivative is denoted by α =
α̇ − Lα + α tr (L) with α̇ as total time derivative. In a moving frame Ohms
law is given by e j f = Ωee = j f , where Ω is the conductivity tensor and
positive definite. We further write

∇x × e
e = −b ⇒ ∇x × e = −∂t b,

∇x × h
e=e j f + d ⇒ Ω−1 ∇x × h = e
e + 0 Ω−1 ∂t e ⇒ ∇x × h = Ωe
e.

Note that 0 ∼ 10−12 and for a conductor |Ω| ∼ 107 . This means the product
0 Ω−1 ∼ 10−19 and thus negligible.
The conservation laws of mass, linear momentum, angular momentum
and energy are:

ρ̇ + ρ∇x · ẋ = 0, ∇x · σ + ρf b = ρġ, skw (σ) = skw (ρg ⊗ ẋ), (3)


0
ρu̇ + ∇ · q − σ L : L − e
jf · e f · ḃ − ρrh = 0.
e+m (4)

The generalized momentum density is given by


0 0
g = ẋ + e×b=v+ e×b (5)
ρ ρ
and the total stress is decomposed as
0 0
σ = σL + σM (6)
0
where, σ L is the work conjugate of the velocity gradient L and the general-
0
ized Maxwell stress σ M is given by,
   
M0 b⊗b b · b 0 e · e
σ = − m ⊗ b + 0 e ⊗ e + 0 e × b ⊗ ẋ − + − m · b I. (7)
µ0 2µ0 2

4
2.1 Thermodynamic framework and constitutive equa-
tions
The constitutive responses with respect to a potential ψ(F , b, s, {ζ}), where
s is the entropy and the set {ζ} represents the collection of internal state
variables, are obtained through Coleman and Noll procedure [70],
0
σ L = ρ∂F ψF T , µ0 m = −ρ∂b ψ, s = −∂T ψ, −∂ζi ψ · ζ˙i > 0. (8)

After performing Legendre transformations [65, 71], with respect to a Gibbs


free energy G(S E , H, T, {Z}) the following constitutive equations are ob-
tained

E = −ρ0 ∂SE G, µ0 M = −ρ0 ∂H G, s = −∂T G, −ρ∂Zi G · Żi > 0. (9)

Here, H = F T h is the magnetic field, E = 21 (F T F − I) is the Green


0
strain, S E = JF −1 σ E F −t is the work conjugate of E, where σ E = σ L +
µ0 h ⊗ m + µ20 (m · m)I and {Z} is the set of internal variables in B0 . The
Gibbs free energy for MSMA consists with the set of internal state variables
{Z} = {E r , M r , ξ, g}.
In a typical experimental condition, the MSMA sample, initially at austenitic
phase, is subjected to a constant compressive mechanical load. A phase
transformation into the martensitic variant takes place by cooling down the
temperature. This particular martensitic variant, among the other three,
is called stress favored variant or variant-1. Then a magnetic field is ap-
plied along the perpendicular to the loading direction. After a critical value,
field induced variant (or variant-2) nucleates and both variants coexist by
forming a twinned microstructure. Complete variant reorientation to field
induced variant takes place at a sufficiently high applied magnetic field. By
switching from variant-1 to variant-2 with magnetic field thus generates high
actuation strain or reorientation strain with hysteresis.
Variant-1 in the initial configuration only contains 180o domain walls as
there is no remnant magnetization before applying any magnetic field. When
a magnetic field is applied, the magnetization vectors start rotating in each
domain from the easy axis to the hard axis. The associated energy is known
as Magneto-crystalline Anisotropy Energy (MAE). Once the critical field for
the variant reorientation has been reached, the field favored variant nucleates,
and the structural rearrangement takes place to form an intermediate twin
structure forming 900 domain walls. When MAE is sufficient to overcome the
energy required for twin boundary motion, magnetic field favored martensitic
variant (variant-2) grows, and field induced macroscopic shape change is
observed. The applied field for the completion of the reorientation process

5
is around 1.5 [T]. In MSMA, however, the magnetic domain wall motion
appears to be associated with a a minimal amount of dissipation. More
detail could be found in [7, 72].
We denote the reorientation strain tensor by E r and internal magneti-
zation vector due to reorientation by M r . ξ denotes the volume fraction of
the reoriented field-favored martensitic variant (variant-2) and the scalar g
is an internal state variable associated with the interaction (mixing) energy
during reorientation [65]. We assume that the reorientation strain rate Ė r ,
the rate of internal magnetization vector Ṁ r and the rate of mixing energy
g obey the following flow rules
˙
Ė r = Λr ξ, ˙
Ṁ r = γ r ξ, ˙
ġ = f r ξ. (10)

The tensors Λr describes the direction and magnitude of the strain generated
during variant reorientation, γ r takes into account the direction and magni-
tude of the internal magnetization due to reorientation respectively and f r
is the hardening function. Expanding the entropy inequality (9d), we get

π Er : Ė r + π M r · Ṁ r + πξ ξ˙ + πg ġ ≥ 0. (11)

The thermodynamic driving forces are denoted by

π Er = −ρ0 ∂Er G π M r = −ρ0 ∂M r G πξ = −ρ0 ∂ξ G πg = −ρ0 ∂g G.

If the above relations are substituted in to (11), we get

π Er : Λr ξ˙ + π M r · γ r ξ˙ + πξ ξ˙ + πg f r ξ˙ ≥ 0, ⇒ π r ξ˙ ≥ 0,

where the total thermodynamic driving force π r due to reorientation is given


by

π r = π E r : Λ r + π M r · γ r + πξ + πg f r . (12)

The following reorientation function, Φ r , is then introduced,

π r − Y r , ξ˙ > 0
(
r r
Φ := , Φ ≤ 0, (13)
−π − Y , ξ˙ < 0
r r

where Y r is a positive scalar associated with the internal dissipation during


reorientation and can be found from calibration. The proposed transforma-
tion function is similar to the transformation function used with conventional
shape memory behavior [73, 74]. It is assumed that the constraints of the

6
reorientation process follows the principle of maximum dissipation and can
be expressed in terms of the Kuhn Tucker type conditions [75]

Φ ξ˙ = 0 .
r r
Φ ≤ 0, (14)

The martensitic phase of Ni2 MnGa has 10M structure with I4/mmm
space group [76]. The classical point group is 4/mmm (D4h ). The discrete
symmetry restrictions for a specific Gibbs free energy potential, constitutive
equations and evolution equations are elaborately discussed in [65, 66]. The
incremental constitutive equations for single crystal anisotropy can be written
in the following forms:
˙ = [L]bEc
bSc ˙ + [K]Ḣ, (15)
˙ + [K0 ]Ḣ.
Ṁ = [L0 ]bEc (16)

Here [L]6×6 is a mechanical tangent stiffness matrix, [K]6×3 and [L0 ]3×6 are
magneto-mechanical stiffness matrices, and [K0 ]3×3 is a magnetic stiffness
matrix. In a general way, half-vectorization of a n × n symmetric matrix A
is denoted by

vech(A) = [A11 , ..., An1 , A22 , ..., An2 , ..., A(n−1)(n−1) , A(n−1)n , Ann ]T

and we write for simplicity vech(A) = bAc.

3 Small deformation approximation and quasi-


static responses
We reduce the modelh by considering
i infinitesimal strain approximation for
1 T
which E ≈ ε = 2 ∇u + (∇u) , S ≈ σ E , H ≈ h and M ≈ m. Under
E

these conditions, we have the following system of equations:

∇·b = 0, ∇ · d = 0, ∇ × h = Ωe e (17)
∇×e = −∂t b, ∇ · σ = ρv̇, skw (σ) = 0 (18)
σ̇ = [L](ε̇) + [K]ḣ (19)
ṁ = [L0 ](ε̇) + [K0 ]ḣ. (20)

We further write from (6)


0 0
n µ0 o 0
σ = σ L + σ M = σ E − µ0 h ⊗ m − (m · m)I + σ M = σ E + σ h ,
2

7
where, modified Maxwell stress σ h takes the following form
0 µ0
σ h = σ M − µ0 h ⊗ m − (m · m)I
2 i h
h µ0 0 i
= µ0 h ⊗ h − (h · h)I + 0 e ⊗ e − (e · e)I .
2 2
The rate form of the MSMA constitutive relations ((19),(20)) and the com-
ponents of [L], [K], [L0 ] and [K0 ] can be found in [65, 66]. In 2-D
   
1 ν 0  max
 0 −1
E   , [K] = E∆Mε
[L] = 2
ν 1 0 + µ0 hy 0 1  ,(21)
1−ν 1 r(1 + ν)
0 0 2 (1 + ν) 0 0
   
0 0 0 0 0 0 a1 ∆M/r
[L ] = , [K ] = . (22)
0 0 0 0 −b1 ∆M/r
The material parameters a1 = µ0 MxC , b1 = µ0 ∆M = µ0 (M sat − MyC ), where
MxC is the x-component of the magnetization response at forward reorien-
tation start and ∆M = (M sat − MyC ) + µ0 hy is the difference between the
saturation magnetization and the y-component of the magnetization at for-
ward reorientation start . The parameter r = ∂Φ ∂ξ
, is an explicit function
of the internal variable ξ, which depends on the magnetic field hy . The
functional form r(ξ(hy )) depends on the selection of the hardening function.
Thermodynamic force calculation and selection of hardening function are
briefly presented in Appendix A.

3.1 Quasistatic conditions and constitutive responses


A half space MSMA conductor is considered for 1-D approach (Fig. 1). Al-
though the geometry is idealized, its simplicity allows relatively simple semi-
analytic solutions to gain a useful insight into the processes. The x = 0
plane is kept fixed, while a biasing compressive stress σ ∗ is applied at x = L
along the negative x direction. A magnetic field h0y (t) is then applied in the
y-direction such that it produces actuation due to reorientation along the
x-axis. ĥy (x, t) is a perturbation field due to the presence of the electric field
ẽz . However, for a quasistatic case, both ĥy (x, t) and ẽz are zero. We consider
a linearly applied field in the form
h0y (t) = c1 t + c2 .
Considering µ0 hy = HsM2 at t = ts and µ0 hy = HfM2 at t = tf , we can write
HfM2 − HsM2 HsM2 tf − HfM2 ts
µ 0 c1 = , µ0 c2 = . (23)
tf − ts tf − ts

8
Figure 1: Schematic representation of the geometry and assumptions on the
field variables. Infinitely extended y and z planes are considered to avoid
electromagnetic edge effects.

The reduced 1-D constitutive equations become (from (19) and (20)):

(a) (b)

Figure 2: Predictions of (a) magnetization and (b) strain responses for variant
reorientation mechanism (after experimental data correction [77]) with an
incremental method.

9
∆Mεmax
0 = ε̇xx + µ0 ḣy , (24)
 r 
[∆M ][∆M]
ḃy = µ0 1 − µ0 ḣy . (25)
r
We assume that total stress (component σxx ) inside the specimen is con-
stant and is maintained at σ ∗ = −2 MPa to trigger field induced variant
reorientation and so σ̇xx = 0. Equation (24) reads
εmax ∆M εmax ∆M
Z
ε̇xx = ∂x v = −µ0 ḣy , ⇒ εxx = − µ0 dhy (26)
r r
and from (25) we write
Z  
[∆M ][∆M]
by = µ 0 1 − µ 0 dhy . (27)
r
We present the quasistatic responses for magnetization and strain, obtained
in an incremental way, in Fig. 2. Comparisons of predicted responses with
experimental data and data corrections due to demagnetization effect are
discussed in [77].

4 Dynamic effect on MSMA actuation


In this section, we consider the dynamic MSMA responses. The geometry
of the body is same as presented in Fig. 1. We further assume that inertia
effect is negligible and conservation of linear momentum is not taken into
account. A motional emf vby is generated when high actuation is considered.
Consequently, a magnetic field ĥy (x, t) is created due to the presence of eddy
current and is added up with the existing magnetic field. The total field
becomes

hy (x, t) = h0y (t) + ĥy (x, t) = c1 t + c2 + ĥy (x, t). (28)


1
In this context, Faraday’s law and Ampere’s law reduce to

∂x ez = ∂t by [V /m2 ], (29)
∂x hy = Ωeez [A/m2 ]. (30)
Units: It is always helpful to keep track of the units. We have 0 = 8.85 · 10−12 F/m,
1

µ0 = 4π · 10−7 N/A2 and c = 3 · 108 m/s. Note that F=A.s/V and Volt can further be
written as V =N.m/A.s. Electric field e is V/m and magnetic induction b is T=N/A.m.
Magnetization m and magnetic field h have unit A/m. The unit for current density j is
A/m2 . The conductivity Ω is S/m, where Simens S=A/V.

10
Denoting

β1 = c1 /µ0 [A/m.s],
∆Mεmax
β2 = −µ0 [m/A],
 r 
[∆M ][∆M]
µr = µ0 1 − µ0 [N/A2 ],
r

we solve for ez with a given ∂t by (equation (25)) from the following equation

∂x ez = ∂t by = µr ∂t hy , (31)

where ∂t hy = c1 + ∂t ĥy . The velocity due to actuation is obtained from (26)


as

∂x v = β2 ∂t hy . (32)

Knowing the velocity we then compute eez = ez + vby , from which we find ĥy
from (30)

∂x ĥy = Ωe
ez . (33)

We solve all the above equations numerically in a staggered way, and the
outline is presented in Tab. 2. Material constants are given in Table. 1. The

µ0 HsM2 = 0.6 [T], µ0 HfM2 = 0.9 [T], µ0 HsM1 = 0.75 [T], µ0 HfM1 = 0.17 [T]

M sat =742 [kA/m], K =0.5 [T]−1 , H C = HsM2 , εmax = 5.65%, E=2.0 [GPa]

n1 = 6.6, n2 = 1.4, n3 = 0.8, n4 = 0.3, (Hardening exponents, Appendix A)

Ω = 1.43 · 107 [S/m], σ ∗ = −2.0 [MPa], ρ = 1.206 · 104 [kg/m3 ]

Table 1: Used material constants in the analysis. The values of the magnetic
parameters can be found in [66, 77]. The value for conductivity is taken from
[78, 79].

initial-boundary values are

IC : ĥy (x, 0) = 0 x ∈ [0, L]


BC : ez (0, t) = 0, vx (0, t) = 0, ĥy (0, t) = 0 t ∈ [0, +∞) (34)

11
1. Given external applied magnetic field h0y = c1 t + c2 .
(1,n) (1,n) (1,n)
2. Initialize v (1,n) = 0, ez = 0, ĥy = 0, ∆ĥy = 0.

3. For i = 1 : M (space discretization, along length)

4. For n = 1 : N , (time discretization)


(i,n) 0, (i,n) (i,n) (i,n) (i,n)
5. Find: hy = hy + ĥy , ∆hy = c1 ∆t + ∆ĥy .
(i,n)
6. Find: [∂t hy ](i,n) = c1 + ∆ĥy /∆t.
(i,n) (i,n) (i,n) (i,n)
7. Compute: ξ (i,n) , r(i,n) , β1 , β2 , µr . (all are function of hy )
(i,n) (i,n) (i,n)
8. Compute: eez = ez + v (i,n) by .
(i,n+1) (i,n) (i,n) (i,n)
9. Compute: by = by + µr ∆hy ,
(i,n+1) (i,n) (i,n) (i,n)
εxx = εxx + β2 ∆hy .
(i,n)
10. From (32), find: v (i+1,n) = v (i,n) + [β2 [∂t hy ](i,n) ]∆x.
(i+1,n) (i,n) (i,n)
11. From (31), find: ez = ez + [µr [∂t hy ](i,n) ]]∆x.
(i+1,n) (i,n) (i,n)
12. From (33), find ĥy = ĥy + Ωe
ez ∆x.

13. Update: v (i+1,n+1) = v (i+1,n) , ĥ(i+1,n+1) = ĥ(i+1,n)


(i+1,n+1) (i+1,n)
ez = ez .

Table 2: Numerical scheme to solve the coupled electro-magneto-mechanical


problem in a staggered way.

Next, we predict some simulations for dynamic conditions. Our magnetic


range of interest is between reorientation start and finish. Time at HsM2 is
considered as t = 0 and at HfM2 , t = ts−f . The time dependent applied field
rate is considered in the form of
ḣ0y = c(t)
such that
Z
h0y = c(t)dt + c2 . (35)

We consider a constant rate c(t) = c1 for which the values of c1 and c2 are the

12
same as given in (23). We present two cases where two different lengths are
considered without altering the frequency (25 Hz, equivalent to ts−f =0.02 s)

4.1 Case-I: Small length and low frequency

(a) (b)

Figure 3: (a) Volume fraction of the field induced martensitic variant


(variant-2) and (b) velocity distribution due to actuation.

In this case, the length of the bar is considered 25 mm and the frequency is
25Hz. The plot for the field induced volume fraction is presented in Fig. 3(a).
The distribution along the length almost remains uniform for the entire time
period. Note that, along the length, each material point has an electric field,
which creates an induced magnetic field ĥy . The induced magnetic field thus
nucleates more field induced variants and an increase in volume fraction is
observed along the length. The velocity distribution along the time axis is
increasing-decreasing by nature Fig. 3(b). This means, the velocity increases
initially with the rate of nucleation of the new field induced variant and then
gradually slows down at the end. Induced magnetic field due to the motion
is presented in Fig. 4(a) and the Maxwell stress is given in Fig. 4(b). Both of
them are significant. The maximum Maxwell stress turns out to be around
0.4 [MPa], 20% of the -2 [MPa] blocking stress. The eddy current density
along the transverse direction is shown in Fig. 5(a). The predicted current
density is large. The current flows along the width of the bar and the body
is heated up due to the resistance. The Joule heating is given in Fig. 5(b)
and is significant too. A comparison with the standard steel will be provided
as the work progress (see Table. 3). This means the increased temperature

13
(a) (b)

Figure 4: (a) Induced magnetic field due to induced electric field and (b)
Maxwell stress in the x-direction.

(a) (b)

Figure 5: (a) Eddy current density and (b) corresponding Joule heating.

(heat) may facilitate to nucleate thermally induced phase transformation.


However, we did not consider thermal coupling in this study.

4.2 Case-II: Large length and low frequency


We increase the length four times, i.e., 100 mm while keeping the frequency
same. The volume fraction distribution becomes strongly nonuniform due

14
(a) (b)

Figure 6: (a) Volume fraction of the field induced martensitic variant


(variant-2) and (b) velocity distribution due to actuation.

to the increase in length (Fig. 6a). The distribution indicates that complete
reorientation (ξ = 1) takes place almost everywhere after half of the length.
Once the complete reorientation takes place, actuation stops and a sharp
instantaneous drop in velocity to zero is observed (Fig. 6b).

(a) (b)

Figure 7: (a) Induced magnetic field due to induced electric field and (b)
Maxwell stress in the x-direction.
The induced magnetic field becomes high, and it reaches its maximum
value around 0.8 [T] (Fig. 7a). Maxwell stress is given in Fig. 7(b). The eddy

15
current density (Fig. 8b) increases almost twice by increasing the length four
times. The Joule heating also increases than the previous case (Fig. 8b).

(a) (b)

Figure 8: (a) Eddy current density and (b) corresponding Joule heating.

One should notice that a sawtooth is created around x=65 mm for both
current and power responses. This is due to the fact that the velocity becomes
zero after this point and so the motional emf drops. This is mainly the part
Ωvx by that disappears due to zero velocity. However, the presence of applied
field rate contributes in the increasing of jf in spite of zero velocity. The
same argument holds true for the power P .
Note that for ĥy , no sharp drop is observed. As ∂∂xĥy
= Ω(ez + vx by ), it
turns out that the contribution from Ωvx by is very small (in the order of
10−2 ) compared to the static part Ωez .

4.3 Case-III: Small length and high frequency


We observed in the previous cases that the length has a significant effect on
the output variables. Now we keep the length of the body same as Case-I,
25 mm, but consider a frequency 500Hz. The volume fraction and velocity
are plotted in Fig. (9). As the velocity becomes very high, we suspect that
inertia may play an important role in this regime. Recall that we assumed
conservation of linear momentum is trivially satisfied due to constant stress
assumption and negligible inertial force. The linear momentum equation
reads

∂x σxx = ρv̇x = ρ[∂t vx + vx ∂x (vx )]

16
(a) (b)

Figure 9: (a) Volume fraction of the field induced martensitic variant and
(b) velocity distribution due to actuation.

and from the known values of the right hand side, we calculate total stress
distributions at 25 Hz and 500 Hz for L=25 mm (Fig. 10). Note that at

(a) (b)

Figure 10: Total stress distribution due to the frequency (a) 25 Hz and (b)
500 Hz for L = 25 mm.

500 Hz, the maximum stress level (0.6 MPa) is nearly 30% of the assumed
biasing stress (2 M P a). However, at low frequency this approximation holds
good (Fig. 10a).
We perform some parametric studies by varying length up to 100 mm

17
and frequency up to 0.5 KHz. The variation of the maximum stress response
is presented in Fig. 11(a). The white dotted line represents the limit below
which the stress level is less than 5% with respect to the blocking stress.
Considering this range as an admissible domain of L and f , we project this
line on the maximum velocity plot (Fig. 11b), eddy current and Joule heating
plots (Fig. 12).

(a) (b)

Figure 11: (a) Variation of maximum total stress and (b) maximum velocity
at different lengths and frequencies. The dotted lines are the threshold above
which the data start to accumulate spurious values due to constant stress
assumption and lack of satisfying conservation of linear momentum.

It is evident from the above study that we need to consider full coupling of
the conservation of linear momentum equation, coupled magneto-mechanical
constitutive equations and Maxwell equations for accurate predictions in a
wide range of length-frequency combinations. At smaller time scale inertia
will undoubtedly play a crucial role in generating stress waves. Since the
MFIS strongly depends on the stress intensity, a highly nonuniform strain
field may be produced. A proper wave propagation study is required to
conclude about the inertia effect, which is not investigated in this study.
However, the presented analysis gives feasible results only at a specific zone
of interest, as mentioned inFig. 11.
Finally, we present some quantitative data for mechanical waves, elec-
tromagnetic waves and skin depth for MSMA conductors, operating under
fluctuating field. The propagation speed of longitudinal
q mechanical waves
along an elastic MSMA rod (1-D) is given by vm = Eρ = 407.23 m/s. On
the other hand, if we assume linear magnetic MSMA material for which

18
(a) (b)

Figure 12: (a) Maximum eddy current density and (b) Joule heating at
different lengths and frequencies.

µ = max {µr (hy )} ≈ 3.5µ0 , then the plane wave propagation speed is given
by
r sr
ω 0 µr Ω 2
ue = , where κ=ω 1+( ) + 1.
κ 2 0 ω

The values at 1 MHz and 1 kHz frequencies are 1.13 · 103 m/s and 35.88 m/s,
respectively. The skin depth, δ = 1/κ, increases from 0.18 mm to 5.7 mm due
to the decrease in frequency. Further decreasing the frequency, e.g. at 100
Hz, ue = 11.34 m/s and δ = 18.1 mm.

4.4 Comparison with rigid and magnetostrictive ferro-


magnets
In this section, we compare the eddy current and Joule heating due to MSMA
actuation with the conventional rigid magnet and magnetostrictive materials.
We select ‘430 steel’ as a rigid ferromagnet which has saturation magnetiza-
tion M sat = 955kA/m. Such a soft ferromagnetic material reaches saturation
when the applied field is around 2.5 [mT] [80]. This value is significantly small
compared to the operating field range of forward variant reorientation (0.6-
0.9 [T]) for MSMA. We can thus easily assume that magnetization remains
fixed in this range and there is no rate of change of magnetization. On the
other hand, for a magnetostrictive material, such as Terfenol-D, magneti-
zation and strain saturation take place around 0.25 [T] [81]. Almost same

19
electrical conductivity (Ω = 1.66 · 106 [S/m]) of Terfenol-D [82] with the 430
steel suggests that both of them will behave almost same in the considered
range of the magnetic field (0.6-0.9 [T]) .

jz [A/mm2 ]
2.5
6
jz [A/mm2 ]

2
4
1.5

1 2

0.5 0
100
0
100 50 1
×10-3
50 0.01 x [mm] 0
2
t [s]
x [mm] 0 0.02 t [s]
(a) (b)

Figure 13: Induced eddy current at (a) 25 [Hz] and (b) 200 [Hz] for a con-
sidered length of 100 [mm].

The absence of rate of change in magnetization and magnetostriction


gives ∂t by = µ0 hy and eez = ez . We write from equations (29) and (30)
1 1
∂x2 jz = ∂t jz , where γ = . (36)
γ Ωµ0
We solve the above equation with the same initial-boundary values, used for
the MSMA problem. The analytic solution of the above equation is given by
(
0 (x, t) = (0, 0)
jz (x, t) = . (37)
A erf (ζ) + B (x, t) ∈ (0, L] × (0, +∞)

Here ζ = x/2 γt and A, B are arbitrary constants. From jz (0, t) = 0 we get
B = 0. Considering the fact that µ0 h(0, 0+ ) = HsM2 and µ0 h0 (0, tf ) = HfM2 ,
we further write from the relation Ω1 ∂x jz (0, t) = µ0 ∂t hy (0, t)

M2 √
Hf /µ0 tf Ω(HfM2 − HsM2 ) πγ
Z Z
A dt
µ0 dh(0, t) = √ √ =⇒ A = √ . (38)
Hs
M2
/µ0 Ω πγ 0 t 2 tf

20
The plots are given in Fig. 13 for the eddy current at two different finish
time tf = 0.02 [s] and tf = 2.5 · 10−3 [s], which correspond to 25 [Hz] and 200
[Hz], respectively. At low frequency, saturation of eddy current reaches at
the end of the length. However, saturation reaches much earlier for the high
frequency. A similar trend is also observed for the Joule heating. Maximum
values of the eddy current and the Joule heating and a comparison with the
MSMA responses are given in Table 3.

MSMA 430 Steel/ Terfenol-D


25 [mm] 25 [Hz] 200 [Hz] 25 [Hz] 200 [Hz]
jzmax 3.38 19.12 1.25 6.11
P max 5.15 165.85 0.92 22.51
100 [mm]
jzmax 7.95 50.11 2.20 6.11
P max 28.46 1.13·103 2.81 22.51

Table 3: Comparison of MSMA and ferromagnetic/magnetostrictive re-


sponses at different frequencies and lengths.

Comparing the responses in Fig. 13(a) and Fig. 8(a), we observe that
eddy current generation in MSMA is higher than the conventional ferro-
magnetic/magnetostrictive materials (see also Table 3). Presence of rate of
change of magnetization and velocity of the material points during variant
reorientation make MSMA responses significant different. However, for a
non-MSMA materials, responses relax quickly with the evolution of time.
Power loss also follows the similar trend as of the eddy current.
Before conclusion, let us have a quick estimation of temperature change
due to the Joule heating by solving a 1-D steady-state heat equation. Con-
sider Fig. 1 and assume that the surface x = 0 is insulated. The temperature
is uniform, and the system is below the martensitic finish temperature Mf . If
g0 is the energy generation and is assumed to be homogeneous for simplicity,
then the heat conduction equation is written as
1
∂xx T + g0 = 0 (39)
k
where k is the thermal conductivity. Thermal analysis with ambient heat
convection could be found in [83]. Let us assume the temperature at x = L
is maintained at Mf and the other boundary condition is ∂x T = 0 at x = 0.
g0 2
The temperature distribution is then given by T (x) = − 2k [x − L2 ] + Mf ,
where L is the length of the slab. From Table. 3, let us consider the case
L = 25 [mm] and 200 [Hz] for which g0 = P max = 19.12 [mW/mm3 ]. The

21
thermal conductivity of Nickel based alloy is nearly equal to that of Nickel,
i.e., k = 91 W/m·K. For the given thermal conductivity, energy generation,
and geometry the temperature at the adiabatic surface becomes T (0) =
98◦ C and quadratically decreases to T (L) = Mf = 32◦ C (Ni50 Mn28 Ga22
composition, [84]). The austenitic finish temperature for the same chemical
composition is Af = 52◦ C. We are now looking for a range of x for which
g0 2
T (x) = − 2k [x − L2 ] + Mf ≥ Af and the solution for the above mentioned
material parameters is x ≤20.6 [mm]. This implies that more than 80% of
the length will be fully transformed to the austenitic phase and the MFIS will
be decreased due to the presence of austenite [83]. However, this is a gross
estimation, and one needs to solve a coupled magneto-thermo-mechanical
system for more precise information.

5 Conclusion
We found that the time dependent part of the Maxwell equations, which
generally omitted in the quasi-static MSMA analysis, is significant for high
frequency analysis. The high actuation strain rate generated due to variant
reorientation cuts the applied magnetic field and produces motional elec-
tromotive force (emf). The total induced emf due to the rate of change of
magnetic field and motional emf generates an eddy current. Additional mag-
netic field due to eddy current adds up with the existing static magnetic field
and helps in variant reorientation. We solved numerically an Initial Bound-
ary Value Problem by using the existing MSMA constitutive equations in
a staggered way to capture electro-magneto-mechanical coupling responses
of different field variables. We found the eddy current and Joule heating
are significant. A comparison of eddy current and Joule heating between
MSMA and rigid-ferromagnetic/magnetostrictive materials reveals that the
differences are significantly high in MSMA. These could be the two essen-
tial parameters for shape and material optimization in device level MSMA
applications.

Acknowledgments
The authors would like to acknowledge the financial support of the Army
Research Office, Grant no. W911NF-06-1-0319 for the initial stages of this
work, NSF-IIMEC (International Institute for Multifunctional Materials for
Energy Conversion) under Grant No. DMR-0844082, and NSF-NIRT, Grant
no. CMMI: 0709283 for the support of the first author during his graduate

22
study at Texas A&M University.

APPENDIX
A Thermodynamic driving force
The four critical magnetic fields are: the start of forward reorientation, HsM2 ,
the end of forward reorientation, HfM2 , the start of reverse reorientation,
HsM1 , and the end of reverse reorientation, HfM1 . The reduced form of the
thermodynamic force (12) is given by
µ0 2 cur
π r = (σ ∗ + H )ε + µ0 ∆M2 H2 + f r . (A-1)
2 y
The hardening function is chosen as [85]

ξ˙ > 0 ,
( A
r
− 2 (1 + ξ n1 − (1 − ξ)n2 ) + B,
f := (A-2)
− C2 (1 + ξ n3 − (1 − ξ)n4 ) + D, ξ˙ < 0 ,

where, n1 , n2 , n3 and n4 are some given exponents.


We need to know the parameters A, B, C, D, Y r . From the Kuhn Tucker
condition (14) we obtain two conditions at the beginning and two conditions
at the finish of the forward reorientation. They are

π r (σ ∗ , HsM2 ) − Y r = 0, for ξ˙ > 0, at ξ = 0 (A-3a)


π r (σ ∗ , HfM2 ) − Y r = 0, for ξ˙ > 0, at ξ = 1 (A-3b)

Similarly, for reverse reorientation we get two more equations,

π r (σ ∗ , HsM1 ) + Y r = 0, for ξ˙ < 0, at ξ = 1 (A-4a)


π r (σ ∗ , HfM1 ) + Y r = 0, for ξ˙ < 0, at ξ = 0 (A-4b)

The constant stress level is denoted by σ ∗ . The continuity of the hardening


function [73] gives us
Z 1 Z 1
r
f dξ = f r dξ. (A-5)

0 ξ̇>0 0 ξ̇<0

Solving the above five equations (from A-3a to A-5), we get the solutions of
five unknowns, A, B, C, D, Y r . The detailed derivation and the values of the
five parameters can be found in the [66].

23
References
[1] R. C. O’Handley. Model for strain and magnetization in magnetic shape-
memory alloys. Journal of Applied Physics, 83(6):3263–3270, 1998.

[2] R. C. O’Handley, S. J. Murray, M. Marioni, H. Nembach, and S. M.


Allen. Phenomenology of giant magnetic-field-induced strain in ferro-
magnetic shape-memory materials (invited). Journal of Applied Physics,
87(9):4712–4717, 2000.

[3] P. Müllner, V. A. Chernenko, and G. Kostorz. A microscopic approach


to the magnetic-field-induced deformation of martensite (magnetoplas-
ticity). Journal of Magnetism and Magnetic Materials, 267:325–334,
2003.

[4] T. W. Shield. Magnetomechanical testing machine for ferromagnetic


shape-memory alloys. Review of Scientific Instruments, 74(9):4077–
4088, 2003.

[5] S. J. Murray, M Marioni, P. G. Tello, S. M. Allen, and R. C. O’Handley.


Giant magnetic-field-induced strain in Ni-Mn-Ga crystals: Experimental
results and modeling. Journal of Magnetism and Magnetic Materials,
226–230(1):945–947, 2001.

[6] P. Müllner, V. A. Chernenko, M. Wollgarten, and Kostorz. Large cyclic


deformations of a Ni-Mn-Ga shape memory alloy induced by magnetic
fields. Journal of Applied Physics, 92(11):6708–6713, 2002.

[7] R. C. O’Handley, S. M. Allen, D. I. Paul, C. P. Henry, M. Marioni,


D. Bono, C. Jenkins, A. Banful, and R. Wager. Keynote address: Mag-
netic field-induced strain in single crystal Ni-Mn-Ga. in: Proceedings
of SPIE, Symposium on Smart Structures and Materials, 5053:200–206,
2003.

[8] M. A. Marioni, R. C O’Handley, and S. M. Allen. Pulsed magnetic field-


induced actuation of Ni-Mn-Ga single crystals. Applied Physics Letters,
83(19):3966–3968, 2003.

[9] O. Heczko, L. Straka, and K. Ullakko. Relation between structure, mag-


netization process and magnetic shape memory effect of various marten-
sites occuring in Ni-Mn-Ga alloys. Journal de Physique IV France,
112:959–962, 2003.

24
[10] A. A. Likhachev, A. Sozinov, and K. Ullakko. Different modeling con-
cepts of magnetic shape memory and their comparison with some exper-
imental results obtained in Ni-Mn-Ga. Material Science & Engineering
A, 378:513–518, 2004.
[11] R. D. James and M. Wuttig. Magnetostriction of martensite. Philosoph-
ical Magazine A, 77(5):1273–1299, 1998.
[12] T. W. Shield and J. Cui. Magneto-mechanical behavior of a ferromag-
netic shape memory alloy: Fe3 Pd. Proceedings of SPIE, Symposium on
Smart Structures and Materials, 4699:251–262, 2002.
[13] J. Cui, T. W. Shield, and R. D. James. Phase transformation and
magnetic anisotropy of an iron-palladium ferromagnetic shape-memory
alloy. Acta Materialia, 52:35–47, 2004.
[14] J. Cui, T. W. Shield, and M. Wuttig. Magnetostriction of stress-induced
martensite. Applied Physics Letters, 85(8):1–3, 2004.
[15] T. Yamamoto, M. Taya, Y. Sutou, Y. Liang, T. Wada, and L. Sorensen.
Magnetic field-induced reversible variant rearrangement in Fe-Pd single
crystals. Acta Materialia, 52(17):5083–5091, 2004.
[16] A. Fujita, K. Fukamichi, F. Gejima, R. Kainuma, and K. Ishida.
Magnetic properties and large magnetic-field-induced strains in off-
stoichiometric Ni-Mn-Al heusler alloys. Applied Physics Letters,
77(19):3054–3056, 2000.
[17] Y. Sutoua, Y. Imano, N. Koeda, T. Omori, R. Kainuma, K. Ishidab,
and K. Oikawab. Magnetic and martensitic transformations of NiMnX
(X=In,Sn,Sb) ferromagnetic shape memory alloys. Applied Physics Let-
ters, 85:4358–4360, 2004.
[18] D. Y. Cong, Q. Luo, S. Roth, J. Liu, O. Gutfleisch, M. Pötschke,
C. Hürrich, and L. Schultz. Sequence of structural and magnetic tran-
sitions in Ni4 8–Co2 –Mn3 9–Sn1 1 shape memory alloy. Journal of Mag-
netism and Magnetic Materials, 323:2519–2523, 2011.
[19] P. J. Webster, K. R. A. Ziebeck, S. L. Town, and M. S. Peak. Magnetic
order and phase transformation in Ni2 MnGa. Philosophical Magazine
B, 49(3):295–310, 1984.
[20] I. K. Zasimchuk, V. V. Kokorin, V. V. Martynov, A. V. Tkachenko,
and V. A. Chernenko. Crystal structure of martensite in Heusler alloy
Ni2 MnGa. Physics of Metals and Metallography, 69(6):104–108, 1990.

25
[21] V. V. Martynov and V. V. Kokorin. The crystal structure of thermally-
and stress-induced martensites in Ni2 MnGa single crystals. Journal de
Physique III France, 2:739–749, 1992.
[22] K. Ullakko. Magnetically controlled shape memory alloys: A new class of
actuator materials. Journal of Materials Engineering and Performance,
5(3):405–409, 1996.
[23] K. Ullakko, J. K. Huang, C. Kantner, R. C. O’Handley, and V. V. Koko-
rin. Large magnetic-field-induced strains in Ni2 MnGa single crystals.
Applied Physics Letters, 69(13):1966–1968, 1996.
[24] M. Pasquale. Mechanical sensors and actuators. Sensors and Actuators
A, 106:142–148, 2003.
[25] J. Tellinen, I. Suorsa, I. Jääskeläinen, Aaltio, and K. Ullakko. Basic
properties of magnetic shape memory actuators. in: Proceedings of the
8th International Conference ACTUATOR 2002, Bremen, Germany,
10-12 June 2002, pages 566–569, 2002.
[26] N. Sarawate and M.J. Dapino. Experimental characterization of the
sensor effect in ferromagnetic shape memory NiMnGa. Applied Physics
Letters, 88:121923, 2006.
[27] N. N. Sarawate and M. J. Dapino. A continuum thermodynamics model
for the sensing effect in ferromagnetic shape memory NiMnGa. Journal
of Applied Physics, 101(12):123522, 2007.
[28] I. Karaman, B. Basaran, H. E. Karaca, A. I. Karsilayan, and Y. I.
Chumlyakov. Energy harvesting using martensite variant reorientation
mechanism in a NiMnGa magnetic shape memory alloy. Applied Physics
Letters, 88:172505, 2007.
[29] R. Toupin. The elastic dielectric. J. Rational Mech. Anal, 5(6):849–915,
1956.
[30] R. Toupin. Stress tensors in elastic dielectrics. Archive for Rational
Mechanics and Analysis, 5(1):440–452, 1960.
[31] P. Jr. Penfield and H. A. Haus. Electrodynamics of Moving Media. The
MIT Press, Cambridge, Massachusetts, Cambridge, 1967.
[32] K. Hutter, A. A. F. van de Ven, and Ursescu. Electromagnetic Field
Matter Interactions in Thermoelastic Solids and Viscous Fluids. Lecture
Notes in Physics. Springer-Verlag, New York, 2006.

26
[33] A. C. Eringen and G. A. Maugin. Electrodynamics of Continua I —
Foundations and Solid Media. Springer-Verlag, New York, 1990.

[34] A. DeSimone. Energy minimizers for large ferromagnetic bodies. Archive


for Rational Mechanics and Analysis, 125:99–143, 1993.

[35] A. DeSimone and P. Podio-Guidugli. On the continuum theory of de-


formable ferromagnetic solids. Archive for Rational Mechanics and Anal-
ysis, 136:201–233, 1996.

[36] A. Dorfmann and R. W. Ogden. Nonlinear magnetoelastic deforma-


tion. The Quarterly Journal of Mechanics and Applied Mathematics,
57(4):599–622, 2004.

[37] A. Dorfmann and R. W. Ogden. Nonlinear magnetoelastic deformations


of elastomers. Acta Mechanica, 167:13–28, 2005.

[38] R. M. McMeeking and C. M. Landis. Electrostatic forces and stored


energy for deformable dielectric material. Journal of Applied Mechanics,
72:581–590, 2005.

[39] R. M. McMeeking, C. M. Landis, and S. M. A. Jimenez. A principle


of virtual work for combined electrostatic and mechanical loading of
materials. International Journal of Nonlinear Mechanics, 42:831–838,
2007.

[40] David J. Steigmann. Equilibrium theory for magnetic elastomers and


magnetoelastic membranes. Non-Linear Mechanics, 39:1193–1216, 2004.

[41] D. J. Steigmann. On the formulation of balance laws for electromagnetic


continua. Mathematics and Mechanics of Solids, 14:390–402, 2009.

[42] P. P. Guidugli, T. Roubicek, and G. Tomassetti. A thermodynamically


consistent theory of the ferro/paramagnetic transition. Archive for Ra-
tional Mechanics and Analysis, 198:1057–1094, 2010.

[43] S. V. Kankanala and N. Triantafyllidis. On finitely strained magnetorhe-


ological elastomers. Journal of the Mechanics and Physics of Solids,
52:2869–2908, 2004.

[44] J. L. Ericksen. A modified theory of magnetic effects in elastic materials.


Mathematics and Mechanics of Solids, 11:23–47, 2006.

27
[45] R. Bustamante, A. Dorfmann, and R. W. Ogden. On variational formu-
lations in nonlinear magnetoelastostatics. Mathematics and Mechanics
of Solids, 13:725–745, 2008.

[46] C. Miehe, D. Rosato, and B. Kiefer. Variational principles in dissipa-


tive electro-magneto-mechanics: A framework for the macro-modeling
of functional materials. International Journal for Numerical Methods in
Engineering, 84(10):1225–1276, 2011.

[47] C. Miehe, B. Kiefer, and D. Rosato. An incremental variational for-


mulation of dissipative magnetostriction at the macroscopic continuum
level. International Journal of Solids and Structures, 48(13):1846–1866,
2011.

[48] H. E. Karaca, I. Karaman, B. Basaran, Y. I. Chumlyakov, and H. J.


Maier. Magnetic field and stress induced martensite reorientation in
NiMnGa ferromagnetic shape memory alloy single crystals. Acta Mate-
rialia, 54(1):233–245, 2006.

[49] R. Kainuma, Y. Imano, W. Ito, Y. Sutou, H. Morito, S. Okamoto, O. Ki-


takami, K. Oikawa, A. Fujita1, T. Kanomata, and K. Ishida. Magnetic-
field-induced shape recovery by reverse phase transformation. Nature,
439:957–960, 2006.

[50] H. E. Karaca, I. Karaman, B. Basaran, D. C Lagoudas, Y. I


Chumlyakov, and H. J Maier. On the stress-assisted magnetic-field-
induced phase transformation in Ni2 MnGa ferromagnetic shape memory
alloys. Acta Materialia, 55:4253–4269, 2007.

[51] H. E. Karaca, I. Karaman, B. Basaran, , Y. Ren, Y. I. Chumlyakov, and


H. J. Maier. Magnetic field-induced phase transformation in NiMnCoIn
magnetic shape memory alloys-a new actuation mechanism with large
work output. Advanced Functional Materials, 19:1–16, 2009.

[52] K. Haldar, D. C. Lagoudas, and I. Karaman. Magnetic field-induced


martensitic phase transformation in magnetic shape memory alloys:
Modeling and experiments. Journal of the Mechanics and Physics of
Solids, 69:33–66, 2014.

[53] A. DeSimone and R. D. James. A theory of magnetostriction oriented to-


wards applications. Journal of Applied Physics, 81(8):5706–5708, 1997.

28
[54] S. J. Murray, R. C. O’Handley, and S. M. Allen. Model for discontinuous
actuation of ferromagnetic shape memory alloy under stress. Journal of
Applied Physics, 89(2):1295–1301, 2001.

[55] A. DeSimone and R. James. A constrained theory of magnetoelasticity.


Journal of the Mechanics and Physics of Solids, 50:283–320, 2002.

[56] A. DeSimone and R. D. James. Energetics of magnetoelastic domains in


ferromagnetic shape memory alloys. Journal de Physique, 112:969–972,
2003.

[57] L. Hirsinger and C. Lexcellent. Modelling detwinning of martensite


platelets under magnetic and (or) stress actions on NiMnGa alloys. Jour-
nal of Magnetism and Magnetic Materials, 254–255:275–277, 2003.

[58] L. Hirsinger and C. Lexcellent. Internal variable model for magneto-


mechanical behaviour of ferromagnetic shape memory alloys Ni-Mn-Ga.
Journal de Physique IV France, 112:977–980, 2003.

[59] B. Kiefer and D. C. Lagoudas. Magnetic field-induced martensitic vari-


ant reorientation in magnetic shape memory alloys. Philosophical Maga-
zine Special Issue: Recent Advances in Theoretical Mechanics, in Honor
of SES 2003 A.C. Eringen Medalist G.A. Maugin, 85(33-35):4289–4329,
2005.

[60] B. Kiefer, H. E. Karaca, D. C. Lagoudas, and I. Karaman. Characteriza-


tion and modeling of the magnetic field-induced strain and work output
in Ni2 MnGa shape memory alloys. Journal of Magnetism and Magnetic
Materials, 312:164–175, November 2006.

[61] B. Kiefer and D. C. Lagoudas. Modeling the coupled strain and magne-
tization response of magnetic shape memory alloys under magnetome-
chanical loading. Journal of Intelligent Material Systems and Structures,
20:143–170, 2009.

[62] J. Wang and P. Steinmann. A variational approach towards the model-


ing of magnetic field induced strains in magnetic shape memory alloys.
Journal of the Mechanics and Physics of Solids, 60:1179, 2012.

[63] Xue Chen, Ziad Moumni, Yongjun He, and Weihong Zhang. A three-
dimensional model of magneto-mechanical behaviors of martensite re-
orientation in ferromagnetic shape memory alloys. Journal of the Me-
chanics and Physics of Solids, 64:249–286, 2014.

29
[64] Yong Jun He, Xue Chen, and Ziad Moumni. Reversible-strain criteria of
ferromagnetic shape memory alloys under cyclic 3d magneto-mechanical
loadings. Journal of Applied Physics, 112(3):033902, 2012.

[65] K. Haldar and D. C. Lagoudas. Constitutive modeling of magnetic shape


memory alloys with discrete and continuous symmetries. Proceedings
of the Royal Society of London. Series A, Mathematical and Physical
Sciences, 470:20140216, 2014.

[66] K. Haldar, G. Chatzigeorgiou, and D. C. Lagoudas. Single crys-


tal anisotropy and coupled stability analysis for variant reorientation
in magnetic shape memory alloys. European Journal of Mechanics -
A/Solids, 54:53–73, 2015.

[67] I Suorsa, J Tellinen, K Ullakko, and E Pagounis. Voltage generation


induced by mechanical straining in magnetic shape memory materials.
Journal of Applied Physics, 95(12):8054–8058, 2004.

[68] Nickolaus M Bruno, Constantin Ciocanel, Heidi P Feigenbaum, and Alex


Waldauer. A theoretical and experimental investigation of power har-
vesting using the nimnga martensite reorientation mechanism. Smart
Materials and Structures, 21(9):094018, 2012.

[69] Ciro Visone, Daniele Davino, and Amr A Adly. Vector preisach mod-
eling of magnetic shape memory materials oriented to power harvesting
applications. IEEE Transactions on Magnetics, 46(6):1848–1851, 2010.

[70] B. D. Coleman and W. Noll. The thermodynamics of elastic materials


with heat conduction and viscosity. Archive for Rational Mechanics and
Analysis, 13:167–178, 1963.

[71] K. Haldar. Magneto-Thermo-Mechanical Coupling, Stability Analysis


and Phenomenological Constitutive Modeling of magnetic shape memory
alloys. PhD thesis, Texas A&M University, December 2012.

[72] R. Tickle. Ferromagnetic Shape Memory Materials. PhD thesis, Uni-


versity of Minnesota, Minneapolis, MN, May 2000.

[73] D. C. Lagoudas, Z. Bo, and M. A. Qidwai. A unified thermodynamic


constitutive model for SMA and finite element analysis of active metal
matrix composites. Mechanics of Composite Materials and Structures,
3:153–179, 1996.

30
[74] M. A. Qidwai and D. C. Lagoudas. Numerical impementation of a
shape memory alloy thermomechanical constitutive model using return
mapping algorithms. International Journal for Numerical Methods in
Engineering, 47:1123–1168, 2000.
[75] J. C. Simo and T. J. R. Hughes. Computational Inelasticity, volume 7
of Interdisciplinary Applied Mathematics. Springer-Verlag, New York,
1998.
[76] S. J. Murray, M. Marioni, A. M. Kukla, J. Robinson, R. C. O’Handley,
and S. M. Allen. Large field induced strain in single crystalline Ni-
Mn-Ga ferromagnetic shape memory alloy. Journal of Applied Physics,
87(9):5774–5776, 2000.
[77] K. Haldar, B. Kiefer, and D. C. Lagoudas. FE-analysis of the demagneti-
zation effect and stress inhomogeneities in msma samples. Philosophical
Magazine, 91(32):4126–4257, 2011.
[78] Jae-hoon Kim, Fumiaki Inaba, Takashi Fukuda, and Tomoyuki
Kakeshita. Effect of magnetic field on martensitic transformation tem-
perature in ni–mn–ga ferromagnetic shape memory alloys. Acta mate-
rialia, 54(2):493–499, 2006.
[79] MD Lee, NV Nong, NP Thuy, YD Yao, SF Lee, Y Liou, YY Chen, and
CR Wang. Temperature dependence of magnetic properties in ni-mn-ga
shape memory alloys. physica status solidi (c), 1(12):3579–3582, 2004.
[80] Paul Oxley, Jennifer Goodell, and Robert Molt. Magnetic properties of
stainless steels at room and cryogenic temperatures. Journal of Mag-
netism and Magnetic Materials, 321(14):2107–2114, 2009.
[81] S Chakrabarti and MJ Dapino. Fully coupled discrete energy-averaged
model for terfenol-d. Journal of Applied Physics, 111(5):054505, 2012.
[82] BA Cook, JL Harringa, and T Hansen. Electrical and thermal properties
of tb 0.3 dy 0.7 fe 2- x. Journal of Applied Physics, 87(2):776–780, 2000.
[83] Shaobin Zhang, Xue Chen, Ziad Moumni, and Yongjun He. Thermal ef-
fects on high-frequency magnetic-field-induced martensite reorientation
in ferromagnetic shape memory alloys: An experimental and theoretical
investigation. International Journal of Plasticity, 2018.
[84] Chengbao Jiang, Yousaf Muhammad, Lifeng Deng, Wei Wu, and Huibin
Xu. Composition dependence on the martensitic structures of the mn-
rich nimnga alloys. Acta Materialia, 52(9):2779–2785, 2004.

31
[85] D. Lagoudas, D. Hartl, Y. Chemisky, L. Machado, and P Popov. Con-
stitutive model for the numerical analysis of phase transformation in
polycrystalline shape memory alloys. International Journal of Plastic-
ity, DOI:10.1016/j.ijplas.2011.10.009, 2011.

32

View publication stats

You might also like