Sei sulla pagina 1di 28

This article was downloaded by: [Southeast University]

On: 30 June 2015, At: 12:24


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954
Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Philosophical Magazine A
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tpha20

Magnetostriction of martensite
a

R. D. James & Manfred Wuttig

Department of Aerospace Engineering and Mechanics ,


University of Minnesota , Minneapolis, Minnesota, 55455,
USA
b

Department of Materials and Nuclear Engineering ,


University of Maryland , College Park, Maryland,
207422115, USA
Published online: 12 Aug 2009.

To cite this article: R. D. James & Manfred Wuttig (1998) Magnetostriction of martensite,
Philosophical Magazine A, 77:5, 1273-1299, DOI: 10.1080/01418619808214252
To link to this article: http://dx.doi.org/10.1080/01418619808214252

PLEASE SCROLL DOWN FOR ARTICLE


Taylor & Francis makes every effort to ensure the accuracy of all the information
(the Content) contained in the publications on our platform. However, Taylor
& Francis, our agents, and our licensors make no representations or warranties
whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and
views of the authors, and are not the views of or endorsed by Taylor & Francis. The
accuracy of the Content should not be relied upon and should be independently
verified with primary sources of information. Taylor and Francis shall not be liable
for any losses, actions, claims, proceedings, demands, costs, expenses, damages,
and other liabilities whatsoever or howsoever caused arising directly or indirectly in
connection with, in relation to or arising out of the use of the Content.
This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden.
Terms & Conditions of access and use can be found at http://www.tandfonline.com/
page/terms-and-conditions

PHILOSOPHICAL
MAGAZINE
A, 1998, VOL. 77, No. 5 , 1273-1299

Magnetostriction of martensite
By R. D. JAMES
Department of Aerospace Engineering and Mechanics, University of Minnesota,
Minneapolis, Minnesota 55455, USA
and MANFRED
WUTTIG
Department of Materials and Nuclear Engineering, University of Maryland,
College Park, Maryland 20742-21 15, USA
[Received 17 March 1997 and accepted in revised.form 16 September 19971

AESTRACT
A genera1 strategy is described for inducing magnetostriction in ferromagnetic
martensitic materials. An analysis of domain redistribution caused by a magnetic
field is given, and certain relations between material constants that promote this
effect are described. These relations suggest a constrained theory of magnetostriction which is used to predict strain against field in a tetragonal martensite
subject to an orthogonal biaxial magnetic field and uniaxial stress. These
predictions are compared with the corresponding experiments in Fe70Pd30.
Reversible field-induced strains of 0.6% are exhibited in this system.
Microstructural observations confirm that these strains are caused by a fieldinduced redistribution of martensitic twins.

5 1. INTRODUCTION
Magnetostriction is the spontaneous deformation of a solid in response to its
magnetization. Martensitic transformations also produce a spontaneous deformation of a solid, upon lowering the temperature. If, in addition, the martensitic material is also ferromagnetic, there exists the possibility of inducing the martensitic
transformation or of rearranging the variants of martensite, by applying a magnetic
field. Since the spontaneous strain in martensitic materials is commonly one order of
magnitude larger than that of (giant) magnetostrictive materials, the field-induced
strain available from a martensitic or ferromagnetic material is potentially much
larger than giant magnetostrictive materials. The purpose of this paper is to
investigate this phenomenon: placement of transformation and Curie temperatures, desirable domain structures, micromagnetic theory, fundamental criteria
for the existence of this effect. We also report the existence of this effect in the
Fe-Pd system.
In ordinary magnetostrictive materials the magnetostrictive strain has a volume
component and a shear component determined by the existence of the spontaneous
magnetization and its anisotropy respectively. In ferrous materials both are normally
in F e C o alloys (du Trkmolet de
range, but they can be as large as
in the
Lacheisserie 1993). Much higher values of the shear magnetostrictive strain can be
found in the binary rare earth-iron compounds RFe2. This large shear strain is
accompanied by large magnetocrystalline energy, which limits the technical value
of macroscopic magnetostrictive strain that can be obtained. This limitation can
be overcome by designing a solid solution (R1)Fe2-(R2)Fe2 of compounds having
0141-8610/98 $12.00 0 1998 Taylor & Francis Ltd

1274

R. D. James and M. Wuttig

magnetocrystalline anisotropy constants of opposite sign such that the ternary alloy
has nearly zero magnetocrystalline anisotropy. This recognition led to the discovery
of Terfenol-D (Clark 1980) with a magnetostrictive strain of loM3and zero fourthorder magnetocrystalline anisotropy energy at room temperature.
Another interesting alloy system for the study of the magnetostriction of martensite are Heusler alloys near the composition Ni2MnGa, studied recently by David
(1991), Chernenko and Kokorin (1992), K. Ullakko and R. C. OHandley (1996,
private communication) and Ullakko ef al. (1996). Ullakko et al. demonstrate fieldinduced strains of 0.2% in Ni2MnGa. From the available information, the martensitic crystallography and magnetic structure of NizMnGa are similar to those of Fe3Pd,
but with apparently different easy axes. However, preliminary calculations suggest
that the theory to be presented in $5 is applicable. Because of its large saturation
magnetization and favourable anisotropy (see $ 4 and James and Wuttig (1 996)), the
Fe-Pd alloy system was chosen for the studies reported in this paper.
The evolution of the magnetoelastic strains in martensitic materials depends on
the relative positions of the Curie and martensite start temperatures. In $ 2 we
describe qualitatively the different responses possible with the six different relative
positions of transition temperatures. These are field-induced variant rearrangements
and field-induced austenite-martensite transformation.
In $ 3 we take a closer look at the possible domain structures, with the goal of
assessing whether there are low-energy transformation paths between the initial and
final states. A key result states that, even with large deformations (i.e. geometrically
nonlinear strains), the typical strains observed in martensite, together with the typical easy axes observed in ferromagnetic materials, lead to layered domain structures
that are simultaneously mechanically and magnetically compatible. (This is a generalization of a result of James and Kinderlehrer (1993. 5 6.2).) These provide lowenergy transformation paths.
In $ 4 we draw on the results of $ 9 2 and 3 to formulate specific criteria that can
be used to screen promising ferromagnetic shape memory alloys. These criteria are
also useful as a theoretical tool and are satisfied by a constrained theory of micromagnetics introduced recently by DeSimone and James ( 1996). This theory can be
used to predict the effect of constant field and constant stress on behaviour, via
relatively easy calculations. In $ 5 we use the theory to predict the overall strain
against field, and associated energy-minimizing microstructures, in several situations
of interest in the Fe-Pd system.
In $ 6 we report the field-induced rearrangement of martensite variants among
promising alloys in the Fe-Pd system, near the composition Fe70Pd30.These alloys
have a high value of the saturation magnetization at room temperature and undergo
a cubic-to-tetragonal transformation just below room temperature. We show fieldinduced strains of 0.5% at modest fields with a two-field arrangement. There is
conceptual agreement between the measured and predicted dependences of strain
against field, but quantitative discrepancies indicate that the precise mechanism in
this system needs further study.
While our results are presented in the context of ferromagnetic martensites, they
apply equally to ferroelectric martensites. In particular, most of the results of $ $ 2-5
hold with the magnetization m replaced by the polarization p. Exceptions are, in the
few cases where we invoke the evenness of the free energy in m, our results apply
only to electrostrictive materials, and the use of conductive electrodes to apply the
electric field to ferroelectric materials leads to different boundary conditions from

Magnetostriction of martensite

1275

those assumed in $5. The results can be easily modified to account for these differences.
A preliminary version of several of the results presented here was given by James
and Wuttig (1996).
$ 2. MARTENSITE
AND MAGNETISM
The material of interest will have two transformations: a ferromagnetic transition and martensitic transformation. Depending on how the Curie temperature and
the martensitic transformation temperature are ordered, the material can be expected
to have different behaviours. In this section we consider the effects of different
orderings on qualitative behaviours.
We assume that the ferromagnetic transition is second order, but we allow the
martensitic transformation to be either first or second order. (The terms transition
and transformation are chosen here to reflect this distinction.) We first observe that,
in general, martensite and austenite are expected to have different Curie temperatures. From a theoretical viewpoint the Curie temperatures are obtained by linearized theory (in the spirit of the Landau (1965) theory of second-order phase
transitions), while the martensitic transformation is of first order, that is associated
with an exchange of stability among remote energy wells.
There are then three relevant temperatures: the martensitic tramformation?
temperature Oaus-mart, the Curie point O
F for austenite and the Curie point OFrt
for martensite. Excluding equalities, there are six different orderings passible;

To illustrate the qualitative response, let us picture the austenite + martensite


transformation as the square-to-rectangle distortion illustrated in figure 1. The six
different orderings lead to the four different qualitative behaviours shown there. In
the case when the austenite-martensite transformation temperature is the lowest
(figure 1 (a)),the non-ferromagnetic austenite first undergoes a ferromagnetic transition upon cooling to ey. One would then expect divergence-free magnetization as
pictured qualitatively for the square structure in figure 1 (a). There will also be a very
small conventional magnetostriction associated with this magnetic ordering which is
not illustrated (and is not of main interest here). Upon further cooling, the martensitic transformation will take place at eausPmart.
With the ordering of temperatures
shown in figure 1 (a)the martensite will necessarily be ferromagnetic but, owing to its
low symmetry, is expected to have uniaxial magnetic anisotropy, as shown.

t For the purpose of this section we ignore the thermal hysterises of the martensitic transformation, commonly defined by the temperatures M s , Mf, A, and A. For definiteness,
can be considered as the temeprature at which unstressed austenite and martensite
have equal free energies.

R. D. James and M. Wuttig

1276

temperature, 8

pq
4

temperature. 8

Figure 1. Qualitative pictures of domain structures with different orderings of austenitemartensite transformation temperature e,,s~mart,
Curie temperature
of austenite
and Curie temperature OFart of martensite.

OF

Depending on the domain structure, there is the expectation that the easy axes
corresponding to the two variants of martensite point in divergent directions.
Thus by alternately applying a field in various directions it should be possible to
favour one variant over another, leading to a relative change in shape (of the order of
the eigenstrain of the martensite). Alternatively, one can have a stress favour one
variant and a field favour the other, as discussed more precisely in 4 4. An example of
a material corresponding to figure 1 ( a ) is Ni2MnGa.

Magnetostriction of martensite

1277

From figure 2 it can be appreciated that, even though a Curie temperature Oyrt
for martensite exists at a temperature greater than OausPmart, it may not be observable
under ordinary conditions, because (upon heating from a low temperature) the
crystal transforms to austenite before O F t is reached. In such situations it may be
observable under conditions of stress-induced transformation. That is, it is generally
by applying
possible to stabilize the martensite at temperatures greater than eauSPmart
by conventional methods, with the
a suitable stress. Then one can search for
caveat that the crystal has been slightly biased by the stress, for example its response
is expected to be comparable with the ordinary magnetostrictive material heated
under stress through its Curie point, which has been well studied (du Trkmolet de
Lacheisserie 1993).
Note that orderings shown in figure 1 lead to the same multivariant state at low
temperatures. Thus, irrespective of the ordering of temperatures, the potential for
field-induced variant rearrangement is possible. A schematic arrangement for achieving the full eigenstrain, involving the competition between field and stress, is shown
in figure 2 (a). However, there is another method suggested by figure 1. Suppose that

rt

(3

(3

martensite

martensite

(3
(3

h
b)

austenite
martensite

Figure 2. Schematic methods of producing a change in shape: (a) field-induced rearrangement


of martensite variants; (b) field-induced austenite-martensite transformation.

1278

R. D. James and M. Wuttig

there is a jump of the magnetization at the austenite-martensite transformation.


From figure 1 this occurs in the first four orderings of 1. In these cases there is
the potential of causing a field-induced austenite-martensite transformation (see
Kakeshita, et ul. (1995) and references therein). There are several ways to conceptualize such a transformation; one possibility is illustrated in figure 2(6) in the case
when either BauS~mar,<
< Omart or t9aus-mar, < OFrt< O F , the austenite-martensite transformation is first order, and the saturation magnetization of martensite
exceeds that of austenite. Suppose that a compressive stress is applied to the austenite as shown on the left of figure 2 (b).The ordinary magnetomechanical coupling of
the austenite is then likely to lead to the uniaxial domain structure shown. Now
suppose that the field is increased from zero. If the anisotropy constants of austenite
are low, the field will first rotate the magnetization into the direction of the field,
accompanied by ordinary magnetostriction. However, since the saturation magnetization of martensite exceeds that of austenite, it is expected that sufficiently high
fields will induce a transformation from austenite to martensite. Alternatively, if the
anisotropy constants of austenite are high, a direct transformation from austenite to
martensite is expected, not preceeded by magnetization rotation. An estimate of the
field required to transform the sample at a given temperature is given by a ClausiusClapeyron equation derived from a suitable expression for the free energy ($4).

eS

fi 3. DOMAIN
STRUCTURES
The qualitative remarks above suggest that it will be possible to have a fieldinduced rearrangement of martensite variants at low temperatures, caused by the
passage from one state to another state of lower free energy. Whether this actually
will happen depends on wall mobility and crucially on the existence of a low energy
transformation path between the two states, that is, a mechanism. In this section we
show that, even with arbitrarily large deformations and the typical crystallographic
changes exhibited by shape memory alloys, together with the easy axes expected
from ferromagnetism, low-energy transformation paths exist between pairs of single
variant states. This is a generalization to general crystallographies of a result of
James and Kinderlehrer (1 993. 8 6.3) in the cubic-to-trigonal case.
We first describe the strains observed in martensitic materials below the transformation temperature. We use here a finite strain formulation, so as to avoid errors
associated with linearized strains. (As shown by Bhattacharya (1993), the use of
linearized strains can lead to substantial errors in calculations of microstructure,
when dealing with shape memory materials having larger strains, such as NiTi.) In
this format, the elementary distortions associated with n variants of martensite,
measured relative to undistorted austenite. are given by matrices

u , .u2... . . u,,.
that is each U, - 1 is the (finite) Bain strain or eigenstrain associated with variant i.
These matrices can always be chosen to be symmetric and positive-definite. The
number 17 of such matrices is given by the order of the austenite point group divided
by the order of the martensite point group. and the precise forms of the matrices for
various symmetry changes have been given by Ball and James (1992) (see also Pitteri
and Zanzotto (1996a,b) for more details about the structure of these distortions). By
definition, each matrix U, maps lattice vectors of austenite to lattice vectors for the
ith variant of martensite.

Magnetostriction of martensite

1279

Given a pair of variants i an d j, it may be true that Ui = RUjRT, for some 180"
rotation matrix R in the point group of the austenite. In fact, in an overwhelming
number of cases, among all possible symmetry changes, this is true. There are rare
cases in low-symmetry systems, for example some cubic-to-monoclinic transformations (but not that represented by NiTi) and some tetragonal-to-monoclinic transformations (represented, for example by LaNb04), which have pairs of variants not
related in his way. Compatible pairs of variants that do not satisfy this condition
form martensitic domains instead of twins (Li and James 1996). We omit such cases
and assume this condition.
If Uj = RUiRT, for some 180" rotation matrix R in the point group of the
austenite, it is known that the two variants i a n d j are compatible and exhibit a
type I twin and a reciprocal type I1 twin (these can both degenerate to compound
twins in special cases). That is, there is a rotation matrix Q and vectors a and n such
that

QU, - Ui = a 8 n.

(3)

Here, a 18n is the 3 x 3 matrix with components (a 8 n)km= aknm.In fact, given Ui
and Uj, there are precisely two solutions (Q', a', n') and (Q", a", n") of equation (3)
corresponding to the type I and type I1 twins respectively. Formulae for the solutions
are given by

n1= e ,

aI = 2

(lu;'u~:

uie),

~-

Here e is a unit vector on the axis of R: R e = e. For future reference note that, from
equation (4), Rn' = n1 and Rn" = -n". Using equation (4), the corresponding
rotation matrices Q' and Q" can be obtained from equation (3).
The condition (3) is just the Hadamard compatibility condition for a deformation y(x) having alternating gradients Vy(x) = U,/QU,/U,/QU,/U,/QU, on layers
with normal n. For a picture of a deformation of this type see figure 3; this figure
was obtained by integrating Vy(x) = U,/QU,/U,/QU,/U,/QU, to get y(x) and
then plotting the deformed positions y(xl), y(x2), y(x3),. . . of an array of dots,
which in the reference configuration occupied positions xl, x2, x 3 , .. . lying on a
square grid. The particular variants in this case were chosen to be variants 1 and
2 of the cubic-to-tetragonal case, in which U1 = diag ( 1 E ~ 1,+
1 +E ~ )
U2 = diag (1
1 E ~ 1, E ~ and
)
U3 = diag ( 1 E ~ 1,
1 E ~ ) .These two
variants satisfy U2 = RUIRTfor a suitable 180" rotation matrix R and have (compound twinned) solutions given by equation (4).
Now we turn to the magnetic part. For this we must focus on the deformed
configuration (e.g. the pictures of the dots in figures 3(6) and (c)). The variants
represent low-symmetry phases, and in the majority of cases in ferromagnetism
the directions of easy axes are along low-index directions. The low-index directions
in the martensite variants i and j are the eigenvectors (corresponding to the same
eigenvalues) of the matrices U, and U,. More precisely, it follows from Ball and
James (1992, equation (2.42)) that each element of the point group of variant i
maps the eigenspace of U, corresponding to a fixed eigenvalue to itself.

+
+

R. D. James and M . Wuttig


... ................ ..........................................
..........................................
..........................................
..........................................
..........................................
..........................................

........................................
........................................
........................................
........................................

..........................................
..........................................

................... . . .

..........................................
..........................................

..........................................
..........................................

Figure 3 . Illustration of the result that mechanical compatibility implies magnetic compatibility at a large strain ( a ) reference configuration; (h) solution I of equations ( 3 )
and (4); (c) solution I1 of equations (30) and (4). Solutions I and I1 are for two
tetragonal variants of martensite.

Let us assume that this is true, that is that the preferred magnetizations f m t ,
f m 2 , .... f m , 7 corresponding to variants U 1 , Uz,. . . ,U, respectively satisfy
U,m, = Am;, i = 1,2, . . . . n. Consistent with this, we assume crystallographic equivalence of mi and mi; if Uj = RUiRT, then mi = Rm,. The latter assumption follows
from U,m, = Am, if the eigenvalues of U, are distinct; otherwise, it serves harmlessly
to fix a degeneracy. According to these assumptions, when we consider a deformation with gradient Vy(x) = QUi we must simultaneously consider a magnetization
f Q m , and plot the arrow f Q m i on the deformed configuration of variant i. This
was done to make the pictures of magnetic domain structures shown in figure 3. Of
course, the flexibility implied by f allows one to subdivide a martensitic domain into
180 magnetic domains.
Now we consider energetics. The deformation defined by Vy(x) =
U,/QUj/Uj/QUj with corresponding magnetization m(y(x)) = f m , / f Qmj/f
mi/ f Qmj provides a continuous path between the single variant i and the single
variant j , obtained by continuously changing the volume fraction, as is clear
from figure 3. A reasonable assumption consistent with general principles of micromagnetics (James and Kinderlehrer 1993) is that the magnetoelastic energy density is
minimized exactly at the orbits (RU,,
Rm,),i = 1:. . . . n , where R is any rotation
matrix. If we omit 180 walls within a variant, the total exchange energy is proportional to the total area of interfaces between variants. The stray field energy arises
from two sources: firstly from the internal divergence of magnetization,
[*Qmj - (.t)m,]-n, where n is the deformed normal of the interfaces, and
secondly the boundary divergence, f Q m j - n or f m i - n ,where n is the normal to
the deformed boundary of the specimen. The boundary divergence can be reduced by
the choice of appropriate geometries, or by using various yokes of soft magnetic

Magne tostric tion of mar tensite

128 1

material. However, the internal divergence is clearly material dependent, that is


dependent on the distortions Ui and Uj (through the normal and the solution for
Q via equation (4)) and the corresponding easy magnetizations. Thus it is of fundamental interest to know whether, with the assumptions above (which should be
regarded as supremely typical from both the martensitic and magnetic viewpoints),
the internal divergence is zero, that is, whether, with Q and n obtained as solutions
of equations (3) and (4),

Here, we have, without loss of generality, omitted one f,and we have used the
standard formula from continuum mechanics for the deformed normal n'. If one
looks carefully at figure 3, one sees that indeed equation (5) holds for a certain choice
of (f).We show below that, under our assumptions, this is always true.
Before doing so, we note what would be the consequence of not satisfying
equation (5). There would then be poles on every interface in the laminate, which
could produce substantial stray field energy, depending on the actual value of the
left-hand side of equation (5). Assuming also typical values of the exchange constant,
it would be unfavourable for the material to have such a large magnetostatic energy,
and therefore each layer would break into the 180" domains whose density would be
determined by a competition between wall and magnetostatic energies. The energy
then would be the sum of the remaining stray field energy and the total energy of the
180" and the original interfaces. These are not exactly 90" walls, as we are allowing
finite deformations: cf. figure 3 and equation (4). These statements could be made
quantitative by micromagnetic calculations along classical lines.
To show that equation (5) holds under the stated assumptions, we first use the
integral form of equation (5):
m(z) .Vq(z) dz

= 0.

LQl
Here y(s2) is the deformed configuration. The condition (5) is equivalent to equation
(6) holding for all smooth test functions T that are supported on a subregion of
y(l2). Now change the variables in equation (6), z = y(x), and define m(x) =
*mi/ f mi/ f mi/ f rnj on the corresponding layers in s2 with normal n. We get

JQ

= (det Ui)X-'

jQm(x).V<(x)dx.

(7)

Here, we have used the specific form of Vy and the assumptions on the easy axes, and
we put <(x) = q(y(x)). From equation (7) and the definition of m we see that a
condition equivalent to equation (5) is the simpler condition:
[*mj - mi].n

= 0.

(8)

1282

R. D. James and M. Wuttig

Now, using that m, = Rm,, we have the equivalent condition [fRm, - mi].n = 0,
but now recall (cf equation (4))that there is a relation between R and n implied by
mechanical compatibility: that is, Rn = n (for the type I twin) and Rn = -n (for the
type I1 twin). Thus, equation (8) is equivalent to
[&mi - mi].n = 0

(for a type I twin),

[&(-mi) - milan = 0 (for a type 11 twin).

(9)

Hence, by making an appropriate choice of f in each case, equation (9) is satisfied


(and therefore also equation ( 5 ) ) . The result can be stated simply (with some loss of
precision) that, for type I or type I1 twins with typical easy axes, mechanical compatibility implies magnetic compatibility, even with arbitrarily large deformations.
Note that the, result is true even though type I1 twins generally have irrational
interfaces (figure 3, however, shows a case of compound twins).

$4.CRITERIA
FOR SCREENING ALLOYS
From the conceptualization presented above, several features of the desired alloy
become evident. In this section we collect these features and explain how they are
related to measurable material constants. Certain relations among these material
constants promote the behaviour that we seek. For definiteness, we concentrate of
the field-induced rearrangement of martensite variants.
A crucial feature implied by figure 2 ( a )is that values of strain and magnetization
are always associated; when the magnetization changes for some reason, so does the
strain and vice versa. This kind of behaviour can be expected for martensitic and
magnetostrictive materials, because they are governed by magnetoelastic energies
that have energy wells, and the bottoms of these wells are defined by certain special
strain-magnetization pairs, as described in 3. However, to satisfy this condition the
material should have the property that the magnetoelastic energy rises steeply away
from its energy well minima. Below, we translate this statement into conditions on
moduli.
Suppose that this were not true, and reconsider figure 2 (a). Upon application of
the vertical field, a competing possibility would be that the magnetization simply
rotates into the direction of the field, while the shape does not change. This kind of
pure rotation of magnetization at fixed strain is governed partly by magnetic anisotropy constants of the martensite. Magnetic anisotropy constants have been widely
studied for perpendicular recording media, and especially the conditions that promote very high magnetic anisotropy (Victoria and MacLaren 1993). The principal
qualitative finding of this work is that high magnetic anisotropy is favoured by high
crystalline anisotropy (R. H. Victoria 1996, private communication), produced by
anisotropic crystal structure, anisotropic ordering or very-fine-scale multilayer geometry. This qualitative feature. at least, is consistent with our proposal, as nearly all
martensites are of low symmetry. In fact, high-symmetry austenite and low-symmetry martensite also gives many variants of martensite, which is crucial for the shape
memory effect in random polycrystals (Bhattacharya and Kohn 1996), and also
should be beneficial in the present case.
Similar issues arise when considering elastic stiffness at constant magnetization,
which is the consideration of a different path away from an energy well minimum.
Referring to the right of figure 2 (a), a crystal that is too elastically soft would simply
compress under the stress, regardless of its magnetization.

Mugnetostriction qf murtensite

1283

High anisotropy, that is steep energy wells, implies large energy barriers between
the wells. At first, these barriers might be considered to produce such large metastability that it might be regarded as difficult to move twin boundaries. In fact, Clarks
(1980) well known strategy that led to the discovery of the giant magnetostrictive
material Terfenol-D involved partly the mixing of appropriate concentrations of
terbium and dysprosium so as to have a nearly ambiguous easy axis. In terms of
the energy landscape, this corresponds to replacing an energy barrier (or at least a
saddle) by an energy well. The effect of this is to flatten the energy surface on
average. On the other hand, there exist large-deformation shape memory alloys,
with rather hard moduli, whose twin boundaries move under small stress. For example, Cu-14.0 wt% A1-3.9 wt% Ni, which has relatively hard moduli, undergoes a 6%
shear strain under a critical resolved shear stress on twin boundaries of just 1.5 MPa,
for a suitably oriented specimen (Chu 1993). In this particular case, the crystal
essentially jumps between energy wells. Orientation and texture, and metallurgical
factors such as precipitates and dislocations, may also play a crucial role. A predictive understanding on metastability and hysteresis is surely missing for the materials considered here, but particular examples show that it is possible to have the
coexistence of high anisotropy, large strains and mobile twin boundaries.
To formalize these criteria, let us consider micromagnetic theory. To be able to
state criteria in terms of conventional material constants, we choose a conventional
expression for the energy, using geometrically linear strains. Here there are two
caveats: firstly with the larger martensitic strains (e.g. 5-10% range) this can lead
to significant errors, and it does not immediately relate to the finite-strain considerations of Q 3, and secondly, as pointed out by Brown (1966), the conventional smallstrain formulation is missing certain terms that should be present according to a
straightforward linearization starting from geometrically nonlinear theory. With
these reservations, we write the standard micromagnetic energy
F=

JQ

{cp(rn)+~[E-Eo(m)]-CIE-Eo(m)]
- ho+m-ao*E}dx

Here, we have omitted exchange energy and strain gradient energies, as we are
dealing with large bodies (DeSimone 1993). cp is the magnetic anisotropy energy.
The applied field is ho and the applied stress is c0.The strain tensor E is given by the
. potenusual formula in terms of the displacement gradient: E = 4 (Vu + ( V U ) ~ )The
tial Crn in the magnetostatic energy is the unique solution (among potentials with
finite magnetostatic energy) of the magnetostatic equation
div ( -V&,

+ 4nm) = 0,

(11)

which itself arises from the conditions curl h = O(+ h = -VCrn), div b = 0,
b = h + 4nm. In equation (10) the form of cp depends on symmetry, the elastic
modulus tensor C is positive definite, and Eo(m) is the preferred strain tensor
corresponding to the magnetization m; in the cubic case the most general quadratic
form is
Xloo(m 8 m -;I)

+ ( A l l 1 - Xloo)

rfiifij(ei
8 ej)
ii6i

1284

R. D. James and M. Wuttig

where m = m/m,, {el,e2,e 3 }is an orthonormal basis parallel to the cubic axes, and
rii, = m e e , are the direction cosines of the magnetization.
We assume that p has the usual cubic form

v ( m ) = K l (&&

-2 -2 -2
+ ri~irii:+ &+I$) + K2mlm2m3.

(13)

The energy wells are associated with the minimizers of the total anisotropy energy.
These minimizers are the pairs ( E l l&m,), (EZlAm,), . . . (En,
Amn), where
f m , , f m 2 , . .. , f m , are minimizers of p and E, = E,(m,). The numbers and
types of variants are determined by the particular values of K~ and K~ (Bozorth
1978, table 4, p. 582). I t follows from this form of the micromagnetic energy that
pairs of energy wells, that is pairs of variants of the martensite, are related by
E, = QE,QT,m, = Qm,, for some Q belonging to the point group of austenite (of
the Fe-Pd alloys studied in this work the point group m3m) but not belonging to the
point group of the ith variant of martensite. These conditions are the translation to
geometrically linear theory of the conditions relating variants given in 5 3.
As discussed above, a meaningful criterion for screening alloys is that the magnetoelastic energy density grows rapidly as we depart from an energy well in any
direction. It is essential to formulate criteria in terms of dimensionless parameters.
To non-dimensionalize the magnetostatic equation (1 l), we divide by m,. Thus
&,,/m, is dimensionless; so we make the energy (10) dimensionless by dividing the
whole expression by mi.The departure from a well (caused by an applied stress or
field) should be compared with something having the same dimensions. Following
the qualitative argument given in $ 2 and based on figure 2 ( a ) the departure from an
energy well should be small compared with a typical distance between the wells.
To calculate this departure we first rewrite the non-dimensionalized micromagnetic energy (10) in an alternative form by completing the square on the elastic term.
This gives

Consider the term containing square brackets in the integrand and suppose that m is
assumed close to one of the values f m , , f m z , .. . , f m , , . Then the strain E will be
close to the corresponding value in the list E l . E 2 , .. . , E,, if the quantity Q=-Ico/(Eol
is small compared with the unit tensor Eo/lEol. This leads to the criterion

where min C is the minimum elastic modulus and

Magnetostriction of martensite

1285

Now suppose that (15) holds and consider a departure of m from one of the values
f m l , *m2,. . . , *m,. This departure is found by calculating the minimizer of the
quantity cp(m) - h o - m- ao-Eo(m),which gives rise to lengthy expressions that
simplify upon linearization. We find that (to first order in ho and no) the minimizers
of cp - ho.m - no-Eoremain close to the values f m , , &m2,.. . , fm, if the following conditions hold:

where

(Pl, P2) =

1.(

I, lK1 I)

(IKlli

(1%

12x1 + 4)
+ ~ ~ Ib
1 l, + 4)

[loo] easy axes,


[ 1lo] easy axes,
[ 1 1 11 easy axes.

(18)

A calculation based on equations (15) and (17) shows that these conditions are
sufficient to ensure that the magnetoelastic energy density is negligibly perturbed
by the field and stress.
There is a more general possibility consistent with the behaviour shown in figure
2 (a). That is, the anisotropy energy could have certain low energy valleys (in (E, m)
space) leading away from its minima, but with these valleys isolated from each other
by large barriers so as to allow the process shown in figure 2(a) to occur. As an
example, suppose that the criterion (15) holds, but not criterion (17). Suppose also
that Eo(m)exhibits high anisotropy. Then, equation (15) would imply that any state
(E(x),m(x)) having low energy would tend to satisfy E(x) = Eo(m(x)),at least on
most of 52. Hence, even though m(x) might be quite free to range over the constraint
set Im(x)I = m, with little energetic penalty, there would be a strong tendency for
strain and magnetization to be coupled, leading to the possibility of a large fieldinduced strain. This possibility is not essentially different from the one above; it is
just that the energy wells form an extended set. This idea seems to play a role in the
Fe-Pd system (see $6).
Finally, we note that equations (15) and (17) have been derived under the
hypothesis that every stress and field satisfying the restrictions only slightly perturbs
the energy-well minima. If one goes back and puts in a special field and stress, these
conditions can be relaxed considerably. Thus, if the magnetoelastic energy density
rises steeply in certain directions but not others, it may be possible to design special
applied fields and stresses that take advantage of the particular feature, again leading
to the behaviour shown in figure 2(a).
$5. EVOLUTION
OF

VARIANTS UNDER COMBINED STRESS A N D FIELD:

MICROMAGNETICS OF MARTENSITE

The criteria (1 5 ) and (1 2) lead naturally to a constrained theory (DeSimone and


James 1996) which has proved to be extremely useful for predicting the behaviour of
emerging ferromagnetic shape memory alloys. Here, we briefly describe this theory
and then apply it to the behaviour of Fe,Pdl_,(x M 0.7).
If equations (15) and (17) hold, it is natural to expect that, except for small
regions of the specimen, the strain-magnetization pair (E(x),m(x)) will lie close

R. D. James and M. Wuttig

1286

m(x))Ewciis Fconstr
to the energy wells. This leads to the constrained theory min(E(x).
where

-h,-m(x) - oo-E(x)dx

lV<,1* dx

+ excess.

(19)

Here, the wells refer to the strain-magnetization


pairs (El, f m , ) ,
(E2,f m 2 ) ,. . . , ( E l l .fm,,). and F,,,,,, is the constrained free energy. The term
excess is a non-negative and vanishes for a magnetization that is locally divergence
free; excess is the energy of internal poles. In reality, the constraint
(E(m),m(x)) E wells is interpreted in a slightly weaker sense that allows for departures from the well on small transition layers. See DeSimone and James (1996) for
the precise statements and a rigorous derivation of this theory starting from general
micromagnetics. The energy Fc,,,t, is still somewhat difficult to use for general
shapes, but it can be simplified by assuming that R is an ellipsoid. To see this,
note that the first two terms of equation (20) depend only on the average strain
and the average magnetization, since the applied field ho and applied stress are
assumed to be constant. Hence, without loss of generality, one can fix the average
magnetization and minimize just the magnetostatic energy. It is well known that,
with the average magnetization fixed, and with R an ellipsoid, the magnetostatic
energy alone is minimized by a constant magnetization m(x) = (m) =
(l/volumeQ) J, m(x)dx. The value of the minimum is i(m)-D(m)where D is
the demagnetization matrix for the ellipsoid R. Finally, it is possible to show by
explicit constructions involving layered (and layers within layers, etc.) geometries,
together with suitable divergence-free magnetic substructures, that, if the wells are
pairwise compatible, any average magnetization-strain in the convex hull of the
wells is achievable. In other words, the average strain-magnetization is given by
the expressions

+ X2(-mi) + A3m2+ . . + X2,1(-m,r),


(E) = ( X I + X2)Ei + (A, + A4)Ez + . . . + ( A z ~ -+~ XZ,~)E,,

(m) = Alml

(20)

where A,, Az. . . . . Azn take values in the interval (0,l) and C, A, = 1. The fact that
the magnetic substructure can always be constructed using a divergence-free magnetization implies that excess equals zero. Hence, the minimization problem (20)
becomes a quadratic programming problem
min

(Fconstr)
=

(E(x).m(x))Ewelis

min

(-h,-(m) -oo*(E)
+$(m).D(m)), (21)

(E).(m)satisfy (20)

solvable by standard methods. The solution is a set of minimizing weights XI,


Xz, . . . , X z I l ; from these the theory guarantees at least one compatible magnetoelastic
domain structure, which we have drawn in various cases below. Sometimes there is
non-uniqueness of the domain structure; in such cases we have determined a unique
structure by forcing compatibility with a given initial state (see below).
There is a superficial resemblance between the final form (21) and the domain
rotation model of Jiles and Thoelke (1992, 1994). The starting point of the domain
rotation model is quite different, namely a collection of non-interacting particles, and
the resulting energy and predictions are also different. A discussion of the relation
between the two models has been given by DeSimone and James (1996).
Now we specialize the theory to Fe,,Pd,,, and predict its response under various
combinations of applied field and stress. By comparing the predictions of the theory

Magnetostriction of martensite

1287

based on several choices of easy axes with experiments described in $6, it was
inferred that the easy axes of the martensite are the pair of (100) directions perpendicular to the c axis. This results in energy wells defined by

(El,

-G),

(El,

&Il

where

Note that there are two easy axes corresponding to each easy strain. (This does
not cause problems for the theory above, although equation (12) for E,(m) is too
simple to give the wells (23).) We consider three kinds of experiment on a [loo] rod.

(1) Fixed axial field; variable transverse field. Here we fix the axial along [ 1001
and gradually increase from zero the transverse field along [OlO]. For each
pair of fields we calculate the solution of the quadratic programming problem. The input to the calculations is the wells (22) evaluated with the data
given in table 1. All particular graphs below use the data at 40C below M,.
The demagnetization matrix is taken to be that appropriate to an infinitely
long rod with axis [loo], for simplicity. In solving this quadratic programming problem, it is found that the solution may lie in the interior of
the constraint set (all Xi between 0 and l), or on a face (one Xi = 0 or l),
or edge (two of the X i = 0 or l), etc. Thus, at special values of the two fields
there is a change in the type of the solution. These transitions correspond to
either a jump or a change in slope of the graph of strain against field. For the
present case the calculated transitions are potted in figure 4. The region
between the two Vs is the region where demagnetization energy competes,
neither entirely winning nor losing, with the energy of the applied field. It is
found from the solution that, as expected, the volume fractions corresponding to &mf,*mi, -my and -m$ are always zero for any non-zero values of
the fields. We also find that, even though the values of the transition fields
are uniquely determined, the remaining individual volume fractions are not.
(The essential reason for this is the plethora of easy axes available in equation (22)) and the rather symmetric problem chosen.) In particular, there is
freedom to assign arbitrarily the ratio of the volume fractions associated

R. D. James and M. Wuttig

1288

Table 1. Input data for the calculation of strain against field modelled after Fe70Pd30(magnetic data from Kussmann and Jessen (1963) and Matsui and Adachi (1989), and
strains from the lattice parameter measurements of Oshima and Sugiyama (1982)).
m, is taken to be independent of temperature.
Strain F1
on c axis

Temperature interval
("C)below M ,

10
25
40
100

Strain c2
on a axis

Saturation magnetization ms
(emu cm-I)

-0.010

0.0044

-0.015
-0.021
-0.029

0.0055
0.0090
0.01 1

1390
1390
1390
1390

with m';and m:. By checking interface normals during the corresponding


tests (4 6 ) , we found that, upon cooling through the austenite-martensite
transformation in the absence of a field, the specimen exhibited a typical
habit plane transformation, leading to a finely twinned specimen with strains
El and E3. It is well known from the 'crystallographic theory of martensite'
that the expected ratio of volume fractions of these two variants should be
either f or We observed that the subsequent domain structure evolution
seen during the present experiments took place within the already established twinned laminate of martensite. Thus it seems reasonable to put the
ratio of volume fractions corresponding to my and m$ equal to f or 3 and it
was found that, by comparison with the measured curves, provided a better
match. Thus, the curves of strain against field shown in figure 5 were plotted
under this restriction. We emphasize that since the transitions of field (figure
4) are uniquely determined, the only effect of different choices of this volume
fraction is to scale the strain axis shown in figure 5.

3.

h , = -h,

+ 2zm,

h = hI

+ 2nm,

I
h,
Figure 4. Transition fields corresponding to cases ( I ) and (2). The axial field is h l and the
transverse field is hZ.

1289

Magnetostriction of martensite
strain

'

-6000

-4000

-2000

2000

4000

axial field(0e)

6000

Figure 5. Predicted axial strain against axial field at a fixed transverse field of 2000Oe and no
stress.

An evolution of domain structure corresponding to various points on the


graph of strain against field is shown in figure 6. All strains in all directions have been multiplied by a factor of ten so that the macroscopic and
microscopic strains are clearly visible. In figure 6 (as well as in figures 8
and 11 later) the scale of the microstructure should be much smaller than
the dimensions of the specimen, in keeping with the large-body limit
assumed in the theory. Note that the considerations 9 3 are applicable
here.
(2) Fixed transverse jield; variable axial jield. Here, we fix the transverse field
along [OlO] and increase the axial field along [loo] from zero. The discussion
of case (1) applies also to this case. Strain against field is shown in figure
7, while a corresponding evolution of domains is shown in figure 8. The
same ratio for the volume fraction of variant 1 to that of 3 as above was
used here.
( 3 ) Axial compressive stress; axial jield. Here we fix an axial compressive
stress along [loo] and increase the field along [loo] from zero.
Intuitively, this should be the most effective way to produce a large strain
at a small field because a compressive stress prefers strain El, an axial
field prefers magnetizations m3 or m5 (which are associated with E2 and

1290

R. D. James and M. Wuttig

Figure 6. Predicted microstructures and domain structures corresponding to the black full
circles in figure 5, with a fixed transverse field of 2000 Oe, no stress, and axial fields of
( a ) OOe, ( b ) 100OOe. ( c ) 2000Oe and (d)4000Oe: small black full circles, variant 3;
large grey full circles. variant 2; large black full circles, variant I . All strains in all
directions multiplied by a factor of ten.

E3 respectively, both of which have the same strain in the [loo] direction),
no demagnetization energy is associated with mixtures of m3 or m 5 with
the present long-rod idealization, and the largest possible uniaxial strain
e-(E,
-Ej)*e is obtained by choosing e = [ l O O ] ; i = 1 and j = 2 o r 3.
Analogous to figure 4,there is a uniquely determined graph of transition
stress against applied field (figure 9), where a jump of strain occurs on the
graphs of strain against field. Unlike the case above, here the graph of
strain against field is uniquely determined (figure 10). There is still nonuniqueness of some of the volume fractions (i.e. those associated with m 3
and m5 at large fields), but this does not affect strain against field. An
evolution of domain structure corresponding to figure 10 is shown in
figure 11.

Magnetostriction o j martensite

1291

strain

field (

Figure 7. Predicted axial strain against transverse field at a fixed axial field of 2000 Oe and no
stress.

$ 6 . EXPERIMENTS
ON Fe-Pd ALLOYS
In this section we collect our preliminary experimental findings on Fe70Pd30.A
complete magnetomechanical characterization of this alloy, together with a detailed
description of the experimental set-up, will be given in the forthcoming paper by
James et al. (1997), The purpose of this section is to confirm the expected magnetomechanical response in a promising martensite.
Fe-Pd alloys near the composition Fe70Pd30undergo an almost second-order
fcc 4 fct transition with c / a < 1 (Sugiyama et al. 1984) and display transitional
states within about 5" of the transformation temperature (Seto et al. 1990a,b,c),
which in the present alloy was measured to be about - 10C. The lattice parameters
exhibit a rapid variation with temperature in the fct phase (Oshima and Sugiyama
1982), leading to relatively large eigenstrains several tens of degrees Celsius below
transformation. The ordering of transition temperatures is as in figure 1 (a), as both
austenite and martensite are ferromagnetic. In the temperature range of interest the
saturation magnetization is only weakly dependent on temperature but has large
values near 1400e1nucm~~
(Matsui and Adachi 1989). Rather large values of magnetocrystalline anisotropy constant )il of the order of 107-108 e r g ~ m have
- ~ been
reported (Klemmer et al. 1995). Hence, from what is known, this system satisfies the
conditions that favour the effect studied in this paper.

1292

R. D. James and M. Wuttig

Figure 8. Predicted microstructures and domain structures corresponding to the black full
circles in figure 7, with a fixed axial field of 2000 Oe, no stress and transverse fields of
(a) 2000Oe. (b) 5000Oe, ( c ) 8000 Oe and (6)10000Oe: small black full circles, variant
3; large grey full circles, variant 2; large black full circles, variant I . All strains in all
directions multiplied by a factor of ten.

The specimen was a single crystal

of approximate dimensions

1.5 x

1.5 mm x 6.6 mm oriented with [ 1001 pointing parallel to the long axial direction.
The sample was cut from an available crystal (Li et a/. 1993, Saxena and Barsch
1993, Chopra and Wuttig 1995). It was mounted so that three different fields could
be applied to it at various temperatures: an axial field parallel to [loo], a transverse
field parallel to [0, 0.95, 0.311 and a compressive stress parallel to [loo]. Except for
the transverse magnetic field being tilted away from [OIO] and the specimen not being
infinitely long (recall the choice of the demagnetization matrix in $5), the experimental conditions correspond to the situations analysed in 0 5. The strains were

Mugne tos tr iction of mar tensite

1293

stress ( M P a )

Figure 9. Transition stresses corresponding to case (3) with no transverse field.

0.005-

-4000

2000

-2000

-0.005

-0.01

4000

'

axial field(0e)

'

-0 n i q r

.....
.....
.....
E+
.....
E+
.....
.....
.....
.....
.....
.....

!
i
+

Figure 11. Predicted microstructures and domain structures corresponding to the black full
circles in figure 10 with no transverse field: small black full circles, variant 3; large grey
full circles, variant 2; large black full circles, variant 1 . All strains in all directions
multiplied by a factor of ten.

R. D. James and M. Wuttig

1294

measured with an axial strain gauge. The fixed fields were applied by a set of permanent magnets. The microstructural evolution of the martensite variants, upon
application of a magnetic field or stress, was observed from surface relief on the
specimen by an optical microscope with differential interference contrast. Three
experiments are reported here which correspond, respectively, to the predictions in $5.
(1) Fixed transverse magnetic field; variable axial magnetic field; no stress
(figure 12). Here, a transverse field of 2300 Oe was applied during cooling
and during the experiment performed at -36C and the axial field was cycled
about zero, starting from zero. The dependence of the strain on the axial

-12000

-10000

-8000

-6000

-4wO

-Zoo0

ZOO0

4000

6000

8000

10000

H (oe)

Figure 12. Strain against axial field in Fe7,,Pd3, at -36C. The transverse field was fixed a t
2300Oe and the axial field was cycled as shown. No stress was applied.

1295

Magnetostriction of martensite

field is shown in figure 12. Upon cooling through the austenite-martensite


transformation, axial bands of martensite appeared (cf. 9; 5, case (1)). As the
axial field was increased, there was large-scale movement of these bands and
some new bands appeared, all having interface normals consistent with those
shown in figure 6. There was clear evidence of large-scale field-induced
movement of martensitic twins. The left-hand side of the curve in figure
12 conceptually agrees with the corresponding predicted dependence (figure
5). Starting from the saturated state created during cooling in the transverse
field of 2300 Oe, the strain decreases approximately linearly as the axial field
drops below this value. However, the strain does not increase in a symmetrical fashion beyond zero as expected for a perfectly oriented [lOO][OlO]
sample. Instead, it increases asymmetrically towards a different asymptotic
strain possibly reflecting magnetization rotation.
(2) Fixed axial magnetic field, variable transverse magnetic field, no stress
(figure 13). Here, an axial field of 2300 Oe was applied during cooling to

4000

2300 Oe

-8000

/pdoo

4000

-Zoo0

Zoo0

4000

6000

8000

loDw

1: 30

H (oe)

Figure 13. Strain against transverse field in Fe70Pd30at -17C. The axial field was fixed at
2300 Oe and the transverse field was cycled as shown. No stress was applied.

1296

R. D. James and M . Wuttig

-17C and the transverse field, applied parallel to [0, 0.95, 0.311 was cycled
about zero, starting from zero. The observed dependence of the strain
against transverse field is shown in figure 13. Again, there was clear evidence
of magnetic field-induced motion of martensitic twins as shown in figure 14.
Note particularly the change in the volume fraction. The observed curve
shows an approximate linear dependence of the strain on the magnitude
of the axial magnetic field in conceptual agreement with the corresponding
predicted curve (figure 7). However, the observed curve is not symmetrical
with respect to the axial magnetic field; it is shifted to the right. There are
two reasons for this shift. They are, first, the lack of symmetry of the investigated sample and, second, the asymmetry of the field-cooled initial state.
Both asymmetries apparently lead to magnetization rotation as indicated by
the nonlinearity at large positive axial fields. Figure 14 shows the observed
microstructure at two fields corresponding to the graph in figure 13. The
temperature value is nominal. Note the clear evidence of the movement of
martensitic domains, especially the change of volume fractions. If we
account for the relation of the normal (0, -0.31, 0.95) of the observed
face relative to (OOl), the orientations of interfaces shown in figure 14
agree well with those in figure 8. In particular, a comparison of the two
microstructures reveals a decrease in the population of variants whose
traces are approximately perpendicular to the [IOO] sample axis as the
magnetic field is increased from a large negative value, as expected
from figure 8.
( 3 ) Fixed compressive stress; variable axial magnetic field; no transverse magnetic
field (figure 15). Here, a constant compressive stress of -5MPa parallel to
[ 1001 was applied with a soft loading device and held fixed during cooling to
- 1 6 T . Subsequently, an axial magnetic field was applied, also parallel to
[ 1001, and was cycled about zero, starting from zero. As predicted by theory
(see figure lo), the evolving magnetic field induced strain is symmetrical in
the axial magnetic field and saturates. However, the predicted sudden variation in the strain at a critical field is smoothed out and the values of the jump
in strain are significantly smaller than those predicted. The former can
be attributed to the variations in the internal magnetic field due to
the inhomogeneous demagnetization field. The latter may reflect the
same inhomogeneity, giving rise to local rotation propagating through
the sample.
In summary, there is conceptual agreement between theory and experiment in
all three cases. The field-generated strain is symmetrical with the respect to the
applied varying field when the sample-field configuration is symmetrical and it is
asymmetrical otherwise. In addition, the field-generated strain is linear over significant ranges of the generating field as predicted. The major limitation of the
present experiments has to do with the lack of symmetry of the sample chosen
because of its ready availability. This lack of symmetry will lead to a significant
torque on the spontaneous magnetization which has not been accounted for in
the theory. The torque will lead to magnetization rotation which will compete
with the elastic twinning and lead to non-linearities in the field-generated strain.
Future studies of approximately cut single-crystal ellipsoids will eliminate rotation.

Magnetostriction of martensite

1297

Figure 14. Twin structure of the Fe70Pd30sample at two transverse fields as indicated by the
arrows in figure 13: (a) -6500 Oe; (b)4800 Oe. The temperature is nominal; the field of
view is 1.9 mm x 1.7 mm. The upper and lower borders of the sample are parallel to its
[loo] axis.

1298

R. D. James and M. Wuttig


Field-strain at T= -16C, 5 MPa axial load

-8ooo

4000

-4000

2000

-2000

4000

6ooo

8000

1' 00

-500

-1500

-2000

H (Oe)

Figure 15. Strain against axial field in Fe-Pd at - 16C. A stress of 5 MPa was applied and the
axial field was cycled as shown. No transverse field was applied.
ACKNOWLEDGEMENTS
The authors gratefully acknowledge the experimental assistance of T. Shield, P.
Schumacher, R. Tickle and Y. Zhang. This study was supported by Office Naval
Research-Advanced Research Projects Agency (for M.W. grants N00014-95-11071
and NOOO14-93-10506; for R.D.J., grants NOOO14-95-1-1145 and NOOO14-91-5-4034).
It also benefited from the support of the Air Force Office Scientific Research (for
R.D.J., grant 49620-96-1-0057), the Army Research Office (for M.W., grant
DAAL03-92-G-0121) and the Natural Science Foundation (for M.W., grant
DMR-93-21185; for R.D.J. grant DMS-9505077).

Magnetostriction of martensite

1299

REFERENCES
BALL,J. M., and JAMES,R. D., 1992, Phil. Trans. R . SOC.A, 338, 389.
BHATTACHARYA,
K., 1993, Contrib. Mech. Thermodyn., 5, 205.
BHATTACHARYA,
K., and KOHN,R. V., 1996, Acta mater., 44,529.
BOZORTH,
R. M., 1978, Ferromagnetism (Piscataway, New Jersey: IEEE).
BROWN,
W. F., 1966, Magnetoelastic Interactions, Springer Tracts in Natural Philosophy, Vol.
9, edited by C. Truesdell (New York: Springer).
1992, Proceedings of the hternational Conference on
CHERNENKO,
V. A., and V. V. KOKORIN,
Martensitic Transformations, edited by C. M. Wayman and J. Perkins (Monterey, CA:
Institute for Advanced Studies), p, 1205.
H. D., and WUTTIG,M., 1995, J. Phys., Paris, C8-157.
CHOPRA,
CHU,C., 1993. Thesis, University of Minnesota.
CLARK,
A. E., 1980, Ferromagnetic Materials, Vol. 1, edited by E. P. Wohlfarth (Amsterdam:
North-Holland), p. 53 1.
DAVID,S., 1991, Thesis, Laboratoire de Magnetisme Louis Neel, Grenoble.
DESIMONE,
A., 1993, Arch. Ration. mech. Anal., 125, 99.
R. D., 1996, preprint.
DESIMONE,
A., and JAMES,
DE LACHEISSERIE,
E., 1993, Magnetostriction-Theory and Applications of
D u TREMOLET
Magnetoelasticity (Boca Raton, Florida: CRC Press).
D., 1993, Phil. Mag. B, 68, 237.
JAMES,
R. D., and KINDERLEHRER,
JAMES,R. D., SCHUMACHER,
P., SHIELD,T. W., TICKLE,R., WUTTIG,M., and ZHENG,Y.,
1997 (to be published).
JAMES,R. D., and WUTTIG,M., 1996, Mathematics and Control in Smart Structures, Proceedings of SPIE, Vol. 2715, edited by V. V. Varadan and J. Chandra (Bellingham,
Washington: SPIE), p. 420.
JILES,D. C., and THOELKE,
J. B., 1992, J . Magn. magn. Mater., 104-107, 1453; 1994, ibid.,
134, 143.
T., and SHIMIZU,
K., 1995, J. Phys., Paris, 5, C8-367.
KAKESHITA,
T., SABURI,
D., OKUMURA,
H., ZHANG,
B., and SOFFA,
W. A., 1995, Scripta
KLEMMERR,
T., HOYDICK,
metall. Mater., 33, 1793.
KUSSMAN,
A., and JESSEN,K., 1963, Z. Metallk., 54, 504.
L. D., 1965, Collected Papers of L. D. Landau edited by D. TerHaar (New York.
LANDAU,
Gordon and Breach (Oxford: Pergamon).
LI, J., CHOPRA,H. D., and WUTTIG,M., 1993, Proceedings of the Conference on Shape
Memory Alloys, Trans. Mater. Res. Soc. Japan, 18B, 821.
LI, J., and JAMES,R. D., 1997, Acta mater (submitted).
K., 1989, Physica B, 161, 53.
MATSUI,M., and ADACHI,
M., 1982. J. Phys., Paris, 42, C4-383.
OSHIMA,
R., and SUGIYAMA,
M., and ZANZOTTO,
G., 1996a, preprint; 1996b, preprint.
PITTERI,
A., and BARSCH,
G. R., 1993, Physica D, 66, 195.
SAXENA,
Y., 1990a, Mater. Sci. Forum, 56, 77; 1990b. J. phys. Soc.
SETO,H., NODA,Y., and YAMADA,
Japan, 57, 3668; 1990c, ibid., 59.
SUGIYAMA,
M., OSHIMA,O., and FUJUTA,E., 1984, Trans. Japan Inst. Metals, 25, 585.
J. M., 1993, Phys. Rev. B, 47, 11 583.
VICTORIA,
R. H., and MACLAREN,
ULLAKKO,
K., HUANG,
J. K., JANTNER,
C., O'HANDLEY,
R. C., and KOKORIN,
v. V., 1996.
Appl. Phys. Lett., 69, 1967.

Potrebbero piacerti anche