Magnetoelastic stresses in rare-earth thin films and superlattices
We report on the study of the magnetoelastic behavior of some rare-earth based thin films and superlattices (SL`s). Magnetoelastic stress (MS) measurements (by using a cantilever capacitive technique) within a wide range of temperatures (10-300 K) and magnetic fields (up to 12 T) have been performed...
Gespeichert in:
Datum: | 2001 |
---|---|
Hauptverfasser: | , , , , , , , , , |
Format: | Artikel |
Sprache: | English |
Veröffentlicht: |
Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України
2001
|
Schriftenreihe: | Физика низких температур |
Schlagworte: | |
Online Zugang: | http://dspace.nbuv.gov.ua/handle/123456789/130011 |
Tags: |
Tag hinzufügen
Keine Tags, Fügen Sie den ersten Tag hinzu!
|
Назва журналу: | Digital Library of Periodicals of National Academy of Sciences of Ukraine |
Zitieren: | Magnetoelastic stresses in rare-earth thin films and superlattices / J.I. Arnaudas, M. Ciria, C. de la Fuente, L. Benito, A. del Moral, R.C.C. Ward, M.R. Wells, C. Dufour, K. Dumesnil, A. Mougin // Физика низких температур. — 2001. — Т. 27, № 4. — С. 342-363. — Бібліогр.: 54 назв. — англ. |
Institution
Digital Library of Periodicals of National Academy of Sciences of Ukraineid |
irk-123456789-130011 |
---|---|
record_format |
dspace |
spelling |
irk-123456789-1300112018-02-05T03:02:34Z Magnetoelastic stresses in rare-earth thin films and superlattices Arnaudas, J.I. Ciria, M. de la Fuente, C. Benito, L. del Moral, A. Ward, R.C.C. Wells, M.R. Dufour, C. Dumesnil, K. Mougin, A. Низкотемпеpатуpная магнитостpикция магнетиков и свеpхпpоводников We report on the study of the magnetoelastic behavior of some rare-earth based thin films and superlattices (SL`s). Magnetoelastic stress (MS) measurements (by using a cantilever capacitive technique) within a wide range of temperatures (10-300 K) and magnetic fields (up to 12 T) have been performed. We derive expressions relating the cantilever curvatures and the magnetoelastic stresses in anisotropic thin films and SL`s (for cubic symmetry) deposited onto crystalline substrates. The magnetoelastic energy associated to the interfaces and the epitaxial stress dependence of the volume MS has been investigated by studying the basal plane MS in Hon/Lu₁₅ and in Ho₁₀/Ym SL`s: we obtain interface MS even higher than the volume ones and the effect in bulk`s MS of the epitaxial strain is large. In Dy/Y and Er/Lu SL's we also deduce the MS contributions but, for Er/Lu, incomplete saturation leads to inconclusive results. Although the latter case also happens in Ho/Tm SL`s, MS clearly shows anisotropy competition. In TbFe₂ (t) / YFe₂(1000 Å) (300 Å < t < 1300 Å) epitaxial bilayers, we determine all the MS allowed by the symmetry and show that epitaxial stress strongly modifies the tetragonal MS. The thermal dependence of MS parameters is also analysed. 2001 Article Magnetoelastic stresses in rare-earth thin films and superlattices / J.I. Arnaudas, M. Ciria, C. de la Fuente, L. Benito, A. del Moral, R.C.C. Ward, M.R. Wells, C. Dufour, K. Dumesnil, A. Mougin // Физика низких температур. — 2001. — Т. 27, № 4. — С. 342-363. — Бібліогр.: 54 назв. — англ. 0132-6414 PACS: 75.70., 75.80.+q http://dspace.nbuv.gov.ua/handle/123456789/130011 en Физика низких температур Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України |
institution |
Digital Library of Periodicals of National Academy of Sciences of Ukraine |
collection |
DSpace DC |
language |
English |
topic |
Низкотемпеpатуpная магнитостpикция магнетиков и свеpхпpоводников Низкотемпеpатуpная магнитостpикция магнетиков и свеpхпpоводников |
spellingShingle |
Низкотемпеpатуpная магнитостpикция магнетиков и свеpхпpоводников Низкотемпеpатуpная магнитостpикция магнетиков и свеpхпpоводников Arnaudas, J.I. Ciria, M. de la Fuente, C. Benito, L. del Moral, A. Ward, R.C.C. Wells, M.R. Dufour, C. Dumesnil, K. Mougin, A. Magnetoelastic stresses in rare-earth thin films and superlattices Физика низких температур |
description |
We report on the study of the magnetoelastic behavior of some rare-earth based thin films and superlattices (SL`s). Magnetoelastic stress (MS) measurements (by using a cantilever capacitive technique) within a wide range of temperatures (10-300 K) and magnetic fields (up to 12 T) have been performed. We derive expressions relating the cantilever curvatures and the magnetoelastic stresses in anisotropic thin films and SL`s (for cubic symmetry) deposited onto crystalline substrates. The magnetoelastic energy associated to the interfaces and the epitaxial stress dependence of the volume MS has been investigated by studying the basal plane MS in Hon/Lu₁₅ and in Ho₁₀/Ym SL`s: we obtain interface MS even higher than the volume ones and the effect in bulk`s MS of the epitaxial strain is large. In Dy/Y and Er/Lu SL's we also deduce the MS contributions but, for Er/Lu, incomplete saturation leads to inconclusive results. Although the latter case also happens in Ho/Tm SL`s, MS clearly shows anisotropy competition. In TbFe₂ (t) / YFe₂(1000 Å) (300 Å < t < 1300 Å) epitaxial bilayers, we determine all the MS allowed by the symmetry and show that epitaxial stress strongly modifies the tetragonal MS. The thermal dependence of MS parameters is also analysed. |
format |
Article |
author |
Arnaudas, J.I. Ciria, M. de la Fuente, C. Benito, L. del Moral, A. Ward, R.C.C. Wells, M.R. Dufour, C. Dumesnil, K. Mougin, A. |
author_facet |
Arnaudas, J.I. Ciria, M. de la Fuente, C. Benito, L. del Moral, A. Ward, R.C.C. Wells, M.R. Dufour, C. Dumesnil, K. Mougin, A. |
author_sort |
Arnaudas, J.I. |
title |
Magnetoelastic stresses in rare-earth thin films and superlattices |
title_short |
Magnetoelastic stresses in rare-earth thin films and superlattices |
title_full |
Magnetoelastic stresses in rare-earth thin films and superlattices |
title_fullStr |
Magnetoelastic stresses in rare-earth thin films and superlattices |
title_full_unstemmed |
Magnetoelastic stresses in rare-earth thin films and superlattices |
title_sort |
magnetoelastic stresses in rare-earth thin films and superlattices |
publisher |
Фізико-технічний інститут низьких температур ім. Б.І. Вєркіна НАН України |
publishDate |
2001 |
topic_facet |
Низкотемпеpатуpная магнитостpикция магнетиков и свеpхпpоводников |
url |
http://dspace.nbuv.gov.ua/handle/123456789/130011 |
citation_txt |
Magnetoelastic stresses in rare-earth thin films and superlattices / J.I. Arnaudas, M. Ciria, C. de la Fuente, L. Benito, A. del Moral, R.C.C. Ward, M.R. Wells, C. Dufour, K. Dumesnil, A. Mougin // Физика низких температур. — 2001. — Т. 27, № 4. — С. 342-363. — Бібліогр.: 54 назв. — англ. |
series |
Физика низких температур |
work_keys_str_mv |
AT arnaudasji magnetoelasticstressesinrareearththinfilmsandsuperlattices AT ciriam magnetoelasticstressesinrareearththinfilmsandsuperlattices AT delafuentec magnetoelasticstressesinrareearththinfilmsandsuperlattices AT benitol magnetoelasticstressesinrareearththinfilmsandsuperlattices AT delmorala magnetoelasticstressesinrareearththinfilmsandsuperlattices AT wardrcc magnetoelasticstressesinrareearththinfilmsandsuperlattices AT wellsmr magnetoelasticstressesinrareearththinfilmsandsuperlattices AT dufourc magnetoelasticstressesinrareearththinfilmsandsuperlattices AT dumesnilk magnetoelasticstressesinrareearththinfilmsandsuperlattices AT mougina magnetoelasticstressesinrareearththinfilmsandsuperlattices |
first_indexed |
2025-07-09T12:41:28Z |
last_indexed |
2025-07-09T12:41:28Z |
_version_ |
1837173197788676096 |
fulltext |
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3, p. 342–363 A rn auda s J. I. , Cir ia M., de la Fu ente C., Ben it o L., del Mo ral A. , War d R. C. C., Wells M. R., Dufou r C., Dume snil K., an d Mou gin A. Mag net oelastic str esses in ra re-ea rth t hin film s and sup erlat ticesA rna udas J. I., Ciria M., de la Fue nte C., Benito L. , del Mor al A., War d R. C. C., Wells M. R. , Du four C. , Du mesn il K., and Moug in A.Mag neto elastic str esses in rar e-ear th th in f ilm s a nd supe rlatt ic es
Magnetoelastic stresses in rare-earth thin films
and superlattices
J. I. Arnaudas, M. Ciria, C. de la Fuente, L. Benito, and A. del Moral
Departamento de Magnetismo de So′lidos, Departamento de Fisica de la Materia Condensada and ICMA,
Universidad de Zaragoza and CSIC, 50071 Zaragoza, Spain
E-mail: arnaudas@posta.unizar.es
R. C. C. Ward and M. R. Wells
Dept. of Physics, Clarendon Laboratory, Oxford OX1 3PU, U.K.
C. Dufour, K. Dumesnil, and A. Mougin
Laboratoire de Me′tallurgie Physique et de Science des Mate′riaux, Universite′ Henri Poincare′, Nancy, Franc
\
e
Received October 26, 2000
We report on the study of the magnetoelastic behavior of some rare-earth based thin films and
superlattices (SL’s). Magnetoelastic stress (MS) measurements (by using a cantilever capacitive
technique) within a wide range of temperatures (10–300 K) and magnetic fields (up to 12 T) have been
performed. We derive expressions relating the cantilever curvatures and the magnetoelastic stresses in
anisotropic thin films and SL’s (for cubic symmetry) deposited onto crystalline substrates. The
magnetoelastic energy associated to the interfaces and the epitaxial stress dependence of the volume MS
has been investigated by studying the basal plane MS in Ho
n
/Lu
15
and in Ho
10
/Y
m
SL’s: we obtain
interface MS even higher than the volume ones and the effect in bulk’s MS of the epitaxial strain is
large. In Dy/Y and Er/Lu SL’s we also deduce the MS contributions but, for Er/Lu, incomplete
saturation leads to inconclusive results. Although the latter case also happens in Ho/Tm SL’s, MS
clearly shows anisotropy competition. In TbFe
2
(t)/YFe
2
(1000 A° ) (300 A° < t < 1300 A° ) epitaxial bi-
layers, we determine all the MS allowed by the symmetry and show that epitaxial stress strongly
modifies the tetragonal MS. The thermal dependence of MS parameters is also analysed.
PACS: 75.70.–i, 75.80.+q
1. Introduction
The study of the magnetoelastic (ME) properties
of magnetic thin films and layered nanostructures,
where magnetic layers are interleaved by non-mag-
netic ones or by other having different magnetic
properties, presents considerable interest, not only
because of fundamental reasons, but also due to the
technological applications of these artificial sys-
tems. For example, the success of molecular beam
epitaxy (MBE) of rare-earth (RE) superlattices
(SL’s) has allowed the investigation of the novel
magnetic properties exhibited by the well known
RE metals when thin layers of a rare earth are
interleaved with non-magnetic layers (e.g., Lu or
Y) or with layers made from a different rare earth.
Although much research has been performed to date
in relation with different magnetic properties [1–3]
of these nanostructures, it was only recently that
direct magnetoelastic stress measurements were per-
formed in RE-based superlattices [4–6]. The know-
ledge of the magnetoelastic behavior of RE SL’s is
important because of the influence of the ME en-
ergy in the spin configuration and magnetic proper-
ties of such systems. Several RE/Y and RE/Lu
SL’s have been studied and some of the differences
in the magnetic behavior observed on having Y or
Lu as interleaving layers are currently attributed to
the different epitaxial strain, tensile or compressive,
respectively, occuring in both cases. For instance,
in bulk Ho the basal plane ME energy is one of the
© J. I. Arnaudas, M. Ciria, C. de la Fuente, L. Benito, A. del Moral, R. C. C. Ward, M. R. Wells, C. Dufour, K. Dumesnil, and
A. Mougin, 2001
agents driving the magnetic structure from helical
to a ferro-cone phase [7]; interestingly, in Ho/Y
SL’s this ferro-cone phase is suppressed, unlike in
Ho/Lu systems, where the effect of the com-
pressive stress stabilizes the ferromagnetic phase
[3,8], the FM transition temperature increasing for
decreasing Ho fraction. Regarding the RE SL’s, in
this paper we will review the magnetoelastic stress
studies we carried out in two series of Ho/Lu and
Ho/Y superlattices, as well as in Dy/Y and
Er/Lu SL’s and in the more complex SL system,
Ho/Tm.
On the other hand, and concerning nanostruc-
tured systems suitable for applications, some RE
alloys are, in principle, the logic choice. So, new
devices, such as microsystems actuators working at
room temperature, should be based on alloys with
very large magnetostriction at moderately low
fields. The well-known REFe2 (RE: rare earth or
rare-earth alloy) Laves phases appear as good candi-
dates to fulfil the necessary requirements for appli-
cations. The magnetostriction of bulk single-crystal
and polycrystalline REFe2 was thoroughly studied
in the 1970’s and a noticeable effort has been made
in the last years to produce amorphous and polycry-
stalline RE-TM (TM: transition metal) films and
spring-magnet type multilayers of with improved
magnetoelastic properties [9,15]. However, the in-
trinsic magnetoelastic behavior of thin-film samples
had not been compared with that of crystalline bulk
materials due to the lack of REFe2 single-crystal
films. The recent success in growing by MBE high
quality epitaxial films of these Laves phase magnets
[16] has opened the possibility of perform such kind
of studies [17,18]. In this article we present magne-
toelastic stress measurements performed in a series
of TbFe2/YFe2 (110) single-crystal bilayers.
First of all we will describe in detail the founda-
tions of the experimental technique employed, a
cantilever method. This is important because the
analysis of the cantilever deflections, in the case of
crystalline films and substrates, differs markedly
from that performed for isotropic plates, from
which it is obtained the following expression relat-
ing the cantilever deflection and the magnetostric-
tion [19,20]:
∆ = 3λshms
L
hsubs
2
Ems
Esubs
(1 + νsubs)
(1 + ν
ms
)
,
where λsubs is the magnetostriction; L, the length of
the plate; hsubs and hms , the substrate and film
thicknesses, respectively; E and ν are the Young
moduli and Poisson ratios; the subscripts ms and
subs refer to film and substrate, respectively. For
films or SL’s having hexagonal symmetry an adequ-
ate formula, expressed in terms of the magnetoelastic
stresses Bp
µl , instead of in terms of the magnetos-
triction coefficients λjk
µ , has been derived [4,21].
The case of anisotropic substrates and the boundary
conditions imposed by the experimental situation,
have also been considered [4] and will be outlined
below for films having cubic symmetry. The ana-
lysis of the hexagonal symmetry case can be found
elsewhere [6].
2. Determination of magnetoelastic stresses in
anisotropic thin films and superlattices: case
of cubic symmetry
In the following we deal with a situation slightly
more complicated than the occurring in hexagonal
systems with cylindrical symmetry. In the cubic
case, to be able to determine all the relevant mag-
netoelastic stress parameters from the experiments,
we should perform measurements with the sample
cut along different crystalline directions, and conse-
quently, we need to obtain the corresponding ex-
pressions relating the magnetoelastic stresses and
the plate curvatures.
2.1. Evaluation of the elastic energy of a thin
plate
We consider a thin plate where the length l and
width w are much smaller than the thickness h (in
our samples h is 10−2 times the other dimensions).
Since, in our particular case, the samples were
grown in the (110) plane, we have cut the samples
using two types of configurations, which will be
considered to get practical expressions. In the first
configuration, our coordinate system axes y′ and z′
are parallel to the sample sides of dimensions l1 and
w1 , respectively, and x′ axis is normal to the
growing plane, being y′ axis || [1
__
10], z′ axis || [001]
and x′ axis || [110]. In the second one, our coordi-
nate system axes y′′ and z′′ are parallel to the new
sides of dimensions l2 and w2 , respectively, and
x′′ = x′, being y′′ axis || [1
__
11] and z′′ axis || [11
__
2]. To
save space, we will label the coordinate systems
OX’Y’Z’ as S’ and OX"Y"Z" as S"; also the axes
and superscripts corresponding to both systems will
be denoted with a single label, s, for a more
compact notation, being s = ′ or ′′, respectively,
unless otherwise specified. The boundary conditions
on the plate’s surfaces are σiknk = Pi , where nk are
the components of the normal to the plate surface, σik
are the stress tensor components and Pi are the
components of the external pressure applied to the
plate. Because of the strength of the internal co-
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 343
hesive forces, much higher than the external forces,
we can consider that σisxs = 0 (is = xs, ys and zs) on
the plate surface. Within a thin plate, these compo-
nents of the stress tensor are small compared with
the other ones, and can also be neglected [22]. We
will now assume a pure bending of the plate, i.e.,
bending without torsion; in this case, it is easy to
show that the longitudinal strains are given by [23]
ε
y
s
y
s =
xs
Rys
, εzszs =
xs
Rz
s
, εy
szs = 0 . (1)
In Eq. (1), xs is the coordinate of the plane
where the strains are evaluated (xs = 0 is the neu-
tral plane); Ris (i
s = ys, zs) (with s = ′ and ′′) are
the radii of the curvature of the neutral plane of the
plate in the planes: y′x′ and z′x′ planes in the first
case, and y′′x′′ and z′′x′′ planes for the second one.
The elastic energy density for the plate in our both
situations:
eelas
s =
1
2
σ
i
s
j
s ε
i
s
j
s (2)
is obtained by writing the stresses in terms of the
curvatures and the compliance coefficients sisjsksls ,
in the way
σ
y
s
y
s =
xs
(ŝ1s1s ŝ2s2s − ŝ1s2s
2 )
ŝ2s2s
Ry
s
−
ŝ1s2s
Rz
s
, (3)
σ
z
s
z
s =
xs
(ŝ1s1s ŝ2s2s − ŝ1s2s
2 )
ŝ1s1s
Rz
s
−
ŝ1s2s
Ry
s
, (4)
where
ŝ1s1s = s1s1s −
s1s6s
2
s6s6s
, ŝ2s2s = s2s2s −
s2s6s
2
s6s6s
;
ŝ1s2s = s1s2s −
s1s6s s2s6s
s6s6s
, (5)
and where the conventional abbreviated subscripts
have been used for the compliance coefficients.
Thus, applying the boundary conditions and substi-
tuting Eqs. (1) and (3), (4) in Eq. (2), we obtain
eelas
s =
1
2
Cy
s
y
s
Ry
s
2 +
2Cyszs
Rys Rz
s
+
Cz
s
z
s
Rz
s
2
xs2 , (6)
where
Cy
s
y
s =
ŝ2s2s
ŝs
, Cz
s
z
s =
ŝ1s1s
ŝs
, Cyszs = −
ŝ1s2s
ŝs
(7)
with ŝs = ŝ1s1s ŝ2s2s − ŝ1s2s
2 .
2.2. Determination of the magnetoelastic stresses
from the curvature of the plate
We should now evaluate the energy density of
the plate, i.e., the magnetic film, or SL, plus the
substrate. This energy is contributed by: a magne-
toelastic one associated to the magnetic sample,
which is the origin of the bending of the plate under
an applied magnetic field, and by the elastic en-
ergies from the substrate and from the magnetic
sample.
2.2.1. Magnetoelastic and elastic density en-
ergies of the magnetic sample. To first approxima-
tion it is assumed that the magnetoelastic hamilto-
nian terms are quadratic in the spin components and
linear in the strains. The magnetoelastic and elastic
energy are then written considering only single-ion
contributions and taking into account the point-
symmetry group, 4⁄m 3
__
2⁄m, for the REFe2 interme-
tallic compounds, in the form [13]
e
mel+el
s = b0(ε
xsx
s + ε
y
s
y
s + ε
z
s
z
s) + b1(α
x
s
2 ε
x
s
x
s + αys
2 ε
y
s
ys + α
z
s
2 ε
z
s
z
s) + b2(α
x
sα
y
s εx
s
y
s + α
x
sα
z
s εx
s
z
s + α
z
sα
y
s εz
s
y
s) +
+
1
2
c1s1s (εx
s
x
s
2 + ε
y
s
y
s
2 + εzs
z
s
2 ) + c1s2s (εx
s
x
s εz
s
z
s + ε
y
s
y
s εz
ss + εxsxs εy
s
y
s) +
1
2
c4s4s (εx
s
ys
2 + ε
x
s
z
s
2 + ε
zsy
s
2 ) . (8)
In Eq. (8) the represented energy densities emel′ and
emel′′ are written in the S′ and S′′ reference systems,
respectively; αis denote the direction cosines of the
macroscopic magnetisation, M; bi (i = 0, 1 and 2)
are the magnetoelastic coupling parameters (or
magnetoelastic stresses); ε’s and c’s are the carte-
sian strain and elastic components, respectively, of
REFe2 compounds, for the corresponding reference
system.
To evaluate the energy density of the system:
substrate-magnetic sample we will assume that the
strains in the film or SL are uniform because of it
small thickness, compared with the substrate one.
This means that, in Eq. (1), the variable xs can be
J. I. Arnaudas et al.
344 Fizika Nizkikh Temperatur, 2001, v. 27, No. 3
substituted by the constant value βhsubs , the dis-
tance between the neutral plane and the bottom of
the film or SL, expressed as a fraction β of the
substrate thickness hsubs .
The boundary conditions, extended to the vol-
ume of the sample, imply the minimization of its
total density energy: emel+el
s with respect to εxsxs ,
εxsys , εxszs for our two measurement configurations.
So, after that, the Eq. (8) can be rewritten, using
Eq. (1), and leave Eq. (8) as a function of the two
principal curvatures: 1/Ry′ and 1/Rz′ , for S′, and
1/Ry′′ and 1/Rz′′ , for S′′, and of the position of the
neutral plane, β (the same for both cases). Thus,
the integration of the total energy densities to the
corresponding volumes of the substrate and mag-
netic film, give rise the total energies: for S′,
E′ mel
tot = (α′ || [010]′) = A ∫
(β−1)h
subs
βh
subs
1
2
Cy′y′
Ry′
2 +
2Cy′z′
Ry′ Rz′
+
Cz′z′
Rz′
2
x′2dx′ +
+ A ∫
βh
subs
βh
subs
+h
ms
b0
x′
Rz′
c11 − c12 + 2c44
c11 + c12 + 2c44
+
x′
Ry′
4c44
c11 + c12 + 2c44
+ b1
x′
Ry′
c12
c11 + c12 + 2c44
+
+ b2
x′
Ry′
c11 + c12
c11 + c12 + 2c44
+
x′
Rz′
c12
c11 + c12 + 2c44
dx′ , (9)
and for S′′,
E′′ mel
tot (α′′ || [010]′′) = A ∫
(β−1)h
subs
βh
subs
1
2
Cy′′y′′
Ry′′
2 +
2Cy′′z′′
Ry′′Rz′′
+
Cz′′z′′
Rz′′
2
x′′2 dx′′ +
+ A ∫
βh
subs
βh
subs
+h
ms
−
b0
3
x′′
Ry′′
c12 − c11 − 10c44
c11 + c12 + 2c44
−
x′′
Rz′′
c12 − c11 − 4c44
c11 + c12 + 2c44
−
b1
9
x′′
Ry′′
c12 − c11 − 10c44
c11 + c12 + 2c44
+
+ 2
x′′
Rz′′
c12 − c11 − 4c44
c11 + c12 + 2c44
+
2b2
9
x′′
Ry′′
4c11 + 5c12 + 4c44
c11 + c12 + 2c44
+
x′′
Rz′′
c12 − c11 − 4c44
c11 + c12 + 2c44
dx′′ , (10)
where A is the area of the plate.
2.2.2. Relation between curvatures and magne-
toelastic stresses parameters. For the magnetic
sample, we have assumed it is uniformly strained,
so: εy′y′ = βhsubs/Ry′ and εz′z′ = βhsubs/Rz′ for S′,
and εy′′y′′ = βhsubs/Ry′′ and εz′′z′′ = βhsubs/Rz′′ for
S′′ performing the integration operation of Eqs. (9)
and (10), we obtain: for S′,
E′
mel
tot (α′ || [010]′) = Ahsubs
3
β2 − β +
1
3
1
2
Cy′y′
Ry′
2 +
2Cy′z′
Ry′Rz′
+
Cz′z′
Rz′
2
+ Ahms
b0
βhsubs
Rz′
c11 − c12 + 2c44
c11 + c12 + 2c44
+
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 345
+
βhsubs
Ry′
4c44
c11 + c12 + 2c44
+ b1
βhsubs
Ry′
c12
c11 + c12 + 2c44
+ b2
βhsubs
Ry′
c11 + c12
c11 + c12 + 2c44
+
βhsubs
Rz′
c12
c11 + c12 + 2c44
,
(11)
and for S′′,
E′′
mel
tot (α′′ || [010]′′) = Ahsubs
3
β2 − β +
1
3
1
2
Cy′′y′′
Ry′′
2 +
2Cy′′z′′
Ry′′Rz′′
+
Cz′′z′′
Rz′′
2
+ Ahms
−
b0
3
βhsubsν
Ry′′
c12 − c11 − 10c44
c11 + c12 + 2c44
−
−
βhsubs
Rz′′
c12 − c11 − 4c44
c11 + c12 + 2c44
−
b1
9
βhsubs
Ry′′
c12 − c11 − 10c44
c11 + c12 + 2c44
+ 2
βhsubs
Rz′′
c12 − c11 − 4c44
c11 + c12 + 2c44
+
+
2b2
9
βhsubs
Ry′′
4c11 + 5c12 + 4c44
c11 + c12 + 2c44
+
βhsubs
Rz′′
c12 − c11 − 4c44
c11 + c12 + 2c44
. (12)
In this situation, the minimization of the energies
given in Eqs. (11) and (12) with respect to the four
independent parameters: β, Ry′′
−1 and Rz′′
−1 for S′ and
Rz′′
−1 for S′′, leads to β = 2/3 and to the following
expressions:
∂E′tot(α || [1
__
10])
∂Ry′
−1
= σ~ ([1
__
10], [1
__
10]) +
+
2
3
b2(c11 + c12) + 2c44(b1 + 2b0)
c11 + c12 + 2c44
= 0 , (13)
∂E′tot(α || [1
__
10])
∂Rz′
−1
= σ~ ([1
__
10], [001]) +
+
(b2 − b1)c12 + b0(c11 − c12 + 2c44)
c11 + c12 + 2c44
= 0 , (14)
∂E′′tot(α || [1
__
11])
∂Rz′′
−1
= σ~ ([1
__
11], [11
__
2]) −
−
2
9
(b1 − b2 + 2b0)
− c11 + c12 − 4c44
c11 + c12 + 2c44
= 0 , (15)
where we have defined the σ~(α, β) MEL stresses as
follows,
σ~ ([1
__
10], [1
__
10]) ≡
1
6
hsubs
2
hms
Cy′y′
Ry′
+
Cz′y′
Rz′
, (16)
σ~ ([1
__
10], [001]) ≡
1
6
hsubs
2
hms
Cz′′z′′
Rz′′
+
Cy′′z′′
Ry′′
, (17)
σ~ ([1
__
11], 11
__
2]) ≡
1
6
hsubs
2
hms
Cz′′z′′
Rz′′
+
Cy′′z′′
Ry′′
. (18)
Notice that we have minimized the total energies
for the three different cases to obtain the minimum
number of equations needed to obtain the three
relevant MEL stresses parameters, b0 , b1 , and b2 .
By solving the system of Eqs. (13), (14), and (15)
we obtain these MEL parameters as a function of
the MEL stresses which can be determined from the
experimental results (see Sec. 3 below). In this way
just with an unique magnetostrictive experiment, it
has been possible to determine all the second order
MEL contributions to the total MEL energy,
which, in fact, is the most relevant one.
Let us compare the above expressions with the
obtained ones for the hexagonal case, when the
magnetization of the sample is parallel to the x axis
(αx = 1, αy = αz = 0). In this situation, the minimi-
zation of the total energy of the system with respect
to the three independent parameters, β, 1/Rx and
1/Ry , leads to the expressions
1
9
Ahsubs
3
Cxx
Rx
+
Cxy
Ry
−
2
3
Ahsubshms
�
+
1
4
Bγ,2
= 0 ,
(19)
J. I. Arnaudas et al.
346 Fizika Nizkikh Temperatur, 2001, v. 27, No. 3
1
9
Ahsubs
3
Cxy
Rx
+
Cyy
Ry
−
2
3
Ah
subshms
�
−
1
4
Bγ,2
= 0 ,
(20)
and β = 2/3, which has been already substituted in
Eqs. (19) and (20). In these equations � represents
a complicated combination of symmetry elastic con-
stants, cjk
µ , and magnetoelastic stresses, Bp
µl , for
the magnetic sample, which reads
� = −
1
2
B1
α,0 −
1
3
B1
α,2
√3 c12
α + c22
α
c11
α + 2/√3 c12
α + 1/3c22
α
−
−
B2
α,0 −
1
3
B2
α,2
√3 c11
α + c12
α
c11
α + 2/√3 c12
α + 1/3c22
α
.
(21)
Note that the volume magnetoelastic constants ap-
pearing in Eq. (21) include the exchange contribu-
tion to the volume strain.
We have obtained the expressions (13), (14),
and (15) and (19) and (20), which relate the
principal curvatures of a free plate with the magne-
toelastic stresses originated upon the application of
a magnetic field. However, this is not the real
situation in our experiments. In the next paragraph
we explain the actual situation as well as details
about the experimental technique and charac-
teristics of the measured samples.
3. Experimental technique and samples
3.1. Experimental arrangement
Our samples, which are rectangular (typical
plane dimensions of about 10 mm × 10 mm), are
clamped along one of their edges (cantilever con-
figuration) (Fig. 1). The exact analytical solution
of this kind of problem is unfeasible, because the
boundary conditions can not be properly imposed.
To take into account the effect of the clamping on
the deflection of the plate, finite element modeling
[24] and approximate analytical solutions [25] have
been proposed. However, we cannot apply the re-
sults obtained in these approaches to the problem,
which are based on the determination of the deflec-
tion of the free end of the clamped plate. In our
experimental set-up, we use a capacitive technique
which allows us to measure the change of capacit-
ance related with the overall curvatures of the
plate. Since the equation representing the plate’s
surface cannot be known, we associate this change
of capacitance to a single curvature of the plate.
Our hypothesis is that the curvature of every line
parallel to the clamping line is much smaller than
the curvature which appears perpendicularly to the
clamping direction. We will neglect such small
curvatures not only close to the clamping line but
also for the whole plate. This approximation overes-
timates the value of the deflection and, hence, of
the magnetoelastic parameters deduced from it,
which can have an error of order 5% [26] for our
approximately squared samples.
Thus, in the case of cylindrical symmetry, for the
sample clamped along the y direction (Fig. 1,a) we
will only consider Eq. (19), with Ry
−1 = 0, i.e.,
∂Etot
∂Rx
−1
=
1
9
Ah
subs
3
Cxx
Rx
−
2
3 Ahsubshms
�
+
1
4
Bγ,2
= 0 ,
(22)
while for experiments with the sample clamped
along x, we will use Eq. (20), with Rx
−1 = 0, which
reads
∂Etot
∂Ry
−1
=
1
9
Ah
subs
3
Cyy
Ry
−
2
3
Ahsubshms
�
−
1
4
Bγ,2
= 0 .
(23)
From Eqs. (22) and (23) it is practical to define the
magnetoelastic stresses acting along the OX and
OY directions,
σ~(x, x) =
�
+
1
4
Bγ,2 , σ~(x, y) =
�
−
1
4
Bγ,2 , (24)
where the first letter within the brackets stands for
the direction of the applied magnetic field and, the
second one, for the direction along which the plate
deforms longitudinally under the effect of the mag-
netoelastic stresses. Subtraction of the above stres-
ses allows us to obtain the magnetoelastic stress
parameter Bγ,2. Thus, we get
Fig. 1. Experimental arrangement of the cantilever: bending
along the direction of the applied magnetic field (a); bending
perpendicular to the direction of the field (b).
a b
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 347
Bγ,2 = 2[σ~(x, x) − σ~(x, y)] =
1
3
hsubs
2
hms
Cxx
Rx
−
Cyy
Ry
.
(25)
Therefore, by performing two curvature measure-
ments, according to the two different configurations
shown in Fig. 1, we can separate the parameter
Bγ,2, associated with the magnetoelastic strain
ε1
γ = 1⁄2 (εxx − εyy) within the (a, b) basal plane of
the hcp structure of the magnetic sample. The
subtraction eliminates, not only the volume strain,
ε1
α , and tetragonal strain, ε2
α , terms but also the
effect of the differential thermal expansion between
magnetic sample and substrate in the zero-field
capacitance. Note that both kind of measurements
should be performed with the field applied along
the magnetic easy direction of the sample, i.e., the
a or b axes of the basal plane for rare-earth-based
samples, provided that their growing plane is the
basal plane.
In the case of cubic symmetry we should perform
at least three curvature measurements to be able to
determine the three relevant MEL stress parame-
ters, as we described in the previous section. Ex-
pressions similar to Eq. (25), in which the only
elastic constants appearing are those of the sub-
strate, are not obtained now. Instead of it, by
solving the system of Eqs. (13), (14), and (15), we
get for b0 , b1 and b2 relationships where the radii
of curvature are multiplied by complex combina-
tions of all the elastic constants, for the substrate
and the magnetic sample.
3.2. Cantilever capacitive technique
The determination of the curvature of the plate is
performed by means a capacitive cell, in which the
metallic magnetic sample deposited on the substrate
acts as one of the electrodes. The rest of the cell was
made in copper, and annealed at 800 K to improve
its behavior under thermal cycling. Measurements
where done either in a three terminal capacitor
configuration with a AEL Ltd. ratio bridge (sen-
sitivity: 10−6 pF), or in a two plates capacitor
configuration with an Andeen–Hagerling 2500 A
capacitance bridge (sensitivity: 5⋅10−7 pF). No dif-
ferences were found between both methods, except
for a small difference in the signal to noise ratio in
favour of the latter case. In this situation, cantil-
ever deflections as small as 10−9 m can be detected
for a typical sample length of 10 mm.
For small deflections, it is straightforward to
show that a curvature R−1 of the cantilever pro-
duces a capacitance change given by the expression
∆C = − C0
2L2/6ε0AR , (26)
where C0 is the capacitance of the parallel plate
capacitor, A and L respectively, are the area and
length of the plate; ε0 is the permittivity of vacuum.
For instance, for hexagonal systems, expression
(25), together with Eq. (26), gives the magnetoe-
lastic stress Bγ,2 in terms of the different experimen-
tal values for both kind of measurements, i.e.,
sample clamped along y and x directions, in the
form
Bγ,2 = −
2ε0hsubs
2
hms
×
×
Cxx
A(x,x)∆C(x,x)
L(x,x)
2 C0(x,x)
2 − Cyy
A(x,y)∆C(x,y)
L(x,y)
2 C0(x,y)
2
, (27)
where Cxx and Cyy are given by expressions similar
to those of Eq. (7) (see Ref. 6), with the com-
pliance coefficients corresponding to the substrate
material, and the labels (x, x) and (x, y) refer to the
two measuring configurations.
In the experimental setup the capacitive cell is
placed within a continuous flow cryostat, allowing
us to measure between 1.7 and 300 K. Magnetic
fields up to 12 T, produced by a superconducting
coil, can be applied parallel to the plane of the
sample.
3.3. Samples characteristics
The magnetoelastic stress experiments reported
in this paper were performed in Ho films and in
Ho/Y, Ho/Lu, Dy/Y, Er/Lu, and Ho/Tm super-
lattices and in different REFe2 films and bilayers.
All the RE superlattices were grown by molecu-
lar beam epitaxy using a Balzers UMS 630 facility.
The rare-earth metals grow epitaxially onto a Nb
metal layer deposited on a sapphire substrate [3].
Both the body-centered-cubic Nb and hexagonal-
close-packed rare-earth metals grow with their re-
spective close-packed atomic planes parallel to the
substrate plane. The epitaxial relationships are
{112
__
0} Al2O3 || {110} Nb || {0001} RE, resulting the
a axis of the rare earth at an angle of 5° with [0001]
Al2O3 . The crystalline structure of the superlat-
tices was investigated using a triple-axis x-ray dif-
fractometer, giving an interface width of ± 2 lattice
planes [3]. The sapphire substrates, of initial thick-
ness of 500 µm, were thinned down to 150 µm to
increase the sensitivity of the cantilever method
(note the high value of the elastic constants of
sapphire, ∼ 400 GPa). After the thinning, the mag-
J. I. Arnaudas et al.
348 Fizika Nizkikh Temperatur, 2001, v. 27, No. 3
netic-sample thickness to substrate thickness ratio
remains very small (∼ 10−3). This allows us to
disregard the terms of type (Ri Rj)
−1, appearing in
Emel
tot , which arise from the elastic energy of the
magnetic sample and which are hma/hsubs times
smaller than the corresponding ones in the elastic
energy of the substrate.
From x-ray diffraction measurements performed
in our Ho/Y and Ho/Lu superlattices, between
RT and 10 K [28], we know that the epitaxy is
good: the misfit between basal plane lattice par-
ameters of Ho and Lu (or Y) in the superlattice is
very small (e.g., for the hexagonal a lattice para-
meter: (aY − aHo)/aHo = 2.1⋅10−3, at 45 K in a
[Ho40/Y15]50 SL).
The REFe2 films studied in this work have been
also grown by MBE. Sapphire substrates were co-
vered with a niobium buffer (the epitaxial re-
lationships being: [11
__
1] Nb || [0001] Al2O3 and
[112
__
] Nb || [101
__
0] Al2O3). After that, a very thin
iron layer of 15 A° thick was deposited at 820 K onto
the (110) niobium plane reacting with it to produce
a NbFe-ϕ alloy on the surface [29]. RHEED surface
analyses have shown a 2D rectangular lattice,
where the lattice parameters were close to those of
a C15 cubic Laves phase in (110) plane, 7.0 ± 0.1 A°
and 4.8 ± 0.1 A° . The TbFe2 layers were obtained by
co-deposition of the rare-earth and iron constitu-
ents, keeping the substrate at 820 K for YFe2 layer
deposition, and reducing the temperature to 620 K
for TbFe2 layer in order to avoid interdiffusion.
At this level the epitaxial relationships are:
[11
__
0] REFe2 || [11
__
0] Nb and
[001] REFe2 || [001] Nb. Finally, all the samples
were overcoated with a 200 A° thick layer of Y to
protect them from oxidation. A small average
roughness (≈ 25 A° ) of the layers was confirmed by
Atomic Force Microscopy [30] in all the samples.
However TbFe2 (300 A° )/YFe2 (1000 A° ) sample
presented islands quite close each other, in fact
quasi-continuously spread on the plane, and elong-
ated along [110] direction. The composition of the
bilayers was checked by microprobe analysis. The
stoichiometric composition was found to be within
± 2% for REFe2 (RE = Tb and Y). The thickness
was estimated by the calibration of the evaporation
rates, using quartz balances and optical sensors with
a 10% of error; x-ray diffraction scattering con-
firmed that all the films grown epitaxially on the
NbFe-ϕ buffer. They exhibited an average ≈ 0.6%
lattice expansion along [110]. The mismatch lattice
parameter between TbFe2 and YFe2 layers induces a
shear strain at the interface, εxy
0 ≈ − 0.61, − 0.58,
− 0.57, and − 0.54% for TbFe2 samples thickness of
300, 600, 1000, and 1300 A° , respectively [30,31].
The dispersion observed along the [220] orientation
corresponds to an average mosaic spread of −∼ 1.5°,
while the Bragg peaks broadening along this orien-
tation also indicates a mean coherence length ran-
ged between 200 and 400 A° . In addition, analyses of
Mo..ssbauer spectra carried out on the samples indi-
cate that [11
__
1] is the magnetization easy direction
in absence of applied magnetic field at room tem-
perature [30]. However, a careful analysis of these
results has also concluded that the magnetization
easy axis would be strongly influenced by the shear
deformation existing in the growing plane, which is
temperature dependent [30].
4. Magnetoelasticity of Ho films
We have measured the magnetoelastic stress in
two epitaxial Ho films, 5000 and 10000 A° thick
[32]. The interest in these measurements is double:
first, they are useful for checking the experimental
system and the validity of the hypotheses assumed
in the analysis explained in Sec. 2, as well as the
clamping conditions; second, the results obtained in
Ho films can be a reference for magnetoelastic stress
measurements in Ho-based superlattices, although
the studied samples, being so thick, are expected to
Fig. 2. Magnetoelastic stress isotherms σ~(b, a) for the
5000 A° Ho film.
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 349
exhibit a bulk-like behavior in their magnetic and
magnetoelastic properties.
In Figs. 2 and 3 we show the magnetoelastic
stress isotherms obtained for the 5000 A° Ho film
between 10 and 140 K. The measurements were
performed applying the magnetic field B along the
b-easy direction of holmium and with the sample
clamped along the b axis, σ~(b, a) in Fig. 2, and
along the a axis, σ~(b, b) in Fig. 3.
The magnetoelastic stress isotherms display
changes in the slope at certain critical fields that
can be ascribed to field induced transitions to dif-
ferent magnetic phases. However, the study of the
phase diagrams and the analysis of the differences
observed in the critical fields when compared with
the bulk Ho values (in the films, the critical fields
are higher [33], are not aims of the present work.
We will only stress that, at the maximum applied
field of 12 T, for temperatures below ∼ 90 K, we
have reached a ferromagnetic phase and we can
safely assure that the sample is saturated along the
easy axis.
Subtraction of the values of both magnetoelastic
stresses (see Eq. (25)), for the different tempera-
tures and at the maximum field of 12 T, gives the
temperature variation of the magnetoelastic stress
parameter Bγ,2, which is plotted in Fig. 4 for the
5000 A° film. We show in the same figure the fit of
the experimental results by using the expression
given by the standard Callen and Callen theory for
the magnetoelastic coupling [27], which reads
Bγ,2(T) = Bγ,2(0)Î5/2 [ � −1(m(T))] , (28)
where m is the reduced magnetization of the Ho
film (m = M(T)/M(0)); Î5/2 is the reduced hyper-
bolic Bessel function, Il+1/2/I1/2 , of order l = 2;
� −1 is the inverse Langevin function. The 0 K value
of the Bγ,2 stress obtained from the fit was
Bγ,2(0) = 0.29 ± 0.02 GPa. Similar results were ob-
tained for the 10000 A° film, for which Bγ,2(0) =
0.28 ± 0.02 GPa. Both values are in good agreement
with the basal-plane-symmetry-breaking magnetoe-
lastic stress parameter determined for bulk Ho, in
which an extrapolation from B = 3 T and T = 70 K
gives Bγ,2(0) = 0.275 GPa [34].
From the above results, i.e., the thermal depend-
ence and the values of the magnetoelastic stress
Bγ,2, we conclude that the origin of the magnetoe-
lastic strain in our Ho films is a single-ion crystal
electric field interaction. In addition, as antici-
pated, this strain is not affected by any surface
effect, likely due to the present high volume to
surface ratio. Then, in this case, Bγ,2 ≡ Bν
γ,2 . Let us
label this volume stress parameter obtained for
bulk-like films as Bν0
γ,2 .
5. Magnetoelastic behavior of Ho/Lu
superlattices
Magnetization and MEL stress measurements
have been performed in a series of [HonHo
/LunLu
]50
superlattices, with nHo ranging from 8 to 85 atomic
planes. Here we only report on the stress measure-
ments in the HonHo
/Lu15 samples, although the
magnetization isotherms and isofields also provide
valuable information about the different field-in-
duced and zero-field magnetic phase transitions
[33]. In this regard, the most remarkable fact is the
effect of the compressive strain in the basal plane of
Ho layers due to the Lu interleaving layers; this
compression, and the corresponding c-axis expan-
sion, stabilizes the ferromagnetic phase [3,8]. More-
over, the ferromagnetic transition temperature in-
creases for decreasing Ho fraction and the helifan
phases, intermediary between the helix and fan
phases, existing in bulk Ho [35] and also observed
by us in the Ho films, do not appear in the Ho/Lu
Fig. 3. Magnetoelastic stress isotherms σ~(b, b) for the 5000 A°
Ho film. Fig. 4. Thermal variation of the basal plane magnetoelastic
stress parameter for the 5000 A° Ho film. The continuous line is
the fit to Bγ,2(0)Î5/2[ � −1(m(T))], with Bγ,2(0) = 0.29 GPa.
J. I. Arnaudas et al.
350 Fizika Nizkikh Temperatur, 2001, v. 27, No. 3
superlattices. In the Ho/Lu superlattices with low
number of atomic planes of Ho, even the fan phase
is absent [33] (Fig. 5).
In Fig. 6 we display, as an example, the σ~(b, a)
and σ~(b, b) magnetoelastic stresses for the thin Ho
layers superlattice Ho8/Lu15 . The isotherms were
done in the temperature range from 10 to 160 K
(the lower limit of 10 K was chosen to avoid the
superconductivity of the Nb layer present in the
samples, appearing below 9 K, which uncontrol-
lably modifies the magnetization and the magneto-
elastic stress measurements). The magnetoelastic
stress isotherms show anomalies similar to those
displayed by the magnetization isotherms. We as-
cribe the abrupt jumps at the lowest temperatures
to transitions to a ferromagnetic state (e.g., see in
Fig. 6 the σ~(b, a) isotherms, at T ≤ 40). Smoother
anomalies at higher temperatures, which are better
detected in the field derivative of the stress iso-
therms, correspond to helix-fan or helix-ferromag-
net transitions. In the paramagnetic phase, T > TN ,
we observe the usual B2 thermodynamic behavior.
The basal plane magnetoelastic stress parameter
Bγ,2, which appears in the magnetoelastic energy
[6], is determined by using Eq. (25). However,
there is a difference with the case of the Ho films:
for a superlattice [AtA
/BtB
]r which has only a
volume fraction FA = tA/(tA + tB) of material A
undergoing magnetoelastic strain, the value that we
actually obtain from the experiment is Bmeas
γ,2 =
= Bγ,2fA . In our Ho/Lu superlattices we consider
that only the Ho layers behave magnetoelastically.
Also, because the c-axis parameters of both Ho and
Lu are very similar, the ratio between thicknesses
fA can be well approximated by the ratio between
numbers of atomic layers. Therefore, for A = Ho
and B = Lu, we have
Bmeas
γ,2 = Bγ,2
n
Ho
nHo + nLu
. (29)
For convenience, as we will see below, we have
plotted in Fig. 7 the values, at 10 K and 12 T, of
Bmeas
γ,2 (nHo + nLu) (with nLu = 15) as a function of
the number of Ho atomic planes in the bilayer
repeat, nHo . In the next section we will see that the
observed non-linear dependence of the product
Bmeas
γ,2 (nHo + nLu) with nHo denotes the presence of
additional terms in the magnetoelastic stress, i.e.,
in the present SL’s Bγ,2 is not simply Bν0
γ,2 (the
volume magnetoelastic stress obtained for the Ho
films). If the latter case would be true we would
obtain the dashed line shown in Fig. 7, as it is
obvious from Eq. (29).
5.1. Analysis of the basal plane cylindrical
symmetry breaking magnetoelastic stress
In a previous work [4], to explain the anomalous
thermal dependence of the basal plane magnetoelas-
tic stress in a [Ho6/Y6]100 superlattice, we showed
that it was necessary to consider that the associated
magnetoelastic parameter Bγ,2 had a linear depend-
Fig. 5. Magnetic phase diagram for the Ho22/Lu15 superlattice.
The critical fields Bc were obtained from magnetoelastic stress
and magnetization measurements; FM and H correspond to the
ferromagnetic and helix phases; AFB is a phase where FM Ho
blocks are antiferromagnetically coupled [8].
Fig. 6. Magnetoelastic stress isotherms for the Ho8/Lu15 super-
lattice.
a
b
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 351
ence on the strain (this implies to have a non-linear
dependence on the strain in the magnetoelastic
energy). In this paragraph we will outline a model
[36] which takes into account this fact, as well as
the effect of the interfaces in the basal plane mag-
netoelastic stress in superlattices.
(a) Let us assume that the volume magnetoelas-
tic parameter is, to first order in the strain,
Bν
γ,2 = Bν0
γ,2 + Dν
γ ε (30)
where Dν
γ = (∂Bv
γ,2/∂ε)ε=0 accounts for a possible
modification of Bν
γ,2 due to the epitaxial strain ε.
This strain occurs when the two materials, A and B
constituting the superlattice adjust their different
bulk lattice parameters to a common one, having a
intermediate value. If the lattice mismatch, defined
in terms of the bulk’s basal-plane a-lattice par-
ameters of both materials as e0 = (aB − aA)/aB, is
not too high (for A = Ho and B = Lu we have
e0 = − 0.017, at 10 K), it is reasonable to assume
that A and B will accommodate their lattice par-
ameters to a single value aA = aB = asl (see end of
Sec. 3) and that no misfit dislocations will appear
[37]. In Eq. (30) we have assumed isotropic basal
plane strains, i.e., εxx = εyy = ε.
With the above hypotheses it is easy to relate the
basal-plane strains of elements A and B through the
mismatch e0 , obtaining
εA =
asl
A − aA
aA =
asl
B − aB + aB − aA
aA −∼
−∼
asl
B − aB
aB +
aB − aA
aB = εB + e0 . (31)
The preceding relation allows to express the misfit
elastic energy of a bilayer A/B in terms of the
epitaxial strain εA. Minimization of that energy,
considering free x − y interfaces for the bilayer,
with no stress component in the z direction, gives us
the equilibrium value of the basal plane epitaxial
strain [33]:
εA = e0
tB
kAkB
−1tB + tA
, (32)
where kA,B are the following combinations of elastic
constants for the A and B elements:
kA,B = c11
α,A,B
1 −
2c13
A,B
c
33
A,B
2
−
−
1
√3
c12
α,A,B
1 −
2c13
A,B
c33
A,B
2
+
1
12
c12
α,A,B
1 +
2c13
A,B
c33
A,B
2
.
Now, Eq. (30) can be written in the form
Bν
γ,2 = Bν0
γ,2 + Dν
γ,2 e0
t
B
kAkB
−1t
B + tA
, (33)
where kHo/kLu is close to 1, for A being a rare-earth
element and B being Y or Lu.
(b) The interfacial contribution to the magnetoe-
lastic stress is now treated similarly to the case of
the interface anisotropy: the symmetry breaking at
the surfaces or interfaces was related with the
appearance of an additional term in the magnetoc-
rystalline anisotropy, this term being proportional
to the surface-to-volume ratio 1/t, where t is the
magnetic film (or block, in superlattices) thickness
[38]. Then, considering the two surfaces or inter-
Fig. 7. Bmeas
γ,2 (nHo + nLu) for the series Hon/Lu15 (●), with
8 < n < 85. The line is the function Bγ,2nHo , where Bγ,2 is given
by Eq. (34), with the fitting parameters Bν0
γ,2 = 0.27 GPa, B̂surf
γ,2 =
– 7 GPa and Dν
γ,2 = – 135 GPa, taking e0 = (aLu − aHo)/aLu =
− 0.017. The dashed line is Bν0
γ,2nHo (a). Same as (a) in a
double-logarithmic plot, including the values for the two Ho
films (❍) (b).
J. I. Arnaudas et al.
352 Fizika Nizkikh Temperatur, 2001, v. 27, No. 3
faces per magnetic block, we add an interfacial term
2Bsurf
γ,2 /tA to the volume magnetoelastic stress of
Eq. (33), obtaining
Bγ,2 = Bν0
γ,2 +
2B̂surf
γ,2
nA
+ Dν
γ,2e0
t
B
kA
kB
−1tB + tA
, (34)
where B̂surf
γ,2 = Bsurf
γ,2 (cA/2)−1 is the value of the inter-
facial magnetoelastic stress expressed in units of
energy per unit volume of the superlattice (cA/2 is
the (a, b)-planes spacing for element A).
For the HonHo
/LunLu
superlattices, substitution
of Eq. (34) in Eq. (29) gives
Bmeas
γ,2 =
=
nHo
nHo + nLu
Bν0
γ,2 +
2B̂surf
γ,2
nHo
+ Dν
γ,2e0
nLu
kHokLu
−1nHo + nLu
with kHo/kLu = 0.95 and e0 = − 0.017. Now, from
Eq. (35), the advantage of plotting Bmeas
γ,2 (nHo + nLu)
vs. nHo becomes clear. In Fig. 7 we show the fitting
of the experimental values to the function Bγ,2nHo ,
where Bγ,2 is given by Eq. (34). The parameters
giving the best fit are Bν0
γ,2 = 0.27 GPa, B̂surf
γ,2 =
= − 7 GPa and Dν
γ,2 = − 135 GPa. In a log–log plot
(Fig. 7) we have included the values for two Ho
films studied, also satisfying Eq. (35). Note that
the obtained Bν0
γ,2 agrees with the value of Bγ,2 for
bulk holmium [34], which supports the proposed
model. Interestingly, we have found a quite high
value of the epitaxial stress increase of the un-
strained volume stress parameter (Dν
γ,2e0nLu/(nLu +
+ nHo) ≈ 5.4⋅Bν0
γ,2 for nHo = 8). Also, we have de-
duced a interfacial contribution to the magnetoelas-
tic stress which is larger and of opposite sign to the
volume stress (2B̂surf
γ,2 ≈ − 25⋅Bν0
γ,2 ; e.g. 2B̂surf
γ,2 /nHo ≈
≈ − 6.4⋅Bν0
γ,2 for nHo = 8). The interfacial stresses
found in the present HonHo
/Lu15 superlattices are
even higher than the highest values previously repor-
ted in Cu/Ni/Cu trilayers (B̂surf
γ,2 ≈ − 6⋅Bν0
γ,2) [39].
5.2. Temperature variation of the magnetoelastic
stresses
The study of the thermal dependence of the
magnetoelastic stresses serves as a separate test of
the previous analysis and the ensuing fitting pa-
rameters. For volume-originated distortions like
Bν0
γ,2 and Dν
γ,2 , the standard Callen and Callen
theory [27] predicts a temperature variation of the
type Î5/2[ � −1(m(T))] (see Eq. (28)) . For the
interface magnetoelastic parameter Bsurf
γ,2 , a depen-
dence type m4(T) at low temperatures is predicted
[40] if we assume that, at the interfaces, the spin
dimensionality is D = 2 (some interface anisotropy
could account for this reduction in dimensionality).
In the high temperature regime both volume and
interface stress parameters should display the same
thermal dependence, type m2(T) [27,40]. Therefore,
we have analyzed the temperature variation of Bγ,2
by using the expression
Bγ,2 = (Bν0
γ,2 + Dν
γ,2ε) Î5/2[ � −1(m)] +
2B̂surf
γ,2
nHo
mα ,
(36)
where α = 4 at low temperatures and α = 2 at
temperatures close to TN .
In Fig. 8 we show the temperature variation of
Bγ,2 at 12 T for two of the studied samples, together
with the theoretical fits obtained by using Eq. (36).
Fig. 8. Thermal dependence of Bγ,2 for the Ho30/Lu15 and
Ho40/Lu15 superlattices. The lines are theoretical fits using Eq.
(36), which includes interface Bsurf
γ,2 and volume Dν
γ,2 terms, the
latter associated to the epitaxial strain.
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 353
Similar satisfactory results were obtained for the
other samples, except for the superlattice with
nHo = 8, were the fit was poorer. We employed the
parameters previously deduced from the analysis of
the Ho-thickness dependence of Bmeas
γ,2 (nHo + nLu)
(see Fig. 7), allowing less than a 9% variation in
their values. The interface contribution, varying as
m4(T), was included below 50, 80, 55, and 40 K, for
the superlattices with nHo = 14, 30, 40, and 45,
respectively. These results support the validity of
the magnetoelastic parameters obtained and confirm
the single-ion crystal electric field origin for the
magnetoelastic coupling in the Ho/Lu superlat-
tices.
6. Magnetoelastic behavior of Ho/Y
superlattices
We have carried out magnetoelastic stress meas-
urements in the superlattices Ho10/YnY
, where
nY = 10, 30, and 40. In this case the interface
contribution to Bγ,2 remains unchanged, since the
Ho layers thickness is kept constant. However, the
influence of the epitaxial strain ε on the volume
magnetoelastic parameter is expected to be in-
creased for increasing values of the Y layers thick-
ness (see Eq. (33). In Fig. 9 we show, as an
example, the magnetoelastic stress isotherms for the
Ho10/Y10 superlattice, where the transitions from
helix to ferromagnetic phase at low temperatures
are observed in the form of relatively abrupt chan-
ges in the stress; at higher temperatures, additional
changes in the isotherm slope are associated with
intermediary fan phases (the magnetization iso-
therms display such kind of changes at the same
fields). Above TN the paramagnetic behavior is
observed.
However, the most striking fact found in the
present Ho/Y series is the change of sign of the
magnetoelastic stress parameter Bγ,2. In Fig. 10 we
show the σ~(b, a) and σ~(b, b) isotherms, at 10 K and
12 T, for the Ho10/Y10 and Ho10/Y40 superlat-
tices. For comparison, the values obtained for the
5000 A° film are also shown. Note that, in order to
compare the experimental values of σ~(i, j) corre-
sponding to samples with different Ho volume frac-
tions, we have plotted in Fig. 10 the measured
stresses multiplied by (tY + tHo)/tHo , according to
Eq. (29) applied to the Ho/Y case (we denote
these normalized stresses as σ~mag(i, j). The mag-
netoelastic stress Bγ,2 = 2[σ~mag(b, b) − σ~mag(b, a)]
changes its sign along the series, as it is shown in
Fig. 11, where the 10 K and 12 T values of Bγ,2 are
plotted against the volume fraction of Y. In this
series, the interface contribution, inversely propor-
tional to the Ho thickness, must be constant. There-
fore, the plot in Fig. 11 reveals the dependence of
the magnetoelastic stress with the epitaxial strain
ε on the Ho blocks, which is proportional to the Y
volume fraction (see Eq.(32)). The line in Fig. 11
is a fit to the model Eq. (34), where A = Ho and
B = Y, nHo = 10, kHo/kY = 1.05 and e0 = 0.023. The
parameters obtained from the fit are Bν0
γ,2 = 0.27 GPa,
B̂surf
γ,2 = 2.16 GPa, and Dν
γ,2 = − 56 GPa (the values
Fig. 9. Magnetoelastic stress isotherms σ~(b, a) and σ~(b, b) for
the Ho10/Y10 superlattice.
Fig. 10. Magnetoelastic stress isotherms σ~mag(b, a) ( ), and
σ~mag(b, b) (- - -) at 10 [4] K for the Ho10/Y10 and Ho10/Y10
superlattices and for the 5000 A° Ho film.
J. I. Arnaudas et al.
354 Fizika Nizkikh Temperatur, 2001, v. 27, No. 3
for bulk Ho and for the 5000 A° Ho film have not
been included in the fit, although they are plotted
in Fig. 11 to emphasize the change of the sign of
Bγ,2 in the Ho10/YnY
superlattices, for the film,
tY = 1000 A° , the Y seed thickness). For this series
the constant interfacial contribution to the magne-
toelastic stress is near two times the unstrained
volume magnetoelastic stress (2B̂surf
γ,2 /nHo ≈ 1.6⋅Bν0
γ,2 ,
with nHo = 10). The epitaxial stress contribution to
the unstrained volume stress parameter depends on
the Y thickness but it always has opposite sign to
Bν0
γ,2 (e.g., Dν
γ,2e0nY/(nY + nHo) ≈ − Bν0
γ,2 , with
nHo = 10 and for nY = 40).
The explanation of the change of sign of Bν
γ,2 is
another confirmation of the validity of the model
assumptions, mostly the existence of a strain de-
pendence of the volume magnetoelastic parameter,
represented by the term Dν
γ,2ε in Eq. (34). The
parameter Dν
γ,2 is negative for both the Ho/Lu and
Ho/Y studied series; however, the different sign of
the lattice mismatch e0 in both cases makes possible
to observe the change of sign of Bγ,2 only in the
Ho/Y series.
In order to understand the origin of the interface
contribution to the magnetoelastic stress, a model
of crystal electric field including screening due to
the conduction band electrons has been proposed
[41]. In this model, a Hartree-Fock-type dielectric
constant and a Gaussian distribution, of half-width
b, for the electronic charge density of the Ho3+ ions
are assumed. The calculated values of Bν0
γ,2 and
Bsurf
γ,2 , for the Ho/Lu superlattices, agree in sign
and magnitude with those determined in our experi-
ments, indicating that both the volume and, more
interestingly, the interface magnetoelastic stresses
have their origin in the single-ion interaction with
the distorted crystal electric field. The experimental
values can be well reproduced if the Fermi energy
of the interface, before epitaxy, is smaller than the
volume one and if we take a density of states close
to the obtained one from the bulk RAPW band-
structure calculations for Dy. The difference be-
tween the Fermi-energy values is simply, but quan-
titatively, explained assuming the formation of
Ho3+ interface weakly localized electron states close
to the Ho/Lu ideal contact plane [41]. However,
two cautions are suggested: small systematic vari-
ations with nHo of the volume contribution Bν0
γ,2
(and of Dν
γε) will result in variations of Bsurf
γ,2 ,
although such an effect cannot be ascertained with
our available experimental data; however, the sign
and order of magnitude should probably be pre-
served, as both our volume and interface MEL
parameters must be single-ion properties. Therefore,
a more refined theory of the conduction electron
band structure for the whole superlattice could
predict changes in the Ho volume MEL stress ex-
tended to a certain number of layers within the Ho
block.
7. Competing anisotropies in the
magnetoelastic behavior of
Ho/Tm superlattices
In this section we present a study of two Ho/Tm
SL’s: Ho8/Tm16 and Ho30/Tm16 , nearly isomor-
phous to the corresponding Hon/Lu15 samples ana-
lysed in Sec. 5. The differences when substituting
Ho for Lu are, however, of fundamental import-
ance. Now, the Tm «spacers» between Ho blocks
are, not only magnetic, but also have the magneti-
zation easy axis along c. In fact, bulk thulium has
its magnetic structure longitudinally modulated
along the c axis below TN = 58 K [42]. Its strong
axial anisotropy hardly leaves that its magnetic
moments been tilted out of the c axis when a
magnetic field is applied perpendicularly (within
the basal-plane). This deviation is lesser than a 2%
of the total moment at low temperatures and high-
fields, −∼ 10 T [43]. However, a magnetic field ap-
plied parallel to the c axis can break the modulated
structure along this axis, and then gets a fully
aligned magnetization state, −∼ 2700 emu/cm3 at 0 K,
for magnetic field higher than 2.8 T [43].
Thus, having the Ho/Lu system as a well-
known reference, we have chosen Ho/Tm SL’s to
study the presumable competition between the ani-
sotropies of holmium and thulium layers in a SL
system. By doing magnetization and MEL stress
measurements we have analysed the differences
which are observed when comparison with the hol-
Fig. 11. Dependence of Bmag
γ,2 with nY/(nY + nHo). The line is a
fit to Eq. (34) (see text for details). The values for bulk Ho
(❍) and for the 5000 A° Ho film (∆∆) are included.
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 355
mium-lutecium results is made [44]. In this Ho/Tm
SL’s, the analysis of the magnetization gives some
insight about the kind of interaction existing be-
tween the Ho and Tm layers. However, the MEL
stress measurements prove much better than magne-
tization whether the holmium layers or thulium
ones dominate or not the magnetic behavior in
Ho/Tm SL’s. This is because, the MEL stress is a
power of the magnetization [27,40]. In this way, a
competition between the anisotropies of Ho and Tm
ions can be clearly revealed by determining the
MEL stresses in the Ho/Tm SL’s. The different
sign of the Ho and Tm anisotropies makes opposite
the contribution to the magnetostriction from both
ions, in particular, for the basal-plane cylindrical
symmetry breaking MEL stress that we determine
in these hexagonal-symmetry systems.
In Fig. 12 we show the σ~(b, b) and σ~(b, a) MEL
stress isotherms, for the Ho8/Tm16 and Ho30/Tm16
superlattices. The MEL stress isotherms for the
Ho8/Lu15 SL were shown in Fig. 6, and those for
the Ho30/Lu15 are similar in their field dependence
(saturation at relatively small magnetic fields,
H ≥ 0.5 T, and being more saturated when decreas-
ing the temperature), although they have different
saturation values.
As a first approach, and for comparison with the
Ho/Lu SL’s, we can analyse the MEL stress data
obtained in the Ho/Tm samples assuming that the
Tm moments do not leave their magnetization easy
axis, even under the action of the maximum applied
field of 12 T. In this way we obtain the Bγ,2 MEL
parameter, which is shown as a function of the
temperature in Fig. 13, for the two Ho/Tm SL’s
studied, together with the Bγ,2 values for the iso-
morphous Ho/Lu samples.
The values of Bγ,2 for the Ho/Tm SL’s are
positive at high temperatures and very close to the
Ho/Lu ones, showing a broad peak at around
80–90 K. Below 80 K, Bγ,2 start to decrease, reach-
ing a maximum slope on decreasing the temperature
at −∼ 60 K, and undergoing a change of sign at a
certain temperature, which is 60 K for Ho8/Tm16
and 20 K for Ho30/Tm16 . Considering the single-
ion CEF origin of the Bγ,2 MEL stress, this strong
deviation towards negative values is interpreted as
a result of the competition of the Ho and Tm
anisotropies. Moreover, the effect of the Tm layers
is much more intense in Ho8/Tm16 SL than in
Ho30/Tm16 , where the basal-plane anisotropy of
holmium blocks dominates the Bγ,2 MEL stress,
making the values more positive for this SL, where
the thickness of holmium layers is high enough with
respect to the thulium ones. This is in good agre-
ement with the magnetization results too [44]. So,
our MEL stress experiments clearly indicate that
the Tm moments are tilted out from the c axis when
a magnetic field is applied within the basal plane of
the hcp structure (otherwise the values of Bγ,2
would be nearly the same as for the Ho/Lu SL’s),
this effect being larger in the case of the
Ho30/Tm16 sample than for the Ho8/Tm16 SL.
Nevertheless, experiments in higher magnetic fields
are needed to attempt the full saturation of these
competing anisotropy SL’s, and, in this way, to be
able to obtain separately the Tm and Ho MEL stress
parameters [44].
8. Magnetoelastic behavior of Dy/Y and
Er/Lu superlattices
In dysprosium based SL’s the determination of
the easy basal plane cylindrical symmetry breaking
Fig. 13. Thermal dependence of Bγ,2 for the Ho8/Tm16 and
Ho8/Lu15 SL’s (a) and the Ho30/Tm16 and Ho30/Lu15 SL’s (b).
Fig. 12. Magnetoelastic stress isotherms for the Ho8/Tm16 and
Ho30/Tm16 superlattices.
J. I. Arnaudas et al.
356 Fizika Nizkikh Temperatur, 2001, v. 27, No. 3
magnetoelastic stress, Bγ,2, is of paramount interest,
inasmuch as it decisively helps to drive the sponta-
neous helical (H) ferromagnetic (F) transition in
bulk Dy, at Tc > 85 K [7]. However, it was
observed than in Dy/Y superlattices, grown along
the c axis, the Y block tensile stress suppresses the
H-F transition [1].
In bulk Er the c axis is easy, and the magnetic
structure changes from paramagnetic to sinusoidal
at TN1 > 85 K, to the (a, c) plane elliptically
cycloidal structures at TN2 > 52 K and to ferro-
conical phase of wave vector (5/21) c∗ at Tc > 20 K
[45,46]. To our knowledge no magnetic phase diag-
rams have been traced for Er/Lu superlattices.
We have carried out magnetoelastic stress measu-
rements in the superlattices (Dyn/Y15) and
(Erm/Lu10), with n = 5, 15, 25, and m = 10, 20, 30
atomic planes, from 10 K and in applied magnetic
fields up to B = 12 T; the magnetic field was
applied within the basal plane, along the a easy
axis, for both series.
In Fig. 14 we present, as an example, the
σ~(a, b) and σ~(a, a) isotherms for B || a, for the
[Dy25/Y15]50 SL. We notice changes in slope for
certain fields that we tentatively ascribe to the
H-fan (FN) and to the FN-F transitions. These
fields correspond well with the observed ones on the
magnetization isotherms. As we may observe, sa-
turation is practically accomplished at 12 T for
T < TN . This indicates that we are measuring a MS
of crystal electric field origin, once the sample is F.
For the Er SL’s in Fig. 14,b we present, as an
example, the same kinds of isotherms for the
Er30/Lu10]40 SL, where we observe: three changes
in slope for T < Tc , two for Tc < T < 40 K, and no
features for 40 K < T < TN1 , if we tentatively
ascribe the same phases of the bulk to the SL
[45,46]. Similar changes in slope were observed in
the magnetization isotherms at about the same fields for
T < Tc , although for the range 40 K < T < TN1 , a
slope change is observed at about 2 T.
According to bulk neutron diffraction measure-
ments and mean field calculations [45,46] we tenta-
tively ascribe the transitions for T < Tc as follows:
The transition at Hc1 ≅ 3 T could be to a conical-fan
structure of wave vector (1/4) c∗, which, at Hc2 ≅
≅ 8 T, transforms to an FN structure around the a
axis. Finally, the transition at Hc3 ≅ 12 T could be
toward an F one along the a axis. Notice that only
the transition at Hc1 is abrupt. For the range
Tc < T < 40 K, we observe changes of slope at
about 1, 2, 5, and 12 T. If we tentatively translate
from the bulk [45] these fields would respectively
correspond to transitions to commensurable cycloi-
dal structures of wave vectors (in c∗ units) 6/23
and 4/15, a fan 2/7 about the basal plane, and
toward an F phase along B.
If we analyse the thickness dependence of the
magnetoelastic stress of the Dy/Y SL’s as we did
for the Ho/Lu SL’s, we obtain the values: Bν0
γ,2 =
= 0.85 GPa, Dν
γ = 5.2⋅102 and Bsurf
γ,2 /(c/2) =
= − 22.8 GPa [5] (note that the SL with n = 5 can
only be included in the analysis duplicating the
value of Dν
γ , which indicates that the SL’s with
very thin Dy blocks are more sensitive to epitaxial
strain variations). The effective interface magnetoe-
lastic stress 2Bsurf
γ,2 tDy is strong compared to the
volume term, up to about one order of magnitude
larger and of the opposite sign.
Fig. 14. Magnetoelastic stress isotherms for the SL’s
[Dy25/Y15]50 (a) and [Er30/Y10]40 (b), σ~a ≡ σ~(a, b) and
σ~b ≡ σ~(a, a) correspond to SL clamping along the a and b axes.
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 357
Another test of the CEF origin for the Bγ,2 is to
check if it scales with the reduced saturation (at
B = 12 T) magnetization m = M(T)/M(0) as m3,
according to the standard theory of magneto-
striction [27]. In Fig. 15 we compare for the
Dy25/Y15 SL the thermal variation of Bγ,2 (12 T)
with the one for m3 (12 T); the agreement is
satisfactory. For the Ern/Lu10 SL’s we have done a
study fully similar to the one above for the Dy/Y
SL’s. However, as it can be seen in the isotherms
plotted in Fig. 14,b the magnetoelastic stresses are
far away from saturation, as it happens with the
magnetization [33]. Thus, the analysis of the ex-
perimental results in this case is inconclusive (see
Fig. 15 and Ref. 5). Magnetoelastic stress experi-
ments in larger fields, up to 30 T, which are needed
to saturate this SL’s within the growing plane, are
in progress.
9. Magnetoelastic behavior of (110)
TbFe
2
/YFe
2
bilayers
For the TbFe2/YFe2 bilayers, the MEL stress
measurements were performed between 10 K and
RT and the magnetic field, up to 12 T, was applied
within the growing plane (temperatures below 10 K
were avoided because of the diamagnetic behavior
of the Nb buffer). In Fig. 16 we show, as an
example, isotherms corresponding to the TbFe2
(600 A° )/YFe2 (1000 A° ) bilayer, for the stresses:
σ~([1
__
10], [001]), σ~([1
__
10], [1
__
10]) and σ~([1
__
11], [11
__
2])
(see Eqs. (16), (17), and in Sec. 2.2.2). For the
sake of clarity, and because of the similarity be-
tween the different isotherms for a given sample, we
have only plotted the isotherms obtained at the
lowest and the highest temperatures of the measur-
ing range. At the beginning, and before starting
each measurement, we have demagnetized the
sample. Due to the high Curie temperature of
TbFe2 , of about 696 K, it was unfeasible with our
experimental set-up to get any virgin state by heat-
ing up first the sample above 700 K and, then,
cooling it down to get the initial demagnetized
state. Moreover, at these high temperatures the
bilayer structure would be destroyed due to the
diffusion between the layers. Instead of this, we
have performed, at the measuring temperature, hys-
teresis cycles with decreasing amplitude of the
maximum applied magnetic field in each cycle, to
get a macroscopically demagnetized sample. From
the MEL stress isotherms, we can distinguish two
kinds of magnetoelastic behavior, depending on the
applied magnetic field direction. For those measure-
ments where the magnetic field is parallel to [1
__
11]
direction (see Fig. 16,c), the MEL isotherms show
a large hysteresis with coercive field values very
close to those observed in the magnetisation meas-
urements [18]. On the contrary, when the field is
applied along [1
__
10] direction, the MEL curves
hardly show hysteresis, the fields for zero MEL
stress crossing being even smaller than the corre-
sponding ones obtained from magnetisation hys-
teresis measurements. In all the cases the coercive
fields increase on reducing the thickness of TbFe2
layer. In spite of these differences, it is important to
note that all the loops have approximately the same
Fig. 15. Comparison of Bγ,2(T) with m3(T) at 12 T, where m is
the reduced magnetization, for Dy25/Y15 and Er30/Lu10 SL’s.
Fig. 16. The magnetoelastic stress isotherms for
TbFe2 (600 A° )/YFe2 (1000 A° ) bilayer: σ([1
__
10], [001]) (a);
σ([1
__
10], [1
__
10]) (b), and σ([1
__
11], [11
__
2]) (c). For the sake of
simplicity we only represent the RT and 10 K isotherms.
J. I. Arnaudas et al.
358 Fizika Nizkikh Temperatur, 2001, v. 27, No. 3
«closing field», which has very large values (≈ 6 T
at RT). Moreover, at the maximum applied field of
12 T, the σ~(α, β) MEL stresses do not show satura-
tion at all. This lack of saturation even occurs when
the magnetic field is applied along the easy [1
__
11]
direction, which, however, seems to show the best
approach to the saturation. To some extent, magne-
tostriction is more sensible to show the lack of the
full saturation than magnetization, a usual fact.
Therefore, since the magnetization is indeed satu-
rated at 12 T [18], we will consider the MEL stress
values at 12 T as corresponding to a saturated
regime.
9.1. Magnetoelastic stress parameters
We have obtained the b0 , b1 , and b2 MEL
stress parameters as explained in Sec. 2.2.2. Their
thermal dependences for the different TbFe2(t)/YFe2
bilayers are shown in Fig. 17 (the t = 300 A° sample
is the only one of the series which does not show a
measurable MEL stress). We can see a smooth but
monotonous decrease of all the parameters on rising
up the temperature, and also that the values of b0
are about one order of magnitude smaller than those
of b1 and b2 . Notice that, because of the selected
applied field and MEL stress directions, it has been
possible to determine directly all the second order
MEL stress parameters in the present bilayers.
However, b0 was never determined before in bulk
TbFe2 , and therefore, we cannot compare its value
with those ones for the TbFe2/YFe2 bilayers.
We also can observe that b1 , for
TbFe2 (1300 A
° )/YFe2 (1000 A
° ) sample and
TbFe2 (1000 A
° )/YFe2 (1000 A
° ) sample are very
close in value, −∼ – 0.4 GPa at 10 K and 12 T, but
for TbFe2 (600 A
° )/YFe2 (1000 A
° ) it is about 25%
smaller than in the others. This could indicate that
b1 tends to saturate for increasing TbFe2 block
thickness in the bilayer. However, from anisotropic
magnetostriction, λt
s , measurements in polycrystal-
line TbFe2 and from measurements in crystalline
TbFe2 , and using the relationship λt
s = (3/5)λ[100] +
+ (9/10)λ[11
__
1] [47–51], we have estimated a 0 K
value of b1 −∼ – 0.11 GPa for bulk TbFe2 , which is
less than half of those obtained for the t = 1000 A°
and 1300 A° studied bilayers. The origin of this
enhancement of MEL stress in the bilayers will be
analysed in more detail later.
It is interesting to note that b2 is 74% of the
bulk’s value for TbFe2 (1300 A
° )/YFe2 (1000 A
° ),
at RT and 12 T, whereas it is only a 55% of it in the
case of TbFe2 (1300 A
° ) on NbFe–ϕ [4]. This would
indicate that the use of NbFe–ϕ (15 A° ) as buffer
instead of YFe2 (1000 A
° ) decreases the value of the
MEL stress, b2 , in the (110) TbFe2 (1300 A
° ) thin
films. The reason why this happens is not very well
understood at the present, although the epitaxial
strain should be playing an important role. The
strain originated by the epitaxial growth of the
1300 A° thick TbFe2 layer onto the 1000 A° thick
YFe2 buffer is −∼ – 0.54%, at RT [31] and, when the
buffer is a 15 A° thick NbFe–ϕ layer, the epitaxial
strain increases up to −∼ – 0.64% [30,52]. Therefore
such a small misfit strain reduction changes sub-
stantially the rhombohedral, b2 , MEL parameter.
This could also happen with the observed increase of
|b2| (see Fig. 16) when t increases because, as the
TbFe2 block thickness increases, the shear strain,
εxy , decreases [31]. Now, and since we have infor-
mation about the in-plane strains in the bilayers, we
can perform the same kind of analysis as we did for
the RE SL’s, although modified for the current
cubic symmetry, to obtain information about the
origin of the MEL coupling in the present systems.
So, for b2 we can write
b2 = b2
v +
b2
s
t
= (b2
v0 + d2
v εxy
0 ) +
b2
s
t
, (37)
where b2
v0 is the main volume contribution, d2
v =
= (∂b2/∂εxy
0 ) taking into account the non-linear ef-
fects on the MEL energy of Eq. (8). This strain,
Fig. 17. The temperature dependence of the MEL stress par-
ameters: b0 (isotropic MEL stress) (■); b1 (tetragonal MEL
stress) (●), and b2 (rhombohedral MEL stress) (▲), for
TbFe2(t)/YFe2 (1000 A° ) bilayers at different t, A° : 600 (a);
1000 (b); 1300 (c).
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 359
within the elastic regime (see Sec. 5.1), for the
cubic symmetry reads
εxy
0 = ε0
1000 A°
β2t + 1000 A°
, (38)
where ε0 (−∼ – 0.7%)31, is (a⊥
TbFe
2 − a⊥
YFe
2)/〈a⊥〉, being
a⊥
REFe
2 the lattice parameters along [220] and β2 ≡
≡ c44
A /c44
B , where c44
A,B are the shear elastic con-
stants for A ≡ TbFe2 and B ≡ YFe2 . Notice that the
value of β2 −∼ 0.2 has been deduced from the εxy
0
experimental results (only at RT [30,31] by using
Eq. (38) (remember that εxy
0 depend on t and their
values have been given in Sec. 3.3). The small value
obtained for β2 (<< 1) could indicate that the elas-
tic regime assumption that we are taking for
granted would not be valid for large t values (we
should note that, if we consider a possible disloca-
tion energy, we would arrive to a dependence with
the thickness similar to that appearing in Eq. (38)
[53,54]; so, both contributions, elastic and disloca-
tion energy, should be taken into account and this
could change the value obtained for β2). The last
term in Eq. (37) is the Ne′el interface contribution
to the stress. In Fig. 18,b we plot the RT values of
b2 versus TbFe2 thickness, as obtained from Fig. 17.
The small value of b2 for the TbFe2 (300 A
° )/YFe2
(1000 A° ) bilayer [17], and the bulk TbFe2 value,
b2
v0 = – 0.307 GPa [47,48] are also shown. We
represent in the same Figure the fit of these values
by the model expression given in Eq. (37). The fit
is good regardless the reduced number of experi-
mental points available. The parameters giving the
best fit are: d2
vε0 = 0.0158 GPa and b2
s = 98 GPa⋅A° ,
and so d2
v = – 2.3 GPa. For instance, for t = 1000 A° ,
the contributions to b2 amount: b2
v = – 0.313 GPa
and b2
s/t = – 0.1 GPa. We see that the overall
volume MEL stress is practically the same as the
bulk one and three times the interfacial contribu-
tion. Clearly non-linear MEL effects on b2 (rhom-
bohedral striction) are small for our bilayers.
The same kind of analysis can be applied to b1 ,
i.e.,
b1 = b1
v +
b1
s
t
= (b1
v0 + d1
vεxx
0 ) +
b1
s
t
, (39)
where d1
v = (∂b1/∂εxx
0 )eq with εxx
0 the misfit strain.
This strain is
εxx
0 =
+
ε0
2
− e0
1000 A°
β1t + 1000 A°
−
ε0
2
� , (40)
being e0 the lattice mismatch for the bulk TbFe2
and YFe2 (e0 −∼ – 0.19%) and � ≡ 1000 A° t [c11
A c12
B −
− c11
B c12
A ]/{[t(c11
A + 2c12
A ) + 1000 A° (c11
B + 2c12
B )][t(c11
A −
− c12
A ) + 1000 A° (c11
B − c12
B )]} and β1 ≡ (c11
A +
+ 2c12
12
A )/(c11
B + 2c12
B ), where c1,j
A,B (j = 1, 2) are the
elastic constants for A ≡ TbFe2 and B ≡ YFe2 .
From the data available in the literature [47], we
obtain that � −∼ 0 and β1 −∼ 0.7. In Fig. 18,a we
present the RT values of b1 for our bilayers and for
bulk TbFe2 , b1
v0 = – 49.5 MPa [47,48,51]. Notice
that, for the TbFe2 (300 A° )/YFe2 (1000 A° ) bi-
layer, b1 is negligible compared with the other b1
values. Although the value for t = 300 A° could not
be experimentally determined, we should notice
that b2 in this bilayer is rather small and that
b1 < b2 for bulk TbFe2 and for the rest of the
bilayers. Again the fit by model Eq. (39) is good in
the available region of thickness. Now the fitting
parameters are: d1
ν(+ε0/2 − e0) = – 0.32 GPa and
b1
s = 110 GPa⋅ A° . At RT, +ε0/2 − e0 = – 0.16% and
then, d1
v = 200 GPa. For instance, for t = 1000 A° ,
the different contributions to b1 are: b1
ν = – 0.24 GPa
and b1
s/t = 0.11 GPa, both being of comparable
size, but of opposite sign. This is clearly reflected in
the minimum exhibited by b1 at t −∼ 1000 A° , because
of the competition of b1
ν and b1
s . For b2 both
contributions, b2
ν and b2
s , have the same sign and,
so, the competition is absent. It is also clear that for
Fig. 18. TbFe2 block thickness dependence of the tetragonal
MEL stress parameter, b1 (a), and the rhombohedral MEL
stress, b2 (b), at RT. The continuous lines represent the best
fits using the Eqs. (39) and (37). The horizontal dotted lines
are the values of b1 and b2 for bulk TbFe2 . Notice the asymp-
totic approximation to these values of b1 and b2 for the bilayers.
J. I. Arnaudas et al.
360 Fizika Nizkikh Temperatur, 2001, v. 27, No. 3
b1 the misfit strain strongly affects the bulk’s value,
b1
ν0 , although keeping its sign.
The same kind of fitting was done for the other
measurement temperatures. The temperature vari-
ations of the fitted parameters d1
ν(+ε0/2 − e0) and
d2
νε0 are shown in Fig. 19,a. Due to the lack of
experimental data about ε0 , we cannot separate the
MEL parameters d1
ν and d2
ν for the whole range of
temperatures, as we did at RT. The thermal vari-
ation of d1
νεxx
0 and d2
νεxy
0 is rather different: |d1
νεxx
0 |
decrease with increasing temperature and |d2
νεxy
0 |
remains constant. In Fig. 19,b we show the thermal
variation of the interface MEL parameters b1
s and
b2
s . Noticeable is that b1
s and b2
s are almost the
same, although they are associated to quite different
MEL stresses modes, b1 (tetragonal stress) and b2
(rhombohedral stress). This means that ∂b2/∂εxy
0 and
∂b1/∂εxx
0 are quite sensitive to the magnetostriction
mode. This is not surprising since both are axial
modes.
To ascertain the spin dimensionality, D, for the
different MEL contributions to the MEL stress
parameters we analysed their thermal dependence in
terms of Callen and Callen standard theory [27,40]
of single-ion crystal electric field magnetostriction.
Thus, for volume spins (D = 3) we expect: bi
ν(T) =
= bi
ν,0(T) + (di
ν(0 K)εi
0(T, t))Î5/2[ � −1(m(T))] (i = 1,
2 and ε1
0 = εxx
0 , ε2
0 = εxy
0 ), where m(T) is the reduced
magnetization of Tb sublattice (m(T) = MTb(T)/MTb(0));
� −1 is the inverse Langevin function; Î5/2 is the
Bessel reduced function of first kind. In Fig. 19,a
we plot the thermal dependences of d1
νεxx
0 and
d2
νεxy
0 and the fit by the above formula, where we
have used the values of bi
ν,0(T) existing in the
literature for bulk TbFe2 [47–51]. The agreement is
good, indicating Heisenberg spins.
On the other hand, for the interfacial MEL
stresses we tried two scalings: bi
s(T) = bi
s(0 K)mα(T)
(i = 1, 2) with α = 4 at low temperatures and α = 2
in the high temperature range. This is the scaling
predicted for spin dimension D = 2; a scaling with
α = 3 for low temperatures indicates an spin dimen-
sion D = 3.
The reduced magnetization, m(T), was deter-
mined by subtracting from the Ms,TbFe2
(T), the
weighted iron sublattice magnetization, which was
assumed to be the same as for YFe2 [18]. In Fig. 19,b
we plot, as an example, the scaling of the interfacial
MEL stress parameter bi
s(T) (i = 1, 2) with m4(T)
and m3(T) (notice that the temperatures are suffi-
ciently low and the m2(T) regime does not appear).
The agreement is perhaps better with the m3(T)
scaling at higher temperatures, but at the lower
ones the scalings are equally satisfactory. Therefore
it is difficult to decide about the spin dimension-
ality at the interfaces; and whether the spins are
confined or not to the growing plane.
10. Conclusions
The relevance of the magnetoelastic energy asso-
ciated to the interfaces and the epitaxial stress
dependence of the volume magnetoelastic stress,
which implies a non-linear behavior of the volume
magnetoelastic energy, has been investigated by
studying the magnetoelastic behavior of two series
of Ho-based superlattices. Measurements of the basal
plane magnetoelastic stress isotherms in HonHo
/Lu15
and in Ho10/YnY
SL’s, between 10 K and above the
Ne′el temperatures, allowed us the determination of
interface magnetoelastic stresses which are of the
order, even higher, than the volume ones. The
modification of the bulk’s magnetoelastic stress due
to the epitaxial strain is also found to be important
in both series. Moreover, the thermal dependence of
the magnetoelastic stress is in agreement with a
single-ion crystal field character for the magneto-
elastic interaction in the studied superlattices. In
Dy/Y and Er/Lu superlattices we have also per-
formed MEL stress measurements, obtaining the
Fig. 19. Thermal dependence of (a) the non-linear MEL
stresses: d1
νεxx
0 (■) and d2
νεxy
0 (❐), and (b) interface MEL
stresses: b1
s (●) and b2
s (❍). All of them were obtained from the
best fit to b1 and b2 . The dotted lines are the fits according to
Î5/2 Callen and Callen law [27]. The dash-dotted line is the
thermal dependence for the 2D-spin model [40].
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 361
different MEL stress contributions for the Dy/Y
samples, but not for the Er/Lu, where incomplete
saturation leads to inconclusive results. The latter
situation also takes place for Ho/Tm SL’s; never-
theless, MEL stresses clearly show up the competi-
tion between the axial and in-plane anisotropies of
Tm and Ho layers. We have measured the magne-
toelastic stress in TbFe2(t)/YFe2 (1000 A° ) (with
t = 300, 600, 1000 and 1300 A° ) epitaxial bilayers.
The sample with t = 300 A° is the only one which
does not show a measurable MEL stress. For the
rest of the bilayers, we have been able to determine
all the parameters accounting for the different MEL
distortions allowed by the cubic symmetry. We
have shown that the epitaxial stress strongly
modifies the value of the MEL tetragonal stress,
which, due to this, increases more than three times
its value as compared with the bulk TbFe2 one for
a given range of t values. An interface contribution
to the MEL stresses has been also determined. As
concerns to the origin of both volume and interface
contributions, the thermal dependence of the MEL
parameters pointed out to a single-ion crystal-elec-
tric-field origin for the volume stress contribution
and a not well-defined 2D behavior for the spins at
the interface.
Acknowledgments
We acknowledge the financial support of Span-
ish CICYT under grants MAT95-1539, MAT97-
1038 and CSIC-CNRS project HF1997-0074.
1. R. W. Erwin, J. J. Rhyne, M. B. Salamon, J. A. Borchers,
S. Shantanu, R. R. Du, J. E. Cunningham, and C. P.
Flynn, Phys. Rev. B35, 6808 (1987); J. A. Borchers, M. B.
Salamon, R. W. Erwin, J. J. Rhyne, R. R. Du, and C. P.
Flynn Phys. Rev. B43, 3123 (1991); R. S. Beach, J. A.
Borchers, A. Matheny, R. W. Erwin, M. B. Salamon, B.
Everitt, K. Pettit, J. J. Rhyne, and C. P. Flynn, Phys.
Rev. Lett. 70, 3502 (1993).
2. C. F. Majkzrak, J. Kwo, M. Hong, Y. Yafet, Doon Gibbs,
C. L. Chien, and J. Bohr, Adv. Phys. 40, 99 (1991).
3. D. A. Jehan, D. F. McMorrow, R. A. Cowley, R. C. C.
Ward, M. R. Wells, N. Hagmann, and K. N. Clausen,
Phys. Rev. B48, 5594 (1993).
4. M. Ciria, J. I. Arnaudas, A. del Moral, G. J. Tomka, C. de
la Fuente, P. A. J. de Groot, M. R. Wells, and R. C. C.
Ward, Phys. Rev. Lett. 75, 1634 (1995).
5. A. del Moral, M. Ciria, J. I. Arnaudas, R. C. C. Ward,
and M. R. Wells, J. Appl. Phys. 81, 5311 (1997).
6. J. I. Arnaudas, A. del Moral, M. Ciria, C. de la Fuente, R.
C. C. Ward, and M. R. Wells, in: Frontiers in Magnetism
of Reduced Dimension Systems, V. G. Baryakhtar, P. E.
Wigen, and N. A. Lesnik (eds.), Kluwer Acad. Publishers,
Dordrecht, pp. 525–552 and References therein (1998).
7. A. del Moral, and E. W. Lee, J. Phys. C: Solid State
Phys. 8, 3881 (1975).
8. P. P. Swaddling, D. F. McMorrow, J. A. Simpson, M. R.
Wells, R. C. C. Ward, and K. N. Clausen, J. Phys.:
Condens. Matter 5, L481 (1993).
9. T. Honda, Y. Hayashi, K. I. Arai, K. Ishiyama, and M.
Yamaguchi, IEEE Trans. Magn. MAG-29, 3126 (1993);
ibid. p. 3129.
10. P. I. Williams, D. G. Lord, and P. J. Grundy, J. Appl.
Phys. 75, 5257 (1994).
11. E. Quandt, J. Appl. Phys. 75, 5653 (1994).
12. T. Honda, K. I. Arai, and M. Yamaguchi, J. Appl. Phys.
76, 6994 (1994).
13. P. J. Grundy, D. G. Lord, and P. I. Williams, J. Appl.
Phys. 76, 7003 (1994).
14. N. H. Duc, K. Mackay, J. Betz, and D. Givord, J. Appl.
Phys. 79, 973 (1996).
15. E. Quandt, A. Ludwig, J. Betz, K. Mackay, and D.
Givord, J. Appl. Phys. 81, 5420 (1997).
16. V. Oderno, C. Dufour, K. Dumesnil, P. H. Mangin, and
G. J. Marchal, J. Crystal Growth 165, 175 (1996).
17. M. Ciria, J. I. Arnaudas, C. Dufour, V. Oderno, K.
Dumesnil, and A. del Moral, J. Appl. Phys. 81, 5699
(1997).
18. C. de la Fuente, J. I. Arnaudas, M. Ciria, A. del Moral, C.
Dufour, A. Mougin, and K. Dumesnil, to be publ. (2000).
19. E. Klokholm, IEEE Trans. Magn. 12, 819 (1976).
20. E. du Tre′molet de Lacheisserie, and J. C. Peuzin, J. Magn.
Magn. Mater. 136, 136 (1994).
21. E. du Tremolet de Lacheisserie, Phys. Rev. B51, 15925
(1995).
22. L. D. Landau and E. M. Lifshitz, Theory of Elasticity, 2nd
ed., Pergamon Press, Oxford (1970).
23. S. Timoshenko and S. Woinowsky-Krieger, Theory of
Plates and Shells, 2nd ed., MacGraw-Hill, New York
(1959).
24. R. Watts, M. R. J. Gibbs, W. J. Karl, and H. Szymczak,
Appl. Phys. Lett. 70, 2607 (1997).
25. V. Iannotti and L. Lanotte, J. Magn. Magn. Mater. 202,
191 (1999) .
26. M. Ciria, J. I. Arnaudas, del A. Moral, M. R. Wells, and
R. C. C. Ward, App. Phys. Lett. 72, 2044 (1998).
27. E. R. Callen and H. B. Callen, Phys. Rev. 139, A455
(1965).
28. C. de la Fuente, et al., to be publ.
29. V. Oderno, C. Dufour, K. Dumesnil, Ph. Mangin, G. Mar-
chal, L. Hennet, and G. Patrat, J. Cryst. Growth 165, 175
(1996).
30. A. Mougin, Ph. D. Thesis, Institut National Polytechnique
de Lorraine (France) (1999).
31. A. Mougin, private commun.
32. M. Ciria , J. I. Arnaudas, A. del Moral, M. R. Wells, and
R. C. C. Ward, App. Phys. Lett. 72, 2044 (1998).
33. M. Ciria, Ph. D. Thesis, Univ. of Zaragoza (1997).
34. J. J. Rhyne, S. Legvold, and E. T. Rodine, Phys. Rev. 154,
266 (1967).
35. J. Jensen and A. R. Mackintosh, Phys. Rev. Lett 64, 2699
(1990).
36. A. del Moral, M. Ciria, J. I. Arnaudas, M. R. Wells, R. C.
C. Ward, and C. de la Fuente, J. Phys. : Condens. Matter
10, L139 (1998).
37. J. H. van der Merwe, J. Appl. Rhys. 34, 123 (1963).
38. L. Ne′el, J. Phys. Radium 15, 225 (1954).
39. C. H. Lee, H. He, F. J. Lamelas, W. Vavra, C. Uher, and
R. Clarke, Phys. Rev. B42, 1066 (1990).
40. E. Callen, J. Appl. Phys. 53, 8139 (1982).
41. A. del Moral, M. Ciria, J. I. Arnaudas, and C. de la
Fuente, Phys Rev. B57, R9471 (1998).
J. I. Arnaudas et al.
362 Fizika Nizkikh Temperatur, 2001, v. 27, No. 3
42. W. C. Koehler, J. W. Cable, E. O. Wollan, and M. K.
Wilkinson, J. Appl. Phys. 33, 1124 (1962); W. C. Koehler,
J. Appl. Phys. 36, 1078 (1965).
43. D. B. Richards and S. Legvold, Phys. Rev. 186, 508
(1969).
44. C. de la Fuente, J. I. Arnaudas, L. Benito, M. Ciria, A. del
Moral, R. C. C. Ward, and M. R. Wells, to be publ.
(2000).
45. B. Coqblin, The Electronic Structure of Rare-Earth Metals
and Alloys: the Magnetic Heavy Rare Earths, Academic,
London (1977); J. Jensen and A. R. Mackintosh, Rare
Earth Magnetism, Oxford Univ. Press, Oxford (1991).
46. D. A. Jehan, D. F. McMorrow, J. A. Simpson, R. A.
Cowley, P. P. Swaddling, and K. N. Clausen, Phys. Rev.
B50, 3085 (1994) and References therein.
47. E. Clark, in: Handbook of the Physics and Chemistry of
Rare Earth, K. A. Gschneidner and L. Eyring (eds.),
North-Holland, Amsterdam, vol. 2 (1982).
48. A. E. Clark, in: Ferromagnetic Materials, E. P. Wohlfarth
(ed.) North-Holland, Amsterdam, vol. 1 (1980).
49. E. W. Lee, Proc. Phys. Soc. 72, 249 (1958).
50. E. W. Lee and J. E. L. Bishop, Proc. Phys. Soc. 89, 661
(1966).
51. A. del Moral and A. Vela, Anales de Fisica (Madrid) 265,
72 (1976).
52. A. Mougin, C. Dufour, K. Dumesnil, N. Maloufi, and Ph.
Mangin, Phys. Rev. B59, 5950 (1999).
53. J. H. Basson and C. A. B. Ball, Phys. Status Solidi A46,
707 (1978).
54. K. Ha, M. Ciria, R. C. O’Handley, P. W. Stephens, and S.
Pagola, Phys. Rev. B60, 13780 (1999).
Magnetoelastic stresses in rare-earth thin films and superlattices
Fizika Nizkikh Temperatur, 2001, v. 27, No. 3 363
|