You are on page 1of 9

3 December 2001

Physics Letters A 291 (2001) 175183


www.elsevier.com/locate/pla

Shape memory effect due to magnetic field-induced thermoelastic


martensitic transformation in polycrystalline NiMnFeGa alloy
A.A. Cherechukin a , I.E. Dikshtein a , D.I. Ermakov a , A.V. Glebov a , V.V. Koledov a ,
D.A. Kosolapov a , V.G. Shavrov a, , A.A. Tulaikova a , E.P. Krasnoperov b , T. Takagi c
a Institute of Radio Engineering and Electronics, Russian Academy of Sciences, Moscow, 101999, Russia
b Russian Research Center Kurchatov Institute, Moscow, 123182, Russia
c Institute of Fluid Science, Tohoku University, Sendai, Japan

Received 10 September 2001; accepted 26 September 2001


Communicated by V.M. Agranovich

Abstract
In Heusler-type alloy Ni2+xy Mn1x Fey Ga, partial substitution of Mn for Ni causes the temperatures of structural
(martensitic) TM and magnetic TC (Curie point) phase transitions to converge. Close to the crossover of TM and TC , we have
observed the strong strains (l/ l 24%) induced by the external magnetic field. This effect could be classed with colossal
magnetostriction. The system exhibits the magnetic field-controlled one-way shape memory effect at fixed temperature as a
result of the magnetic field-induced martensite to austenite structural phase transition. 2001 Elsevier Science B.V. All rights
reserved.
PACS: 64.70.Kb; 75.30.Kz; 75.50.Cc; 81.30.Bx; 75.80.+q
Keywords: Martensitic phase transition; Curie point; Shape memory effect; Colossal magnetostriction

1. Introduction
Heusler-type alloy Ni2+x Mn1x Ga exhibits a structural (martensitic) phase transition (SPT) from hightemperature cubic phase (austenite) to low-temperature tetragonal phase (martensite) in a ferromagnetic
state [1]. The combination of magnetic ordering and
shape memory makes this alloy promising in the
search for the possibility of controlling the shape of
a sample by varying temperature, pressure, and external magnetic field [213]. This can be of a great im-

* Corresponding author.

E-mail address: shavrov@mail.cplire.ru (V.G. Shavrov).

portance in design of new magneticfieldcontrolled


smart materials.
In the martensitic phase the contraction of about 6%
of the cubic cell along one [100] axis and the extension
of about 2% along the other two are observed. The cubic symmetry admits the existence of three different
tetragonal phases called variants depending on the direction of the contraction axis. On the one hand, the
first-order SPT presupposes the coexistence of hightemperature austenite and low-temperature martensite.
This phase coexistence that is called the intermediate state occurs in the specific temperature range close
to the SPT due to the elastic strains that accompany
the nucleation of the martensitic inclusions in austenite phase. On the other hand, in the martensitic phase

0375-9601/01/$ see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 5 - 9 6 0 1 ( 0 1 ) 0 0 6 8 8 - 0

176

A.A. Cherechukin et al. / Physics Letters A 291 (2001) 175183

far from the SPT the tetragonal distortions of the cubic lattice occur with the same probability along each
of the crystallographically equivalent axes of the [100]
type in the absence of the external forces. The coherent and partially coherent conjugation of the phases
on the twin boundaries between martensitic variants
results in the quasiperiodic structure of the martensitic
domains [14].
There are two different physical mechanisms that
can be responsible for the magnetic field-induced
strains and magnetic shape memory effect [9], namely,
the rearrangement of martensite variants far from the
SPT or austenitemartensite domains formation close
to the reversible field-induced SPT. Recent experiments [513] concerned the former one in Ni2 MnGa
single crystals. These experiments demonstrated colossal magnetic field-induced deformation of about
6%, which is considerably greater than that of the
giant magnetostrictive material of (Tb, Dy)Fe2 . The
reason for this effect is the growth of the preferable
twin variant volume under the applied magnetic field
H < 10 kOe through the reversible twin boundary motion, provided that each variant has a strong magnetic
anisotropy in which the easy axis is aligned with the
c-axis.
Contrary to the previous studies our goal here is to
investigate the colossal magnetostriction and the magnetic shape memory effect as a result of the austenite to
martensite SPT driven by the external magnetic field.
However, the realization of such a magnetically controlled reversible SPT in Ni2 MnGa is hindered by the
weak field dependence of the SPT temperature. Recent
studies [15,16] showed that the partial substitution of
Mn for Ni results in the increase in the SPT temperature TM and the decrease in the Curie point temperature TC up to their coincidence at x = xC 0.19 and
room temperature. Close to the crossover of TM and
TC , the magnetic field influence on SPT is enhanced,
and the SPT temperature is shifted to higher temperature range proportionally 1 K per 10 kOe [1619]. In a
magnetic field of about 100 kOe, the shift of the hysteresis loop exceeds its width. This makes it possible
to observe the magnetically controlled reversible SPT
at the fixed temperature [18]. However, due to the high
brittleness, loaded Ni2+x Mn1x Ga samples subjected
to heat cycling are of limited use.
In this Letter we report measurements of the colossal magnetostriction and the one-way magnetic shape

memory effect as a result of the martensite to austenite


SPT in polycrystalline Ni2+xy Mn1x Fey Ga close to
crossover of TM and TC . In experiment the colossal
(practically plastic) strains are accounted for by the
change in phase volumes. Low doping of Fe allows
to enhance the alloy plasticity without sacrificing its
magnetic and thermoelastic properties.

2. Samples and procedure


Polycrystalline ingots of Ni2+xy Mn1x Fey Ga alloys with diameter 1 cm, length 10 cm and weight 40 g
were prepared by arc melting in an Ar atmosphere on
a cold hearth. For homogenization we annealed the ingot materials at 1100 K for 9 days. Then the ingots
were quenched in ice water and cut into 0.5 mm thick
plates.
Optical measurements were performed to determine
the field dependence of the SPT temperature with a
real resolution over the sample. To this end, the samples were polished at T > TMA in austenite phase
with TMA , the martensite stability loss temperature.
Then a sample in a two-loop thermostat with transparent windows was placed in the field of a Bitter magnet 0150 kOe. For T < TAM (with TAM , the austenite stability loss temperature), the mechanical stresses
resulted from the formation of martensiteaustenite
domain walls as well as martensitic (structural) twin
boundaries form a relief on the sample surface. The
sample surface was observed under a microscope with
oblique illumination. For the H T phase diagram construction the domain structure evolution was recorded
with a video camera.
In experiment a mechanical load P was applied to
the free end of the polycrystalline plate supported in
a cantilever manner. Then the sample was trained by
repeated heat cycling. Optical system included an inclined mirror and half a lens that made it possible to
measure the bending of the loaded sample and to observe the austenitemartensite domain nucleation and
annihilation simultaneously. Maximal strains e (the
extension of the upper plate surface and the contraction of the lower one) were determined through the
maximal plate bending. In the course of heat cycling
the video film was made that enabled us to construct
the plot e(T , P ).

A.A. Cherechukin et al. / Physics Letters A 291 (2001) 175183

Fig. 1. The experimental and theoretical field dependences of


TAM (H ) and TMA (H ) for x = 0.19, y = 0.04. The theoretical
curves are shown as thin solid lines.

177

(a)

3. Phase diagram
For x = 0.19, y = 0.04, the field dependences of
TAM (H ) and TMA (H ) are depicted in Fig. 1. The functions TAM (H ) and TMA (H ) are approximately linear,
with dTAM /dH dTMA /dH 0.078 K/kOe. The
width of the heat hysteresis loop T = TMA (H )
TAM (H ) equal to 6 2 K. It reduces slightly as the
field increases. The measurement accuracy is determined by the inhomogeneity of the polycrystalline
sample. In a magnetic field of about 100 kOe, the shift
of the hysteresis loop exceeds its width T . Hence,
for H 100 kOe the transition from high-temperature
(austenite) to low-temperature (martensite) phase will
occur at a higher temperature TAM (H ) than the temperature of the reverse transition TMA (0) in the absence of the magnetic field. Then, switching the magnetic field on and off at the fixed temperature within
the interval TMA (0) < T < TAM (H ) makes it possible
to induce a reversible SPT. In the system of interest
the temperature interval T = TAM (H ) TMA (0) that
will be called the magnetic control interval for the SPT
was about of 12 K.

4. Thermoelastic properties
Fig. 2(a) presents the eT curves of the sample
loaded by the weight P = 3 N in the austenite phase
for some successive cycles of cooling and heating. At
cooling, the loaded sample is strongly deformed in the
martensite phase and recovers its initial shape in the

(b)
Fig. 2. eT curves of the sample loaded by the weight 3 N (a) and
the unloaded sample (b). The sample was trained for N successive
cycles of cooling and heating.

austenite phase at heating doing the work against the


external force. It means that the system in study exhibits the one-way shape memory effect. Simple ideas
may help to understand this effect. As the mechanical
load is applied, the SPT temperatures TAM and TMA
increase. The sample is deformed inhomogeneously
under the load, with maximal extension and contraction strains being respectively at the upper and lower
plate surfaces close to the clamping line. Hence, at
cooling run, the austenitemartensite domains are initially formed near this line. The contraction of the
lower surface leads to the domain formation with the
contraction axis c preferentially perpendicular to the
clamping line. The upper surface extension is responsible for the domain nucleation with the axis c parallel to the clamping line. Such domain distribution

178

A.A. Cherechukin et al. / Physics Letters A 291 (2001) 175183

is energetically favorable. In the intermediate phase


(f )
(s)
(P ), the plate is being considerTAM (P ) < T < TAM
(f )
(s)
ably bent, with TAM (P ) and TAM (P ) are the start and
finish temperatures of the austenite to martensite SPT,
respectively. In the heating run, the reverse SPT occurs
at T = TMA (P ), and the initial plate shape recovers.
The direct observation of the austenitemartensite
domain nucleation and annihilation supports the conclusion that the plate deformations result from the
SPT, with the austenitemartensite domains being initially formed in the most strained regions at cooling
and annihilated in the least strained regions at heating.
In the course of heat cycling (for 1, 5, 12 and 18 heat
cycles), the sample is deformed progressively, and the
heat hysteresis loop is slightly narrowed down and is
shifted to higher temperature range. At the point with
the most curvature, the restored deformation e was
24%.
The eT curves of an unloaded sample trained for
5, 12 and 18 heat cycles without an external magnetic
field are shown in Fig. 2(b). In the cooling run, the
unloaded sample is deformed in the martensite phase
and recovers its original shape in the austenite phase in
the heating run. This means that the system of interest
demonstrates the two-way shape memory effect with
transformation strains about 0.2%.
The results above are consistent with the observation in Ni2 MnGa [2,20].
Fig. 3. The temporal dependences of H , e, and T during the experiment.

5. Colossal magnetostriction and shape memory


effect
The one-way shape memory effect in a magnetic
field will be further considered. The samples suited
for this purpose were trained for 1, 5, 12 and 18 heat
cycles. Fig. 1 illustrates the experiment on the shape
memory effect controlled by a magnetic field. The
evolution of sample state is shown by the ABCDE
trajectory on T H phase diagram. An initial state of
the trained sample at room temperature is marked by
the point A. Then the magnetic field was switched on
(point B), and the sample was heated in the magnetic
field up to the temperature T0 within the magnetic
control interval T (point C). When switching the field
off at T = T0 , the reverse SPT in the austenite phase

occurs (point D), and the initial plate shape (point E)


recovers (the plate becomes straight).
Fig. 3 presents the temporal dependence of the
field H , the strain e, and the temperature T during the
experiment. Reduction of the sample temperature due
to absorption of the first-order SPT latent heat can be
a reason for temporal delay of the shape restoration.
Photos of the sample made in the course of this experiment are shown in Fig. 4. At H = 100 kOe and T =
315 K, the sample (Fig. 4(a)) is bent with the maximal deformation e = 3%. When the field is decreasing the sample becomes straight (e = 0). Straightening
process lasts for 15 s, after the field is switched off.
The resulting photo of the straight sample is shown in
Fig. 4(b), right.

A.A. Cherechukin et al. / Physics Letters A 291 (2001) 175183

179

(a)

(b)
Fig. 4. The photos of the sample in the course of the experiment: the field is switched on (a); the field is switched off (b).

It follows from Figs. 3 and 4 that NiMnFeGa


exhibits the strong strains (l/ l 24%) induced by
the external magnetic field that may be classified as
colossal magnetostriction.

6. Discussion
The coherent and partially coherent conjugation of
the phases on the austenitemartensite twin bound-

180

A.A. Cherechukin et al. / Physics Letters A 291 (2001) 175183

aries brings about the elastic strains e in addition to the


strains inherent in the SPT. The additional energy e
associated with these strains and the twin boundaries
surface energy s must be taken into account for the
calculation of the phase energy. The effect of a magnetic field on the SPT temperatures TAM and TMA can
be estimated using the ClapeironClausius thermodynamic equation [15,16,21]. At T = TAM the condition
of metastable phase equilibrium can be written as
M (TAM , H ) A (TAM , H )
= (TAM ) (MM VM MA VA )H
+ eAM + sAM = 0,

(1)

where A,M are the thermodynamic potentials of austenitic and martensitic phases, respectively; MA,M
and VA,M are the magnetizations and the volumes of
austenitic and martensitic phases for T = TM , H = 0
and e = 0; (T ) = Q(T TM )/TM , Q and TM are
the difference of the thermodynamic potentials, the
hidden heat of SPT and the SPT temperature for H = 0
and e = 0. For direct and reverse SPT, the elastic and
surface energies e and s are distinguished. For the
reverse SPT T = TMA the conditions of metastable
phase equilibrium are given by
A (TMA , H ) M (TMA , H )
= (TMA ) (MA VA MM VM )H
+ eMA + sMA = 0.

(2)

The difference in TMA and TAM is responsible for the


heat hysteresis of the SPT T , with


T = TM eAM + sAM + eMA + sMA Q.
(3)
From Eqs. (1) and (2) one can determine the field
dependence of TAM and TMA :


TAM,MA = TM 1 + (MM VM MA VA )H

 
eAM,MA + sAM,MA Q .
(4)
To construct the H T phase diagram of Ni2+xy
Mn1x Fey Ga ferromagnet, we need to evaluate the
magnitudes of TM , MM and MA . To do this the phenomenological model based on Landau expansion of
the free energy [15,16,18] can be used. We consider
a cubic ferromagnet of space group Oh experiencing magnetic and structural phase transitions at cooling. The order parameters responsible for the SPT are

the combinations of the deformation tensor components eik . The Curie point is described by the magnetization M. After minimization with respect to the
components of the deformation tensor which are not
responsible for the SPT, the thermodynamic potential
for the system of interest can be written as

1 
= 0 + a e22 + e32
2

 1 
2
1
+ be3 e32 3e22 + c e22 + e32
3 
4



1 
1 
+ B2 e2 m21 m22 + e3 3m23 m2
6
2


2
1
1  2
+ m1 + m22 + m23 + m21 + m22 + m23
2
4


+ K m21 m22 + m22 m23 + m23 m21 H3 M3 .
(5)
Here ei are the linear combinations the deformation

tensor components eik : e2 = (exx eyy )/ 2, e3 =

(2ezz exx eyy )/ 6; a, b, and c are linear combinations of the second-, third-, and fourth-order elastic moduli, respectively,
a = c11 c12 , b = (c111

c112 + c123 )/ 216, c = (c1111 + c1112 3c1122


8c1123)/48, m = M/M0 , M0 is the saturation magnetization, = 0 (T TC0) and are exchange constants,
0 = /TC0 , B is the relativistic magnetostriction constants, K1 is the first cubic anisotropy constant. The
elastic modulus a tends to zero, while approaching the
SPT temperature, and close to the SPT. It is given by
a = a0 (T TM0 ). The third-order terms presented in
the free energy give rise to a first-order SPT.
For the system in hand the inequalities b > 0, c > 0,
and K < 0 are valid. Then in the absence of a magnetic field, five different phases: the paramagnetic and
ferromagnetic cubic phases, the paramagnetic tetragonal phase, and the ferromagnetic angular and collinear
tetragonal phases can be realized [15,16,18]. Due to
paraprocess an arbitrary small magnetic field suppresses the second-order magnetic phase transition.
In addition, a magnetic field leads to the tetragonal
distortion of a cubic lattice owing to magnetoelastic
coupling of magnetization and elastic deformations.
Therefore in strong magnetic fields H in excess of
the magnetic anisotropy field HA = 2K/M0 , two ferromagnetic phases with nonzero distortion can be realized, because the symmetry is broken already by the
applied field at any temperature. However, phases with
the strong and weak tetragonal distortions can be

A.A. Cherechukin et al. / Physics Letters A 291 (2001) 175183

181

treated as the martensitic and austenitic phases, respectively.


The minimization of the free energy with respect to
the variables e2 , e3 , m1 , m2 , and m3 makes it possible to determine all feasible structural and magnetic
phases. As a result for H > HA the following states
of the ferromagnet and their stability conditions are
found:
m1 = m2 = 0,
(6)
e2 = 0,


2

m3 + m3 + 2Kme = M0 H,


e3 a + be3 + ce32 + 3m = 0,
(7)



2m33 + M0 H a + 2be3 + 3ce32 8B22 m33  0, (8)

where K me = 2/3 B2 e3 and 3m = 2/3 B2 m23 are


the effective magnetic anisotropy and magnetic
stress due to magnetostriction interaction (see below), respectively.
To search the composition, temperature and field
dependence of m3 and e3 and to construct the phase
diagram of the magnet in the T H xy coordinates,
we assume (as in Refs. [15,16]) that both TM0 and TC
linearly depend on concentration:

(a)

(b)

TM0 = TM0 + x x y y,
TC0 = TC0 x x + y y,

(9)

with i and i are the proportionality coefficients.


From Eqs. (7), one can determine the concentration
dependence of TM , eA,M and mA,M = MA,M /M0 .
Since many parameters of the problem are unknown
at the present time, a comparison of the results of the
theoretical simulations with the experimental data in
Fig. 1 will only be qualitative. The best agreement
between them is reached for the following values of
the parameters: TC0 = 375 K, TM0 = 200 K, a0 =
b2 /(100c), 0 = /TC0 , 2.5 108 ergs/cm3 ,
c/b 11.1, b 3.8 1010 ergs/cm3 , B2 3
107 ergs/cm3 , M0 = 300 Oe, x = 175 K, y = 400 K,
x = 700 K, and y = 1100 K.
Insertion of magnitudes TM and mA,M = MA,M /M0
into Eq. (4) yields the field dependence of TAM and
TMA . For x = 0.19, y = 0.04, TM = 314.25 K, eM =
0.0561, eA = 0.0042, mM = 0.374,
mA = 0.351,
2 V e 2 V )/2 2/3 B (e V
Q = TM [a0 (eM
M
2 M M
A A
eA VA )/TC0 ], VM = VA = V , and T 6 K, the
phase diagram of the system under study is shown as

Fig. 5. The theoretical temperature dependences of the reduced


magnetization m = m3 (a) and the strain e = e3 (b).

thin solid lines in Fig. 1. It can be seen that the experimentally obtained linear field dependence of TAM and
TMA (Fig. 1) is in a good agreement with the calculated ones.
The temperature and field dependences of m3 and
e3 are shown in Figs. 5 and 6, respectively. These
figures illustrate the general tendency of m3 and |e3 |
to increase with an increase in H and a decrease in T .
The field dependence of m3 and |e3 | exhibits the sharp
jump at TAM and TMA . The reason for these effects is
the twofold role of the magnetoelastic energy
i Mk e/M
2 (T = 0)
Eme = BM
M M ,
e K
m

me

(10)

i Mk /M 2 (T = 0), the effective magwith m = BM


= B e/M
netic stress, K
2 (T = 0), the effective
me
magnetic anisotropy. On the one hand, for TC
TM the magnetization produced by the external mag-

182

A.A. Cherechukin et al. / Physics Letters A 291 (2001) 175183

fields H in excess of the magnetic anisotropy field HA .


There are the fundamental differences between the results of current and previous [313] studies. The latter were concerned with the shape memory effect and
the colossal magnetic field-induced strains due to the
twin-boundary motion driven by the Zeeman energy
difference between the adjacent twin variants. These
effects can be realized only in single crystals in the
martensitic phase in H below saturation (H < HA ).
(a)

Acknowledgements
The authors would like to thank Prof. A.N. Vasilev
for useful discussions. This work was supported by the
Russian Foundation for the Basic Research, projects
99-02-18247, 01-02-06053, 01-02-06056.

References

(b)
Fig. 6. The theoretical field dependences of the reduced magnetization m = m3 (a) and the strain e = e3 (b).

netic field (paraprocess) induces the effective magnetic stress m that in turn shifts the SPT to higher
temperatures. On the other hand, large strains associated with the SPT produce the jump of the effec
tive magnetic anisotropy K me that leads to the sharp
jump in M3 . The theory justifies the existence of the
reversible SPT driven by the external magnetic field
for some interval of the temperatures and magnetic
fields as well as the colossal magnetic field-induced
strains in the vicinity of the SPT (Fig. 6).
In summary, we report the results of measurements of the magnetically controlled shape memory
effect and the colossal magnetic field-induced strains
in NiMnFeGa alloy close to the crossover of TM
and TC . These effects are associated entirely with the
reversible magnetic field-induced austenite to martensite transformation and can be observed in both single and polycrystalline materials in strong magnetic

[1] P.J. Webster, K.R.A. Ziebeck, S.L. Town, M.S. Peak, Philos.
Mag. B 49 (1984) 295.
[2] V.A. Chernenko, E. Cesari, V.V. Kokorin, I.N. Vitenko, Scr.
Metall. Mater. 33 (1995) 1239.
[3] R.D. James, M. Wuttig, Proc. SPIE 2715 (1996) 420;
R.D. James, M. Wuttig, Philos. Mag. A 77 (1998) 1273.
[4] K. Ullakko, J. Mater. Eng. Perform. 5 (1996) 405.
[5] K. Ullakko, J.K. Huang, C. Kantner, R.C. OHandley, V.V.
Kokorin, Appl. Phys. Lett. 69 (1996) 1966.
[6] K. Ullakko, J.K. Huang, V.V. Kokorin, R.C. OHandley, Scr.
Mater. 36 (1997) 1133.
[7] R.C. OHandley, J. Appl. Phys. 83 (1998) 3263.
[8] S.J. Murray, M. Farinelli, C. Kantner, J.K. Huang, S.M. Allen,
R.C. OHandley, J. Appl. Phys. 83 (1998) 7297.
[9] R. Tickle, R.D. James, J. Magn. Magn. Mater. 195 (1999) 627.
[10] G.H. Wu, C.H. Yu, L.Q. Meng, J.L. Chen, F.M. Yang, S.R.
Qi, W.S. Zhan, Z. Wang, Y.F. Zheng, L.C. Zhao, Appl. Phys.
Lett. 75 (1999) 2990.
[11] R.C. OHandley, S.J. Murray, M. Marioni, H. Nembach, S.M.
Allen, J. Appl. Phys. 87 (2000) 4712.
[12] W.H. Wang, G.H. Wu, J.L. Chen, C.H. Yu, Z. Wang, Y.F.
Zheng, L.C. Zhao, W.S. Zhan, J. Phys. Condens. Matter 12
(2000) 6287.
[13] Y. Ma, S. Awaji, K. Watanabe, M. Matsumoto, N. Kobayashi,
Appl. Phys. Lett. 76 (2000) 37.
[14] A.L. Roytburd, in: H. Ehrenreich, F. Seitz, D. Turnbell (Eds.),
Solid State Physics: Advances in Research and Applications,
Vol. 33, Academic Press, New York, 1978, p. 317.
[15] A.N. Vasilev, A.D. Bozhko, V.V. Khovailo, I.E. Dikshtein,
V.G. Shavrov, V.D. Buchelnikov, M. Matsumoto, S. Suzuki,
T. Takagi, J. Tani, Phys. Rev. B 59 (1999) 1113.

A.A. Cherechukin et al. / Physics Letters A 291 (2001) 175183


[16] A.D. Bozhko, A.N. Vasilev, V.V. Khovailo, I.E. Dikshtein,
V.V. Koledov, S.M. Seletski, A.A. Tulaikova, A.A. Cherechukin, V.G. Shavrov, V.D. Buchelnikov, JETP 88 (1999) 954.
[17] I. Dikshtein, V. Koledov, V. Shavrov, A. Tulaikova, A. Cherechukin, V. Buchelnikov, V. Khovailo, M. Matsumoto, T.
Takagi, J. Tani, IEEE Trans. Magn. 35 (1999) 3811.
[18] I.E. Dikshtein, D.I. Ermakov, V.V. Koledov, L.V. Koledov,
T. Takagi, A.A. Tulaikova, A.A. Cherechukin, V.G. Shavrov,
JETP Lett. 72 (2000) 373.

183

[19] K. Inoue, K. Enami, Y. Yamaguchi, K. Ohoyama, Y. Morrii, Y.


Matsuoka, K. Inoue, J. Phys. Soc. Jpn. 69 (2000) 3485.
[20] V.V. Martynov, V.V. Kokorin, J. Phys. III 2 (1992) 739.
[21] M.A. Krivoglaz, V.D. Sadovski, Fiz. Met. Metalloved. 18
(1964) 502.

You might also like