Você está na página 1de 5

Materials Science and Engineering C 26 (2006) 1151 1155

www.elsevier.com/locate/msec

Annealing effects on the structural and magnetic properties of FeAl silica


nanocomposites prepared by sequential ion implantation
M.A. Tagliente a,*, M. Massaro a, C. de Julian Fernandez b, G. Mattei b, P. Mazzoldi b
a

ENEA, Centro Ricerche Brindisi, UTS MAT-COMP, SS. 7 Appia km 714, I-72100 Brindisi, Italy
b
Dipartimento di Fisica, Universita` di Padova, via Marzolo 8, I-35131, Padova, Italy
Available online 8 November 2005

Abstract
The nanostructural and magnetic properties of FeAl SiO2 granular solids prepared by sequential ion implantation have been investigated as a
function of the annealing atmosphere (either oxidizing or reducing) and implantation order. Nanoscopic particles with a bcc structure were found
in both as-implanted samples. In the sample Al Fe prepared by implanting first the Al ions and later the Fe ions, the lattice parameter indicates the
presence of practically pure iron nanoparticles. On the other hand, in sample Fe Al with the implantation order inverted, the lattice parameter is
consistent with the presence of an iron rich iron silicon alloy. The magnetic data confirm the presence of the pure Fe and the Fe Si alloy in the
as-implanted samples and the absence of FeAl intermetallic compounds. The annealing in Ar/H2 promotes the growth of the clusters and increases
the Si content in the particles in both samples. In Fe Al sample, this induces a disorder order phase transition from the disordered Fe Si solid
solution to the Fe3Si phase and the coprecipitation of the ordered FeSi phase. The magnetic moment increases after the annealing in Ar/H2 due to
the incorporation of the iron atoms dispersed in the matrix and to the higher crystalline order. The annealing in air is responsible essentially of the
growth of the Fe Si clusters in both samples. On the other hand, in sample Al Fe the oxygen interacts with the pure iron clusters by promoting
the Fe2O3 formation.
D 2005 Elsevier B.V. All rights reserved.
Keywords: Magnetic particles; Ion implantation; Glancing incidence X-ray diffraction; Silica composites

1. Introduction
Nanocomposite materials formed by metallic nanoparticles
embedded in dielectric matrices have attracted much attention
due to their peculiar optical, magnetic and catalytic properties.
Among the preparation procedures, single or sequential ion
implantation is currently used as a suitable technique for
synthesizing monoelemental and bimetallic nanoclusters embedded in glass, respectively [1 4]. In particular, in sequential
ion implantation the energy and dose of the two elements can
be tailored so as to maximize the overlap between the
implanted species and to control their local relative concentration. Moreover, the as-implanted composite can be treated by a
proper combination of treatments such as ion irradiation, heat
treatments in controlled atmosphere or pulsed laser irradiation,
in order to promote the alloying or the separation of the
implanted species [5 7].
* Corresponding author. Tel.: +39 831 507399; fax: +39 831 507656.
E-mail address: antonella.tagliente@brindisi.enea.it (M.A. Tagliente).
0928-4931/$ - see front matter D 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.msec.2005.09.076

In this paper, we present a study on the effect of thermal


annealing of Fe50Al50 silica nanocomposites prepared by
sequential ion implantation of iron (Fe) and aluminium (Al)
ions in silica in order to form bimetallic FeAl nanoparticles in
silica. Bulk Fe Al alloys are magnetic up to an Al
concentration of 35 at.%, after that the alloy becomes nonmagnetic [8,9]. The magnetic order of this alloy can be
ferromagnetic, antiferromagnetic or spin glass depending on
the composition and composition order [10 12]. In a previous
work, we have investigated the nanostructure and magnetic
properties of silica double implanted with Fe and Al ions [13].
It was evidenced the formation of complex core-shell clusters
with a structure and composition that in some cases were not
clearly defined. The aim of this work is to investigate the
evolution of the structural and magnetic properties of the Fe
Al silica nanocomposite with annealing in several atmospheres
(either oxidizing or reducing) in order to clarify the structure
and composition. The microstructural characterization was
performed by Glancing incidence X-ray diffraction (GIXRD).
The magnetic properties of the composites were investigated

M.A. Tagliente et al. / Materials Science and Engineering C 26 (2006) 1151 1155

Fe(211)

a)

Fe 2O 3(116)

Fe(200)

Al-Fe samples

Fe 2O 3(024)

Fused silica (Herasil type II by Heraeus) slides were


implanted by using a 200 kV high current DANFYSIK 1090
implanter at the ENEA Ion Implantation Laboratory (Brindisi, Italy). Al+ and Fe+ ions were sequentially implanted at
room temperature at current densities lower than 2 AA/cm2
in order to avoid possible heating during the implantations.
Ion beam energies (150 keV for iron and 50 keV for
aluminium, respectively) were chosen to get the same
projected range of the implanted species, namely about 70
nm. The implantation dose was 1.5  1017 ions/cm2 for both
ions in order to have the average local composition Fe50Al50
in the matrix. Moreover, the implantations were performed
by considering both implantations orders. In the following,
Fe Al samples indicate the composites obtained by implanting first the Fe ions and later the Al ions whereas in the
Al Fe samples the implantation order was reversed. Also, a
sample implanted only with Fe+ ions was prepared with the
same implantation dose and energies (only-Fe sample). The
as-implanted slides were then heat treated in a conventional
furnace at 900 -C and 1 h in air or in an Ar (90%) H2
(10%) gas mixture.
GIXRD experiments were carried out by using a Philips
MPD PW1880 X-ray diffractometer in Parallel Beam geometry
which employs CuKa radiation (k CuKa = 0.154186 nm) and
operated at 40 kV, 40 mA. The X-ray beam incident on the
sample is collimated by a 1/30- divergence slit. The reflected
beam is collimated by a 0.1 mm receiving slit followed by a
parallel plate collimator and a flat graphite monochromator
which reflects the diffracted beam in a proportional counter.
The measurements were performed by fixing the incident beam
at a 1- incidence angle and by rotating the detector in the 2h
range between 5- and 105- with a 2h step of 0.02-, where 2h is
the diffraction angle. In order to get a more quantitative
information, a detailed peak profile analysis was accomplished
to each diffraction peak. In particular, the diffraction line
profiles were fitted by a pseudo-Voigt (pV) function (in
practice, a sum of two pseudo-Voigt functions in order to take
into account the a1 a2 doublet) and it was assumed that the
background in proximity of the reflection was parabolic. The
diffraction peaks were fitted by using a least-squares refinement procedure in order to determine parameters defining the
position, breadth and shape of the Bragg reflections [14]. Then
the volume averaged coherent domain size D was determined
by considering that all the inclusions have the same lattice
parameter from the total integral breadth of the simulated
diffraction profiles by using the Scherrer formula. The
magnetic characterization was carried out using a Cryogenic
S600 superconducting quantum interference magnetometer.
The hysteresis loops were measured at the temperature of 3 K
and applying a maximum magnetic field of 6 T in the plane of
the samples.

Fig. 1a compares the experimental GIXRD spectrum of the


as-implanted Al Fe sample with the spectra of the same
sample after the annealing at 900 -C in Ar/H2 or in air. The
curves are shifted along the vertical axis for clarity reasons.
The pattern of the as-implanted sample exhibits a very broad
diffraction peak near the (110) reflection of the a-Fe phase
(ICDD card # 87-0722 SG = Im3m, a = 0.28608(4) nm).
Diffraction line profile analysis performed on this peak,
assuming a bcc phase as found by TEM experiments in [13],
gives a lattice parameter a = 0.285(1) nm which is comparable
with the bulk a-Fe lattice parameter. The corresponding size of
the coherent domains was 1.5(4) nm. The pattern of the sample
annealed in Ar/H2 exhibits three diffraction peaks which are
near the (110), (200), and (211) reflections of the a-Fe phase.
In comparison with the as-implanted case, the peaks are sharper
and more intense suggesting that the average clusters size was
increased. However, diffraction line profile analysis demonstrates that each peak is the convolution of two reflections from
two bcc phases with lattice parameters slightly different. In
particular, we have found two phases with the following lattice

Fe 2O 3(104)
Fe 2O 3(110)
Fe(110)

2. Experimental procedure

3. Experimental results

X-Ray Diffracted Intensity (a.u.)

by measuring the magnetization hysteresis loops at 3 K. The


results obtained, peculiarly related to the atmosphere of
annealing and implantation order, are discussed.

Air
Ar/H2
As
20

40

60

80

100

Diffraction angle 2 (degrees)


X-Ray Diffracted Intensity (a.u.)

1152

b)
Fe(110)

Ar/H2
40

42

44

46

48

50

Diffraction angle 2 (degrees)


Fig. 1. (a) Experimental GIXRD spectra of the as-implanted (as) and annealed
in Ar/H2, air Al Fe samples. (b) Experimental GIXRD spectrum (dotted line)
and line profile fit (solid line) of the sample annealed in Ar/H2. The lower
curves indicate the separate contribution of the two bcc phases to the whole
simulation.

M.A. Tagliente et al. / Materials Science and Engineering C 26 (2006) 1151 1155

20

40

80

60

100

X-Ray Diffracted Intensity (a.u.)

Diffraction angle 2 (degrees)

b)
Fe 3Si(220)

FeSi(210)

FeSi(211)

Ar/H 2

40

42

44

46

48

50

52

54

Diffraction angle 2 (degrees)


Fig. 2. (a) Experimental GIXRD spectra of the as-implanted (as) and annealed
in Ar/H2, air Fe Al samples. (b) Experimental GIXRD spectrum (dotted line)
and line profile fit (solid line) of the sample annealed in Ar/H2. The lower
curves indicate the separate contribution of the Fe3Si (dashdotted line) and FeSi
(solid line) phases to the whole simulation.

parameters: a 1 = 0.2866(2) nm which corresponds to the a-Fe


phase and a 2 = 0.2836(2) nm which is smaller than the bulk
iron lattice parameter ( 0.87%). The fit of the diffraction
profile of the principal peak is shown in Fig. 1b. The calculated
size of the coherent domains is 10.6(5) nm for the iron phase
and 11.6(5) nm for the other phase. The spectrum of the sample
annealed in air exhibits the diffraction peaks in the vicinity of
two crystalline phases: the a-Fe phase and the Fe2O3 phase
(hematite, ICDD card # 87-1166, SG = R3c). Again, here the
peaks in proximity of the a-Fe phase reflections are sharper
and more intense with respect to the as-implanted sample and
each peak consists in a couple of peaks with the following
lattice parameters: a 1 = 0.2865(1) nm which corresponds to the
bulk a-Fe lattice parameter and a 2 = 0.2841(2) nm which is
smaller than the a-Fe phase lattice constant ( 0.69%). The
corresponding size of the coherent domains was D 1 = 13.4(5)
nm and D 2 = 8.0(1) nm.
Fig. 2a compares the experimental GIXRD patterns of the
as-implanted Fe Al sample with the patterns of the same
sample after the annealing at 900 -C in Ar/H2 or in air. In the
as-implanted sample, a broad diffraction peak near the (110)
reflection of the bcc a-Fe phase is visible. Diffraction line
profile analysis performed on this peak, assuming a bcc phase

Only-Fe samples

Fe(211)

As

Fe(200)

Fe3Si FeS i Fe Si
3
Ar/H 2

Fe(110)

Air

as found by TEM experiments in [13], gave a lattice parameter


a = 0.284(1) nm which is smaller ( 0.88%) than the bulk a-Fe
lattice parameter. The pattern of the Fe Al sample annealed in
Ar/H2 exhibits several peaks which are consistent with the
presence of two crystalline phases. In fact, line profile analysis
indicates that the peak at 2h ; 45- is really a couple of peaks at
the angles 2h = 45.456(2)- and 45.329(2)-. In this way, all the
peaks occurring in the pattern can be indexed with the
reflections of the phases Fe3Si (ICDD card # 45-1207 SG
Fm3m a = 0.56533 nm) and FeSi (ICDD card # 76-1748
SG = P213, a = 0.4467 nm). Fig. 2b shows the fit of the two
principal peaks and the separated contributions of the two
phases to the whole simulation. It is evident that the
contribution of the Fe 3 Si phase is predominant. The
corresponding size of the coherent domains was D 1 = 7.5(1)
nm and D 2 = 15.1(3) nm. The spectrum of the Fe Al sample
annealed in air exhibits diffraction peaks near the (110), (200)
and (211) reflections of the a-Fe phase. In comparison with the
as-implanted case, the peaks are sharper and more intense by
indicating that the average coherent domains size was
increased. According to the quantitative analysis of the peaks
profile, the corresponding lattice parameter is a = 0.2835(1) nm
which is smaller than the bulk iron parameter ( 0.90%).
Fig. 3 compares the experimental GIXRD patterns of the asimplanted only-Fe sample with the patterns of the sample
annealed in Ar/H2 or in air. The spectrum of the as-implanted
sample exhibits the a-Fe (110) reflection from which a lattice
parameter a = 0.2864(3) nm was calculated, by indicating the
presence of pure iron clusters. After the annealing in Ar/H2, the
peaks of the a-Fe phase are more intense and sharper. In the
pattern of the only-Fe sample annealed in air, the diffraction
peaks of the Fe phase disappeared and the Fe2O3 phase
reflections appeared, demonstrating the total oxidation of the
metallic clusters.
The magnetic behavior of the samples was investigated by
measuring the hysteresis loops at 3 K. The diamagnetic
contribution of the silica was independently measured and
subtracted from the measured magnetization and the magne-

Fe2O3(104)
Fe2O3(110)

FeSi Fe3Si
FeSi

a)

X-Ray Diffracted Intensity (cps)

Fe(211)

Fe(200)

X-Ray Diffracted Intensity (a.u.)

Fe(110)

Fe-Al samples

1153

Air
Ar/H2
As

20

40

60

80

2 angle (degrees)
Fig. 3. Experimental GIXRD spectra of the as-implanted (as) and annealed in
Ar/H2, air only-Fe samples.

M.A. Tagliente et al. / Materials Science and Engineering C 26 (2006) 1151 1155

Al-Fe samples
2x10

-5

Ar/H2
As

Normalized magnetization

1x10

Air

-5

-1x10

-2x10

-5

-5

-2.0

-0.10

-1.5

-1.0

-0.5

0.0

-0.05

0.5

0.00

1.0

0.05

1.5

0.10

2.0

Applied field (T)


Fig. 4. Hysteresis loops measured at the temperature of 3 K of the Al Fe
samples. Inset: detail of the low field region.

tization was normalized to the sample mass. The comparison


among the resulting normalized hysteresis loops of the asimplanted and thermal treated Al Fe and Fe Al samples is
shown in Figs. 4 and 5, respectively. The insets in the
respective figures show the low field region of the magnetization curves. It is evident that the main differences among the
different samples are found in the values of the magnetization
at high magnetic fields, while the values of the coercive fields
are very similar. By considering that the Fe amount does not
change with the thermal treatments, the magnetic moments per
Fe atom at 6 T are 2, 2.5 and 1.2 AB (T10%) for the asimplanted, annealed in Ar/H2 and air Al Fe sample, respectively. The corresponding values for the sample Fe Al are 1,
1.5 and 2 AB (T 10%). The maximum coercive field is 28 mT
(as-implanted Fe Al sample), and the minimum is 11 mT (asimplanted Al Fe sample and Fe Al sample annealed in air).
In the Al Fe samples the thermal treatment produces an
increase of the coercive fields with respect to that of the asimplanted sample, while it occurs just the opposite in the Fe
Al sample.
4. Discussion
According to the experimental results, the as-implanted Al
Fe sample contains mostly nanoscopic Fe crystalline aggregates dispersed in the silica matrix. On the other hand, in the
Fe Al sample, the phase is different from pure Fe because the
corresponding lattice parameter is contracted with respect to
the a-Fe phase. By considering that the bulk bcc intermetallic
Fe1 x Alx alloys have a larger lattice parameter than the a-Fe
phase, the formation of FeAl alloys nanoparticles in both
samples was ruled out. Moreover, the observed pattern does not
match any Fe, Al oxides or silicate phases. On the other hand,
the iron rich iron silicon alloys (Fe1 x Six ) have a bcc structure
and a lattice parameter which starts from the a-Fe value and

decreases linearly with the Si molar fraction x [15]. The


structure of the Fe1 x Six alloys corresponds to a disordered
solid solution, in which the Si atoms are in substitutional
random positions in the bcc Fe lattice. Therefore, the calculated
lattice parameter in Fe Al sample is compatible with the
presence of bcc iron rich Fe1 x Six alloys clusters. From the
calculated lattice parameter, the corresponding Si molar
fraction x in the alloy is 10 at.% [15]. The magnetic data
confirm the presence of the a-Fe and FeSi alloy clusters in the
as-implanted samples and the absence of FeAl intermetallic
compounds. The a-Fe and Fex Si1 x compounds are ferromagnetics at 3 K. The Fe atom magnetic moments of the a-Fe and
Fe1 x Six alloy with a Si content smaller than 10% [15] are very
similar and equal to 2.2 AB. The smaller is the Fe content in the
iron silicon alloy, the smaller is the magnetic moment of the
Fe atoms [16]. This explains the value of the magnetic moment
found in the as-implanted Al Fe sample (2.0 AB) and in the
sample Fe Al (1.0 AB), respectively.
The only-Fe sample does not exhibit the Fe1 x Six alloy
clusters, but only Fe clusters. Therefore, the formation of the
alloyed Fe Si colloids in the sample Fe Al is activated in our
systems by the Al+ ion implantation. For Fe and Al, the
respective Gibbs energy of oxide formation are negative and
very high [17], so the formation of oxides is expected.
Therefore, the implantation with Al+ ions forms preferentially
compounds of the Al atoms with the oxygen atoms of the
matrix which come from the broken Si O bonds, while Si
atoms remain unbonded. As a consequence in sample Fe Al,
the first implantation with Fe+ ions produces metallic iron
clusters. After the implantation with Al+ ions, the Al atoms
bond to the oxygen atoms of the matrix while the Si atoms are
incorporated in iron clusters. A possible mechanism for the
Fe1 x Six alloy formation is a radiation enhanced diffusion of
silicon in small iron clusters during the second implantation.
Fe-Al samples
Ai r
1x10

Ar/H 2

-5

As

Normalized magnetization

1154

-1x10

-5

-0.10

-2.0

-1.5

-1.0

-0.5

0.0

-0.05

0.5

0.00

1.0

0.0 5

1.5

0.10

2.0

Applied field (T)


Fig. 5. Hysteresis loops measured at the temperature of 3 K of the Fe Al
samples. Insets: detail of the low field region.

M.A. Tagliente et al. / Materials Science and Engineering C 26 (2006) 1151 1155

In all the three implanted samples, the subsequent annealing in


Ar/H2 at 900 -C is able to drive the Si and Fe atoms which
were dispersed in the matrix in separated or oxidized forms
toward the already formed clusters, increasing their size and/or
their Si content. In Al Fe sample, this mechanism favors the
presence of two phases: the a-Fe phase and the Fe1 x Six alloy
with a Si molar fraction x = 10%. On the other hand, in the
sample Fe Al, the annealing in Ar/H2 increases the Si content
in the alloy up to a threshold of 15% which is the limit of the
solubility of Si and a-Fe and corresponds to the metastable
Fe3Si phase, beyond which there is the coexistence with
another ordered phase which is the FeSi phase [15]. Therefore,
the increase in the Si content induces a structural disorder
order phase change of the solid solution to the metastable Fe3Si
phase and the coprecipitation of the FeSi compound.
As concerns the magnetic properties, in both Al Fe and
Fe Al samples, the magnetic moment increases after the
annealing in Ar/H2. The point is to understand why the
magnetization of the as-implanted samples, which are composed of grains with high magnetic moment (practically of Fe
or of a Fe-rich alloy) are smaller than those of the thermal
treated samples. One reason, as was shown by Perez et al. [18]
for single Fe implants, can be the incorporation in the already
formed grains of the Fe atoms dispersed in the matrix. As the
receiving grains are ferromagnetic, this incorporation increases
the magnetic moment. Another mechanism is related with the
structure of the nanoparticles. Indeed, as in the iron silicides
the Fe magnetic moment depends on the chemical and
crystalline order [15,19,20], typically decreases as disorder
increases, upon annealing an increasing of the magnetic
moment should be expected in our samples according with
the formation of the ordered Fe3Si phase. Indeed, the Fe3Si
compound is ferromagnetic at 3 K and has a magnetic moment
of 1.8 AB [21] while FeSi is paramagnetic [22]. As we
measure a magnetic moment of 1.5 T 0.1 AB, this result
indicates that this sample is mainly formed by the Fe3Si
phase in agreement with the GIXRD data.
The annealing in air promotes the growth of the Fe1 x Six
clusters present in the as-implanted Fe Al and Al Fe samples
and the oxidation of the metallic iron clusters occurring in the
as-implanted Al Fe and only-Fe samples. The Fe Al samples
annealed in air have larger magnetic moment than that in Ar/
H2 due to the larger magnetic moment associated to the
Fe0.9Si0.1 phase with respect to that of the Fe3Si phase. In the
case of the Al Fe the opposite behavior occurs, the sample
annealed in reducing gas has a larger magnetic moment than
the sample annealed in air due to the presence in this last
sample of hematite which is antiferromagnetic at the temperature of 3 K.
The coercive field is typically proportional to the magnetic
anisotropy and inversely proportional to the magnetization
saturation. The magnetic anisotropy of the silicide is smaller
than the Fe [15,23] as the Si content is larger, being the Fe3Si
phase as a very soft phase [24]. Considering the different
values of anisotropy and magnetization, we do not found that
the ratio between these two properties reproduces the differences in the coercive field between all the samples. This can be

1155

explained considering that this description is of first order, and


a more precise description of the nanostructure is required, in
particular, the induced anisotropy could be mainly due to
magneto-elastic contribution arising at the surface of the
Fe1 x Six grains.
Acknowledgments
The authors would like to thank Daniela Carbone for the
valuable technical support in thermal annealings experiments.
This work was financially supported by ENEA in the
framework of the PROTEMA project of Intesa ENEAMIUR under project 4335/04 and by the Italian MIUR
(National Projects).
References
[1] M.A. Tagliente, G. Mattei, L. Tapfer, M. Vittori Antisari, P. Mazzoldi,
Phys. Rev., B 70 (2004) 75418.
[2] F. Gonella, P. Mazzoldi, Handbook of Nanostructured Materials and
Nanotechnology, Vol. 4: Optical Properties, Academic Press, San Diego,
2000, Chapter 2: Metal Nanocluster Composite Glasses, p. 81.
[3] A. Meldrum, L.A. Boatner, C.W. White, Nucl. Instrum. Methods, B 178
(2001) 7.
[4] G. Mattei, Nucl. Instrum. Methods, B 191 (2002) 323.
[5] G. Mattei, G. De Marchi, C. Maurizio, P. Mazzoldi, C. Sada, V. Bello,
Phys. Rev. Lett. 90 (8) (2003) 85502.
[6] F. Gonella, G. Mattei, P. Mazzoldi, C. Sada, G. Battaglin, E. Cattaruzza,
Appl. Phys. Lett. 75 (1) (1999) 55.
[7] A. Miotello, G. De Marchi, G. Mattei, P. Mazzoldi, C. Sada, Phys. Rev., B
63 (2001) 75409.
[8] H.P.J. Wijn (Ed.), Physical Properties and Other Data, vol. III, Pt. 19b,
Springer, Landolt-Bornstein, Berlin, 1987, pp. 295 356.
[9] J.S. Kouvel, J.H. Westbrook, R.L. Fleischer (Eds.), Intermetallic Compounds: Principles, vol. 1, Wiley, New York, 1994, pp. 935 940.
[10] G.R. Caskey, J.M. Franz, D.J. Sellmyer, J. Phys. Chem. Solids 34 (1973)
1179.
[11] A. Hernando, X. Amils, J. Nogues, S. Surinach, M.D. Baro, M.K. Ibarra,
Phys. Rev., B 58 (1998) R11864.
[12] B.V. Reddy, S.C. Deevi, F.A. Reuse, S.N. Khanna, Phys. Rev., B 64
(2001) 132408.
[13] C. de Julian Fernandez, M.A. Tagliente, G. Mattei, C. Sada, V. Bello, C.
Maurizio, G. Battaglin, C. Sangregorio, D. Gatteschi, L. Tapfer, P.
Mazzoldi, Nucl. Instrum. Methods, B 216 (2004) 245.
[14] R.A. Young, The Rietveld Method, Series IUCr, Oxford University Press,
1993.
[15] R.M. Bozorth, Ferromagnetism, D. Van Nostrand Company, New York,
1951, p. 67, and references therein.
[16] C. Chien, in: G. Hadjipanayis, G.A. Prinz (Eds.), Science and Technology
of Nanostructured Magnetic Materials, Plenum Press, 1991, p. 477.
[17] O. Knacke, O. Kubaschewski, K. Hesselmann (Eds.), Thermo-Chemical
Properties of Inorganic Substances, Springer, Berlin, 1991.
[18] A. Perez, M. Treilleux, T. Capra, D.L. Griscom, J. Mater. Res. 2 (1987)
910.
[19] T. Zhou, J. Zhang, J. Xu, Z. Yu, G. Gu, D. Wang, H. Huang, Y. Du, J.
Wang, Y. Jiang, J. Magn. Magn. Mater. 164 (1996) 219.
[20] D. Ruiz, T. Ros-Yanez, L. Vandenbossche, L. Dupre, R.E. Vandenberghe,
Y. Houbaert, J. Magn. Magn. Mater. 290 (2005) 1423.
[21] H. Okumura, D.E. Laughlin, M.E. McHenry, J. Magn. Magn. Mater. 267
(2003) 347.
[22] M.B. Hunt, M.A. Chernikov, E. Felder, H.A. Ott, Z. Fisk, P. Canfield,
Phys. Rev., B 50 (1994) 14933.
[23] E. du Tremolet de Lacheisserie, Magnetisme, Vol.II, Presses Unive.
Grenoble, Grenoble, 1999, p. 96 and references therein.
[24] G. Herzer, J. Magn. Magn. Mater. 112 (1992) 258.

Você também pode gostar