You are on page 1of 6

Journal of Magnetism and Magnetic Materials 384 (2015) 235240

Contents lists available at ScienceDirect

Journal of Magnetism and Magnetic Materials


journal homepage: www.elsevier.com/locate/jmmm

Sub-Kelvin magnetic order in Sm3Ga5O12 single crystal


A.Z. Sharma a,1, H.J. Silverstein a,1, A.M. Hallas c, G.M. Luke c,d, C.R. Wiebe a,b,c,n
a

Department of Chemistry, University of Manitoba, Winnipeg, MB, Canada R3T 2N2


Department of Chemistry, University of Winnipeg, 515 Portage Ave, Winnipeg, MB, Canada R3B 2E9
Department of Physics and Astronomy, McMaster University, 1280 Main St. W, Hamilton, ON, Canada L8S 4M1
d
Canadian Institute for Advanced Research, Toronto, Ontario, Canada M5G 1Z7
b
c

art ic l e i nf o

a b s t r a c t

Article history:
Received 29 December 2014
Accepted 3 February 2015
Available online 4 February 2015

The garnets remain one of the most studied classes of materials due to their technological applications.
While many studies have focused on the low temperature properties of rare-earth containing garnets
such as Gd3Ga5O12, there are relatively few studies concerning the low-temperature magnetic properties
of garnets with lighter rare-earth ions such as Sm3 , with the majority of those studies having been done
on powders only. Sm3Ga5O12 was prepared via standard solid-state reaction and a single crystal was
grown using the oating zone technique. Magnetic susceptibility and low temperature heat capacity
measurements show that Sm3Ga5O12 orders below 1 K and behaves like an effective S 1/2 system with
negligible anisotropy.
& 2015 Published by Elsevier B.V.

Keywords:
Garnet
Hyperkagome
Magnetic
Low temperature
Transition
Heat capacity
Susceptibility
Long range order
Single crystal
Sm3Ga5O12

1. Introduction
Geometric frustration can have an enormous inuence on the
magnetic ground state of a material. This phenomenon occurs
when the spins fail to simultaneously minimize their energy solely
due to the lattice geometry [1]. For example, Ising spins residing
on the vertices of an equilateral triangle cannot minimize their
energy through antiferromagnetic ordering; if two spins were to
order antiferromagnetically, the third is unable to minimize its
energy as it cannot simultaneously lie antiparallel to both [1]. As a
result, the system must make energetic compromises that result in
intriguing new magnetic phases beyond simple antiferromagnetic
order. In two dimensions, physicists often tend to study systems
with stacked triangular or kagom lattices resulting in materials
that fail to order in the low temperature limit [25]. In three dimensions the lattices become more complex, but this is not necessarily the case for the associated magnetic ground state. This is
exemplied by the pyrochlore systems that can exhibit, for example, the spin ice state that is well understood despite the
complexity of the crystal lattice [68]. Spin ice phases have been
predicted (although not yet experimentally reported) for other
n
Corresponding author at: Department of Chemistry, University of Manitoba,
Winnipeg, MB, Canada R3T 2N2.
E-mail address: ch.wiebe@uwinnipeg.ca (C.R. Wiebe).
1
These authors contributed equally to this work.

http://dx.doi.org/10.1016/j.jmmm.2015.02.003
0304-8853/& 2015 Published by Elsevier B.V.

three-dimensional lattices such as those blanketed under the


hyperkagom term [911], wherein corner-linked trimers are
bent or twisted in the third dimension.
The garnets are a class of materials with a hyperkagom lattice
[9]. Y3Ga5O12 (YAG) is perhaps the most widely known garnet for
its use as a host material for solid state lighting. The primary
reason why the garnets are so widely studied is their structural
tolerance; it is possible to substitute and dope the structure using
most of the elements on the periodic table, much like the perovskites. A whole family of magnetic garnets can be synthesized
with d-block ions such as Fe3 [12] that show technological promise. This study, however, is motivated by the relative lack of
fundamental experimental research into the magnetic ground
states of garnets that do not contain d-block ions. This is especially
intriguing considering the wealth of theory available for the hyperkagom lattice [911,1315]. Nd3Ga5O12 [16], Gd3Ga5O12 [17
21], Tb3Ga5O12 [22,23], Dy3Ga5O12 [24,25], Ho3Ga5O12 [26] and
Yb3Ga5O12 [27] each have features indicative of a Nel transition
below 1 K and appear to exhibit a wide variety of magnetic ground
states. Here, the low temperature magnetic properties of
Sm3Ga5O12 [28,29] are studied.

2. Methods
Sm3Ga5O12 and Y3Ga5O12 were synthesized by combining

236

A.Z. Sharma et al. / Journal of Magnetism and Magnetic Materials 384 (2015) 235240

stoichiometric amounts of Sm2O3 (or Y2O3) and Ga2O3. The mixture was compressed in the form of pellets and heated to 1200 C
for 24 h and 1400 C for 48 h in air with additional grinding and
pelleting after each 24 h cycle. The nal product was a dark yellow-brown colour. The powders were then pressed into rods of
about 40 mm in diameter by 6 cm in length. A single crystal of
Sm3Ga5O12 was grown using an IR Image Furnace (Quantum Design) in air equipped with two 650 W halogen lamps. The hot zone
was translated up the rods at a speed of 5 mm/h with rod rotation
speeds of 15 rpm to maintain sample homogeneity. The crystal
was then annealed in O2(g) at 900 C. The translucent yellow
crystal was aligned using a multiwire Laue diffractometer. A diffraction pattern of both the powder and crushed crystal was obtained using a Siemens D5000 X-ray diffractometer with a Cu(s)
anode. All renements were performed using the FullProf Suite
[30]. DC magnetic susceptibility and heat capacity measurements
were made on a 35.1 mg piece of single crystal aligned with the
eld parallel and perpendicular to the 111 direction using the
Dynacool Physical Property Measurement System (Quantum Design) and a Magnetic Property Measurement System equipped
with a 3He insert (Quantum Design). Susceptibility measurements
were performed between 300 K and 0.5 K under a 0.1 T eld. For
all other elds, measurements were only performed to 2 K. For all
elds, measurements were made once while cooling in a eld and
once cooling under zero-eld. Select measurements were made
while warming the system to check for the presence of thermal
hysteresis although none was detected. The heat capacity was
measured in 0 and 9 T elds down to 0.35 K using the He-3 heat
capacity option (Quantum Design). Samples were placed on a
sample stage and xed using Apiezon N-grease. The heat capacity
of the stage and grease were subtracted from the nal
measurement.

3. Results and discussion


Fig. 1 shows the renement of the garnet structure to our
powder X-ray data. The atomic coordinates, bond distances, and
bond valence sums from the renement are contained in Table 1. A
small Sm4Ga2O9 impurity was found in the powders, which is
consistent with studies of the reaction mechanism in other garnets
[29]. The impurity peaks become less intense as the system is
annealed for longer periods of time, indicating that the main
garnet phase is thermodynamically stable as opposed to kinetically metastable. However, as one moves to incorporate larger

Table 1
Atomic coordinates, bond valence sums and exchange map bonds and angles found
in the renement of Sm3Ga5O12. This material crystallizes in the space group Ia3 d
(S.G. #230) with lattice constant a 12.4296(6) . The Debye Waller factors could
not be rened due to peak shape issues. Estimates of the error in the bond distances were multiplied by 10 to reect the uncertainties in the atomic coordinates
of O. 2 2.00, Rp 12.5, Rwp 17.1.
Atomic coordinates
Atom

Wyckoff

Sm
Ga1
Ga2
O

24c
16a
24d
96h

1/8
0
3/8
0.036(1)

0
0
0
0.060(1)

1/4
0
1/4
0.652(1)

Distance ()

Multiplicity

Total sum

2.46(1)
2.43(1)
2.08(1)
1.81(1)
2.46(1)
2.43(1)
2.08(1)
1.81(1)

4
4
6
4
1
1
1
1

3.06

Bond valence sums


Atom
Bond Type
Sm

SmO

Ga1
Ga2
O

GaO
GaO
OSm
OSm
OGa
OGa

2.33
3.22
1.96

rare-earth ions into the garnet structure, it becomes harder to


prepare the garnet, resulting in longer reaction times. This trend is
reected in the literature, with the bulk of the garnet studies focusing on Gd, Tb, Dy, Ho, and Yb-containing garnets rather than Pr,
Nd, and Sm. From the renements, it is estimated that the impurity phase in this sample accounts for less than ve percent of
the overall phase composition. Like other rare-earth garnets such
as Gd3Ga5O12 [17] and Ho3Ga5O12 [26], it is possible to grow single
crystals of Sm3Ga5O12 [31,32]. No impurities were detected in the
X-ray diffraction pattern of the crushed crystals.
The garnets crystallize in the space group Ia3 d. For Sm3Ga5O12,
there are three inequivalent cationic sites accounting for 8-fold
coordinated SmO decahedra, 6-fold coordinated GaO octahedra,
and 4-fold coordinated GaO tetrahedra. Each GaO octahedron is
corner linked to six GaO tetrahedra, while sharing six of its
twelve edges with SmO decahedra (not shown) in order to form a
very convoluted network. Bond valence sum analysis indicates
that GaO octahedra are overbonded while GaO tetrahedra are
underbonded. While this may indeed be the case due to a higher
degree of covalency for these polyhedra, the bond valence sums

Fig. 1. (a) Renement of Sm3Ga5O12 to the x-ray diffraction data (powder sample); (b) Depiction of the two enantiomeric Sm sublattices (blue and orange) is shown in the
representative unit cell (both are generated by symmetry from the 24c site). (c) A view down the [111] axis of the hyperkagom structure. (For interpretation of the
references to colour in this gure legend, the reader is referred to the web version of this article.)

A.Z. Sharma et al. / Journal of Magnetism and Magnetic Materials 384 (2015) 235240

237

Fig. 2. (a) DC-susceptibility of Sm3Ga5O12 taken using powder and an aligned single crystal with a 0.1 T eld lying parallel to the [111]; Inset: Low temperature DCsusceptibility showing a transition at 0.9 K on crystals with the eld aligned parallel and perpendicular to the [111]; (b) Inverse susceptibility (black square symbols) of the
single crystal with a 0.1 T eld aligned along the [111]. The CurieWeiss law was used to t the data (red lines) at various regions within the susceptibility; Inset: The
susceptibility between 5 and 50 K can be t to the doublet-doublet model (yellow line) [50] described within the text; (c) Magnetization measurements versus eld show no
saturation of the moment; (d) The data taken at 10 K cannot be modelled using the Brillouin function for J 1/2 or J 5/2. (For interpretation of the references to colour in
this gure legend, the reader is referred to the web version of this article.)

more likely reect the lack of precision in pinpointing the atomic


coordinates of oxygen since only X-ray diffraction was used here.
An alternative way to depict the garnet structure is to connect
the Sm ions (Fig. 1b, c). In this case, one observes two identical
sublattices of corner linked and twisted trimers, commonly referred to as the hyperkagom lattice. However, there is considerable confusion over the blanket term hyperkagom; the confusion stems from the degrees of freedom in which each of the trimers can be bent in space. Such hyperkagom structures can be
found in a variety of systems outside of the garnet structure, including but by no means limited to -Mn [36], the trillium lattice
[37], and the kagom staircase lattice [38]. Recent studies have
focused on another hyperkagom structure derived from the
popular pyrochlore lattice; it is possible to extend the phase diagram of spin ices, rst proposed by dan Hertog and Gingras
[11,39], to the hyperkagom pyrochlore lattice, which is exhibited
by such materials as Na4Ir3O8 [40].
The magnetic susceptibility and magnetization as a function of
eld are showcased in Fig. 2 and are in general agreement with
previous results on single crystals [41]. These measurements were
performed on a single crystal aligned with the eld both parallel
and perpendicular to the 111 direction (Fig. 2a). The inverse
susceptibility is plotted in Fig. 2b. This system does not appear to
display any signicant anisotropy in the magnetic interactions as
there is no statistical difference between the parallel and

perpendicular curves. Susceptibility measurements were made in


elds of 0.1, 1, 3, 5, 7, and 9 T but no obvious deviations from 0.1 T
occurred as the eld was increased. A peak in the susceptibility
appears at 1 K followed by a precipitous drop indicating antiferromagnetic order (Fig. 2a inset). Usually an application of the
CurieWeiss law as a means to t the linear susceptibility is
standard practice. One can then obtain the effective magnetic
moment and a measure of the mean-eld approximated interactions within the sample. For frustrated systems, the Weiss temperature, CW, is used to index the magnitude of the frustration [1]
by comparing it to the temperature at which magnetic order ensues, To: the larger the ratio between |CW| and To, the more frustrated the material.
However, tting the CurieWeiss law to the susceptibility of
materials where Sm3 is the only magnetic species is quite challenging. Sm3 exhibits strong Van-Vleck paramagnetism that is
responsible for mixing excited state multiplets with the ground
state upon the application of a eld [42]. There have been attempts
to salvage the CurieWeiss law for Sm-containing materials
[42,43]. There are even some experimental reports of Sm-containing metals such as SmCo2 adhering to the CurieWeiss law
[44] although this is quite rare. For the garnet, the application of
the CurieWeiss law was made on a variety of temperature regions
(Fig. 2b). In general, one should try and t the susceptibility at high
temperatures where the mean-eld approximation is valid.

238

A.Z. Sharma et al. / Journal of Magnetism and Magnetic Materials 384 (2015) 235240

Between 50 and 250 K, thermal excitations populate the crystal


elds resulting in an articially large Weiss temperature for most
rare earth containing materials. Even at lower temperatures, between 20 and 50 K, rhombohedral crystal elds are still able to
split the 6H5/2 multiplet. It is not until an upper limit of 4 K is used
that adherence to the CurieWeiss law is best observed yielding
physical values. In the low temperature regime, the extracted effective moment is meff 0.5(1) mB with CW  2.5(1) K. Both values
are comparable to other Sm-containing materials such as
Sm2Ti2O7 and Sm2Zr2O7, both of which do not show magnetic
order as the temperature is lowered to 50 mK [45,46] and 670 mK
respectively [45]. While the effective magnetic moment is lower
but comparable to the free ion value of 0.83 mB, the Weiss temperature is harder to estimate, although it is on the same order of
magnitude as other garnets like Ho3Ga5O12, despite no crystaleld effects being taken into account in the t presented here [26].
Yet extreme caution must be taken in interpreting these values
due to the close proximity of the magnetic transition. The comparison between meff and the free ion value, and the agreement
between CW in Sm3Ga5O12 with other rare earth garnets are likely
little more than a coincidence.
An attempt has been made to include higher order multiplets
between the 6H5/2 states to the 6H15/2 states using spectroscopic
data on Y3Ga5O12:Sm3 and isostructural Y3Al5O12:Sm3 (the
experimental and calculated crystal eld schemes for both are

reproduced in Fig. 3a) [47,48]. This yields a peak in the susceptibility at about 400 K, well above the maximum range of temperatures measured here (not shown). The calculated susceptibility corresponds nicely with data from other frustrated Sm
systems such as the Sm3Ga2.63Al2.37SiO14 langasite where a broad
peak in the susceptibility was observed near the same temperature [49]. It also corresponds to previous theoretical calculations
derived from the hydrogen atom using a screening constant [48].
In the langasite, weakly interacting Sm3 ions occupy a distorted
kagom sublattice and no magnetic ordering was detected down
to 20 mK. But the calculated multiplet susceptibility is only good in
the high-temperature limit; it cannot account for the shape of the
curve between 2 and 300 K for Sm3Ga5O12. Sm3 ions have much
stronger interactions in this system than in the langasite, evidenced by the onset of magnetic order. Instead, it is believed that
the curve can be approximated by crystal eld splitting of the
Kramers doublets separated by a small energy gap, inducing an
effective J 1/2 (doublet) ground state. A modied t based on the
low-temperature doublet-doublet crystal-eld ground state of
Tb2Ti2O7 found through second-order perturbation theory [50] is
shown in Fig. 2b inset:

3k B 1 + e ( / T )
1

+ c
=
x
(g B )2 (a/T ) + b

Fig. 3. (a) The crystal eld scheme of Sm3 doped Y3Ga5O12 (YGG:Sm3 ) and Y3Al5O12 (YAG:Sm3 ). Relative calculated and experimental values from Grnberg et al. are
marked with n while absolute experimental values for YAG:Sm3 are marked with ^ [47]; (b) The heat capacity of Sm3Ga5O12 (black) and nonmagnetic lattice standard
Y3Ga5O12 (red) scaled by 1.7. The inset displays a closer view of the lambda anomaly observed at 0.75 K with a t to the equation described in the text (red line); (c) The
magnetic entropy found from a lattice-subtracted portion of the heat-capacity denoted Cres and plotted as Cres/T vs. T (inset). Integrating this peak yields the magnetic
entropy release. Integrating all the way to 300 K yields erroneous results due to the presence of low-lying crystal elds that are absent in Y3Ga5O12. The entropy was linearly
extracted to 0 J/K at 0 K; (d) Logarithmic t of the heat capacity to the equation described in the text. (For interpretation of the references to colour in this gure legend, the
reader is referred to the web version of this article.)

A.Z. Sharma et al. / Journal of Magnetism and Magnetic Materials 384 (2015) 235240

where kB is the Boltzmann constant, g is the Land splitting factor,


mB is the Bohr magneton, is the gap between the doublets, a is
the exchange between states of the ground state doublet, b incorporates other terms including exchange between the higher
energy states and c is the diamagnetic contribution. Here, the
susceptibility is modelled based on two doublet states that are
separated by an energy gap on the order of 13(3) K. The t is able
to reproduce the susceptibility between the temperature interval
of 5 and 50 K with a 0.061(9), b0.02(1) K  1, and c 45(2) molSm/emu. The t is quite good and yields values that reect a
physically reasonable trend; approximating a as the Curie constant, an effective moment of 0.69 mB is extracted while the temperature interval of b is small compared to kBT, indicating easily
populated higher states. Turning to the magnetization data shown
in Fig. 2c at 2, 5, 10, and 100 K, no eld-dependent hysteresis was
observed in any measurement. Fig. 2d shows the magnetization at
10 K compared against the Brillouin function for J 1/2. The
experimental magnetization is only a fraction of that predicted
by the Brillouin function for either J 1/2 or J 5/2, which assumes
that the moments are non-interacting. One can observe deviations
away from this behaviour as the interactions between Sm3 spins
becoming stronger when TN is approached from above. Fits at
higher temperatures failed due to the mixing of low-lying excited
states.
Turning to heat capacity measurements (Fig. 3bd) one observes a lambda anomaly occurring at 0.75 K which was previously
observed in Sm3Ga5O12 as well as in other garnets [51]. The peak
also agrees with the magnetization measurements (Fig. 3b, inset).
A nonmagnetic lattice standard, Y3Ga5O12, was used to approximate the Debye contribution. Cres is the resultant heat capacity
taken by subtracting off the lattice, although at such low temperatures, the Debye contribution is approximately zero in accordance with its T3 dependence anyway. The lattice standard
matches quite well with Sm3Ga5O12 at temperatures below 20 K,
even without the use of a scale factor (not shown), although one
was employed anyway to match the heat capacities at higher
temperatures (Fig. 3b, outset). At temperatures above 20 K, the
heat capacities of the two systems begin to diverge, which is
probably indicative of low-lying crystal-eld effects rather than
acoustic modes or a lattice mismatch. Integrating out the lambda
anomaly observed in Cres/T (Fig. 3c, inset), it is possible to extract
the magnetic entropy release (Fig. 3c, outset), which is estimated
to be 88% of that expected for J 1/2, saturating at 7 K. The heat
capacity appears to tend towards 0 at 0 K, indicating no evidence
of Pauling residual entropy. Assuming Heisenberg spins with net
antiferromagnetic exchange, one might expect a magnon con
tribution to the heat capacity of CM  T where 3. A simple
power law t yielded 4.5, but this value does not make sense
physically. For gapped states, the low temperature heat capacity
can be t to

Cres = Ae / T T
where is energy of the gap [52]. The t to the low temperature
heat capacity is displayed in Fig. 3b, inset. The t converged with
0.85(5) K xing to 3, implying that the ordered state is an
antiferromagnet, as expected from E  2.5 K (Fig. 3d). The
presence of critical behaviour was assumed to be negligible in
the t since TN occurs at 0.75 K.
Of course this raises the question: what is the nature of this
phase? We can draw an analogy to two other rare-earth garnets:
Gd3Ga5O12 [1721] and Ho3Ga5O12 [26]. Like Sm3Ga5O12, both
systems have CW and magnetic phase transitions that occur at
similar temperatures. Both systems had phase transitions that
indicated partly disordered magnetic phases that coexistent with
magnetic Bragg peaks and diffuse scattering. However, Gd3 is an

239

S-ion (J S 7/2) making the interactions perhaps more isotropic


than Sm3 . The partially ordered phase in Gd3Ga5O12 was attributed to a spin-liquid phase surrounding ordered domains onset by
lattice defects such as impurities or site mixing [53]. The longrange order exhibited by this material is incommensurate with the
nuclear phase, although this is expected of this particular system
according to one study [21]. Ho3Ga5O12 (J 8) also exhibits partial
order with the coexistence of diffuse scattering surrounding
commensurate magnetic Bragg peaks [26]. Clearly, one of the best
tools used to discover the true nature of these magnetic systems is
neutron scattering. Unfortunately, Sm has a very large neutron
absorption cross section although a study using isotopically pure
Sm would be welcomed in the future. As of now, the nature of the
magnetic ordering remains elusive.

4. Conclusion
Powder Sm3Ga5O12 was prepared using a standard solid state
reaction and a single crystal was grown using the oating zone
technique. Magnetic susceptibility, magnetization, and heat capacity measurements were made and indicate a magnetic phase
transition occurring at approximately 0.9 K, in-line with other
rare-earth containing garnets. Although the true nature of the
magnetic ground state is still unknown, it is predicted that like the
Gd3Ga5O12 and Ho3Ga5O12, Sm3Ga5O12 will also have a partially
ordered magnetic ground state. Neutron scattering studies using a
sample containing isotopically pure Sm are in order.

Acknowledgements
We would like to thank NSERC (National Sciences and
Engineering Research Council of Canada (Application No. 3862682010)), ACS Petroleum Fund (The American Chemical Society
Petroleum Research Fund (PRF No. 51374-UR10)), and the CFI
(Canada Foundation for Innovation (Project No. 26849)) for funding this project. C.R.W. thanks the Canada Research Chair (Tier II)
program while A.M.H. and H.J.S. both graciously thank
the Vanier Canada Graduate Scholarship Program (NSERC) for
additional funding. H.J.S. also acknowledges support from the
Manitoba Graduate Scholarship and the University of Manitoba.
All authors would like to thank P.M. Sarte for helping with the
operation of the Physical Property Measurement System.

References
[1] For a review on geometric frustration, please refer to Greedan J E 2001 J.
Mater. Chem. 11 3753.
[2] Y. Ishiguro, et al., Nat. Commun. 4 (2013) 2022.
[3] T.-H. Han, et al., Nature 492 (2012) 406410.
[4] L. Clark, et al., Phys. Rev. Lett. 110 (2013) 207208.
[5] H.D. Zhou, et al., Phys. Rev. Lett. 106 (2011) 147204.
[6] T. Fennell, et al., Phys. Rev. B 72 (2005) 224411.
[7] H.D. Zhou, et al., Nat. Commun. 2 (2011) 478.
[8] A.M. Hallas, et al., Phys. Rev. B 86 (2012) 134431.
[9] T. Yoshioka, A. Koga, N. Kawakami, J. Phys. Soc. Jpn. 73 (2004) 18051811.
[10] T.E. Redpath, J.M. Hoopkinson, Phys. Rev. B 82 (2010) 014410.
[11] Redpath T. E. et al. (2013) Dipolar Hyperkagome Spin Ice. Oral presentation
given at the American Physical Society March Meeting (Baltimore, MD) Abstract ID: T15.00006.
[12] C. Suchomski, et al., Chem. Mater. 25 (2013) 25272537.
[13] C.L. Henley, J. Phys.: Conf. Ser. 145 (2009) 012022.
[14] E.J. Bergholtz, A.M. Luchli, R. Moessner, Phys. Rev. Lett. 105 (2010) 237202.
[15] R.R.P. Singh, J. Oitmaa, Phys. Rev. B 85 (2012) 104406.
[16] W. Wang, J. Appl. Phys. 102 (16) (2007) 063905.
[17] J.A. Quillam, Phys. Rev. B 87 (17) (2013) 174421.
[18] T. Yavorskii, M. Enjalran, M.J.P. Gingras, Phys. Rev. Lett. 97 (18) (2006) 267203.
[19] P.P. Deen, et al., Phys. Rev. B 82 (19) (2010) 174408.
[20] P. Schiffer, I. Daruka, Phys. Rev. B 56 (20) (1997) 13712.

240

A.Z. Sharma et al. / Journal of Magnetism and Magnetic Materials 384 (2015) 235240

[21] P. Schiffer, A.P. Ramirez, D.A. Huse, A.J. Valentino, Phys. Rev. Lett. 73 (21)
(1994) 2500.
[22] U. Lw, et al., Eur. Phys. J. B 86 (22) (2013) 87.
[23] K. Kamazawa, et al., Phys. Rev. B 78 (23) (2008) 064412.
[24] H. Kimura, H. Maeda, M. Sato, J. Mater. Sci. 23 (1988) 809.
[25] W. Wang, Y. Yue, J. Alloys Compd. 488 (2009) 2326.
[26] H.D. Zhou, et al., Phys. Rev. B 78 (R) (2008) 140406.
[27] J.A. Hodges, et al., J. Phys.: Condens. Matter 15 (2003) 4631.
[28] B.Y. Sokolov, Phys. Solid State 39 (1997) 12611262.
[29] J.C. Kim, et al., J. Am. Ceram. Soc. 90 (2007) 641644.
[30] J. Rodriguez-Carvajal, Physica B 192 (1993) 5569.
[31] Y. Katoh, N. Sugimoto, A. Tate, Jpn. J. Appl. Phys. 33 (1994) L726L729.
[32] S.I. Khartsev, A.M. Grishin, J. Appl. Phys. 101 (2007) 053906.
[36] J.R. Stewart, et al., Phys. Rev. B 82 (2010) 144439.
[37] J.M. Hopkinson, H.- Y. Kee, Phys. Rev. B 74 (2006) 224441.
[38] Y. Chen, et al., Phys. Rev. B 74 (2006) 014430.
[39] B.C. den Hertog, M.J.P. Gingras, Phys. Rev. Lett. 84 (2000) 34303433.
[40] N. Rogado, G. Lawes, D.A. Huse, A.P. Ramirez, R.J. Cava, Solid State Commun.
124 (2002) 229233.

[41]
[42]
[43]
[44]
[45]
[46]
[47]

[48]
[49]
[50]
[51]
[52]
[53]

M. Schieber, L. Holmes, J. Appl. Phys. 36 (1965) 11591160.


A.M. Stewart, J. Phys.: Condens. Matter 4 (1992) 68596861.
A.M. Stewart, Phys. Rev. B 6 (1972) 19851998.
A.M. Stewart, J. Phys. C: Solid State Phys. 17 (1983) 15571574.
S. Singh, et al., Phys. Rev. B 77 (2008) 054408.
Sarte P. M. et al. (2014) Private Communication. Results are currently being
prepared for publication.
M. Malinowski, R. Wolski, Z. Frukacz, T. Lukasiewicz, Z. Luczynski, J. Appl.
Spectrosc. 62 (1995) 840843;
P. Grnberg, S. Hfner, E. Orlich, J. Schmitt, J. Appl. Phys. 40 (1969) 15011502.
S. Arajs, Phys. Rev. 120 (1960) 756759.
Zorko A. et al. (2012) arXiv:cond-mat-1210.8187.
M.J.P. Gingras, et al., Phys. Rev. B 62 (2000) 64966511.
D.G. Onn, H. Meyer, Phys. Rev. 156 (1967) 663670.
J.A. Quilliam, et al., Phys. Rev. Lett. 99 (2007) 097201.
O.A. Petrenko, et al., Phys. Rev. Lett. 80 (1998) 4570.

You might also like