Está en la página 1de 7

Journal of Magnetism and Magnetic Materials 458 (2018) 164–170

Contents lists available at ScienceDirect

Journal of Magnetism and Magnetic Materials


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

Research articles

Hydrothermal synthesis and magnetic properties of multiferroic


RMn0.5Fe0.5O3 (R = Tb, Dy, and Ho)
Li Guo, Zhiqiang Zhou ⇑
College of Science, Northeast Forestry University, Harbin 150040, PR China

a r t i c l e i n f o a b s t r a c t

Article history: Pure-phased orthorhombic RMn0.5Fe0.5O3 (R = Tb, Dy, and Ho) were successfully synthesized via the low-
Received 1 February 2018 temperature hydrothermal technique. The nucleation and crystal growth of RMn0.5Fe0.5O3 (R = Tb, Dy,
Received in revised form 9 March 2018 and Ho) were influenced mainly by medium alkalinity, reaction temperature and time. All the three sam-
Accepted 10 March 2018
ples are of orthorhombic structure with +3 valences of Mn and Fe. They present morphologies of spheres
Available online 12 March 2018
that are assembled by smaller tabular crystals. Spin reorientation transitions were observed in
RMn0.5Fe0.5O3 (R = Tb, Dy, and Ho), and magnetization reversal was observed in TbFe0.5Mn0.5O3. The spin
Keywords:
reorientation temperatures show a linear relation with the magnetic field intensity, and the spin reorien-
Multiferroic
Manganite
tation and magnetization reversal disappear when the applied field is big enough. Antiferromagnetic
Spin reorientation transition of R3+ was observed in TbFe0.5Mn0.5O3 and DyFe0.5Mn0.5O3 at lower temperatures. All the three
Magnetization reversal samples show the characteristics of spiral magnetic ordering at 2 K and show residual magnetism at both
2 K and 300 K.
Ó 2018 Elsevier B.V. All rights reserved.

1. Introduction expected, which is worthy of research. As we know, multiferroic


occurs generally accompanied by magnetic phase transition
In recent years, people have done a lot of work in the synthesis [23–26]. The research of magnetic phase transitions may give us
and research of single-phased multiferroic materials that may theoretical and experimental guidance in the search for multifer-
integrate fine magnetism and ferroelectricity, and therein, some roic materials.
orthorhombic composite oxides such as rare-earth manganites Orthorhombic TbMnO3 [2], DyMnO3 [3] and HoMnO3 [4] are
(RMnO3) [1–4], ferrites (RFeO3) [5–12], etc. have become the typical multiferroic candidates. In this work, orthorhombic
research focus. Traditionally, ferroelectricity is considered to be RMn0.5Fe0.5O3 (R = Tb, Dy, and Ho) were prepared via the
forbidden in these centrosymmetric orthorhombic oxides. But fer- hydrothermal technique, and the magnetic phase transitions were
roelectricity induced by special spontaneous polarization mecha- analyzed, which will provide some fundamental data for further
nisms has been discovered in some of these oxides, for example, research in this field.
ferroelectricity induced by spiral magnetic ordering in orthorhom-
bic DyMnO3 [13–19], induced by exchange contraction of Dy and
Fe moments in orthorhombic DyFeO3 has been reported [6,20–22]. 2. Experimental
In the orthorhombic, i.e. perovskite, structure, B-site ions of dif-
ferent electron configurations may provide diverse compositions The starting materials were Fe(NO3)39H2O, R(NO3)36H2O
and interactions, and therein, Fe and Mn are typical B-site mag- (R = Tb, Dy and Ho), KMnO4, MnCl2, and KOH. They were all analyt-
netic elements. Fe3+ and Mn3+ have the same charge number and ically pure. For ensuring intensive mixing of the starting materials,
similar ionic radius, making Fe3+/Mn3+ co-doping in BO6 octahedra the salts were used in the form of aqueous solutions. KOH was used
possible. A variety of exchange interactions such as Fe3+–Fe3+, as a mineralizer in its initial form of pellets. Here, we take
Mn3+–Mn3+, Fe3+–Mn3+, R3+–Fe3+, R3+–Mn3+, R3+–R3+ may exist in TbMn0.5Fe0.5O3 for example to demonstrate the synthesis route.
the B-site Fe3+/Mn3+ co-doped rare-earth composite oxide Fe(NO3)3, Tb(NO3)3, and KMnO4 solutions were mixed at room
RMn1xFexO3, and so more magnetic phase transitions may be temperature, whose dosages were 10.00 mL, 5.00 mL and
3.33 mL, respectively, and whose concentrations were 0.40 M,
0.40 M and 0.12 M, respectively. KOH pellets were then gradually
⇑ Corresponding author. added into the mixture under strong stirring to reach a concentra-
E-mail address: zhouzq_nefu@163.com (Z. Zhou). tion of 20 M (molar number of KOH/initial solution volume). When

https://doi.org/10.1016/j.jmmm.2018.03.019
0304-8853/Ó 2018 Elsevier B.V. All rights reserved.
L. Guo, Z. Zhou / Journal of Magnetism and Magnetic Materials 458 (2018) 164–170 165

the above mixture reached room temperature, 2.86 mL MnCl2 alkalinity and higher temperature benefit the formation of
(0.56 M) was quickly added into it under stirring. The RMn0.5Fe0.5O3 (R = Tb, Dy, and Ho) while the reaction time produce
as-prepared mixture was transferred into a Teflon-lined stainless no significant impact provided it is no less than 48 h. The reaction
steel autoclave, and the filling degree was 80%. The autoclave temperature and time should be no lower than 240 °C and no
was kept at 240 °C for 48 h to complete crystal nucleation and shorter than 48 h, or else, the main products shall be R(OH)3,
growth under autogenous pressure. And then, the autoclave was Fe2O3 and KxMnO2yH2O.
naturally cooled to room temperature. Ultrasonic separation and The XRD patterns of the as-prepared samples are shown in
filtration were used to collect solid compounds from the autoclave. Fig. 1. High degree of crystallization is indicated by the strong
The solid compounds were thoroughly washed with deionized intensities and sharp shapes of the XRD patterns. From the embed-
water to remove the soluble matters, and then dried in air at ded figure, we can see that the diffraction peaks shift to the high
60 °C to present the final samples. The synthesis routes of other angle direction along with the decrease of R3+ radius. Indexation
samples were similar to that of TbMn0.5Fe0.5O3. of the XRD data shows that all the three samples are of orthorhom-
A X-ray diffractometer (Rigaku D/Max 2500 V/PC) was bic structure with space group Pnma. The lattice parameters were
employed to do powder X-ray diffraction (XRD), and the operation got through Pawley refinement, as shown in Table 1. Fig. 2 gives
parameters were as follows: room temperature, 50 kV, 200 mA, the Pawley refinement patterns of HoMn0.5Fe0.5O3 as an example.
Cu-Ka radiation (k = 1.5418 Å), step scanning, increments of The compositions of the samples were analyzed via EDS, as shown
0.02°, and angle range 20°  2h  80°. A scanning electron micro- in Fig. S1 and Table S1. It can be seen that the molar ratios of R:Mn:
scope (SEM, JEOL JSM-6700F) operated at 5 kV was employed to Fe in the samples accord with the designed ratios in the starting
observe the morphologies. Energy dispersive spectroscopy (EDS, materials.
attached to the above-said scanning electron microscope) was XPS was used to analyze the valence states of Mn and Fe in the
used to analyze the chemical compositions of the samples. X-ray samples. Fig. 3 gives the Mn2p and Fe2p patterns of DyMn0.5Fe0.5O3.
photoelectron spectroscopy (XPS, Kratos AXIS Ultra DLD spectrom- The binding energy was corrected with standard C1s (284.6 eV in
eter) was used to analyze the element valences. Quantum Design room temperature). In Fig. 3(b), we can find two peaks at
MPMS-XL (SQUID) was used to record the zero-field-cooled (ZFC) 710.8 eV and 724.3 eV which are ascribed to Fe2p3/2 and Fe2p1/2
and field-cooled (FC) curves with different density of sampling from split of Fe2p. With reference to the standard database, these
point. two peaks are consistent with that of Fe2O3, therefore, the valence
of Fe was determined to be +3. Similarly, the valence of Mn was also
determined to be +3. TbMn0.5Fe0.5O3, DyMn0.5Fe0.5O3 and
3. Results and discussion HoMn0.5Fe0.5O3 were prepared under the same conditions, and so
we have the reason to believe that Mn and Fe in all the three
Many factors may influence the nucleation and crystal growth samples show valence of +3 [27].
of RMn0.5Fe0.5O3 (R = Tb, Dy and Ho), and the most dominant ones The SEM images of RMn0.5Fe0.5O3 (R = Tb, Dy, and Ho) are given
are medium alkalinity, reaction temperature and time. Higher in Fig. 4. TbMn0.5Fe0.5O3 and DyMn0.5Fe0.5O3 take on morphologies

Fig. 1. The XRD patterns of TbMn0.5Fe0.5O3 (a), DyMn0.5Fe0.5O3 (b) and HoMn0.5-
Fe0.5O3 (c); the insert is the magnified main peaks. Fig. 2. Pawley Refinement Patterns of HoMn0.5Fe0.5O3.

Table 1
Lattice parameters obtained from the powder XRD data.

a b c V hai D
TbMn0.5Fe0.5O3 5.5919(1) 7.5897(3) 5.3161(4) 225.63 5.4236 0.02044
DyMn0.5Fe0.5O3 5.6155(1) 7.5396(3) 5.2979(2) 224.31 5.4130 0.024596
HoMn0.5Fe0.5O3 5.6005(1) 7.5727(5) 5.2756(1) 223.75 5.4058 0.023182
P pffiffiffi pffiffiffi ð1=3Þ
Note: a, b, c are unite cell parameters; D denotes the orthorhombic distortion, D ¼ ð1=3Þ i jai  haij=hai, ai denoting a, c and b= 2, hai ¼ ðabc= 2Þ .
166 L. Guo, Z. Zhou / Journal of Magnetism and Magnetic Materials 458 (2018) 164–170

Fig. 3. XPS patterns of Mn2p (a) and Fe2p (b).

Fig. 4. SEM images of TbMn0.5Fe0.5O3 (a, b), DyMn0.5Fe0.5O3 (c, d) and HoMn0.5Fe0.5O3 (e, f).

of spheres of 15 lm and 25 lm in diameter, respectively. From the the morphology takes on spheres and tabulars, suggesting that it
high magnification images, we can see that they are spheres may be just on the way of forming into spheres. It can be specu-
assembled from smaller tabular crystals. As for HoMn0.5Fe0.5O3, lated that the formation of the products underwent two processes:
L. Guo, Z. Zhou / Journal of Magnetism and Magnetic Materials 458 (2018) 164–170 167

first, tabular crystals formed, and then they assembled into recognized as Mn–O–Mn < Fe–O–Mn < Fe–O–Fe, making the TN of
spheres. We tried to learn the morphological formation by investi- RFe0.5Mn0.5O3 (R = Tb, Dy, and Ho) higher than that of RMnO3
gating the influence of reaction conditions on the morphology of (R = Tb, Dy, and Ho) and lower than that of (R = Tb, Dy, and Ho).
HoMn0.5Fe0.5O3, and found that reaction temperature has a signif- Weak ferromagnetism was observed in samples, probably ascribed
icant influence. When the reaction temperature was elevated to to the presence of single ion anisotropy or antisymmetric
260 °C, no change of phase was observed, but the tabular crystals Dzyaloshinski-Moriya super-exchange interaction due to the large
got smaller and their agglomeration got severer, as shown in the tilt of the BO6 octahedra [29]. The Weiss constants h were
embedded figure in Fig. 4(f). It can be explained as that the reaction
rate rose with the elevation of temperature, making that the quan-
tity of initial nucleuses increased and the sizes became smaller.
Smaller sizes led to increase of specific surface area and surface
free energy, and therefore, the smaller nucleuses intended to
agglomerate to lower the surface free energy [28]. It can be
expected that, with further elevation of reaction temperature,
HoMn0.5Fe0.5O3 would formed into spheres assembled from
tabulars.
The ZFC and FC magnetization curves of RFe0.5Mn0.5O3 (R = Tb,
Dy, and Ho) recorded in an applied field of 100 Oe under 2–350
K are given in Fig. 5. The ionic radius of Fe3+(3d5) is nearly the same
with that of Mn3+(3d4), facilitating the random distribution of Fe3+
and Mn3+ in the crystal lattice sites. Fe3+ and Mn3+ are distributed
at the B-sites half-and-half at random, producing coexistence of
Fe3+–O2–Fe3+, Mn3+–O2–Fe3+ and Mn3+–O2–Mn3+ super-
exchanges. The curves of ZFC magnetization reciprocal against
temperature (1/M–T) of RFe0.5Mn0.5O3 (R = Tb, Dy, and Ho) are
given in Fig. 6. From Fig. 6, the antiferromagnetic (AFM) transition
temperatures TN of RFe0.5Mn0.5O3 (R = Tb, Dy, and Ho) were deter-
mined to be 333 K, 327 K and 347 K, which are higher than that of
RMnO3 (R = Tb, Dy, and Ho) and lower than that of RFeO3 (R = Tb, Fig. 7. ZFC magnetization curves of Tb0.5Fe0.5Mn0.5O3 in different fields: (a) 70 Oe,
Dy, and Ho). The order of super-exchange intensities are (b) 100 Oe, (c) 500 Oe, (d) 1000 Oe, (e) 3000 Oe, (f) 5000 Oe.

Fig. 5. ZFC and FC magnetization curves of RMn0.5Fe0.5O3 in an applied field of 100 Oe: (a) TbMn0.5Fe0.5O3, (b) DyMn0.5Fe0.5O3, (c) HoMn0.5Fe0.5O3.

Fig. 6. 1/M–T curves of RMn0.5Fe0.5O3 in an applied field of 100 Oe: (a) TbMn0.5Fe0.5O3, (b) DyMn0.5Fe0.5O3, (c) HoMn0.5Fe0.5O3.
168 L. Guo, Z. Zhou / Journal of Magnetism and Magnetic Materials 458 (2018) 164–170

determined to be 146, 152, and 123 K respectively via fitting the 65 K, respectively, which is called compensation temperature, that
paramagnetic region above TN. is Tcomp = 87 K and Tcomp = 65 K, and whereafter, the magnetization
All the three samples showed spin reorientation transition got reversed. The spin reorientations of DyFe0.5Mn0.5O3 and
along with further fall of temperature below the Neel points TN, HoFe0.5Mn0.5O3 were observed in the range of 238–252 K and
but the transition nature was not the same, as shown in the inserts 200–280 K. Compared with TbFeO3, DyFeO3 and HoFeO3 [31], the
in Fig. 5. TbFe0.5Mn0.5O3 showed spin reorientation transition and spin reorientation temperature ranges of RFe0.5Mn0.5O3 (R = Tb,
magnetization reversal. With fall of temperature, transition peaks Dy, and Ho) have been elevated. In RFeO3, the spin reorientation
at 169 K and 192 K on the ZFC and FC curves were observed respec- along a certain direction begins when R3+–Fe3+ interaction over-
tively, indicating the beginning of spin reorientation and magneti- takes that of Fe3+–Fe3+ [32–36]. In RFe0.5Mn0.5O3 (R = Tb, Dy, and
zation reversal. Magnetization reversal is considered to be caused Ho), some Fe3+ was replaced by Mn3+ producing interaction of
by the magnetization variety of sublattices under different temper- Mn3+–Fe3+/Mn3+–Mn3+ instead of equivalent interaction of
atures and can be explained by the antisymmetry-based Fe3+–Fe3+. Therefore, at higher temperatures compared with RFeO3,
Dzyaloshinski-Moriya interaction and the anisotropy competition the interaction of R3+–Fe3+ can overtake that of Fe3+/Mn3+–Fe3+/
of single ions [30]. From Fig. 5, it can be found that although the Mn3+ and the lattice vibration, and therefore, the spin reorientation
starting temperatures were 169 K and 192 K on the ZFC and FC begins at a higher temperature. From Fig. 5, we can also see that, at
curves, respectively, the spin reorientation stopped at the same lower temperatures the magnetization of all the three samples
temperature 20 K. Below the starting temperature TSR2, the increased dramatically, which is considered to be caused by the
magnetization on ZFC and FC curves reached zero at 87 K and paramagnetism of R3+.
The direction of magnetization can be reversed by several factors
such as external magnetic field [37], spin injection [38] and circu-
Table 2 larly polarized light [39]. Here, we investigate the relation between
Spin reorientation temperature TSR1 and TSR2 of Tb0.5Fe0.5Mn0.5O3 in different fields. magnetization reversal and the applied external magnetic field by
Field intensity/Oe TSR1/K TSR2/K altering the field intensity. Negative magnetization means the mag-
70 13.24 204 netization is in the reverse direction with that of the magnetic field.
100 19.84 191 The magnetic anisotropy energy can be overtaken by the energy
500 27.14 140 induced by magnetic field if the magnetic field intensity is strong
1000 48 90 enough, then the net moment is parallel to the field, the negative
3000 / /
magnetization disappears, and the initial magnetization direction
5000 / /
is reversed. We measured the ZFC curves of TbFe0.5Mn0.5O3 in fields
of 70, 100, 500, 1000, 3000 and 5000 Oe under 2–300 K, as shown in
Fig. 7. From Fig. 7, we can see that, with the increase of magnetic field
intensity, the original magnetization reversal became weaker and
the spin reorientation temperature region got narrower, and the
former disappeared in a field reached 1000 Oe, and the spin reorien-
tation disappeared in a field of 3000 Oe. Table 2 lists the spin reori-
entation temperatures TSR1 and TSR2 in different fields. The beginning
temperature TSR2 fell with the increase of magnetic field intensity
while the ending temperature TSR1 rose, and so the temperature
region got narrower. A linear relation between spin reorientation
beginning/ending temperature TSR2/TSR1 and field intensity H was
observed, as shown in Fig. 8. The fitting equations are as follows:

For T SR1 : T ¼ 0:0329H þ 13:819 ð1Þ

For T SR2 : T ¼ 0:116H þ 202:002 ð2Þ


From Eqs. (1) and (2), dTSR2/dH = 3.5dTSR1/dH can be derived,
from which we can predict that the spin reorientation will be dis-
appear when the magnetic field reaches 1260 Oe.
Fig. 8. Magnetic field dependence of the spin reorientation temperatures TSR2 and At about liquid helium temperatures, some R3+ ions will have
TSR1 of Tb0.5Fe0.5Mn0.5O3.
transition from paramagnetic to antiferromagnetic state, which is

Fig. 9. FC curves of TbMn0.5Fe0.5O3 (a), DyMn0.5Fe0.5O3 (b), HoMn0.5Fe0.5O3 (c) in an applied field of 100 Oe from 2 K to 10 K.
L. Guo, Z. Zhou / Journal of Magnetism and Magnetic Materials 458 (2018) 164–170 169

denoted by TN(R3+). Fig. 9 gives the FC curves of RMn0.5Fe0.5O3 magnetic transition at about 2 K. No transition or its trend for
(R = Tb, Dy, and Ho) in a field of 100 Oe from 2–10 K. It can be seen HoMn0.5Fe0.5O3 was observed in this temperature region. Fig. 10
that TbMn0.5Fe0.5O3 underwent a transition at about 3 K attributed gives the curves of dM/dT against temperature T of TbMn0.5Fe0.5O3
to the antiferromagnetic transition of Tb3+. As for DyMn0.5Fe0.5O3, in different fields. It can be seen that there exist characteristic tem-
we did not measure it at 2 K and so no obvious inflection point peratures of magnetic transitions attributed to rare-earth ion spin
was observed on the curve. But from 4 K downward a transition ordering in antiferromagnetic arrangement. The transition temper-
trend can be obviously observed, and with reference to ature shows a significant change with field intensity, as shown in
TbMn0.5Fe0.5O3, we may speculate that Dy3+ will have antiferro- the inset of Fig. 10. The temperature rose from 6.12 to 11.43 K with
the increase of field intensity.
We measured the M–H curves for TbMn0.5Fe0.5O3, DyMn0.5Fe0.5-
O3 and HoMn0.5Fe0.5O3 from 30,000 Oe to 30,000 Oe at 2, 70, 150
and 300 K, as shown in Fig. 11. It can be seen that the three sam-
ples show neither typical paramagnetic nor antiferromagnetic
behavior but that of spiral spin configurations [40]. At 2 K, all the
three samples show residual magnetism, which may be ascribed
to that the local Dzyaloshinski-Moriya interaction in frustrated
spin is weak ferromagnetism. At 300 K, all the three samples also
show residual magnetism. This is because below TN,
Dzyaloshinsky-Moriya antisymmetric exchange leads to that the
moment of Mn3+/Fe3+ is not completely parallel to that of the
neighboring Mn3+/Fe3+, but a small inclination angle exists instead,
producing weak ferromagnetism. No residual magnetism was
observed at 70 K or 150 K.

4. Conclusions

Pure-phased orthorhombic RMn0.5Fe0.5O3 (R = Tb, Dy, and Ho)


Fig. 10. dM/dT–T curves of TbMn0.5Fe0.5O3 at different fields; inset shows the were successfully synthesized via the low-temperature hydrother-
transition temperatures T. mal technique.

Fig. 11. Magnetic isotherms of TbMn0.5Fe0.5O3 (a), DyMn0.5Fe0.5O3 (b) and HoMn0.5Fe0.5O3 (c) from 3 T to 3 T at 2 K, 70 K, 150 K and 300 K; the insets: magnetic isotherms
around H = 0.
170 L. Guo, Z. Zhou / Journal of Magnetism and Magnetic Materials 458 (2018) 164–170

(1) Relatively higher reaction temperature and higher alkalinity [8] C.J. Du, H.W. Qin, S.Q. Ren, L. Zhao, M.L. Zhao, W.B. Su, J.F. Hu, Appl. Phys. Lett.
104 (2014) 082415.
benefit the formation of target products, while the reaction
[9] W.Y. Zhao, S.X. Cao, R.X. Huang, Y.M. Cao, K. Xu, B.J. Kang, J.C. Zhang, W. Ren,
time has no significant impact and 48 h is enough. Phys. Rev. B 91 (2015) 104425.
(2) The antiferromagnetic transition temperatures of RMn0.5- [10] X.Y. Zhao, K.L. Zhang, K. Xu, P.W. Man, T. Xie, A.H. Wu, G.H. Ma, S.X. Cao, L.B. Su,
Fe0.5O3 (R = Tb, Dy, and Ho) were determined to be 333 K, Solid State Commun. 231 (2016) 43.
[11] A.H. Wu, B. Wang, X.Y. Zhao, T. Xie, P.W. Man, L.B. Su, A.M. Kalashnikova, R.V.
327 K, and 347 K. The three samples showed spin reorienta- Pisarev, J. Magn. Magn. Mater. 426 (2017) 721.
tion in the regions of 20–169 K, 238–252 K, and 200–280 K, [12] M.C. Silva-Santana, C.A. da Silva, P. Barrozo, E.J.R. Plaza, L. de los Santos
respectively. TbFe0.5Mn0.5O3 showed magnetization reversal. Valladares, N.O. Moreno, J. Magn. Magn. Mater. 401 (2016) 612.
[13] I.A. Scrgicnko, C. Sen, E. Dagotto, Phys. Rev. Lett. 97 (2006) 227204.
(3) The beginning temperature of spin reorientation TSR2 of [14] M. Mostovoy, Phys. Rev. Lett. 96 (2006) 067601.
TbFe0.5Mn0.5O3 fell with the increase of applied magnetic [15] H.S. Nair, S. Elizabeth, J. Cryst. Growth 362 (2013) 24.
field intensity while the ending temperature TSR1 rose, and [16] N. Zhang, S. Dong, J.M. Liu, Front. Phys. 7 (2012) 408.
[17] C.L. Lu, S. Dong, Z.C. Xia, H. Luo, Z.B. Yan, H.W. Wang, Z.M. Tian, S.L. Yuan, T.
a linear relation between TSR2/TSR1 and field intensity was Wu, J.M. Liu, Sci. Rep. 3 (2013) 3374.
observed. [18] H.J. Xiang, P.S. Wang, M.-H. Whangbo, X.G. Gong, Phys. Rev. B 88 (2013)
(4) The three samples showed the characteristics of spiral mag- 054404.
[19] H.J. Xiang, E.J. Kan, Y. Zhang, M.-H. Whangbo, X.G. Gong, Phys. Rev. Lett. 107
netic ordering at 2 K and residual magnetism at both 2 K and (2011) 157202.
300 K. TbFe0.5Mn0.5O3 and DyFe0.5Mn0.5O3 showed the anti- [20] R. Iida, T. Satoh, T. Shimura, K. Kuroda, B.A. Ivanov, Y. Tokunaga, Y. Tokura,
ferromagnetic transition of R3+ at lower temperatures. Phys. Rev. B 84 (2011) 064402.
[21] M. Liu, S.X. Cao, S.J. Yuan, B.J. Kang, B. Lu, J.C. Zhang, Acta Phys. Sin. 62 (2013)
147601.
[22] H. Xie, S.J. Yuan, B.J. Kang, B. Lu, S.X. Cao, J.C. Zhang, J. Inorg. Mater. 29 (2014)
Acknowledgements 77.
[23] I. Levin, M.G. Tucker, H. Wu, V. Provenzano, C.L. Dennis, S. Karimi, T. Comyn, T.
Stevenson, R.I. Smith, I.M. Reaney, Chem. Mater. 23 (2011) 2166.
This work was financially supported by the Fundamental [24] J. Strempfer, B. Bohnenbuck, I. Zegkinoglou, N. Aliouane, S. Landsgesell, M.V.
Research Funds for the Central Universities (2572016BB09, Zimmermann, D.N. Argyriou, Phys. Rev. B 78 (2008) 024429.
2572015CB28), the Scientific Research Fund of Heilongjiang Provin- [25] V.A. Khomchenko, L.C.J. Pereira, J.A. Paixão, J. Appl. Phys. 115 (2014) 164101.
[26] K.L. Yadav, J. Nanosci. Nanotechnol. 11 (2011) 2682.
cial Education Department (12543020), and the Heilongjiang [27] V. Naumkin, A. Kraut-Vass, S.W. Gaarenstroom, C.J. Powell, NIST X-ray
Postdoctoral Financial Assistance (LBH-Z16012). Photoelectron Spectroscopy Database, Version 4.1 (National Institute of
Standards and Technology, Gaithersburg, 2012); http://srdata.nist.gov/xps/.
[28] J. Li, Y.Q. Lin, B.G. Zhao, J. Nanopart. Res. 4 (2002) 345.
Appendix A. Supplementary data [29] D. Delmonte, F. Mezzadri, C. Pernechele, G. Calestani, G. Spina, M. Lantieri, M.
Solzi, R. Cabassi, F. Bolzoni, A. Migliori, C. Ritter, E. Gilioli, Phys. Rev. B 88
(2013) 014431.
Supplementary data associated with this article can be found, in
[30] S.J. Yuan, W. Ren, F. Hong, Y.B. Wang, J.C. Zhang, L. Bellaiche, S.X. Cao, G. Cao,
the online version, at https://doi.org/10.1016/j.jmmm.2018.03.019. Phys. Rev. B 87 (2013) 184405.
[31] Z.Q. Zhou, L. Guo, H.X. Yang, Q. Liu, F. Ye, J. Alloy Compd. 583 (2014) 21.
References [32] R.L. White, J. Appl. Phys. 40 (1969) 1061.
[33] L.T. Tsymbal, Y.B. Bazaliy, V.N. Derkachenko, V.I. Kamenev, G.N. Kakazei, F.J.
Palomares, P.E. Wigen, J. Appl. Phys. 101 (2007) 123919.
[1] T. Kimura, T. Goto, H. Shintani, K. Ishizaka, T. Arima, Y. Tokura, Nature 426 [34] S.R. Brown, I. Hall, J. Phys.: Condens. Matter. 5 (1993) 4207.
(2003) 55. [35] G.H. Guo, H.B. Zhang, J. Alloy Compd. 429 (2007) 46.
[2] S.W. Cheong, M. Mostovoy, Nat. Mater 6 (2007) 13. [36] L.G. Marshall, J.G. Cheng, J.S. Zhou, J.B. Goodenough, J.Q. Yan, D.G. Mandrus,
[3] O. Prokhnenko, R. Feyerherm, E. Dudzik, S. Landsgesell, N. Aliouane, L.C. Phys. Rev. B 86 (2012) 064417.
Chapon, D.N. Argyriou, Phys. Rev. Lett. 98 (2007) 057206. [37] P. Mandal, A. Sundaresan, C.N.R. Rao, A. Iyo, P.M. Shirage, Y. Tanaka, C. Simon,
[4] N. Lee, Y.J. Choi, M. Ramazanoglu, W. Ratcliff II, V. Kiryukhin, S.-W. Cheong, V. Pralong, O.I. Lebedev, V. Caignaert, B. Raveau, Phys. Rev. B 82 (2010) 100416.
Phys. Rev. B 84 (2011) 020101. [38] J. Stöhr, H.C. Siegmann, Magnetism: from fundamentals to nanoscale
[5] J. Wang, J.B. Neaton, H. Zheng, V. Nagarajan, S.B. Ogale, B. Liu, D. Viehland, V. dynamics, Springer Series in Solid State Sci. 152 (2006) 656.
Vaithyanathan, D.G. Schlom, U.V. Waghmare, N.A. Spaldin, K.M. Rabe, M. [39] C.D. Stanciu, F. Hansteen, A.V. Kimel, A. Kirilyuk, A. Tsukamoto, A. Itoh, T.
Wuttig, R. Ramesh, Science 299 (2003) 1719. Rasing, Phys. Rev. Lett. 99 (2007) 047601.
[6] Y. Tokunaga, S. Iguchi, T. Arima, Y. Tokura, Phys. Rev. Lett. 101 (2008) 097205. [40] D.V. Tac, V.O. Mittova, I.Y. Mittova, Inorg. Mater. 47 (2011) 521.
[7] Y. Tokunaga, N. Furukawa, H. Sakai, Y. Taguchi, T. Arima, Y. Tokura, Nat. Mater.
8 (2009) 558.

También podría gustarte