8.A Comparative Study of Ultra-Low-Temperature Thermal Conductivity of Multiferroic

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

A comparative study of ultra-low-temperature thermal conductivity of multiferroic

orthoferrites RFeO3 (R = Gd and Dy)


J. Y. Zhao, Z. Y. Zhao, J. C. Wu, H. S. Xu, X. G. Liu, X. Zhao, and X. F. Sun

Citation: AIP Advances 7, 055806 (2017); doi: 10.1063/1.4973293


View online: http://dx.doi.org/10.1063/1.4973293
View Table of Contents: http://aip.scitation.org/toc/adv/7/5
Published by the American Institute of Physics

Articles you may be interested in


Thermal conductivity of ferrimagnet GdBaMn2O5.0 single crystals
AIP Advances 7, 055807 (2016); 10.1063/1.4973294

Nanopatterning spin-textures: A route to reconfigurable magnonics


AIP Advances 7, 055601 (2016); 10.1063/1.4973387

Identification of the low-energy excitations in a quantum critical system


AIP Advances 7, 055701 (2016); 10.1063/1.4972802

Anisotropic electronic states in the fractional quantum Hall regime


AIP Advances 7, 055804 (2016); 10.1063/1.4972854
AIP ADVANCES 7, 055806 (2017)

A comparative study of ultra-low-temperature thermal


conductivity of multiferroic orthoferrites
RFeO3 (R = Gd and Dy)
J. Y. Zhao,1 Z. Y. Zhao,1,2 J. C. Wu,1 H. S. Xu,1 X. G. Liu,1 X. Zhao,3,a
and X. F. Sun1,4,5,b
1 HefeiNational Laboratory for Physical Sciences at Microscale, University of Science
and Technology of China, Hefei, Anhui 230026, People’s Republic of China
2 Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou,

Fujian 350002, People’s Republic of China


3 School of Physical Sciences, University of Science and Technology of China, Hefei,

Anhui 230026, People’s Republic of China


4 Key Laboratory of Strongly-Coupled Quantum Matter Physics, Chinese Academy of Sciences,

Hefei, Anhui 230026, People’s Republic of China


5 Collaborative Innovation Center of Advanced Microstructures, Nanjing, Jiangsu 210093,

People’s Republic of China


(Presented 3 November 2016; received 18 September 2016; accepted 13 October 2016;
published online 23 December 2016)

Ultra-low-temperature thermal conductivity (κ) of GdFeO3 and DyFeO3 single crys-


tals is studied down to several tens of milli-Kelvin. It is found that the κ is purely
phononic and has strong magnetic-field dependence, indicating a strong spin-phonon
coupling. Moreover, the low-T κ(H) with H k c show rather different behaviors in these
two materials. In particular, the κ of GdFeO3 can be strongly enhanced in several tesla
field and becomes weakly field dependent in higher fields up to 14 T; whereas, the κ of
DyFeO3 is continuously suppressed with increasing field and does not show any signa-
ture of recovery at 14 T. The results can be well understood by the difference in the spin
anisotropy of Gd3+ and Dy3+ ions. © 2016 Author(s). All article content, except where
otherwise noted, is licensed under a Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/). [http://dx.doi.org/10.1063/1.4973293]

Multiferroicity induced by spin order has attracted much attention due to its large magnetoelectric
(ME) coupling. The spin-current model or the inverse Dzyaloshinsky-Moriya (DM) interaction can
explain well the production of electric polarization in the non-collinear spin systems.1–4 When the
spins are aligned collinearly, the electric polarization can also be formed through the exchange
striction mechanism, for example, in the rare-earth-based orthoferrites RFeO3 (R = Gd and Dy).5–7
In this case, the interaction between adjacent Fe3+ and R3+ layers drives R3+ ions to displace along
the c axis so as to induce the ferroelectric polarization along the c axis.5,6 In these materials, the spin
structures are playing a key role in the ME coupling and the formation of the spontaneous electric
polarization.
Due to their key role in the ME coupling, the spin structures and spin re-orientations of rare-earth
orthoferrites have been intensively studied.8–13 The rare-earth orthoferrite has a distorted perovskite
structure with an orthorhombic unit cell (Pbnm).14 It is known that Fe3+ spins form an antiferro-
magnetic (AF) order at rather high temperature (TNFe > 600 K) with a Gx Ay F z spin configuration in
Bertaut’s notation;15 that is, the main component of the magnetic moment is along the a axis, accom-
panied with weak ferromagnetism along the c axis.16,17 On the other hand, the rare-earth moments
order antiferromagnetically at much lower temperatures.

a
Electronic mail: xiazhao@ustc.edu.cn
b
Electronic mail: xfsun@ustc.edu.cn

2158-3226/2017/7(5)/055806/7 7, 055806-1 © Author(s) 2016


055806-2 Zhao et al. AIP Advances 7, 055806 (2017)

In GdFeO3 , a Gx -type spin structure of Gd3+ moments is formed below the Néel temperature
TNGd = 2.5 K.6,15,18 At the same time, the ferroelectric polarization appears and is originated from the
spin-exchange striction.5,6,19 With applying magnetic field along the c axis, the magnetic structure
changes from phase I (Gx Ay F z for Fe3+ spins and Gx Ay for Gd3+ moments) to phase IV (Fe3+ : Gx Ay F z ;
Gd3+ : F z ) at ∼ 2.5 T due to the spin-polarization transition of Gd3+ moments.6 Whereas, the electric
polarization decreases to zero as long as the Gd3+ moments are polarized at high field.6
In DyFeO3 , with lowering temperature, the Fe3+ spins undergo a Morin transition at T M = 50 K,20
where the spin configuration changes from Gx Ay F z to Ax Gy C z .17 Moreover, in the c-axis field, the
Fe3+ spin structure can change back to Gx Ay F z .5,21 With further lowering temperature, the Dy3+ spins
Dy
develop a long-range AF order below TN = 4.2 K.5 In the AF state, the Dy3+ spin configuration
can be expressed as Gx Ay , for which the Dy3+ spins are confined in the ab plane with the Ising axis
deviating about 33◦ from the b axis,22,23 and the Fe3+ sublattice was believed to have the Ax Gy C z
Dy
structure.5 The spin-induced multiferroicity was observed only at T < TN and when the spin flop of
Fe3+ moments is introduced by a c-axis field.5 However, our recent work indicated a different ground
state from Ax Gy C z when cooling the sample in zero field.24,25
Low-temperature heat transport is an important physical property of solids and is useful for
probing elementary excitations, such as phonons and magnons. In magnetic insulators, the thermal
conductivity (κ) can show drastic changes at the magnetic transitions, due to either magnon transport
or magnon-phonon scattering.26–33 Our earlier works on the low-T heat transport of GdFeO3 and
DyFeO3 revealed that the field-dependencies of κ are strong and show anomalies at magnetic transi-
tions.24,34 Moreover, the low-T κ(H) isotherms with H k c show some peculiar irreversibilities, which
have different origins for two materials.24,34 Here, we report a study on the thermal conductivity of
GdFeO3 and DyFeO3 single crystals at very low temperatures down to several tens of milli-Kelvin
and in magnetic fields up to 14 T. The main finding in this work is that the heat transport has rather
different field-dependence in two materials, which is originated from the significant difference in the
spin anisotropy of the rare-earth ions.
High-quality RFeO3 (R = Gd, Dy, and Y) single crystals were grown by the floating-zone tech-
nique in flowing oxygen-argon mixture with the ratio of 1:4 or pure oxygen. The samples for the κ
measurements were cut along the crystallographic axes into long parallelepiped after orientation by
using the x-ray Laue photographs. Thermal conductivity was measured by using a “one heater, two
thermometers” technique and three different cryostats:24,28–34 (i) in a 3 He-4 He dilution refrigerator at
temperature regime of 70 mK–1 K; (ii) in a 3 He refrigerator at 0.3–8 K, and (iii) in a pulse-tube refrig-
erator for zero-field data at T > 5 K. Magnetization (M) was measured by a SQUID-VSM (Quantum
Design).
Figure 1 shows the representative data of M(T ) and M(H) with H k c for the GdFeO3 and DyFeO3
single crystals. These results are in good consistency with the earlier works.5,6,17 For GdFeO3 , the
M(T ) curve has a weak transition at ∼ 2.5 K, which is known to be due to the Néel transition of Gd3+
moments.6 At 2 K, the c-axis field can easily polarize the Gd3+ moments, as the M(H) curve indicates.
For DyFeO3 , the M(T ) curve measured in H = 5000 Oe, shown in Fig. 1(c), has two transitions at
5.2 and 4 K, which are the transition of the Fe3+ structure from Gx Ay F z to Ax Gy C z and the Néel
transition of Dy3+ moments, respectively.17 It can be seen from the 2 K magnetization curve that 7 T
is still too weak to polarize the Dy3+ moments, because they have strong anisotropy and are confined
in the ab plane.5,22,23 Therefore, the c-axis field can hardly to change either the Néel transition or the
Dy3+ spin orientation, whereas the two transitions at ∼ 2.5 and 3.5 T in the M(H) curve are due to
the magnetic transitions of Fe3+ spins.24
Figure 2 shows the temperature dependencies of the c-axis thermal conductivity of GdFeO3 and
DyFeO3 down to 70 mK. For comparison, the data for YFeO3 single crystal are also taken in the
same temperature regime. Note that the Y3+ ions are nonmagnetic and there is only AF order of Fe3+
ions. YFeO3 shows a simple and pure phonon heat transport behavior at low temperatures. First,
the κ(T ) curve exhibits a very large phonon peak at about 20 K, with the magnitude of 520 W/Km,
indicating high quality of the single crystal. Second, the temperature dependence of κ is roughly
T 2.7 at subKelvin temperatures, which is close to the T 3 boundary scattering limit of phonons.35 The
DyFeO3 data are rather comparable to those of YFeO3 , including the similar T dependence at the
055806-3 Zhao et al. AIP Advances 7, 055806 (2017)

FIG. 1. Magnetization of GdFeO3 and DyFeO3 measured in magnetic field along the c axis. These data were taken after
cooling the samples to 2 K in zero field. In panel (b), the M(H) curve is reversible between 0 and 7 T for GdFeO3 . In panel
(d), the M(H) curve shows a low-field hysteresis for DyFeO3 . The first transition at 2.5 T disappears when sweeping down
field from 7 T. The arrows indicate the direction of changing field.

FIG. 2. Temperature dependencies of the c-axis thermal conductivity of GdFeO3 , DyFeO3 , and YFeO3 . The data with a 14 T
field (along the c axis) of GdFeO3 are also shown. The dashed line shows a T 2.7 dependence. Inset: temperature dependence
of the phonon mean free path l divided by the averaged sample width W.
055806-4 Zhao et al. AIP Advances 7, 055806 (2017)

lowest temperatures. However, the κ(T ) curve has a clear concavity structure at 0.3–3 K. It is clear
Dy
that at T < TN , the magnon excitations from the Dy3+ spin system can have a significant scattering
on phonons and result in a downward deviation from the T 2.7 behavior. With lowering temperature
further, the κ recovers to the T 2.7 dependence at T < 300 mK. It is likely that the magnon spectra has
a finite energy gap, which prevents the low-energy magnons from being thermally excited at very
low temperatures. It is compatible with the strong anisotropy of Dy3+ spins.
GdFeO3 shows a rather different behavior of κ(T ) at very low temperatures.34 First of all, the zero-
field curve also exhibits a concavity feature below 2 K, which should be due to the magnon-phonon
scattering when the Gd3+ moments order antiferromagnetically at 2.5 K.6 However, the κ(T ) data
show a distinct deviation from the T 3 law, which indicates stronger magnetic scattering of phonons in
this material.35 This result can be understood from the fact that the Gd3+ spins have weak anisotropy6
and the magnon gap is negligibly small, in contrast to the strong anisotropy of Dy3+ spins.
The effect of magnetic field on κ(T ) is also tested for GdFeO3 . As shown in Fig. 2, when 14 T
magnetic field is applied along the c axis, the κ at T < 2 K become larger and the concavity feature
disappears, which clearly indicates the negative effect of magnons on the heat transport. A T 2.7
dependence of the 14 T data indicates that magnetic scattering on phonons is almost smeared out in
14 T field. Apparently, 14 T is strong enough to polarize the Gd3+ spins and the low-energy magnons
are hardly thermally excited. It is consistent with the M(H) data shown in Fig. 1.
It is useful to make an estimation of the mean free path of phonons at low temperatures. The
phononic thermal conductivity can be expressed by the kinetic formula κ ph = 13 C3p l,35 where C = βT 3
is the phonon specific heat at low temperatures, 3 p is the average velocity and l is the mean free path of
phonons. The β value can be obtained from the specific-heat data, which is 1.67 × 10−4 J/K4 mol and

FIG. 3. (a) Magnetic-field dependencies of the c-axis thermal conductivity of GdFeO3 single crystal at very low temperatures.
The magnetic field is applied along the c axis. All the data are measured after zero-field cooling. (b–d) The low-field data show
a hysteresis behavior. As indicated by the arrows, the data shown with solid symbols are measured in the field-up process,
while the open symbols show the data in the field-down process.
055806-5 Zhao et al. AIP Advances 7, 055806 (2017)

1.28 × 10−4 J/K4 mol for GdFeO3 and DyFeO3 , respectively.24,34 Then, the phonon velocity can be
calculated and finally the mean free path is obtained
√ from the κ.30,34 The inset to Fig. 2 shows the ratio
between l and the averaged sample width W = 2 A/π, 30,34,35 where A is the area of cross section, for
DyFeO3 in zero field and GdFeO3 in 14 T. It can be seen that l/W increases with lowering temperature
and becomes larger than one at lowest temperatures, which indicates that all the microscopic phonon
scatterings (including magnon scattering) are negligible and the boundary scattering limit is actually
established.35
For GdFeO3 , our previous work have studied the magnetic-field dependencies of κ with H k c
and at low temperatures down to 360 mK.34 The main results include: (i) the κ(H) isotherms show
a reduction at low fields followed by a strong enhancement at high fields; (ii) there is a shallow
and broad “dip” of κ(H) at low field; (iii) the κ(H) exhibit hysteresis at subkelvin temperatures. In
present work, the magnetic-field dependencies of κ are studied at even lower temperatures, down to
92 mK. As shown in Fig. 3, the present data show good consistency with the earlier data. However,
there are several notable features. First, with lowering temperature, the high-field enhancement of κ
is apparently becoming smaller, which indicates that the magnon scattering of phonons is weakened.
This is understandable because the magnon excitations should become more difficult unless the
magnon gap (caused by anisotropy) is exactly zero. Second, with lowering temperature, the low-field
broad dip becomes more shallow and finally evolves into a weakly field-dependent feature. Third,
the low-field hysteresis becomes a bit larger with lowering temperature. This hysteresis has been
proposed to be related to the ferroelectric domain scattering in this multiferroic material.34 It should
be noted that in temperature regime of 100 mK, the low-field κ is about only two times smaller than
the 14 T value, which means that the mean free path of phonons in the low field has the same order of
magnitude as the sample size. Therefore, the ferroelectric domain scattering seems to be ineffective
and it is not very clear whether the hysteresis of κ(H) has some other origin at such low temperatures. It
calls for further investigations by using electric polarization measurements at ultra-low temperatures.

FIG. 4. Magnetic-field dependencies of the c-axis thermal conductivity of DyFeO3 single crystal in H k c after zero-field
cooling. The arrows indicate the direction of changing field.
055806-6 Zhao et al. AIP Advances 7, 055806 (2017)

For comparison, the κ(H) isotherms of DyFeO3 at 92–360 mK are shown in Fig. 4. Note that the
low-field data (H ≤ 8 T) have been published elsewhere,24 while Fig. 4 shows data in field up to 14 T.
It has been known that the κ(H) of DyFeO3 display rather complicated irreversibility with lowering
temperature. At very low temperatures, the hysteresis of κ(H) at ∼ 4 T was discussed to be related to a
first-order transition of magnetic structure, with the “dip” field corresponding to the transition field.24
This is essentially different from the irreversibility of GdFeO3 data. In present work, one finding is
that the high-field data of DyFeO3 also behave very differently from those of GdFeO3 . Actually, the
κ is continuously suppressed in high field and does not show any signature of recovering even at 14
T. This phenomenon is consistent with the strong spin anisotropy of Dy3+ . It is known that the Dy3+
moments can hardly be polarized by the c-axis field, as Fig. 1(d) shows.
In summary, ultra-low-T thermal conductivity are studied for GdFeO3 and DyFeO3 single crys-
tals. The magnetic field along the c axis has strong effect on the κ, which can be understood by the
magnon-phonon scattering. However, both the temperature and field dependencies of κ are rather dif-
ferent in these two materials, which is closely related to the magnetic excitations at low temperatures,
determined by their different spin anisotropy of the rare-earth ions. Furthermore, the magnon excita-
tions seem to act only as phonon scatterers and there is no signature for their ability of transporting
heat.
This work was supported by the National Natural Science Foundation of China (Grant Nos.
11374277, 11574286, 11404316, U1532147), the National Basic Research Program of China
(Grant Nos. 2015CB921201, 2016YFA0300103), and the Opening Project of Wuhan National High
Magnetic Field Center (Grant No. 2015KF21).
1 H. Katsura, N. Nagaosa, and A. V. Balatsky, Phys. Rev. Lett. 95, 057205 (2005).
2 I. A. Sergienko and E. Dagotto, Phys. Rev. B 73, 094434 (2006).
3 M. Mostovoy, Phys. Rev. Lett. 96, 067601 (2006).
4 T. Kimura, T. Goto, H. Shintani, K. Ishizaka, T. Arima, and Y. Tokura, Nature 426, 55 (2003).
5 Y. Tokunaga, S. Iguchi, T. Arima, and Y. Tokura, Phys. Rev. Lett. 101, 097205 (2008).
6 Y. Tokunaga, N. Furukawa, H. Sakai, Y. Taguchi, T. Arima, and Y. Tokura, Nat. Mater. 8, 558 (2009).
7 J.-H. Lee, Y. K. Jeong, J. H. Park, M.-A. Oak, H. M. Jang, J. Y. Son, and J. F. Scott, Phys. Rev. Lett. 107, 117201 (2011).
8 A. V. Kimel, A. Kirilyuk, A. Tsvetkov, R. V. Pisarev, and T. Rasing, Nature 429, 850 (2004).
9 A. V. Kimel, A. Kirilyuk, P. A. Usachev, R. V. Pisarev, A. M. Balbashov, and T. Rasing, Nature 435, 655 (2005).
10 A. V. Kimel, B. A. Ivanov, R. V. Pisarev, P. A. Usachev, A. Kirilyuk, and T. Rasing, Nat. Phys. 5, 727 (2009).
11 Y. Tokunaga, Y. Taguchi, T. Arima, and Y. Tokura, Nat. Phys. 8, 838 (2012).
12 K. Yamaguchi, T. Kurihara, Y. Minami, M. Nakajima, and T. Suemoto, Phys. Rev. Lett. 110, 137204 (2013).
13 C.-Y. Kuo, Y. Drees, M. T. Fernández-Dı́az, L. Zhao, L. Vasylechko, D. Sheptyakov, A. M. T. Bell, T. W. Pi, H.-J. Lin,

M.-K. Wu, E. Pellegrin, S. M. Valvidares, Z. W. Li, P. Adler, A. Todorova, R. Küchler, A. Steppke, L. H. Tjeng, Z. Hu, and
A. C. Komarek, Phys. Rev. Lett. 113, 217203 (2014).
14 S. Geller, J. Chem. Phys. 24, 1236 (1956).
15 E. F. Bertaut, Magnetism, Vol. 3, (Academic Press, New York, 1963).
16 A. Berton and B. Sharon, J. Appl. Phys. 39, 1367 (1968).
17 G. Gorodetsky, B. Sharon, and S. Shtrikman, J. Appl. Phys. 39, 1371 (1968).
18 J. D. Cashion, A. H. Coole, D. M. Martin, and M. R. Wells, J. Phys. C 3, 1612 (1970).
19 Y. J. Choi, H. T. Yi, S. Lee, Q. Huang, V. Kiryukhin, and S.-W. Cheong, Phys. Rev. Lett. 100, 047601 (2008).
20 F. J. Morin, Phys. Rev. 78, 819 (1950).
21 L. A. Prelorendjo, C. E. Johnson, M. F. Thomas, and B. M. Wanklyn, J. Phys. C: Solid Stat. Phys. 13, 2567 (1980).
22 L. M. Holmes, L. G. Van Uitert, R. R. Hecker, and G. W. Hull, Phys. Rev. B 5, 138 (1972).
23 L. M. Holmes and L. G. Van Uitert, Phys. Rev. B 5, 147 (1972).
24 Z. Y. Zhao, X. Zhao, H. D. Zhou, F. B. Zhang, Q. J. Li, C. Fan, X. F. Sun, and X. G. Li, Phys. Rev. B 89, 224405 (2014).
25 J. C. Wang, J. J. Liu, J. M. Sheng, W. Luo, F. Ye, Z. Y. Zhao, X. F. Sun, S. A. Danilkin, G. C. Deng, and W. Bao, Phys. Rev.

B 93, 140403(R) (2016).


26 A. V. Sologubenko, K. Berggold, T. Lorenz, A. Rosch, E. Shimshoni, M. D. Phillips, and M. M. Turnbull, Phys. Rev. Lett.

98, 107201 (2007).


27 A. V. Sologubenko, T. Lorenz, J. A. Mydosh, A. Rosch, K. C. Shortsleeves, and M. M. Turnbull, Phys. Rev. Lett. 100,

137202 (2008).
28 X. F. Sun, W. Tao, X. M. Wang, and C. Fan, Phys. Rev. Lett. 102, 167202 (2009).
29 X. M. Wang, C. Fan, Z. Y. Zhao, W. Tao, X. G. Liu, W. P. Ke, X. Zhao, and X. F. Sun, Phys. Rev. B 82, 094405 (2010).
30 Z. Y. Zhao, X. M. Wang, B. Ni, Q. J. Li, C. Fan, W. P. Ke, W. Tao, L. M. Chen, X. Zhao, and X. F. Sun, Phys. Rev. B 83,

174518 (2011).
31 Z. Y. Zhao, X. G. Liu, Z. Z. He, X. M. Wang, C. Fan, W. P. Ke, Q. J. Li, L. M. Chen, X. Zhao, and X. F. Sun, Phys. Rev. B

85, 134412 (2012).


32 Z. Y. Zhao, B. Tong, X. Zhao, L. M. Chen, J. Shi, F. B. Zhang, J. D. Song, S. J. Li, J. C. Wu, H. S. Xu, X. G. Liu, and

X. F. Sun, Phys. Rev. B 91, 134420 (2015).


055806-7 Zhao et al. AIP Advances 7, 055806 (2017)

33 X. Zhao, J. C. Wu, Z. Y. Zhao, Z. Z. He, J. D. Song, J. Y. Zhao, X. G. Liu, X. F. Sun, and X. G. Li, Appl. Phys. Lett. 108,
242405 (2016).
34 Z. Y. Zhao, X. M. Wang, C. Fan, W. Tao, X. G. Liu, W. P. Ke, F. B. Zhang, X. Zhao, and X. F. Sun, Phys. Rev. B 83, 014414

(2011).
35 R. Berman, Thermal Conduction in Solids (Oxford University Press, Oxford, 1976).

You might also like