The coupled system of multi-degrees of freedom, such as charge, spin, orbital and lattice, has recently received much attention due to its potential to improve the magnetocaloric effect (MCE). The exotic inverse MCE was observed in rare-earth tetraborides of Ho1-xDyxB4 (x = 0.0, 0.5, and 1.0), associated with a strong coupling between magnetic dipoles and orbital quadrupoles in the strong spin-orbit coupling and geometric frustration. Here, the magnetism and magnetocaloric effects of ErB4 and TmB4 are investigated. It shows the maximum entropy changes of 11.4 J/kgK, and 12.6 J/kgK with the field of ΔH ≈ 40 kOe (Hc) in ErB4 and TmB4, respectively. The field- and temperature-dependence of the entropy change is found to be quite different from those of the conventional MCE. And the entropy change is also found to have strong correlation with the field induced meta-magnetic transition. Because the field induced transition is due to magnetic moment reorientation, which is strongly coupled with quadrupole moment, the abnormal MCE of ErB4 and TmB4 is attributed to the dipole-quadrupole interaction and magnetic frustration. Thus, it supports the fact that the strong coupling between quadrupole and magnetic dipole moments plays important role in the exotic inverse MCE in rare-earth tetraboride system.

The magnetocaloric cooling technology has been widely studied for the high-efficiency and environmental-friendly refrigeration system.1 For the discovery of new materials, which exhibit a large magnetocaloric effect, an interesting scenario have been proposed that large entropy change is expected in a system with an enormous ground state degeneracy such as geometric frustration.2,3 Rare-earth tetraborides, RB4 (R = rare-earth elements) has been known as compounds of geometrically frustrated magnetic system. The sub-lattice of rare-earth ions in the c-plane forms the Shastry-Sutherland lattice. It has been known that the coexistence of strong spin-orbit coupling and geometrical frustration is led to the interesting magnetic states in the rare-earth tetraborides.4–6 The frustration in DyB4 and HoB4 are caused not only by pure magnetic interactions but also by the quadrupole interactions. These compounds show common features of two successive magnetic transitions, exhibiting magnetic dipole ordering at T = TN1 and quadrupole ordering at T = TN2 (TN1 ≈ 20.5 K, and 7 K and TN2 ≈ 13.0 K, and 5.7 K for DyB4 and HoB4, respectively). The maximum value of positive entropy change is observed near T = TN2 with the values of 19.6 J/kg·K and 22.7 J/kg·K at the critical fields of ΔH ≈ 50 kOe and 25 kOe for DyB4 and HoB4, respectively. It is found that the exotic inverse MCE is due to the interplay of strong spin-orbit coupling and geometric frustration.7 To continue the study of series compounds of RB4 (R = rare-earth elements), we further investigate the magnetic and magnetocaloric properties of ErB4 and TmB4, which show fractional magnetization plateaus in isotherm and anisotropic antiferromagnetic ground state in Shastry-Sutherland lattice. In addition, the rotating magnetocaloric effect was recently reported, involving the interesting feature of the angular dependence of magnetic properties.8,9

Herein, we investigated the magnetocaloric effects of RB4 (R = Er and Tm) single crystals. The entropy change exhibits the maximum values of 11.4 J/kgK and 12.6 J/kgK with field of ΔH = 40 kOe for Hc in ErB4 and TmB4, respectively. It is found that the temperature- and field-dependence of the entropy change is quite different from those of conventional magnetocaloric effects.

Single crystals of RB4 (R = Er and Tm) are prepared with a high-temperature metal flux method. A stoichiometric mixture of rare earth metals (≥ 99.9%, China Rare Metal Material Co., LTD.) and boron pieces (99.9%, RND Korea) are placed in an alumina crucible (99.8%, Samhwa Ceramic Company) together with the Al pellets (99.999%, RND Korea) with a mass ratio of samples to Al = 1:60. The samples are placed in a heated tube furnace with an MoSi2 heating element. The heat treatment is followed by heating up to 1650 °C under a high-purity argon atmosphere after dehydration and cooled slowly at a rate of 4.8 °C per hour to 650 °C. The single crystals are separated from the flux by dissolving the excess Al in NaOH.

The crystal structures of the samples are characterized using X-ray diffraction measurements (XRD; Rigaku D/MAX-2500 with a Cu target) at room temperature. The lattice parameters are determined from LeBail refinements using FULLPROF software. The XRD patterns show a single phase of ErB4 and TmB4 without any observable impurity phases. The crystal structures are in good agreement with the tetragonal symmetry of the ThB4-type structure and space group P4/mbm (No. 127). The refined lattice parameters are a = 7.0647(6) Å and c = 3.9936(6) Å for ErB4 and a = 7.067(5) Å and c = 3.9820(0) Å for TmB4. The refinement values are in a good agreement with previous reports.10 The temperature- and field-dependent magnetizations are measured using a superconducting quantum interference device magnetometer (SQUID; Quantum Design MPMS XL).

The temperature-dependence of magnetization of ErB4 under magnetic field of H = 10 kOe for Hc and Hc is plotted in Fig. 1. It shows the strongly anisotropic behaviour of magnetization. Whereas the broad maximum is observed around T = 30 K for Hc due to the Schottky anomaly associated with crystalline electric-field (CEF) level split,11 the antiferromagnetic transition is observed at TN = 15.5 K for Hc. The inset shows that the paramagnetic phase for both Hc and Hc follows the Curies-Weiss law, MTH=CTθ, where C=N0μeff2/3kB, N0 is Avogadro’s number, kB is the Boltzmann constant, and μeff is the effective magnetic moment. The effective magnetic moments are determined to be 9.27μB, and 9.50μB, where μB is the Bohr magneton, and the Weiss temperatures, θ, are also found to be +11.24 K and -23.26 K for Hc and Hc, respectively. The μeff values are close to the theoretical value of Hund’s rule for the ground state of the isolated Er3+ ions (μeff = 9.59 μB).

FIG. 1.

Temperature-dependence of magnetization with an applied magnetic field, H = 10 kOe, for Hc and Hc of ErB4.

FIG. 1.

Temperature-dependence of magnetization with an applied magnetic field, H = 10 kOe, for Hc and Hc of ErB4.

Close modal

Figure 2 shows the isothermal magnetization data of ErB4 at various temperatures for both Hc and Hc. The field-induced meta-magnetic transition and magnetization plateau are noticeably observed at T = 5 K in a range of 20 kOe ≤ H ≤ 40 kOe for Hc and there is no magnetic hysteresis in the magnetization at T = 2 K, as shown in the inset of Fig. 2(a). The magnetic moment of Er3+ ion at H = 50 kOe is found to be ≈ 5.1 μB. On the other hand, the isothermal magnetizations for Hc at various temperatures show the typical antiferromagnetic behaviour and the inset shows the data at T = 2 K. The magnetic moment at T = 2 K and H = 50 kOe is found to be 0.3 μB, indicating that the orientation of Er3+ ion moment is aligned along the c-axis.

FIG. 2.

Field-dependence of the isothermal magnetization at various temperatures in a range of 2 K ≤ T ≤ 50 K for (a) Hc and (b) Hc of ErB4.

FIG. 2.

Field-dependence of the isothermal magnetization at various temperatures in a range of 2 K ≤ T ≤ 50 K for (a) Hc and (b) Hc of ErB4.

Close modal

The magnetic entropy change, ΔSM, can be estimated from the Maxwell equation in the approximated form

where Mi+1 and Mi are the experimentally measured values at temperatures Ti+1 and Ti under applied magnetic field, H, respectively, in the interval of magnetic field, ΔHi. Figure 3 shows the temperature-dependence of the magnetic entropy change of ErB4, which is calculated from isothermal magnetization data for Hc and Hc. The positive entropy change, ΔSM, is observed for Hc at T ≈ 13 K, which is significantly lower than TN (= 15.5 K), as shown in Fig. 3(a). The maximum value of entropy change is observed to be +11.4 J/kgK and almost constant with the fields of ΔH = 30, 40, and 50 kOe. The negative entropy change at higher temperatures is likely due to the crystal field degeneracy and follows field-dependence of typical magneto-caloric effect. Figure 3(b) shows also a positive entropy change for Hc with the maximum value of +11.13 J/kgK at T = TN with ΔH=50 kOe.

FIG. 3.

The magnetic entropy change under various magnetic fields of ΔH = 10, 15, 20, 30, 40 and 50 kOe, for (a) Hc and (b) Hc of ErB4.

FIG. 3.

The magnetic entropy change under various magnetic fields of ΔH = 10, 15, 20, 30, 40 and 50 kOe, for (a) Hc and (b) Hc of ErB4.

Close modal

Figure 4 shows the temperature-dependence of magnetization of TmB4 under magnetic field of H = 10 kOe for Hc and Hc. It also shows the strongly anisotropic behaviour of magnetization as like ErB4. Whereas the broad maximum is found near T = 90 K for Hc due to the influence of crystal electric field,12 there are two successive magnetic transition at TN1 = 11.9 K and TN2 = 9.2 K for Hc. It is already known that there is incommensurate-incommensurate magnetic transition with magnetic fluctuation below T = TN1 and that the fluctuation disappears and the commensurate magnetic structure is developed below T = TN2.13 The inset of Fig. 4 shows that the paramagnetic phase for both Hc and Hc follows the Curies-Weiss law. The effective magnetic moment is determined to be 7.49μB and 7.70μB and the Weiss temperature to be + 40.7 K and - 63.5 K for Hc and Hc, respectively. The μeff values are close to the theoretical value of Hund’s rule for the ground state of the isolated Tm3+ ions (μeff = 7.55 μB).

FIG. 4.

Temperature-dependence of magnetization with an applied magnetic field, H = 10 kOe, for Hc and Hc of TmB4.

FIG. 4.

Temperature-dependence of magnetization with an applied magnetic field, H = 10 kOe, for Hc and Hc of TmB4.

Close modal

Figure 5 shows the isothermal magnetization data of TmB4 at various temperatures for Hc and Hc. The meta-magnetic transitions and magnetization plateaus are observed at T = 5 K in a range of 15 kOe ≤ H ≤ 40 kOe for Hc. The inset shows the similar isothermal magnetization at T = 2 K and the hysteresis at the transition near H ≈ 16 kOe, which is not found in ErB4, may be an indication of lattice distortion due to electron-lattice coupling. The magnetic moment of the Tm3+ ion at T = 2 K and H = 50 kOe is 7.5 μB, which is close to the full saturation moment of Tm3+ ion (7.55 μB). The isothermal magnetization for Hc in Fig. 5(b) shows typical antiferromagnetic behaviour due to strongly anisotropy, as like ErB4. Comparison of the isothermal data of ErB4 and TmB4 for Hc indicates that the meta-magnetic transition is a common feature for two compounds and that ErB4 would have the second transition at higher fields of H > 50 kOe.

FIG. 5.

Field-dependence of the isothermal magnetization at various temperatures in a range of 2 K ≤ T ≤ 100 K for (a) Hc and (b) Hc of TmB4.

FIG. 5.

Field-dependence of the isothermal magnetization at various temperatures in a range of 2 K ≤ T ≤ 100 K for (a) Hc and (b) Hc of TmB4.

Close modal

Figure 6 shows the temperature-dependence of the magnetic entropy change for TmB4, which is calculated from isothermal magnetization data for both Hc and Hc. The field-dependent entropy change is observed to be maximum with a value of + 12.6 J/kgK under the field of ΔH = 40 kOe at T ≈ 8 K, which is close to T = TN2. With further increase of magnetic field, the entropy change decreases to a value of + 7.7 J/kg·K with the field of ΔH = 50 kOe at T ≈ 7 K. The negative entropy change is observed with the maximum value of ΔSM = - 10.1 J/kg·K with ΔH = 50 kOe at TTN2, which shows the conventional magnetocaloric behaviour. On the other hand, the magnetic entropy change for Hc is negligibly small, as shown in Fig. 6(b).

FIG. 6.

The magnetic entropy change under various magnetic fields of ΔH = 10, 15, 20, 30, 40 and 50 kOe, for (a) Hc and (b) Hc of TmB4.

FIG. 6.

The magnetic entropy change under various magnetic fields of ΔH = 10, 15, 20, 30, 40 and 50 kOe, for (a) Hc and (b) Hc of TmB4.

Close modal

The entropy changes of ErB4 and TmB4 for Hc show common interesting peculiar features in terms of field-dependence and correlation with magnetic transition. The maximum entropy change occurs at significantly lower temperatures of T ≈ 12.5 K for ErB4, while TN = 15.5 K, and T ≈ 8 K for TmB4, while TN = 12 K. In addition, the maximum value of ErB4 is found with the field of ΔH = 30 kOe and kept to be almost constant even with further increase of field. And that of TmB4 is found with the field of ΔH = 40 kOe and decreases with the field of ΔH = 50 kOe. This abnormal field-dependence of entropy change is quite different from that of conventional magneto-caloric effect. So, the positive large entropy changes in ErB4 (Fig. 3(a)) and TmB4 (Fig. 6(a)) is not likely to be related with the antiferromagnetic transitions. Indeed, tetraboride compounds (RB4; R = rare-earth elements) are known to have hidden quadrupole ordering and spin reorientation below T = TN due to strong spin-orbit coupling.14–17 The meta-magnetic transitions in isothermal magnetization for Hc are the manifestation of the magnetic coupling with quadrupole ordering. Thus, the abnormal entropy changes of ErB4 and TmB4 is likely the results of degeneracy release in the field-induced meta-magnetic transition. The peculiar field dependence of entropy change can be understood in terms of the strength of spin-orbit coupling. Similar entropy change in Ho1-xDyxB4 system and Landau free energy model for that are recently reported.7 

Structural distortion in ErB4 was found in the measurements of high resolution X-ray diffraction and specific heat.18,19 Although the origin of the distortion is not fully understood at present time, it is conjectured that the contribution of the distortion to entropy change would be negligible because the temperature- and field-dependence of entropy change is of second-order nature. In addition, the distortion temperature is quite higher than the temperatures for maximum entropy change.

In this study, the anisotropic and inverse magnetocaloric effects are investigated in rare-earth tetraboride system. ErB4 and TmB4 are successfully synthesized as the single crystals without any significant impurity phases. These compounds show common features of strongly anisotropic magnetism and field-dependent entropy change. The positive entropy change, i.e., inverse magnetocaloric effect, is observed for Hc with the maximum values of ΔSm= +11.4 J/kgK, and +12.6 J/kgK at ΔH = 30 kOe and 40 kOe for ErB4 and TmB4, respectively. With further increase of field, no more increase of ΔSM for ErB4 and even decrease of ΔSm for TmB4 are observed. The abnormal entropy change is also found to come from magnetic degeneracy release in the field induced meta-magnetic transition. Because the meta-magnetic transition is influenced by competition between Zeeman effects and frustration in strong spin-orbit coupling regime, the peculiar characteristics of MCE in ErB4 and TmB4 is believed due to the multipolar degrees of freedom which is lain in a geometrically frustrated lattice.

This work was supported by the National Research Foundation of Korea (NRF), funded by the Ministry of Science, ICT & Future Planning (No. NRF-2015M3A9B8032703 and No. NRF- 2017R1A2B2008538).

1.
K. A.
Gschneidner
, Jr.
and
V. K.
Pecharsky
,
Int. J. Refrigeration
31
,
945
961
(
2008
).
2.
M. E.
Zhitomirsky
,
Phy. Rev. B
67
,
104421
(
2003
).
3.
Y.
Taguchi
,
H.
Skai
, and
D.
Choudhury
,
Adv. Mater.
29
,
1606144
(
2017
).
4.
Z.
Fisk
,
M. B.
Maple
,
D. C.
Johnston
, and
L. D.
Woolf
,
Solid State Comm
39
,
1189
1192
(
1981
).
5.
D.
Okuyama
,
T.
Matumura
,
H.
Nakao
, and
Y.
Murakami
,
J. Phys. Soc. Jpn.
74
,
2434
(
2005
).
6.
T.
Matsumura
,
D.
Okuyama
,
T.
Mouri
, and
Y.
Murakami
,
J. Phys. Soc. Jpn.
80
,
074701
(
2011
).
7.
M. S.
Song
,
K. K.
Cho
,
B. Y.
Kang
,
S. B.
Lee
, and
B. K.
Cho
,
Sci. Rep.
10
,
803
(
2020
).
8.
M.
Orendáč
,
S.
Gabáni
,
E.
Gažo
,
G.
Pristáš
,
N.
Shitsevalova
,
K.
Siemensmeyer
, and
K.
Flachbart
,
Sci. Rep.
8
,
10933
(
2018
).
9.
M.
Orendáč
,
S.
Gabáni
,
E.
Gažo
,
G.
Pristáš
,
N.
Shitsevalova
,
K.
Siemensmeyer
, and
K.
Flachbart
,
J. Mat. Mag. Mater.
482
,
186
191
(
2019
).
10.
J.
Etourneau
and
P.
Hagenmuller
,
Philos. Mag. B
52
,
589
(
1985
).
11.
V. V.
Novikov
,
A. V.
Morozov
,
A. V.
Matovnikov
,
N. V.
Mitroshenkov
,
D. V.
Avdashchenko
,
S. V.
Kuznetosov
,
B. I.
Kornev
,
O. A.
Marakhina
,
V. V.
Novikova
, and
E. O.
Bordacheva
,
J. Alloys. Compd.
581
,
431
(
2013
).
12.
V. V.
Novikov
,
A. V.
Matovnikov
,
N. V.
Mitroshenkov
, and
A. K.
Tolstosheev
,
J. Alloys. Compd.
666
,
98
(
2016
).
13.
S.
Michimura
,
A.
Shigekawa
,
F.
Iga
,
T.
Takabatake
, and
K.
Ohoyama
,
J. Phys. Soc. Jpn.
78
,
024707
(
2009
).
14.
R.
Watanuki
,
T.
Kobayashi
,
R.
Noguchi
, and
J.
Kazuya Suzuki
,
Physics: Conf. Ser.
150
,
042229
(
2009
).
15.
S.
Yoshii
,
T.
Yamamoto
,
M.
Hagiwara
,
S.
Michimura
,
A.
Shigekawa
,
F.
Iga
,
T.
Takabatate
, and
K.
Kindo
,
Phys. Rev. Lett.
101
,
087202
(
2008
).
16.
H.
Sim
,
S.
Lee
,
K.-P.
Hong
,
J.
Jeong
,
J. R.
Zhang
,
T.
Kamiyama
,
D. T.
Adroja
,
C. A.
Murray
,
S. P.
Thompson
,
F.
Iga
,
S.
Ji
,
D.
Khomskii
, and
J.-G.
Park
,
Phys. Rev. B
94
,
195128
(
2016
).
17.
D.
Okuyama
,
T.
Matsumura
,
T.
Mouri
,
N.
Ishikawa
,
K.
Ohoyama
,
H.
Hiraka
,
H.
Nakao
,
K.
Iwasa
, and
Y.
Murakami
,
J. Phys. Soc. Jpn.
77
,
044709
(
2008
).
18.
Z.
Heiba
,
W.
Schäfer
,
E.
Jansen
, and
G.
Will
,
J. Phys. Chem. Solids
47
,
651
658
(
1986
).
19.
S.
Michimura
,
A.
Shigekawa
,
F.
Iga
,
M.
Sera
,
T.
Takabatake
,
A.
Kikkawa
,
Y.
Tanaka
,
K.
Katsumata
, and
J.
Mag
,
Magn. Mater.
310
,
e446
e447
(
2007
).