Doping-controlled spin reorientation transition and spin switching of Tm1−xPrxFeO3 single crystals

Jingxiu Liu , Zhijie Yang , Zeru Liu , Chenfei Shi , Haohuan Peng , Baojuan Kang , Rongrong Jia , Xiaoxuan Ma , Shixun Cao

Front. Phys. ›› 2026, Vol. 21 ›› Issue (3) : 035202

PDF (3203KB)
Front. Phys. ›› 2026, Vol. 21 ›› Issue (3) : 035202 DOI: 10.15302/frontphys.2026.035202
RESEARCH ARTICLE

Doping-controlled spin reorientation transition and spin switching of Tm1−xPrxFeO3 single crystals

Author information +
History +
PDF (3203KB)

Abstract

Rare-earth orthoferrites (RFeO3), which are canted antiferromagnets exhibiting weak ferromagnetism, are potential materials for spintronic devices. The temperature of spin reorientation transition (SRT) in Tm1xPrxFeO3 (x = 0, 0.15, 0.25, 0.5) single crystals shifts to a lower region with the increase of x, attributing to the exchange energy of Pr3+Fe3+ being less than that of Tm3+Fe3+. Both type-I of spin switching (SSW-I) and the type-II of spin switching (SSW-II) are observed along the a-axis when x = 0.15 and 0.2. The changes of magnetic interaction are analyzed through structural distortion and transforming Curi−Weiss fitting. A rare SSW-II phenomenon is observed along the c-axis during both the cooling and warming processes when x = 0.15, 0.25, 0.5. In addition, a weak ferromagnetic moment (WFM) component is observed despite the antiferromagnetic behavior at high temperature in Tm0.5Pr0.5FeO3 single crystal. The WFM decreases gradually with the increase in temperature, indicating the occurrence of SRT. And the field-induced spin switching (SSW) is observed along the a-axis at 70 K when x = 0.5. The magnetic properties of Pr3+ doped TmFeO3 single crystals are obviously different from TmFeO3 and PrFeO3 single crystals, indicating the complex magnetic interaction among Pr3+, Tm3+ and Fe3+. The complex and abundant magnetic phenomena in Tm1xPrxFeO3 single crystals offer significant potential for studying altermagnet and magneto-optical coupling, and are expected to become a new generation of spintronic devices.

Graphical abstract

Keywords

spin reorientation transition / spin switching / Curie−Weiss law / rare-earth orthoferrites / spintronic

Cite this article

Download citation ▾
Jingxiu Liu, Zhijie Yang, Zeru Liu, Chenfei Shi, Haohuan Peng, Baojuan Kang, Rongrong Jia, Xiaoxuan Ma, Shixun Cao. Doping-controlled spin reorientation transition and spin switching of Tm1−xPrxFeO3 single crystals. Front. Phys., 2026, 21(3): 035202 DOI:10.15302/frontphys.2026.035202

登录浏览全文

4963

注册一个新账户 忘记密码

1 Introduction

Due to the cancellation of adjacent antiparallel magnetic moments, the antiferromagnetic material has the advantages of no stray field and the intrinsic frequency up to terahertz order etc., making them ideal candidates for high-density and high-speed storage. As a special antiferromagnet, rare-earth orthoferrites (RFeO3, R = Y, La−Lu) have been extensively studied in quantum optics [1, 2], spin transport [3] and spin mechanics [4, 5], etc., and still has the potential to explore future applications, which is expected to become the next generation of spintronic devices. Recently, the discovery of collinear magnetic structure known as altermagnets [69] has opened the way to explore the third kind of magnetic order, which combines the advantages of ferromagnetic and antiferromagnetic, providing a promising platform for high-speed information processing and terahertz communication, showing a unique transport phenomenon. RFeO3, as a canted antiferromagnet, also possesses both antiferromagnetic properties and weak ferromagnetic characteristics, which still holds great research potential.

The structure of RFeO3 is derived from the distortion of ideal perovskite structure. Fe3+ and six O2− form FeO6 octahedra, and R3+ is located in the voids between FeO6 octahedra. Due to the large radius of R3+, in order to maintain the structural stability, the FeO6 octahedron will undergo a certain degree of twisting and tilting, and R3+ may leave their ideal positions, thereby causing structural distortion and changing the cubic crystal system space group from Pm3m to Pbnm [10]. The R3+ and the Fe3+−magnetic sublattice [11] of RFeO3 gives rise to three magnetic interactions: R3+R3+ interaction, R3+−Fe3+ interaction, and Fe3+−Fe3+ interaction. Since the far distance between two R3+ or Fe3+, they need to achieve indirect super-exchange interaction through O2− [12]. According to the Goodenough−Kanamori−Anderson rule [1315], the strength of the super-exchange interaction is maximized when Fe3+ and O2− bond angle is 180° [16], and minimized when the angle approaches 90°. Variations in Fe−OI bond length and Fe−OI−Fe bond angle will influence the overlap of Fe3+ and O2− electron clouds, thereby affecting the strength of the super-exchange interaction. Meanwhile, the Dzyaloshinskii−Moriya interaction [17, 18] results in an inclination angle of the magnetic moments of Fe3+, causing the magnetic moments become canted antiferromagnetic with weak ferromagnetic moment (WFM) [19]. The distortions of the lattice and the complex magnetic interactions in RFeO3 have triggered a variety of complex magnetic phenomena, such as ultrafast photomagnetic effects [5], spin reorientation transitions (SRT) [2024], spin switching (SSW) [2528], and ferroelectricity [29]. In our study, we mainly focus on the SRT and the SSW.

The first Neel temperature (TN1) of the RFeO3 is high, typically ranging from 600 K to 700 K. In the high-temperature region below TN1, R3+ is paramagnetic state (above second Neel temperature (TN2 < 10 K)), and the Fe3+−O2−−Fe3+ interaction is the strongest. The WFM arising from the spin-canted state of Fe3+ aligns along the c-axis of the crystal, while the antiferromagnetic moment (AFM) aligns along the a-axis, corresponding to Γ4 (Gx, Ay, Fz) spin configuration. With the decrease of temperature, the R3+−Fe3+ interaction enhanced, and the AFM undergoes a flip, causing the spin configuration to change and the occurrence of SRT. Since RFeO3 only has three spin configurations [10], the temperature-induced SRT can be mainly classified into two types: (i) Γ4 (Gx, Ay, Fz) → Γ2 (Fx, Cy, Gz), as observed in PrFeO3 [21], SmFeO3 [22], TmFeO3 [27], etc.; (ii) Γ4 (Gx, Ay, Fz) → Γ1 (Ax, Gy, Cz), such as CeFeO3 [20] and DyFeO3 [24]. When the temperature is above the R3+ magnetic ordering temperature (TN2), the magnetic moments of R3+ and Fe3+ are arranged in a parallel or antiparallel coupling state. In some RFeO3, at specific temperature or external magnetic field (H), due to the flipping of the magnetic moments of R3+ or the simultaneous flipping of the magnetic moments of R3+ and Fe3+, the magnetization undergoes a sudden change, and this change is called SSW. If the magnetic moments of R3+ and Fe3+ flip simultaneously, resulting in a reversal of the magnetization sign, this is type-I of SSW (SSW-I), such as NdFeO3 [25], ErFeO3 [26] and Nd0.8Pr0.2FeO3 [28]. If only R3+ magnetic moment reverses, it is type-II of SSW (SSW-II), as seen in PrFeO3 [21], TmFeO3 [27] and Tm0.8Sm0.2FeO3 [23].

PrFeO3 single crystal exhibits unique SRT properties. Due to the complex coupling effect between the 4f electrons of Pr3+ and the 3d electrons of Fe3+, SRT is highly sensitive to the H and gradually gets suppressed as the H increases. SRT can be observed at 6.5 K only at a small external field (H = 5 Oe) [21]. Furthermore, in the undoped RFeO3 system, only TmFeO3 [27] and PrFeO3 [21] single crystals have observed SSW-II in both field-cooled cooling (FCC) and field-cooled warming modes along the c-axis. Besides the SSW-II observed along the c-axis, SSW-II is also observed along the a-axis in TmFeO3 [27] single crystal in the FCC mode, and the temperature range of SRT (TSRT) is 82 K−95 K. The peculiar interaction among Tm3+, Pr3+ and Fe3+ leads to more interesting phenomena when Tm3+ and Pr3+ are doped with other R3+.

In our study, high-quality Tm1−xPrxFeO3 (x = 0, 0.15, 0.25, 0.5) single crystals are grown by the optical floating zone method, and the zero-field-cooled warming (ZFC), FCC curves and the isothermal magnetization (MH) curves are measured. Unlike PrFeO3 single crystal which are only found at low temperature and low field, with the increase of x, TSRT shifts towards lower temperature, moving from 82 K−95 K to 72 K−92 K. When x = 0.15 and 0.25, SSW-I is observed along the a-axis, which is absent in TmFeO3 and PrFeO3 single crystals. However, when the doping ratio reached 50%, SSW-I disappears. We discuss the changes in magnetic interactions based on the structural variations, and then analyze the reasons for the appearance of SSW-I. Furthermore, the transforming Curie−Weiss law is employed to fit the data before and after the SSW-I in the Tm0.85Pr0.15FeO3 single crystal at H = 120 Oe in the FCC mode, and the data after SRT in the Tm0.5Pr0.5FeO3 single crystal at H = 100 Oe is also fitted. The magnetic moment of Fe3+ and the internal field arising from the magnetic moment tilt of Fe3+ at the rare earth sites are calculated. The changes in magnetic interactions during the SSW-I transition are further analyzed. Apart from a-axis, SSW-II is also observed along the c-axis, and triggering temperature of both SSW-I and SSW-II can be regulated by H. Finally, we calculated the weak ferromagnetic saturation magnetization of MH curve in the Tm0.5Pr0.5FeO3 single crystal along the a-axis. The WFM part gradually decreases while the antiferromagnetic contribution gradually increases, also indicating the occurrence of SRT.

2 Experimental method

The medicinal powder is weighed according to the ratio specified in Eq. (1) and thoroughly mixed,

1x2Tm2O3+x6Pr6O11+12Fe2O3Tm1xPrxFeO3+x6O2.

x represents the content of Pr3+. RFeO3 polycrystalline powders can be synthesized though the solid-state reaction method followed by multiple sintering steps. Large-sized single crystals can be grown with the diameter of 5 mm and the length exceed 50 mm using the optical float zone method, as shown in Fig.1(a). The single crystals are characterized by X-ray diffraction and structural refinement. The weighted profile R-factors (Rwp = 10.9%, 3.88%, 4.21%, and 6.38%) obtained from the refinement [Fig.1(a)] confirm the high quality of the samples. In addition, the Fe−OI bond length and Fe−OI−Fe bond angle [Fig.1(b)] are obtained through refinement to analyze the changes in the super-exchange interaction. The single crystal is cut into test cube using a Laue camera with a cutting machine. The Laue back-reflection patterns confirm the accurate orientation and high quality of the cut crystals [Fig.1(c)]. Subsequently, the ZFC, FCC and MH curves are tested using the magnetic property measurement system (MPMS 3) from Quantum Design. Selected magnetic measurement results are analyzed using transforming Curie–Weiss fitting to investigate the complex magnetic interactions in RFeO3.

3 Results and discussion

The magnetic characteristics of Tm1−xPrxFeO3 (x = 0, 0.15, 0.25, 0.5) single crystals along the three crystal axes are shown in Fig.2. The different magnetic behaviors along the three axes highlight the presence of magnetic anisotropy in these single crystals. In the high-temperature region (TN2 < T < TN1), the antisymmetric exchange interaction between R3+ and Fe3+ is relatively weak, allowing the Fe3+−Fe3+ interaction to dominate (in the following text, R3+ = Tm3+ and Pr3+). With the decreasing temperature, the R3+−Fe3+ interaction gradually strengthens. and becomes sufficiently strong to overcome the magnetic anisotropy energy. The AFM reorients from the Gx to Gz [30], and the WFM shifts from the Fz to Fx, resulting in the SRT of Γ4→Γ2. The Pr3+ doping does not alter the type of SRT, which is the same as TmFeO3 single crystal [27]. However, with the increase of x, TSRT shifts to lower temperature, moving from 82 K−95 K to 72 K−92 K. According to the phase diagram of the RFeO3 summarized by Li et al. [21], the TSRT of PrFeO3 single crystal is much lower than TmFeO3, suggesting that the Pr3+−Fe3+ exchange interaction energy is weaker than Tm3+−Fe3+. As a result, PrFeO3 can overcome the anisotropic Fe3+−Fe3+ interaction at a lower temperature. Therefore, the increase of Pr3+ content gradually diluted the Tm3+−Fe3+ exchange interaction energy [31], resulting in the decrease of TSRT. In the TmFeO3 single crystal, a magnetic compensation phenomenon is observed at 15 K [as shown in Fig.2(a)]. As the temperature increases to 15 K, the magnetic moment of Tm3+ decreases while the magnetic moment of Fe3+ gradually increases, resulting in a reduction of the magnetization. When the temperature reaches 15 K, the magnetic moment of partial Tm3+, under the induction of the Fe3+ sublattice, flips and couples parallel to the magnetic moment of Fe3+, causing the increase of magnetization. This negative magnetization trend may lead to the emergence of SSW-I. This phenomenon is also observed in TmFe0.8Mn0.2O3 single crystal [23]. As shown in Fig.2(b) and (c), when x = 0.15 and x = 0.25, the SSW-I and SSW-II are observed simultaneously along the a-axis in the FCC mode, which is not present in the parent phase TmFeO3 single crystal [27]. In addition, SSW-I in the ZFC and FCC modes exhibit a butterfly-shape [26]. concerning T0 symmetry.

In the ZFC mode, SSW-I is observed along the a-axis when x = 0.15 and x = 0.25, resulting in a butterfly-shaped SSW-I Compared with TmFeO3 and PrFeO3 single crystals (the PrFeO3 data are from the standard CIF file), Tm0.85Pr0.15FeO3 exhibits a longer Fe−OI bond length and a greater deviation of the Fe−OI−Fe bond angle from 180° [Fig.1(b)], leading to the weakening of the Fe3+−Fe3+ interaction. As shown in Fig.3(a), in the ZFC mode, the isotropic R3+−Fe3+ interaction results in an intrinsic parallel or antiparallel alignment between the R3+ and Fe3+ magnetic moments. Meanwhile, before SSW-I, the molecular field of Fe3+ induces the polarization of the R3+ magnetic moment, resulting in a WFM parallel to H [16]. Therefore, the coupling mode is ↑↓↓ (the arrows from left to right represent the magnetic moment directions of Tm3+, Pr3+, and Fe3+ respectively). Due to weakening of the Fe3+−Fe3+ interaction, the R3+−Fe3+ interaction can be further enhanced with the increase of temperature. The R3+ magnetic moment further decreases, resulting in zero magnetization (T0) and negative magnetization. However, with the temperature further rises, the Fe3+−Fe3+ interaction will be strengthened, and R3+−Fe3+ interaction will be unable to resist the influence of H, resulting in the R3+ and Fe3+ magnetic moments to flip simultaneously. The coupling mode change from ↑↓↓ to ↓↑↑. For example [Fig.3(a)], in the ZFC mode, when H = 180 Oe, the magnetization along the a-axis in the Tm0.85Pr0.15FeO3 single crystal changes from −0.36 emu/g to 0.44 emu/g at TSI-2 = 78 K. When x = 0.25, butterfly-shaped SSW-I is also observed. Although the distortion degree is smaller compared to TmFeO3 single crystal, it is larger than PrFeO3 single crystal [Fig.1(b)], thus explains the occurrence of butterfly-shaped SSW-I. The variations in x = 0.15 and x = 0.25 single crystals reflect the complex magnetic interactions between R3+ and Fe3+. It is worthy to further exploration through various methods. When x = 0.5, SSW-I disappears. Compared with other single crystals, the Fe−OI−Fe bond angle becomes 163.93(5)° [Fig.1(b)], the distortion degree of the octahedron is smaller, and the Fe3+−O2−−Fe3+ super-exchange interaction is enhanced, inhibiting the generation of SSW-I. In the FCC mode, it also shows SSW-I. When SSW-I occurs, the magnetic moment reversal of R3+ and Fe3+ is opposite to the ZFC mode. Moreover, it can be clearly seen that TSI-1/ TSI-2 has a positive/negative correlation with H. In the FCC mode, TSI-1 becomes larger with the increase of H. Tm3+ is in a paramagnetic state above TN2, and the magnetic moment direction tends to be parallel to H. Therefore, with the increase of H, it will flip at higher the temperature. In the ZFC mode, a larger H also leads to the earlier flipping of the Tm3+ magnetic moment.

In addition to SSW-I, SSW-II is also observed in the FCC mode along the a-axis when x = 0.15, 0.25, 0.5. The Pr3+ magnetic moment is antiparallel to Fe3+ below 13 K. In contrast, the Tm3+ magnetic moment is antiparallel to Fe3+ below 80 K. Therefore, the occurrence of SSW-II is attributed to the flipping of the Pr3+ magnetic moment under the influence of Fe3+ and Tm3+ magnetic moments. When x = 0.25, SSW-II exhibits multi-step characteristics. In Fig.3(d), at 40 Oe, the first magnetization jump occurred at 30 K. At this point, Pr3+ magnetic moment partially flipped and the system was not in a steady state. As the temperature was further decreased, Pr3+ magnetic moment continued to flip, causing fluctuations in the magnetization, until the system reached a steady state. In addition, with the increase of H, TSII-1 gradually moves towards low temperature and is suppressed at H = 120 Oe (x = 0.15) and 100 Oe (x = 0.25). When x = 0.5, the situation of SSW-II is similar (as shown in Fig. S1). The triggering temperatures of both SSW-I and SSW-II can be significantly regulated by H and are expected to be applied in spin switching devices.

Structural changes is one of the many reasons contributing to the complex magnetic changes. The transforming Curi−Weiss law is applied to verify the magnetic interaction further. Eq. (2) obtains the experimental Curie constant [32-34],

χ=χ0+CexpTθ.

χ0 is a parameter independent of temperature, Cexp is the experimental Curie constant and θ is the Weiss temperature. The theoretical Curie constant is obtained through Eq. (3) [32],

C=Nμeff23kB.

C, N, kB are the theoretical Curie constant, Avogadro constant and Boltzmann constant respectively. The theoretical effective magnetic moment (μeff) is given by Eq. (4) [32],

μeff=(1x)μTm2+xμPr2+μFe2.

μTm, μPr, μFe is Tm3+, Pr3+ and Fe3+ theoretical effective magnetic moment, respectively. The difference between the Cexp and C is negligible (Fig.4), further confirming the high quality of the single crystals and the reliability of the calculated results. Subsequently, the magnetic interactions in the FCC mode below TSRT are analyzed according to Eq. (5) [35, 36],

M=MFe+Cexp(Hi+Ha)Tθ.

MFe is Fe3+ canted magnetic moment, Hi is internal field generated by Fe3+ canted magnetic moment at R site and Ha is applied magnetic field. The results of x = 0.15 and x = 0.5 are shown in Fig.4. The FCC curves of Tm0.85Pr0.15FeO3 single crystal along the a-axis at 120 Oe are fitted before and after SSW-I. Before SSW-I [the pink curve of Fitting1 in Fig.4(a)], MFe1 = 498.90 emu/mol, Hi1 = −3597.63 Oe, θ1 = −15.44 K. MFe1 > 0 indicates that the magnetic moment of Fe3+ is aligned with H [37]. Hi1 < 0 suggests that the R3+ total magnetic moment is opposite to H. Meanwhile, |Hi1| > Ha indicates that the R3+ total magnetic moment is antiparallel to Fe3+, and θ1 < 0 reveals the antiferromagnetic interaction between R3+ and Fe3+. Based on the phase diagram [21], the magnetic moments of Tm3+ and Fe3+ are antiparallelly coupled, while those of Pr3+ and Fe3+ are parallelly coupled. These correspond to the coupling mode of ↓↑↑. With the decrease of temperature, the magnetization decreases and the negative magnetization strengthens, corresponding to |Hi1+Ha| > H i2+Ha. |Hi1+Ha| > |H i2+Ha| induces to the enhancement of the R3+−Fe3+ interaction [37], leading to the negative magnetization. After completing SSW-I, MFe2 = −458.15 emu/mol, Hi2 = 3417.17 Oe. The Fe3+ magnetic moment is opposite to H, while the R3+ total magnetic moment aligns with H. leading to the coupling mode of ↑↓↓. In the Tm0.5Pr0.5FeO3 single crystal [Fig.4(b)], after SRT, fitting results are MFe = −461.58 emu/mol, Hi = 5626.34 Oe, θ = −12.41 K. The Fe3+ magnetic moment remains antiparallel to H, while the R3+ total magnetic moment keeps aligning with H. Therefore, SSW-I is not observed when x = 0.5.

In addition to the SSW observed along the a-axis, SSW-II is also observed along the c-axis. As shown in Fig.5(a)−(c), SSW-II are observed in the FCC mode along the c-axis for x = 0.15, 0.25, and 0.5. In the high-temperature region, the R3+ total magnetic moment is intrinsically parallel to the Fe3+ magnetic moment corresponding to the coupling mode of ↑↑↑. With the temperature decrease, the R3+ effective magnetic moment increases, leading to an increase in the total magnetization. As the temperature further decreases, the Tm3+ magnetic moment tends to be antiparallel to the Fe3+ magnetic moment. Therefore, the Tm3+ magnetic moment flips, resulting in SSW-II, and the coupling mode shifts to ↓↑↑. With the increase of H, TSII-2 gradually moves towards the lower temperature, and SSW-II is suppressed at 80 Oe, 70 Oe, and 50 Oe respectively. In the high-temperature region, the Tm3+ magnetic moment is aligned with H. Therefore, with the increase of H, Tm3+ magnetic moment is more difficult to flip. When H becomes sufficiently large to maintain the Tm3+ magnetic moment aligned with H, SSW-II disappears. Therefore, in the FCC mode, with the increase of H, Tm3+ will flip at a lower temperature, leading to the decrease of TSII-2, which is regulated by the magnetic field. Due to the reduction of Tm3+ content, the suppressed H of SSW-II decreases with the increase of x. In the ZFC mode [Fig.5(d)], with the increase of temperature, magnetization along the c-axis of Tm0.5Pr0.5FeO3 single crystal fluctuates upward, resulting in a weak SSW-II. The continuous flipping of Tm3+ causes the coupling mode changes from ↓↑↑ to ↑↑↑. With the increase of H, the trigger temperature of SSW-II gradually decreases until 70 Oe. SSW-II along the c-axis at x = 0.15 and 0.25 is similar to x = 0.5 in ZFC mode (as shown in Fig. S2).

In order to further understand the anisotropic magnetic behavior, the MH curves are measured at different temperatures (as shown in Fig. S3). The transition between the weak ferromagnetic state and the antiferromagnetic state indicates that Tm0.5Pr0.5FeO3 single crystal is undergoing SRT when TSRT = 72 K−92 K, which is correspond to Fig.2(d). As previously discussed, RFeO3 is not a strictly antiferromagnetic material. Due to the tilt induced by the distortion of the FeO6 octahedral structure, the incline of Fe3+ magnetic moment causes the WFM in RFeO3. Weak ferromagnetism along the a-axis of Tm0.5Pr0.5FeO3 single crystal is observed in the antiferromagnetic state, according to Eq. (6) [35],

σs=MχAFMH.

σs is the weak ferromagnetic saturation magnetization, and χAFM is the antiferromagnetic contribution. At a higher temperature (95 K), the curve shows a tilted straight line (Fig. S3), presenting obvious antiferromagnetic behavior. As shown in Fig.6(e), after deducting χAFM, the curve exhibits weak ferromagnetic behavior. At low temperature, after deducting χAFM, the magnetization remains essentially unchanged, indicating that the sample has no antiferromagnetic component at 70 K. With the increase of temperature, σs gradually decreases [as shown in Fig.6(f)], indicating the decrease of WFM, and the AFM along the a-axis gradually domains, proving the occurrence of SRT. In addition, as the temperature increases, the coercive field (HC) gradually increases from 20 Oe to 100 Oe [as shown in Fig.6(f)], which is attributed to the hindered rotation of Fe3+ spin with the increase of temperature. The field-induced SSW is also observed along the a-axis in Tm0.5Pr0.5FeO3 single crystal. For example, at 70 K and 110 Oe [as shown in Fig.6(a)], the magnetization changed from −0.94 emu/g to 0.97 emu/g, showing a distinct jump.

4 Conclusion

Tm1−xPrxFeO3 (x = 0, 0.15, 0.25, 0.5) single crystals are successfully grown by optical float zone method and their magnetic properties are measured. With the increase of x, TSRT shifts to lower temperature. SSW-I was absent in TmFeO3 and PrFeO3 single crystals. SSW-I and SSW-II is observed along the a-axis when x = 0.15 and 0.25, but SSW-I disappears at x = 0.5. The variations of magnetic interactions in x = 0.15 and 0.5 single crystals are analyzed through structural distortion and transforming Curi−Weiss fitting, further discussing the reasons for the generation of SSW-I. A rare phenomenon of SSW-II existing simultaneously during cooling and warming process along the c-axis of Tm1−xPrxFeO3 (x = 0.15, 0.25, 0.5) single crystals. Both SSW-I and SSW-II can be regulated by H. Although Tm0.5Pr0.5FeO3 single crystal exhibits antiferromagnetic behavior at higher temperatures along the a-axis, there is still a WFM. With the increase of temperature, the weak ferromagnetic part gradually decreases, proving the occurrence of SRT. Meanwhile, field-induced SSW is observed along the a-axis in Tm0.5Pr0.5FeO3 single crystal. The complex and rich magnetic phenomena observed in Tm1−xPrxFeO3 (x = 0.15, 0.25, 0.5) single crystals hold significant potential for advancing research in altermagnetism and magneto-optical coupling, making them promising candidates for the development of next-generation spintronic devices.

References

[1]

X. W. Li, M. Bamba, N. Yuan, Q. Zhang, Y. G. Zhao, M. L. Xiang, K. Xu, Z. M. Jin, W. Ren, G. H. Ma, S. X. Cao, D. Turchinovich, and J. Kono, Observation of Dicke cooperativity in magnetic interactions, Science 361(6404), 794 (2018)

[2]

D. Kim, S. Dasgupta, X. X. Ma, J. M. Park, H. T. Wei, X. W. Li, L. Luo, J. Doumani, W. T. Yang, D. Cheng, R. H. J. Kim, H. O. Everitt, S. Kimura, H. Nojiri, J. G. Wang, S. X. Cao, M. Bamba, K. R. A. Hazzard, and J. Kono, Observation of the magnonic dicke superradiant phase transition, Sci. Adv. 11(14), eadt1691 (2025)

[3]

S. Das, A. Ross, X. X. Ma, S. Becker, C. Schmitt, F. van Duijn, E. F. Galindez-Ruales, F. Fuhrmann, M. A. Syskaki, U. Ebels, V. Baltz, A. L. Barra, H. Y. Chen, G. Jakob, S. X. Cao, J. Sinova, O. Gomonay, R. Lebrun, and M. Klaui, Anisotropic long-range spin transport in canted antiferromagnetic orthoferrite YFeO3, Nat. Commun. 13(1), 6140 (2022)

[4]

J. A. de Jong, A. V. Kimel, R. V. Pisarev, A. Kirilyuk, and T. Rasing, Laser-induced ultrafast spin dynamics in ErFeO3, Phys. Rev. B 84(10), 104421 (2011)

[5]

Z. Q. Zhang, Y. C. Chien, M. T. Wong, F. Y. Gao, Z. J. Liu, X. X. Ma, S. X. Cao, E. Baldini, and K. A. Nelson, Terahertz stimulated parametric down-conversion of a magnon mode in an antiferromagnet, Sci. Adv. 11, eadv3757 (2025)

[6]

C. Song, H. Bai, Z. Y. Zhou, L. Han, H. Reichlova, J. H. Dil, J. W. Liu, X. Z. Chen, and F. Pan, Altermagnets as a new class of functional materials, Nat. Rev. Mater. 10(6), 473 (2025)

[7]

L. D. Yuan, Z. Wang, J. W. Luo, E. I. Rashba, and A. Zunger, Giant momentum-dependent spin splitting in centrosymmetric low-Z antiferromagnets, Phys. Rev. B 102(1), 014422 (2020)

[8]

Z. Zhou, X. Cheng, M. Hu, R. Chu, H. Bai, L. Han, J. Liu, F. Pan, and C. Song, Manipulation of the altermagnetic order in CrSb via crystal symmetry, Nature 638(8051), 645 (2025)

[9]

M. Naka,S. Hayami,H. Kusunose,Y. Yanagi,Y. Motome,H. Seo, Nature Commun. 10, 4305 (2019)

[10]

R. L. White, Review of recent work on the magnetic and spectroscopic properties of the rare-earth orthoferrites, J. Appl. Phys. 40, 1061 (1969)

[11]

D. Treves, Studies on orthoferrites at the Weizmann institute of science, J. Appl. Phys. 36(3), 1033 (1965)

[12]

H. Kramers, L’interaction entre les atomes magnétogènes dans un cristal paramagnétique, Physica 1(1−6), 182 (1934)

[13]

P. W. Anderson, Antiferromagnetism. Theory of superexchange interaction, Phys. Rev. 79(2), 350 (1950)

[14]

J. Kanamori, Theory of the magnetic properties of ferrous and cobaltous oxides (II), Prog. Theor. Phys. 17(2), 197 (1957)

[15]

J. B. Goodenough, Theory of the role of covalence in the perovskite-type manganites [La, M(II)]MnO3, Phys. Rev. 100(2), 564 (1955)

[16]

T. Yamaguchi, Theory of spin reorientation in rare-earth orthochromites and orthoferrites, J. Phys. Chem. Solids 35(4), 479 (1974)

[17]

T. Moriya, Anisotropic superexchange interaction and weak ferromagnetism, Phys. Rev. 120(1), 91 (1960)

[18]

I. Dzyaloshinsky, A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics, J. Phys. Chem. Solids 4(4), 241 (1958)

[19]

J. Ma, J. Q. Yan, S. O. Diallo, R. Stevens, A. Llobet, F. Trouw, D. L. Abernathy, M. B. Stone, R. J. McQueeney, and Role of magnetic exchange energy on charge ordering in R1/3Sr2/3FeO3 (R=La, Pr, and Nd), Phys. Rev. B 84(22), 224115 (2011)

[20]

S. J. Yuan, Y. M. Cao, L. Li, T. F. Qi, S. X. Cao, J. C. Zhang, L. E. DeLong, and G. Cao, First-order spin reorientation transition and specific-heat anomaly in CeFeO3, J. Appl. Phys. 114(11), 113909 (2013)

[21]

E. Y. Li, Z. J. Feng, B. J. Kang, J. C. Zhang, W. Ren, and S. X. Cao, Spin switching in single crystal PrFeO3 and spin configuration diagram of rare earth orthoferrites, J. Alloys Compd. 811, 152043 (2019)

[22]

S. X. Cao, H. Z. Zhao, B. J. Kang, J. C. Zhang, and W. Ren, Temperature induced spin switching in SmFeO3 single crystal, Sci. Rep. 4(1), 5960 (2014)

[23]

J. X. Liu, H. Song, Z. R. Liu, Z. J. Yang, H. H. Peng, J. Chen, B. J. Kang, R. R. Jia, G. H. Wang, J. Ma, X. X. Ma, and S. X. Cao, J-T distortion tuned spin switching of Fe-site Mn-doped TmFeO3 single crystal, J. Alloys Comp. 1035, 181465 (2025)

[24]

S. X. Cao, L. Chen, W. Y. Zhao, K. Xu, G. H. Wang, Y. L. Yang, B. J. Kang, H. J. Zhao, P. Chen, A. Stroppa, R. K. Zheng, J. C. Zhang, W. Ren, J. Íñiguez, and L. Bellaiche, Tuning the weak ferromagnetic states in dysprosium orthoferrite, Sci. Rep. 6(1), 37529 (2016)

[25]

G. B. Song, J. J. Jiang, B. J. Kang, J. C. Zhang, Z. X. Cheng, G. H. Ma, and S. X. Cao, Spin reorientation transition process in single crystal NdFeO3, Solid State Commun. 211, 47 (2015)

[26]

X. X. Ma, W. C. Fan, G. Zhao, H. Y. Chen, C. K. Wang, B. J. Kang, Z. J. Feng, J. Y. Ge, W. Ren, and S. X. Cao, Low field control of spin switching and continuous magnetic transition in an ErFeO3 single crystal, Phys. Chem. Chem. Phys. 24(2), 735 (2022)

[27]

H. Song, W. C. Fan, R. R. Jia, Z. Q. Sun, X. X. Ma, W. T. Yang, S. Zhu, B. J. Kang, Z. J. Feng, and S. X. Cao, Low field spin switching of single crystal TmFeO3, Ceram. Int. 49(13), 22038 (2023)

[28]

Z. Q. Sun, H. Song, S. Zhu, X. X. Ma, W. T. Yang, C. F. Shi, B. J. Kang, R. R. Jia, J. K. Bao, and S. X. Cao, Pr-doping effect on spin switching in Nd0.8Pr0.2FeO3 single crystal, J. Phys. Chem. C 127(35), 17592 (2023)

[29]

J. H. Lee, Y. K. Jeong, J. H. Park, M. A. Oak, H. M. Jang, J. Y. Son, and J. F. Scott, Spin-canting-induced improper ferroelectricity and spontaneous magnetization reversal in SmFeO3, Phys. Rev. Lett. 107(11), 117201 (2011)

[30]

R. X. Huang, S. X. Cao, W. Ren, S. Zhan, B. J. Kang, and J. C. Zhang, Large rotating field entropy change in ErFeO3 single crystal with angular distribution contribution, Appl. Phys. Lett. 103, 162412 (2013)

[31]

Z. Ren, L. Cheng, G. Sergei, L. Nadzeya, J. T. Li, J. M. Shang, B. Sergei, A. H. Wu, K. Alexandra, Z. W. Ma, C. Zhou, and Z. G. Sheng, Terahertz spectroscopy study of doping and magnetic field induced effects on spin reorientation in Ho1–xYxFeO3 single crystals, Acta Phys. Sin. 69(20), 207802 (2020)

[32]

B. B. Dash and S. Ravi, Structural, magnetic and electrical properties of Fe substituted GdCrO3, Solid State Sci. 83, 192 (2018)

[33]

A. H. Cooke, D. M. Martin, M. R. Wells, and Magnetic interactions in gadolinium orthochromite, GdCrO3, J. Phys. C 7(17), 3133 (1974)

[34]

C. Shi, Z. Lin, Q. Liu, J. Lyu, X. Xu, B. Kang, J. H. Yang, Y. Liu, J. Zhang, S. Cao, and J. K. Bao, Weak low-temperature ferromagnetism and linear magnetoresistance in Lu0.75Fe6Sn6 with a disordered HfFe6Ge6-type structure, J. Magn. Magn. Mater. 615, 172793 (2025)

[35]

J. J. Guo, Y. L. Su, C. F. Shi, G. S. Gong, X. R. Cheng, X. Zhu, H. Y. Hu, and Y. Q. Wang, Temperature dependence of the microstructure and magnetic properties of polycrystalline GdCrO3, J. Supercond. Nov. Magn. 35(3), 711 (2022)

[36]

L. H. Yin, J. Yang, X. C. Kan, W. H. Song, J. M. Dai, and Y. P. Sun, Giant magnetocaloric effect and temperature induced magnetization jump in GdCrO3 single crystal, J. Appl. Phys. 117(13), 133901 (2015)

[37]

C. F. Shi, Y. L. Su, J. J. Guo, J. Y. Zhang, G. S. Gong, H. Y. Hu, and Y. Q. Wang, The microstructure and magnetic properties of Ca2+ ion doped GdCrO3, Ceram. Int. 47(8), 10887 (2021)

RIGHTS & PERMISSIONS

Higher Education Press

AI Summary AI Mindmap
PDF (3203KB)

Supplementary files

fop-26030-of-caoshixun_suppl_1

712

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/