Recently, double perovskite (DP) oxides denoted A2B′B″O6 (A being divalent or trivalent metals, B′ and B″ being heterovalent transition metals) have been attracting much attention owing to their wide range of electrical and magnetic properties. Among them, rhenium (Re)-based DP oxides such as A2FeReO6 (A = Ba, Sr, Ca) are a particularly intriguing class due to their high magnetic Curie temperatures, metallic-like, half-metallic, or insulating behaviors, and large carrier spin polarizations. In addition, the Re-based DP compounds with heterovalent transition metals B′ and B″ occupying B sites have a potential to exhibit rich electronic structures and complex magnetic structures owing to the strong interplays between strongly localized 3d electrons and more delocalized 5d electrons with strong spin–orbit coupling. Thus, the involved physics in the Re-based DP compounds is much richer than expected. Therefore, there are many issues related to the couplings among the charge, spin, and orbitals, which need to be addressed in the Re-based DP compounds. In the past decade, much effort has been made to synthesize Re-based DP compounds and to investigate their crystal structures, structural chemistry, and metal–insulator transitions via orbital ordering, cationic ordering, and electrical, magnetic, and magneto-transport properties, leading to rich literature in the experimental and theoretical investigations. This Review focuses on recent advances in Re-based DP oxides, which include their synthesis methods, physical and structural characterizations, and advanced applications of Re-based DP oxides. Theoretical investigations of the electronic and structural aspects of Re-based DP oxides are also summarized. Finally, future perspectives of Re-based DP oxides are also addressed.

Perovskite oxides have been widely investigated during the past 70 years owing to their interesting multifunctional properties and versatile technical applications.1–3 They possess a wide range of electrical properties such as insulating, semiconducting, metallic, and half-metallic behaviors, on the one hand, and ferromagnetic, ferrimagnetic, and antiferromagnetic behaviors, on the other hand.4–6 In addition, they may demonstrate ferroelectric,7 magnetic–dielectric,8 and multiferroic behaviors.9 Thus, perovskite oxides have promising applications in electronic/spintronic devices,10 fuel cells,11 solar cells,12 and so on.13,14 The above promising properties can be attributed to the high chemical and structural flexibilities of the perovskite structure. In an ideal cubic perovskite compound with the formula ABO3, A-sited cations with a 12-fold coordination occupy the corners of a cubic unit cell and B-sited transition metal cations sit at the center of the octahedron. Normally, the BO6 octahedra are required to distort and/or tilt in various possible directions to accommodate the size mismatches generated by the substituted cations at A and/or B sites. Therefore, almost all elements covered in the periodic table can be presented at the A or B site in the perovskite structure, opening up vast possibilities for the formation of perovskite compounds.15,16 In recent years, with the research and development of quantum electronic/spintronic devices without dissipation, intensive research studies are focused on double perovskite (DP) oxides with the general formula A2B′B″O6 (A being divalent or trivalent metals, B′ and B″ being 3d, 4d, or 5d transition metal cations) because of their intriguing magnetic properties [e.g., ferromagnetic insulator (FMI) at high temperatures and spin polarized transport above room temperature]. DP oxides exhibit rich physical properties owing to a wide variety of combinations of the metal ions at B′ and B″ sites, which offers much tunable B′–O–B″ magnetic interactions via superexchange.10 The first study of DP oxides was reported in the early 1950s, and it was widely carried out at the end of 1950s.17,18 In the mid 1970s, a wide variety of DP oxides were reported. In 1998, Kobayashi et al.5 reported a low-field room temperature magnetoresistance (MR) in an ordered Sr2FeMoO6 DP compound; this finding stimulated the research studies on searching for new ferromagnetic compounds with a DP crystallographic structure.19–22 Among the various DP oxides, rhenium (Re)-based DP oxides (e.g., A2BReO6, Sr2(Fe1−xCrx)ReO6, Sr2Fe1+xRe1−xO6) are a particularly intriguing class owing to their high magnetic Curie temperatures, diverse electrical behaviors, huge magnetic anisotropy, and large carrier spin polarizations.23–25 Moreover, the Re-based DP compounds containing 3d and 5d transition metal ions at the B′ and B″ sites have rich electronic structures and complex magnetic structures, which are ascribed to the strong interactions between strongly localized 3d electrons and more delocalized 5d electrons with strong spin–orbit coupling. Thus, the involved physics in the Re-based DP compounds is much more complex than expected. Therefore, there are many issues related to the couplings among the charge, spin, and orbitals, which need to be addressed in the Re-based DP compounds.23–25 In the past decade, many attempts have been made to study the structural chemistry, microstructure, and metal-to-insulator (MIT) transition via orbital ordering, cationic ordering, and electrical, magnetic, and magneto-transport properties of Re-based DP oxides, forming rich literature in the experimental and theoretical aspects. However, we note that only a few review articles of these compounds have been published. For example, Serrate et al. reviewed the FeMo-based and Re-based DP compounds with a focus on their magnetic properties and MIT behaviors,19 whereas Karppinen and Yamauchi discussed their chemical aspects with emphasis on oxygen stoichiometry, cation ordering, and redox chemistry.26 In this Review, a comprehensive review of the recent advances in Re-based DP oxides is provided, which includes their syntheses, physical and structural characterizations, and advanced applications of Re-based DP oxides. Theoretical investigations of the electronic and structural aspects of Re-based DP oxides are also summarized. Finally, the summary and outlook of Re-based DP oxides are presented. It is expected that this Review will attract more interest in Re-based DP oxides and more researchers to enter this emerging field.

In the Re-based DP oxides of A2MReO6 (A = alkaline earth metals such as Ba, Sr, Ca; M = transition metals such as Sc, Cr, Mn, Fe, Co, Ni, Zn), their rich substitutional properties are advantageous for fabricating DP oxide compounds in the form of bulk single crystals, polycrystalline ceramics, thin films, or nanocrystals. In this section, we will elaborate some methods used for synthesis of Re-based DP oxides.

Re-based DP bulk oxides can be simply obtained by direct annealing the mixture of corresponding metal oxides at high temperatures, where the solid powders undergo both physical and chemical reactions to enable the thermal diffusion of ions or molecules. The solid-state reaction method is widely utilized to fabricate Re-based DP bulk oxides, which is environmentally friendly owing to without releasing toxic gases during the synthesized process. However, it has some disadvantages such as the final products often with impurities and poor chemical homogeneity, non-distributed particle sizes, long reaction time required, and high annealing temperature (e.g., over 1000 °C). These issues can be alleviated by modifying the processing parameters of the solid-state reaction method (e.g., repeatedly grinding the starting materials; increasing annealing temperature and reaction time). For example, Kato et al.27 synthesized the polycrystalline Re-based DP oxides such as Sr2MReO6 and Ca2MReO6 (M being transition metals) by solid-state reactions, where SrO, CaO, MOx, Re2O7, and Re powders were used as the raw materials. The synthesized samples were almost of pure phase despite a small amount of impurities (less than 2% in fraction) except Ca2NiReO6 (about 8%). In another case of the Sr2−xLaxFe1+x/2Re1−x/2O6 (x = 0.0–1.8) compounds, they were synthesized by the solid-state reaction method under an inert atmosphere, where La2O3, SrCO3, Fe2O3, ReO3, and Re were used as the starting materials. They were weighted based on the designed ReO3/Re ratio to the theoretical chemical valence (5+) of Re and calcined at 1100 °C for 2 h in a stream of Ar gas.28 To obtain a single-phase structure, the mixed powders were repeatedly sintered at 1200 °C for 3 h several times in the same atmosphere. To increase the density of bulk ceramics of Re-based DP oxides, Lim et al.29 utilized the spark plasma sintering method to fabricate the bulk ceramic samples (e.g., Sr2FeReO6 and Sr2CrReO6). Spark plasma sintering technique is a high speed consolidation of the powders by using uniaxial pressure and pulsed electrical current under low atmospheric pressure.30,31 The remarkable features of the spark plasma sintering method make it suitable for fabricating Re-based DP ceramic oxides with high densification.

In recent years, Re-based DP thin films have attracted much attention owing to their high Tc value, large spin polarization, and half-metallicity, which have promising applications in spin electronic devices. However, to date, only a few works reported on Re-based DP thin films, unlike their bulk counterpart. Recently, Sohn et al.25 grew epitaxial Sr2Fe1+xRe1−xO6 (−0.2 ≤ x ≤ 0.2) thin films on single-crystal (001) SrTiO3 substrates by pulsed laser deposition (PLD). They demonstrated that the compositional ratio and cation ordering of the Sr2Fe1+xRe1−xO6 thin films were closely related to the oxygen pressure (PO2). The thin films exhibited a ferromagnetic insulator (FMI) state with high-Tc ferromagnetism and high saturation magnetization (MS ≈ 1.8 µB/f.u.). In addition, their room-T sheet resistance was about three order larger as compared with that of metallic films. The stable FMI state was found in the cation-ordered Fe-rich films due to the formation of extra Fe3+–Fe3+ and Fe3+–Re6+ bonding states. In another study, highly ordered epitaxial Sr2CrReO6 thin films were also deposited on single-crystal (001) SrTiO3 substrates with an epitaxial buffer layer of Sr2CrNbO6 (30 nm–48 nm in thickness) by off-axis magnetron sputtering.32 The Sr2CrReO6 films had 99% Cr/Re ordering and high-quality crystallinity. Their electrical resistivity was strongly sensitive to the oxygen partial pressure. 18 000% modulation in electrical resistivity was achieved at a temperature of 7 K (60% modulation at 300 K) as the oxygen partial pressure had 1% modulation during film growth. The results indicate that the electrical properties of Sr2CrReO6 thin films are closely related to the content of oxygen vacancies. Orna et al.33 also grew epitaxial Sr2CrReO6 films by PLD, which exhibited high crystallinity and large cationic ordering. Their MS value at 300 K was ∼1.0 µB/f.u. and the electrical resistivity (ρ) was 2.8 mΩ cm at 300 K, similar to those reported previously for the sputtered epitaxial Sr2CrReO6 films.

Large magnetoresistances are reported in the polycrystalline Re-based DP oxides such as Sr2FeReO6 and Sr2CrReO6, which resulted from the intergrain tunneling magnetoresistance (ITMR) effect controlled by the grain boundary magnetization.34 It is also reported that the formation of nanosized particles can significantly enhance the ITMR effect in Re-based DP powders.35–37 However, the synthesis of Re-based DP powders with high-purity and homogeneity is proven to be challenging due to the strong refractive nature of the reduced rhenates (i.e., Re5+/Re6+). Recently, molten salt synthesis (MSS) is used to synthesize Re-based DP oxide nanoparticles (e.g., Sr2FeReO6, Ba2FeReO6, and Sr2CrReO6), and the eutectic mixture of NaCl–KCl salts acted as the reaction medium. Fuoco et al.37 reported the influence of annealing temperature, holding time, cooling rates, and molten salt fluxes on the particle sizes, ordering degree at the B′/B″ site, and the physical properties of Sr2FeReO6, Ba2FeReO6, and Sr2CrReO6 powders. They found that high purity and homogeneity were achieved in the Re-based DP oxide powders (e.g., Sr2FeReO6, Ba2FeReO6, and Sr2CrReO6) synthesized by the MSS method. The particle sizes of the Re-based DP powders were in the range from ∼50 nm to >1 µm, which were dependent upon the different synthesized conditions and the compositions of DP oxides. Generally, small-sized particles were synthesized under short reaction time, which had much large concentrations of grain boundaries, leading to strong temperature dependent resistivity. This new synthetic route allows one to deeply investigate the effects of the particle size and the ordered degree at the B′/B″ site on the magnetic properties of Re-based DP powders.

Spintronics is becoming a candidate technology based on the spin of the electron to carry and store information as 0 bits (spin up ↑) and 1 bit (spin down ↓). This new technology has several advantages over the current charge-based technologies.38–40 To realize the full commercial spintronic devices, the essential components such as half-metallic spin injectors and magnetic semiconductors are highly required.41–43 However, the main challenge to this new technology is the discovery of suitable candidates that have ferromagnetic ordering and large carrier spin polarizations over room temperature. Re-based DP oxides such as Sr2CrReO6 are considered the potential materials for spintronics owing to their high Tc value (∼635 K), half-metallicity, and large MS (=1.0 μB/f.u.).44–48 Following the successful increase in the TC value in the Mo-based DP oxides by electron doping, similar research works were also carried out in the Re-based DP oxides.49 However, the expected enhancement of TC with lanthanide addition in the Sr2CrReO6 DP oxides was not realized. This means the magnetic interaction strength is decreased with lanthanide doping. The converse electron doping scenarios observed in Sr2CrReO6 and Sr2FeMoO6 may be ascribed to the presence of secondary phases in the final products of Sr2-xLnxCrReO6 (Ln being La, Nd, or Sm). To confirm this assumption, Blasco et al.50 fabricated two sets of polycrystalline Sr2−xLnxCrReO6 and Sr2−xLnxCr1+x/2Re1−x/2O6 samples (x = 0.0–0.5; Ln = La, Nd, Sm) by the solid-state reaction method. They found that the synthesis of the Sr2−xLnxCrReO6 (Ln = La, Nd, or Sm) single phase was impossible via conventional synthetic routes because of the formation of competitive Sr2−xLnxCr1+x/2Re1−x/2O6 and Sr11Re4O24 phases. This was the reason why it is significantly hard to synthesize electron-doped DP oxides in the (Cr, Re)-based systems. Single-phase compounds were only obtained with the stoichiometric formula of Sr2−xLnxCr1+x/2Re1−x/2O6, which crystallized in a fcc cubic cell with a shrinkage when the sizes of Ln ions were decreased. These samples exhibited spontaneous magnetization and large coercive field at room temperature, and their Curie temperature was weakly affected by Ln addition, whereas the MS value was decreased owing to the appearance of anti-site (AS) defects. The electronic states of both Cr3+ and Re5+ ions were not affected by replacing Sr by Ln, as confirmed by x-ray absorption spectroscopy (XAS) at the Cr K edge and Re L1,2,3 edges. Recently, the first-principles calculations demonstrate that the La-doped Ba2−xLaxFeReO6 (x = 0.0, 0.5, and 1.0) compounds undergo a transition from half-metallic to insulating behavior as the La-doping content (x) is increased up to x = 1.0.51 The injections of electrons mainly go to Re t2g orbitals, whereas the Fe 3d orbitals almost remain unchanged when increasing the La-doping content. Therefore, Fe exists as 3+ in doped and un-doped samples, while Re has different chemical states in Ba2FeReO6 (5+), Ba1.5La0.5FeReO6 (4.5), and BaLaFeReO6 (4+) compounds. The remanent magnetization is decreased when increasing the La-doping content.

In the case of homo-valent dopants, polycrystalline Re-based DP Sr2LnReO6 (Ln = Y, La, Nd, Sm–Gd, and Dy–Lu) samples are also synthesized by the standard solid-state reaction, where the transition metals such as Fe, Cr were used to replace lanthanide (such as trivalent Y, La, Nd, Sm–Gd, and Dy–Lu) at the B′ site.52 It is observed that the Ln3+ and Re5+ ions are orderly distributed at the B′ and B″ sites of the DP structure. The Sr2LnReO6 compounds exhibit an antiferromagnetic behavior at 2.6 K–20 K. This can be ascribed to the magnetic interactions between Re5+ ions. In the Sr2YbReO6 compound, the magnetic orderings of Yb3+ and Re5+ ions appear at 20 K, whereas, at around 90 K, Re5+ ions are magnetically ordered in Sr2DyReO6. By furthermore reducing the temperature, Dy moments are ferromagnetically ordered at 5 K.

The structural characterizations of Re-based DP oxides include determinations of crystal structures, chemical compositions, and morphologies. There are several conventional methods for characterizing the crystal structures, which are based on the inactions between the incident x ray (or high-energy electron beam) and the investigated samples. X-ray diffraction (XRD) is a very common method to determine the phase structures of all kinds of matter ranging from fluids, to powders, and crystals. By analyzing the XRD data, the critical features such as crystal structures, crystallite sizes, lattice parameters (ac), lattice volume, and strain can be determined. Infrared and Raman spectra are two complementary techniques, which provide the characteristic fundamental vibrations used for determination and identification of the crystal structures at a molecular level. Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) make use of a fine electron beam to reveal the microstructures of the investigated samples, which work in electron back-scattering and transmission modes, respectively. Two kinds of high-resolution imaging modes named high-resolution TEM (HRTEM) and scanning transmission electron microscopy (STEM) are often used to reveal microstructures (or structural defects) of Re-based DP oxides at an atomic level or even at a sub-Å level. Selected area electron diffraction (SAED) is widely utilized to determine the phase structure of the samples, while convergent beam electron diffraction (CBED) has a unique ability of determining the space group of crystals due to its great advantages of obtaining reliable experimental data from a single-domain area. The spectral analyzed techniques such as energy dispersive x-ray spectroscopy (EDS), electronic energy loss spectroscopy (EELS), and x-ray photoelectron spectroscopy (XPS) are used to determine the chemical compositions and chemical bonding states of the Re-based DP oxide. In this section, we shortly review the recent advantages in the microstructural characterizations of Re-based DP oxides.

To date, many Re-based DP bulk oxides have been synthesized by different methods, and their phase structures are first examined by XRD, from which the lattice parameters (ac), unit cell volume (V), and density of the sample in theory can be determined. As an example, XRD patterns of Ba2FeReO6 and Ca2FeReO6 DP compounds synthesized by the solid-state reaction are demonstrated in Fig. 1. As shown in Fig. 1(a), the XRD pattern of Ba2FeReO6 can be well indexed based on a cubic cell (Fm3m), and the lattice constant (a) is determined to be about 8.054 Å. Ca2FeReO6 [Fig. 1(b)] is found to crystallize in a monoclinic structure with the space group of P21/n and lattice parameters of a ≈ 5.396 Å, b ≈ 5.522 Å, c ≈ 7.688 Å, and β = 90.4°.53 It was also noticed that the impurity phase was not observed in the XRD patterns, implying the formation of a pure compound in each case. Figure 2 demonstrates the room temperature XRD patterns of a series of Ca-doped Sr2−xFeReO6 (x = 0.0–2.0) samples.54 It was observed that the samples with x = 1.5 and 2.0 exhibited a monoclinic structural distortion and crystallized with the space group of P21/n, whereas the samples with x = 0.5 and 1.0 had a pseudo-cubic structure. The lattice parameters (ac) determined from the XRD patterns are displayed in Fig. 3. In order to express the overall behavior of the investigated series, pseudo-cubic parameters, aps (=a2) and bps (=b2), are used rather than the lattice parameters (a and b) of the tetragonal and monoclinic samples. To describe the structural changes undergone in the series of Re-based A2MReO6 DP compounds (A being Sr, Ca; M being Mg, Sc, Cr, Mn, Fe, Co, Ni, Zn), the tolerance factor (t) is introduced, which is written as

(1)

where rA, rM, rRe, and rO represent the ionic radii of A, M, Re, and O ions, respectively. Figure 4 shows the tolerance factor dependent upon the cell volume, monoclinic structural distortion expressed by ∣β-90°∣ (β, the monoclinic angle), and averaged M-O-Re bond angles of the Re-based A2MReO6 compounds.27 It is noticed that the Sr-based compounds (e.g., Sr2MReO6) have larger cell volumes and M-O-Re angles (close to 180°) in comparison with the Ca-based compounds (e.g., Ca2MReO6). In addition, for each A-site ion, with the decreasing tolerance factor t in the A2MReO6 compounds, cell volumes continuously increased, whereas the M-O-Re angles monotonously decreased. For the compounds with t < 0.98, their crystal structures are monoclinically distorted and the distortion degree of ∣β-90°∣increases with the reducing t value.

FIG. 1.

X-ray powder diffraction patterns of double-perovskites of (a) Ba2FeReO6 and (b) Ca2FeReO6. Reprinted with permission from Prellier et al., J. Phys.: Condens. Matter 12, 965 (2000). Copyright 2000 IOP Publishing Ltd.

FIG. 1.

X-ray powder diffraction patterns of double-perovskites of (a) Ba2FeReO6 and (b) Ca2FeReO6. Reprinted with permission from Prellier et al., J. Phys.: Condens. Matter 12, 965 (2000). Copyright 2000 IOP Publishing Ltd.

Close modal
FIG. 2.

XRD patterns of the CaxSr2−xFeReO6 compounds with x = 0.0–2.0 measured at room temperature. From bottom to top, the value of x increases from 0.0 to 2.0 with an interval of 0.5. The inset shows the local XRD patterns (2θ = 22°–30°), showing the (111) diffraction peak and confirming the appearance of the P21/n structure. Reprinted with permission from De Teresa et al., Phys. Rev. B 69, 144401 (2004). Copyright 2004 American Physical Society.

FIG. 2.

XRD patterns of the CaxSr2−xFeReO6 compounds with x = 0.0–2.0 measured at room temperature. From bottom to top, the value of x increases from 0.0 to 2.0 with an interval of 0.5. The inset shows the local XRD patterns (2θ = 22°–30°), showing the (111) diffraction peak and confirming the appearance of the P21/n structure. Reprinted with permission from De Teresa et al., Phys. Rev. B 69, 144401 (2004). Copyright 2004 American Physical Society.

Close modal
FIG. 3.

Lattice parameters (ac) of CaxSr2−xFeReO6 compounds obtained from the room temperature XRD patterns plotted as the average radius (rA) of the A-site in A2FeReO6 (open circle) and A2FeMoO6 (solid triangle). To demonstrate the overall behaviors of the tetragonal and monoclinic samples (x = 0), pseudo-cubic parameters aps (=a2) and bps = (b2) are used to replace the lattice parameters a and b. The solid lines just show the trends. Reprinted with permission from Serrate et al., J. Phys.: Condens. Matter 19, 023201 (2007). Copyright 2007 IOP Publishing Ltd.

FIG. 3.

Lattice parameters (ac) of CaxSr2−xFeReO6 compounds obtained from the room temperature XRD patterns plotted as the average radius (rA) of the A-site in A2FeReO6 (open circle) and A2FeMoO6 (solid triangle). To demonstrate the overall behaviors of the tetragonal and monoclinic samples (x = 0), pseudo-cubic parameters aps (=a2) and bps = (b2) are used to replace the lattice parameters a and b. The solid lines just show the trends. Reprinted with permission from Serrate et al., J. Phys.: Condens. Matter 19, 023201 (2007). Copyright 2007 IOP Publishing Ltd.

Close modal
FIG. 4.

Tolerance factor dependent upon the cell volume, the monoclinic distortion expressed by |β-90°∣ (β, monoclinic angle), and the averaged M–O–Re bond angles of the ordered A2MReO6 DPs (A being Sr, Ca and M being Ca, Mg, Sc, Cr, Mn, Fe, Co, Ni, and Zn). Reprinted with permission from Kato et al., Phys. Rev. B 69, 184412 (2004). Copyright 2004 American Physical Society.

FIG. 4.

Tolerance factor dependent upon the cell volume, the monoclinic distortion expressed by |β-90°∣ (β, monoclinic angle), and the averaged M–O–Re bond angles of the ordered A2MReO6 DPs (A being Sr, Ca and M being Ca, Mg, Sc, Cr, Mn, Fe, Co, Ni, and Zn). Reprinted with permission from Kato et al., Phys. Rev. B 69, 184412 (2004). Copyright 2004 American Physical Society.

Close modal

During the growth of Re-based double perovskite oxides, AS defects are also formed in Sr2FeReO6 (e.g., Fe occupies the normal Re site, FeRe, and Re occupies the normal Fe site, ReFe) and Sr2CrReO6 (e.g., Cr occupies the normal Re, CrRe, and Re occupies the normal Cr site, ReCr), respectively, which are dependent upon the synthesis conditions (e.g., annealing temperature, holding time, and pressure) in the spark plasma sintering process. The formation of AS defects is harmful to magnetization. Therefore, it is critical to suppress the formation of AS defects during the growth process of the DP samples. In general, diffraction techniques are used to detect the AS defects (quantifying the amount of misplaced magnetic ions) since some specific Bragg peaks such as (111), (113), and (331) diffraction peaks would emerge owing to the cationic ordering formed at the B′ and B″ superlattices.55 Recently, local cationic ordering in the pristine and Re-excess Sr2FeReO6 samples was comparatively investigated by high-angle annular dark-field (HAADF)-STEM.56 It is found that cationic ordering appears not only in the pristine sample but also in the Re-excess 15 mol. % Sr2FeReO6 sample. The HAADF-STEM images obtained from the pristine Sr2FeReO6 sample taken along [001] and [110] directions are demonstrated in Figs. 5(a) and 5(b), respectively. It is noticed that only the Fe and Re columns (orange sphere) and Sr column (green sphere) are resolved in the [001] projection, whereas, in the [110] projection, the Fe column (yellow sphere), Re column (blue sphere), and Sr column (green sphere) are clearly resolved. This confirms the Fe/Re ordering in the pristine Sr2FeReO6 sample. Figure 5(c) displays three kinds of intensity profiles of atomic columns, which are obtained from the pristine Sr2FeReO6 sample (marked by a red line), the Re-excess Sr2FeReO6 sample (marked by a blue line), and the simulated Sr2FeReO6 image (marked by a green line). Thus, the Fe/Re cationic ordering at the B′ and B″ sites in the pristine and Re-excess Sr2FeReO6 samples is well confirmed. However, in the pristine Sr2FeReO6 sample, AS defects are also found, as revealed in Fig. 6(a). In column intensity profiles, the Re atomic columns with much higher intensities are clearly distinguishable, and between two adjacent Re columns, there are three Fe atomic columns with almost equal intensities, which are indicated by the red arrow. It is noticed that some Fe atomic columns exhibit much lower intensities as compared with their neighbors, which are indicated by the green or yellow arrow. The appearance of Fe atomic columns with unusual strong intensities indicates the formation of AS defects. Figures 6(b)–6(d) present a series of [110] projections of simulated HAADF-STEM images, which have the Fe–Re exchanging ratios at 10%, 15%, and 20%, respectively. It is observed that the intensity profiles of the Fe atomic columns marked by the green and red arrows indicate more AS defects, which match well with the simulated images with 15% and 20% Fe–Re exchanges. Besides the AS defects, the antiphase boundary (APB) was also observed in the Re-excess Sr2FeReO6 samples. Figure 7(a) shows the HAADF-STEM image taken from an APB in the [110] direction projection, and Fig. 7(b) displays the intensity profile across the APB boundary. It is observed that the intensity profile in Fig. 7(b) exhibits quite symmetry near the Sr atom, as demonstrated by a red arrow. However, across the APB, the intensities of the Re columns (column nos. 4–7) are found to be reduced gradually, whereas an inverse case appears for the Fe columns (indicated by blue arrows). This means the Fe/Re cationic disordering appears around the APB region due to the formation of strong antiferromagnetic coupling between the magnetic cationic pairs of Fe3+/Fe3+ near the APB, leading to the reduced magnetic spin of Fe across the APB. Lim et al.29 also reported that adding Re is an effective way to reduce the AS defects in Sr2FeReO6 (by 10.4% when increasing the amount of excess Re up to 15 mol. %), but it slightly increases the AS defects (by 0.9% in Sr2CrReO6 when increasing the amount of excess Re up to 10 mol. %).29 Such a discrepancy can be ascribed to their different thermodynamic stabilities between Sr2FeReO6 and Sr2CrReO6 with the excess of Re. The microscopic structures of AS defects in Sr2FeReO6 and Sr2CrReO6 are also examined by HAADF-STEM images. Figure 8(a) reveals the sequential atomic chains of Re–Sr–Fe–Sr–Re along the [001] direction and Re–Fe–Re along the [110] direction since B-sited atoms (Re and Fe) are arranged along the ⟨110⟩ direction. Figures 8(b)–8(e) show the HAADF-STEM images taken from Sr2CrReO6-Re excess 5 mol. %, Sr2CrReO6-Re excess 15 mol. %, Sr2CrReO6-no excess Re, and Sr2CrReO6-Re excess 10 mol. %. In the Sr2CrReO6-Re excess 5 mol. % sample [Fig. 8(b)], antiphase-boundary (APB)-like defects were observed, similar to that reported for Sr2FeReO656 and Sr2FeMoO6.57 However, such defect structures were not observed in the sample of Sr2CrReO6 with Re excess 15 mol. % [Fig. 8(c)], which was ascribed to the reduced sizes of AS defects with a larger amount of excess Re. In the pure Sr2CrReO6 sample [Fig. 8(d)] and Re excess 10 mol. % Sr2CrReO6 sample [Fig. 8(e)], the AS defects were not observed due to their wide distribution in the whole samples, similar to the cases in other oxides.58,59 The Fe/Re cation ordering at the B-sites (B′ and B″) of Sr2FeReO6 was also examined by using EDS element mapping formed by the Fe Kα signal at 6.405 keV and Re Kα signal at 8.652 keV, besides the nanodiffraction patterns along ⟨011⟩pc.60Figure 9(a) is the HAADF-STEM image taken from a perfect region of the Sr2FeReO6 sample, and Figs. 9(b)–9(e) show the corresponding atomic-resolution elemental maps and nanodiffraction pattern. The compositional distributions of the Fe element (red) [Fig. 9(b)] and Re element (blue) [Fig. 9(b)] are stacked alternatively in the {111} planes, giving a composite color map, which matches well with the feature of the Z-contrast image shown in Fig. 9(a). The appearance of 1/2{111}pc super-diffraction spots marked by dotted circles in Fig. 9(e) also confirmed the alternative distributions of Fe and Re atoms in the {111} planes. The quantitative results of EDX spectra demonstrate the Fe/Re ratio of the selected region was 0.51:0.49, close to the statistical measurements. Therefore, the Sr2FeReO6 phase can be further written as Sr2[Fe]B′[Re]B″O6.

FIG. 5.

HAADF-STEM images taken from the pristine Sr2FeReO6 samples along (a) [001] and (b) [110] directions. Simulated HAADF-STEM images are shown as the insets. The Fe and Re columns (orange sphere) and Sr column (green sphere) are clearly resolved in the [001] projection, whereas, in the [110] projection, the Fe column (yellow sphere), Re column (blue sphere), and Sr column (green sphere) are clearly resolved. (c) Intensity profiles for the atomic chains of Re–Sr–Fe–Sr–Re along the [001] direction measured from the HAADF-STEM images taken from the pristine samples (red color) and the Re excess samples (blue color) in the [110] projection plane. The intensity profile of the simulated image is marked by green color. Reprinted with permission from Choi et al., Microsc. Microanal. 19(S5), 25 (2013). Copyright 2013 Cambridge University Press.

FIG. 5.

HAADF-STEM images taken from the pristine Sr2FeReO6 samples along (a) [001] and (b) [110] directions. Simulated HAADF-STEM images are shown as the insets. The Fe and Re columns (orange sphere) and Sr column (green sphere) are clearly resolved in the [001] projection, whereas, in the [110] projection, the Fe column (yellow sphere), Re column (blue sphere), and Sr column (green sphere) are clearly resolved. (c) Intensity profiles for the atomic chains of Re–Sr–Fe–Sr–Re along the [001] direction measured from the HAADF-STEM images taken from the pristine samples (red color) and the Re excess samples (blue color) in the [110] projection plane. The intensity profile of the simulated image is marked by green color. Reprinted with permission from Choi et al., Microsc. Microanal. 19(S5), 25 (2013). Copyright 2013 Cambridge University Press.

Close modal
FIG. 6.

(a) HAADF-STEM image taken from the pristine Sr2FeReO6 samples and their intensity profile of the atomic chains taken from the rectangle marked by red color. Simulated HAADF-STEM images of the Sr2FeReO6 samples with Fe/Re disordering of (b) 10%, (c) 15%, and (d) 20% and the corresponding intensity profiles. Reprinted with permission from Choi et al., Microsc. Microanal. 19(S5), 25 (2013). Copyright 2013 Cambridge University Press.

FIG. 6.

(a) HAADF-STEM image taken from the pristine Sr2FeReO6 samples and their intensity profile of the atomic chains taken from the rectangle marked by red color. Simulated HAADF-STEM images of the Sr2FeReO6 samples with Fe/Re disordering of (b) 10%, (c) 15%, and (d) 20% and the corresponding intensity profiles. Reprinted with permission from Choi et al., Microsc. Microanal. 19(S5), 25 (2013). Copyright 2013 Cambridge University Press.

Close modal
FIG. 7.

(a) HAADF-STEM image of the antiphase boundary (APB) observed in the Re-excess Sr2FeReO6 samples and (b) intensity profile across the APB along the [001] direction. Reprinted with permission from Choi et al., Microsc. Microanal. 19(S5), 25 (2013). Copyright 2013 Cambridge University Press.

FIG. 7.

(a) HAADF-STEM image of the antiphase boundary (APB) observed in the Re-excess Sr2FeReO6 samples and (b) intensity profile across the APB along the [001] direction. Reprinted with permission from Choi et al., Microsc. Microanal. 19(S5), 25 (2013). Copyright 2013 Cambridge University Press.

Close modal
FIG. 8.

(a) HAADF-STEM image obtained from Sr2FeReO6 with excess 15 mol. % Re in the [110] projection, where Re (brown)–Sr (green)–Fe (gray)–Sr (green)–Re(brown) atomic chains along the [001] direction are well resolved from their Z contrast. The inset shows a schematic diagram of the atomic distributions in the projection plane of [110]. (b)–(e) HAADF-STEM images taken from Sr2FeReO6-5 mol. % excess Re, Sr2FeReO6-15 mol. % excess Re, Sr2CrReO6-0 mol. % excess Re, and Sr2CrReO6-10 mol. % excess Re samples, respectively. The inset of (b) shows an enlarged AS defect region indicated by the yellow rectangle, exhibiting an apparent difference from the well-ordered structure demonstrated in (a). Reprinted with permission from Lim et al., Sci. Rep. 6, 19746 (2016). Copyright 2016 Nature Publishing Group.

FIG. 8.

(a) HAADF-STEM image obtained from Sr2FeReO6 with excess 15 mol. % Re in the [110] projection, where Re (brown)–Sr (green)–Fe (gray)–Sr (green)–Re(brown) atomic chains along the [001] direction are well resolved from their Z contrast. The inset shows a schematic diagram of the atomic distributions in the projection plane of [110]. (b)–(e) HAADF-STEM images taken from Sr2FeReO6-5 mol. % excess Re, Sr2FeReO6-15 mol. % excess Re, Sr2CrReO6-0 mol. % excess Re, and Sr2CrReO6-10 mol. % excess Re samples, respectively. The inset of (b) shows an enlarged AS defect region indicated by the yellow rectangle, exhibiting an apparent difference from the well-ordered structure demonstrated in (a). Reprinted with permission from Lim et al., Sci. Rep. 6, 19746 (2016). Copyright 2016 Nature Publishing Group.

Close modal
FIG. 9.

Structural and compositional characterizations of ordered Sr2FeReO6. (a) HAADF-STEM image viewed in the [110] projection; (b) and (c) atomic-resolution element maps of Fe (red) and Re (blue) elements, respectively. (d) Mixed Fe and Re maps, and (e) nanodiffraction pattern taken along ⟨011⟩pc. Reprinted with permission from Ho et al., Ultramicroscopy 193, 137 (2018). Copyright 2018 Elsevier B.V.

FIG. 9.

Structural and compositional characterizations of ordered Sr2FeReO6. (a) HAADF-STEM image viewed in the [110] projection; (b) and (c) atomic-resolution element maps of Fe (red) and Re (blue) elements, respectively. (d) Mixed Fe and Re maps, and (e) nanodiffraction pattern taken along ⟨011⟩pc. Reprinted with permission from Ho et al., Ultramicroscopy 193, 137 (2018). Copyright 2018 Elsevier B.V.

Close modal

The local electronic structures and geometries around the Cr and Re atoms in the double perovskite Sr2−xLnxCr1+x/2Re1−x/2O6 doped with Ln (=La, Nd, or Sm) were also investigated by x-ray absorption near edge structure (XANES) and extended x-ray absorption fine structure (EXAFS) spectra.50Figure 10(a) presents the XANES spectra of Sr2CrReO6 obtained at the Cr K-edge, where different chemical valences of Cr ions are selected such as the Cr3+ ions in LaCrO3 and Cr2O3, Cr6+ ion in CrO3, and metallic Cr. It is found that this edge is very sensitive to the local structures of the investigated samples. For example, in the CrO3 sample, a sharp peak appears in the pre-edge region at 5.993 keV, which resulted from the dipole-forbidden transition from 1s to 3d. This indicates the tetrahedral coordination of Cr6+ without an inversion center.61 As shown in Fig. 10(a), the positions of absorption edges are dependent upon the Cr oxidation states, which move to higher energy with increasing chemical valences of Cr ions. It is observed that the rising part of the Sr2CrReO6 main edge located at 6.006 keV coincides with that of LaCrO3 due to both compounds having a similar perovskite structure; thus, the chemical valence of the Cr3+ ion is determined for Sr2CrReO6. The absorption spectra for the Sr2−xLnxCr1+x/2Re1−x/2O6 samples (Ln = Nd or Sm) are identical to those of Sr2CrReO6, and their edge positions are also presented [see the inset of Fig. 10(a)], where the parent compound data at 35 K and 295 K are also presented. It is found that the edge positions maintain almost a constant value, indicating the chemical states of Cr ions are stable despite replacing the Sr element by rare earths. This means Cr3+ ions exist in all the investigated samples. To accurately understand the physical mechanisms of the pre-edge peaks of the Sr2−xNdxCr1+x/2Re1−x/2O6 samples, their pre-edge features are comparatively investigated with those of LaCrO3, as shown in Fig. 10(b). The small peak denoted A was observed in all samples with nearly the same energy and intensity, which resulted from pure quadrupole transitions that occurred in related oxides, and was much sensitive to the occupied 3d orbital.62 In this way, the presence of Cr3+ in the double perovskite of Sr2−xNdxCr1+x/2Re1−x/2O6 is confirmed. However, the other pre-edges of DP compounds have much difference from those of LaCrO3. A strong peak (denoted the B peak) was observed in Sr2CrReO6, and its intensity was decreased with the increasing rare earth-doping concentration [see the inset of Fig. 10(b) for details]. The electrical conductivities of these samples were reported to be decreased as the Ln-doped concentration was increased.63 This is consistent with the concomitant weak orbital overlapping between the Cr and oxygen atoms. Thus, the intensity of peak B reflects the covalent bond strength of the Cr–O bond. Recently, the sensitivity of the XANES spectra has been used to check the oxidation states of Cr and Re elements. The chemical shift of the Cr K edge in the XANES spectra revealed the presence of Cr3+ ions for all compounds [Fig. 11(a)], while the XANES spectra at the Re L1,2,3 edges confirm the Re5+ ions exist in all compounds [Fig. 11(b)]. This means the electronic states of both Cr3+ and Re5+ ions are not affected by Sr substitution by Ln elements. The Re–O and Cr–O bond lengths can be deduced from the EXAFS spectra, which fit well with crystallography. These bond lengths do not change along with the Ln-doping concentration, which agrees with the as expected values for Re5+ and Cr3+ oxides. As a consequence, the spectroscopic study confirms that the chemical valences of Cr and Re elements are not changed before and after substituting Sr with Ln elements.

FIG. 10.

(a) Normalized XANES spectra recorded at the Cr K-edge from metallic Cr (dotted-dashed line), Cr2O3 (dashed line), LaCrO3 (dotted line), Sr2CrReO6 (thick line), and CrO3 (thin line), where the pre-edge features in the spectrum of Sr2CrReO6 are denoted A and B, respectively. The data in the inset are the edge positions determined by the maximum of the first derivative of the normalized XANES spectra for the Sr2−xNdxCr1+x/2Re1−x/2O6 (filled squares) and Sr2−xSmxCr1+x/2Re1−x/2O6 (open squares) samples at 295 K. The edge position data for Sr2CrReO6 at 35 K and 295 K are denoted by filled and open symbols, respectively. (b) Pre-edge structures in the normalized XANES spectra of Sr2−xNdxCr1+x/2Re1−x/2O6 (x ≤ 0.5) and LaCrO3 after subtracting the background. The inset shows the area of peak B in Sr2−xNdxCr1+x/2Re1−x/2O6 and Sr2−xSmxCr1+x/2Re1−x/2O6 series, which reflects the covalent bond strength of the Cr–O bond. Reprinted with permission from Blasco et al., Phys. Rev. B 76, 144402 (2007). Copyright 2007 American Physical Society.

FIG. 10.

(a) Normalized XANES spectra recorded at the Cr K-edge from metallic Cr (dotted-dashed line), Cr2O3 (dashed line), LaCrO3 (dotted line), Sr2CrReO6 (thick line), and CrO3 (thin line), where the pre-edge features in the spectrum of Sr2CrReO6 are denoted A and B, respectively. The data in the inset are the edge positions determined by the maximum of the first derivative of the normalized XANES spectra for the Sr2−xNdxCr1+x/2Re1−x/2O6 (filled squares) and Sr2−xSmxCr1+x/2Re1−x/2O6 (open squares) samples at 295 K. The edge position data for Sr2CrReO6 at 35 K and 295 K are denoted by filled and open symbols, respectively. (b) Pre-edge structures in the normalized XANES spectra of Sr2−xNdxCr1+x/2Re1−x/2O6 (x ≤ 0.5) and LaCrO3 after subtracting the background. The inset shows the area of peak B in Sr2−xNdxCr1+x/2Re1−x/2O6 and Sr2−xSmxCr1+x/2Re1−x/2O6 series, which reflects the covalent bond strength of the Cr–O bond. Reprinted with permission from Blasco et al., Phys. Rev. B 76, 144402 (2007). Copyright 2007 American Physical Society.

Close modal
FIG. 11.

(a) Fourier transform of the k-weighted EXAFS spectra recorded at the Cr K edge for the Sr2−xSmxCrReO6 series at 35 K. Solid lines represent the best-fit simulations where the contributions between 1 Å and 4 Å are only considered in the structural analysis. (b) Fourier transform of the k-weighted EXAFS spectra taken at the Re L3 edge for the Sr2−xSmxCrReO6 series at 35 K. Solid lines represent the best-fit simulations where the contributions between 1.2 Å and 3.9 Å are only considered in the structural analysis. Reprinted with permission from Blasco et al., Phys. Rev. B 76, 144402 (2007). Copyright 2007 American Physical Society.

FIG. 11.

(a) Fourier transform of the k-weighted EXAFS spectra recorded at the Cr K edge for the Sr2−xSmxCrReO6 series at 35 K. Solid lines represent the best-fit simulations where the contributions between 1 Å and 4 Å are only considered in the structural analysis. (b) Fourier transform of the k-weighted EXAFS spectra taken at the Re L3 edge for the Sr2−xSmxCrReO6 series at 35 K. Solid lines represent the best-fit simulations where the contributions between 1.2 Å and 3.9 Å are only considered in the structural analysis. Reprinted with permission from Blasco et al., Phys. Rev. B 76, 144402 (2007). Copyright 2007 American Physical Society.

Close modal

Re-based DP thin films have promising applications in the near-future spintronics due to their high Tc value, half-metallicity, and large spin polarization. In addition, during the growth process of thin films, the strain, chemical heterogeneity, and artificial structures can be well controlled to realize the novel or enhanced properties, which are much different from those observed in the form of bulk counterparts.64–68 Recently, Sohn et al.25 grew high-quality epitaxial Sr2Fe1+xRe1-xO6 (−0.2 ≤ x ≤ 0.2) thin films by PLD. Figure 12(a) demonstrates the XRD patterns of θ–2θ scans for the films grown at 775 °C but with different PO2,andFig. 12(b) displays the off-axis XRD θ scans collected near the (111) diffraction peak of Sr2Fe1+xRe1−xO6, where only the (111) peak is observed in the films grown with PO2 ≥ 15 mTorr. The appearance of the (111) diffraction peak confirms the Fe/Re ordering, which is confirmed by the HAADF-STEM image [Fig. 12(c)] taken from the film grown at 20 mTorr in the projection along the [110] direction. In Fig. 12(c), the atomic chains of Re–Sr–Fe–Sr–Re along the [001] direction are clearly distinguishable and the atomic chains of Re–Fe–Re along the [110] direction are also clearly observed, which are sandwiched between two adjacent Sr atomic layers. Such clear atomic ordering is highlighted by the yellow diamond in Fig. 12(c), which is well consistent with the atomic positions in the projected structure of Sr2FeReO6 along the [110] direction [Fig. 12(d)]. The Re and Fe atoms are alternatively distributed in the (111) planes, as marked by blue and red lines, respectively. In another work, Hauser et al.69 grew (001)-oriented and (111)-oriented Sr2CrReO6 epitaxial thin films by magnetron sputtering, which were characterized by XRD, as shown in Fig. 13. Figure 13(a) shows XRD θ–2θ scans of (111)-oriented Sr2CrReO6 thin films grown on SrTiO3 single-crystal substrates, where only the {111} diffraction peaks from the film and substrate are observed, indicating the formation of the pure DP phase. In Fig. 13(b), Laue oscillations near the (004) diffraction peak of (001)-oriented Sr2CrReO6 thin films are clearly observed, indicating a perfect interface between the film and the substrate as well as a smooth film surface. The high crystalline quality of the Sr2CrReO6 films is reflected by their rocking curves that are shown in Fig. 13(c). The full-width-at-half-maximum (FWHM) of the rocking curve for the Sr2CrReO6 (004) peak is only 0.0088°, which is smaller than 0.0093° for the SrTiO3 (002) peak. For the Sr2CrReO6 films grown under optimal conditions, both clear Laue oscillations and sharp rocking curves reveal the high quality of the Sr2CrReO6 films.70 This is also confirmed by HAADF-STEM images. Figure 14(a) displays a low-magnification HAADF-STEM image of the Sr2CrReO6 film taken along the [110] direction, which exhibits flat surface and uniform thickness. The atomic structure of the Sr2CrReO6 film is revealed by a high-magnification HAADF-STEM image, which demonstrates the atomic chains of Re–Sr–Fe–Sr–Re along the [001] direction and the atomic chains of Re–Fe–Re along the [110] direction, confirming the clear Laue oscillations observed in Fig. 13(b). Pronounced STEM image contrast is obviously observed in the inset of Fig. 14(b), where three distinct shadows are clearly resolved, which correspond to Sr, Cr, and Re atoms, respectively. This matches well with the crystal projection of the DP lattice in the [110] direction [Fig. 14(c)]. The direct observations on the Cr/Re ordering in the Sr2CrReO6 films reveal the high quality of the film. Similarly, epitaxial growth of Sr2CrReO6 films on the SrTiO3 (001) substrate by PLD is also reported by Orna et al.33 The epitaxial growth relationship between the Sr2CrReO6 film and the SrTiO3 substrate is confirmed by high-resolution XRD patterns and STEM images, which is expressed as Sr2CrReO6 (001)[100] || SrTiO3 (001)[110]. The AS concentration in this epitaxial film was estimated to be 14%, and the Curie temperature (TC) was determined to be 481 K, smaller than that reported for sputtered samples.71 

FIG. 12.

(a) XRD θ–2θ scans for the epitaxial Sr2Fe1+xRe1−xO6 films grown on SrTiO3 at 775 °C and under different oxygen pressures. Diffraction peaks from the SrTiO3 substrate are denoted by asterisks. (b) Off-axis XRD θ–2θ scans performed at ψ = 54.7° near the Sr2Fe1+xRe1−xO6 (111) peak, confirming the Fe/Re ordering. (c) HAADF-STEM image of the epitaxial Sr2Fe1+xRe1−xO6 films grown under oxygen pressure of 20 mTorr. (d) Schematic diagram illustrating atomic distributions in the projected plane along the [110] direction. The alternating Re and Fe planes along the [111] direction are denoted by blue and red lines, respectively. The black arrow indicates the [111] direction. The yellow diamond outlined in (c) highlights the ordered distributions of Re ions. Reprinted with permission from Sohn et al., Adv. Mater. 31, 1805389 (2019). Copyright 2019 WILEY-VCH Verlag GmbH & Co. KGaA.

FIG. 12.

(a) XRD θ–2θ scans for the epitaxial Sr2Fe1+xRe1−xO6 films grown on SrTiO3 at 775 °C and under different oxygen pressures. Diffraction peaks from the SrTiO3 substrate are denoted by asterisks. (b) Off-axis XRD θ–2θ scans performed at ψ = 54.7° near the Sr2Fe1+xRe1−xO6 (111) peak, confirming the Fe/Re ordering. (c) HAADF-STEM image of the epitaxial Sr2Fe1+xRe1−xO6 films grown under oxygen pressure of 20 mTorr. (d) Schematic diagram illustrating atomic distributions in the projected plane along the [110] direction. The alternating Re and Fe planes along the [111] direction are denoted by blue and red lines, respectively. The black arrow indicates the [111] direction. The yellow diamond outlined in (c) highlights the ordered distributions of Re ions. Reprinted with permission from Sohn et al., Adv. Mater. 31, 1805389 (2019). Copyright 2019 WILEY-VCH Verlag GmbH & Co. KGaA.

Close modal
FIG. 13.

(a) XRD θ scans of the (111)-oriented Sr2CrReO6 film with a thickness of 1220 nm. The insets show details of the fitting near the Sr2CrReO6 (222) and (444) peaks, which overlap with the substrate peaks. The Cr/Re ordering parameter ξ was determined to be 0.99 by the Rietveld refinements (red or dark gray) of the XRD data of the (111)-oriented film. (b) XRD θ scans of the (001)-oriented Sr2CrReO6 film with a thickness of 190 nm. Laue oscillations near the Sr2CrReO6 (004) peak are clearly observed. The inset shows the local off-axis θ scans around the Sr2CrReO6 (022) peak performed at a tilt angle ψ of 45° in the (001)-oriented film. (c) Rocking curves of the Sr2CrReO6 (004) peak (blue) and the substrate SrTiO3 (002) peak (black) for the (001)-oriented Sr2CrReO6 (001) film with a thickness of 590 nm. Reprinted with permission from Hauser et al., Phys. Rev. B 85, 161201 (2012). Copyright 2012 American Physical Society.

FIG. 13.

(a) XRD θ scans of the (111)-oriented Sr2CrReO6 film with a thickness of 1220 nm. The insets show details of the fitting near the Sr2CrReO6 (222) and (444) peaks, which overlap with the substrate peaks. The Cr/Re ordering parameter ξ was determined to be 0.99 by the Rietveld refinements (red or dark gray) of the XRD data of the (111)-oriented film. (b) XRD θ scans of the (001)-oriented Sr2CrReO6 film with a thickness of 190 nm. Laue oscillations near the Sr2CrReO6 (004) peak are clearly observed. The inset shows the local off-axis θ scans around the Sr2CrReO6 (022) peak performed at a tilt angle ψ of 45° in the (001)-oriented film. (c) Rocking curves of the Sr2CrReO6 (004) peak (blue) and the substrate SrTiO3 (002) peak (black) for the (001)-oriented Sr2CrReO6 (001) film with a thickness of 590 nm. Reprinted with permission from Hauser et al., Phys. Rev. B 85, 161201 (2012). Copyright 2012 American Physical Society.

Close modal
FIG. 14.

(a) and (b) HAADF-STEM images of a (001)-oriented Sr2CrReO6 (SCRO) film with a thickness of 134 nm grown on the SrTiO3 (STO) substrate, viewed along the [110] direction at low-magnification and high-magnification, respectively. The inset in (b) highlights the atomic contrast of Sr, Cr, and Re atoms, matching well with (c) the atomic lattices in the projection plane along the [110] direction. Reprinted with permission from Hauser et al., Phys. Rev. B 85, 161201 (2012). Copyright 2012 American Physical Society.

FIG. 14.

(a) and (b) HAADF-STEM images of a (001)-oriented Sr2CrReO6 (SCRO) film with a thickness of 134 nm grown on the SrTiO3 (STO) substrate, viewed along the [110] direction at low-magnification and high-magnification, respectively. The inset in (b) highlights the atomic contrast of Sr, Cr, and Re atoms, matching well with (c) the atomic lattices in the projection plane along the [110] direction. Reprinted with permission from Hauser et al., Phys. Rev. B 85, 161201 (2012). Copyright 2012 American Physical Society.

Close modal

Recently, the half-metallic A2FeReO6 (A = Sr and Ba) and Sr2CrReO6 powders were synthesized by the MSS method.37Figure 15 demonstrates XRD patterns collected from the flux-synthesized and solid-state-synthesized A2FeReO6 (A = Sr and Ba) and Sr2CrReO6 powders. Rietveld refinements of these XRD data gave the crystal parameters such as unit cell parameters, atomic positions, fractional B′ and B″ site occupancies, and ordering degree. The Fe/Re ordering degree is reflected via the peak intensity of the (111) super-structural reflections in Sr2FeReO6 and Sr2CrReO6, or the analogous (101) super-structural reflection in Ba2FeReO6, and these super-structural reflections are labeled with asterisks in Fig. 15. Morphologies of the flux-prepared A2FeReO6 (A = Sr and Ba) and Sr2CrReO6 powders under different flux-to-product molar ratios were examined by SEM, and the representative SEM images are shown in Figs. 16 and 17. Aggregated block-like particles were observed in the SEM images, and their sizes changed from about 50 nm to over 1.0 µm, which were dependent upon the particular DP oxide compositions and the flux-synthesized conditions. Generally, smaller particles could be synthesized under the larger flux-to-product molar ratio and faster cooling process, or shorter reaction time. Strong temperature dependence of magnetoresistivity was found in the flux-synthesized A2FeReO6 (A = Sr and Ba) and Sr2CrReO6 powders, which are ascribed to the higher concentrations of grain boundaries present, as compared to the solid-state-synthesized samples.

FIG. 15.

XRD patterns and the corresponding Rietveld refinements for (a) Sr2FeReO6 powders synthesized at a 0.5:1 flux ratio for 12 h, (b) Ba2FeReO6 powders synthesized at a 1:1 flux ratio and radiatively cooled, and (c) Sr2CrReO6 powders synthesized at a 1:1 flux ratio. Experimental data are indicated by the circles, and the calculated patterns are denoted by the solid lines. Tick marks below the profiles indicate all the possible Bragg reflections, and the difference diffractogram is shown at the bottom. The main (111) or (101) superstructure reflections are labeled with an asterisk for each phase. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

FIG. 15.

XRD patterns and the corresponding Rietveld refinements for (a) Sr2FeReO6 powders synthesized at a 0.5:1 flux ratio for 12 h, (b) Ba2FeReO6 powders synthesized at a 1:1 flux ratio and radiatively cooled, and (c) Sr2CrReO6 powders synthesized at a 1:1 flux ratio. Experimental data are indicated by the circles, and the calculated patterns are denoted by the solid lines. Tick marks below the profiles indicate all the possible Bragg reflections, and the difference diffractogram is shown at the bottom. The main (111) or (101) superstructure reflections are labeled with an asterisk for each phase. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

Close modal
FIG. 16.

SEM images taken from the Sr2FeReO6 powders synthesized by a molten NaCl/KCl flux at a (a) 0.5:1 ratio for 3 h, (b) 0.5:1 ratio for 12 h, (c) 1:1 ratio for 3 h, and (d) 1:1 ratio for 12 h. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

FIG. 16.

SEM images taken from the Sr2FeReO6 powders synthesized by a molten NaCl/KCl flux at a (a) 0.5:1 ratio for 3 h, (b) 0.5:1 ratio for 12 h, (c) 1:1 ratio for 3 h, and (d) 1:1 ratio for 12 h. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

Close modal
FIG. 17.

SEM images taken from the Ba2FeReO6 powders synthesized by a molten NaCl/KCl flux at (a) a 1:1 ratio for 6 h, (b) a 3:1 ratio for 6 h, and (c) a 3:1 ratio for 6 h and by quenching; Sr2CrReO6 using (d) a 1:1 ratio for 6 h, (e) a 1:1 ratio and slow cooling over 24 h, and (f) a 3:1 ratio for 6 h and quenching. Samples A, B, and D were radiatively cooled. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

FIG. 17.

SEM images taken from the Ba2FeReO6 powders synthesized by a molten NaCl/KCl flux at (a) a 1:1 ratio for 6 h, (b) a 3:1 ratio for 6 h, and (c) a 3:1 ratio for 6 h and by quenching; Sr2CrReO6 using (d) a 1:1 ratio for 6 h, (e) a 1:1 ratio and slow cooling over 24 h, and (f) a 3:1 ratio for 6 h and quenching. Samples A, B, and D were radiatively cooled. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

Close modal

1. Re-based DP bulk oxides

The magnetic properties of Re-based DP bulk oxides are strongly related to the ordering of the B-site ions. Their net magnetizations are decreased owing to the existence of B-site disorder. The possible reasons can be classified as follows: (i) the B-site disorder destroys the spin arrangements in the B′ and B″ sublattices but has no influence on each magnetic moment at B′ and B″ sites and/or (ii) each magnetic moment at B′ and B″ sites decreased because of cation disorder, but the nature of the spin order at the B′ and B″ sublattices is not affected. Since the magnetic properties of Re-based DP oxides have promising applications in spintronics, therefore, their fundamental magnetic properties have received much attention in the past decades, which have been investigated by the superconducting quantum interference device (SQUID), a Quantum Design Physical Property Measurement System (PPMS) with a vibrating sample magnetometer (VSM). For example, under a static magnetic field increasing up to 30 T, high-field magnetizations of Re-based DP bulk oxides of A2FeReO6 (A = Ca, Sr, BaSr) and Sr2CrReO6 were measured, which are shown in Fig. 18.47 It is found that the magnetizations of BaSrFeReO6, Sr2FeReO6, and Ca2FeReO6 compounds measured at 4 K under 5 T are 2.98, 2.81, and 2.25 μB/f.u., respectively. At the maximum magnetic field of 30 T, the magnetization of BaSrFeReO6 [Fig. 18(a)] was measured to be 3.27 μB/f.u. This approached the saturated magnetization (Ms) since this compound was nearly saturated under 30 T, as observed in the inset of this figure. To evaluate the Ms value of this compound, the following expression is used:72,73

(2)

where Ms (expt.) represents the saturated magnetization measured experimentally and Ms represents the saturated magnetization without AS. In the present case, the AS value is 0.5% and Ms (expt.) is 3.27 μB/f.u.; therefore, the Ms value for the BaSrFeReO6 compound is determined to be 3.30 μB/f.u. The M–H loop of the BaSrFeReO6 compound was measured by a SQUID magnetometer up to 5 T, which is shown as an inset of Fig. 18(a) at the upper left corner. Magnetic field dependence of the magnetization of the BaSrFeReO6 compound measured at 4 K and 100 K is shown as an inset of Fig. 18(a) at the lower right corner. The magnetization is expected to be increased when decreasing the temperature, which is confirmed by the slightly increased magnetization at 4 K as compared with that at 100 K, which allows one to ignore any kind of spurious paramagnetic contribution to the measurement at 4 K. The magnetization of the BaSrFeReO6 compound was measured to be 3.21 μB/f.u. at 100 K and under the maximum field of 30 T, which was still much higher than the assumed value of 3.0 μB/f.u. The magnetic data for the Sr2FeReO6 compound are displayed in Fig. 18(b), which are similar to the case of BaSrFeReO6 composition. The magnetizations of the Sr2FeReO6 compound under the maximum field of 30 T were measured to be 3.23 μB/f.u. @4 K and 3.17 μB/f.u. @100 K, respectively, which exhibits a less saturated tendency than BaSrFeReO6 does. It is much more evident that the Ca2FeReO6 sample does not achieve magnetic saturation under 30 T, as demonstrated in Fig. 18(c). Its magnetizations under the maximum field of 30 T were about 3.12 μB/f.u. @ 4 K and 100 K. It is also noticed that the MS value @ 4 K is smaller than that @ 110 K across the magnetic field of 12 T–30 T, which can be ascribed to a structural transition undergone between the two monoclinic crystallographic structures in this compound below 120 K.74–76 The above experimental results clearly demonstrate that the magnetic saturation of these materials can be only achieved by much high magnetic fields. In general, the higher magnetic field required for magnetic saturation and larger coercive fields are, the larger magnetic anisotropy is the compound possesses. Therefore, Re-based double perovskites exhibit a large coercive field and saturation field, which are comparable to those observed in permanent magnets, indicating a high magneto-crystalline anisotropy. In order to understand why the highest Curie temperature (TC ≈ 635 K) is observed in the Sr2CrReO6 compound, recently x-ray magnetic circular dichroism (XMCD), an element-specific magnetic measurement technique based on synchrotron radiation, has been utilized to directly measure each magnetic moment at the B′ and B″ sublattices since XMCD technique has some advantages such as element-specific selectivity and the ability to extract the magnetic contributions from spin and orbital magnetic moments in complex magnetic materials using the magneto-optical sum rules.77,78 As an example, the XMCD signal of Sr2CrReO6 measured as a function of the applied magnetic field at the Re L2 absorption edge and 10 K is displayed in Fig. 19,79 from which the coercive field was determined to be 1.27 T. The intensity of the XMCD signal matches well with the M–H loop obtained by a SQUID at 5 K (see the inset of Fig. 19) and other measurements.27 The MS value was measured to be 0.89 μB/f.u., which is consistent with the theoretical value of 1.0 μB calculated based on a simple ionic model with Cr3+ and Re5+ ion antiferromagnetic coupling. In terms of the data of XMCD at the L2,3 edges and sum rules, the magnetic moments of Re 5d spin and orbital moments in the Sr2CrReO6 compound were determined to be −0.68 μB and +0.25 μB, respectively. The experimental results demonstrate that the Curie temperature of the A2B′B″O6 DPs scales with the magnetic moments of the non-magnetic 3d or 5d transitional metal ions at the B″ site.

FIG. 18.

(a) MH hysteresis loop of the BaSrFeReO6 compound measured at 4 K and the magnetic field up to 30 T. The bottom inset shows a zoom of the magnetization around 30 T measured at 4 K and 100 K. The line marks the expected value for the saturation magnetization without the contribution of orbital magnetization. The top inset shows a comparison between the measured magnetization by a SQUID up to 5 T and the data obtained in the high-field installation (red line). Similar results for (b) Sr2FeReO6, (c) Ca2FeReO6, and (d) Sr2CrReO6. Reprinted with permission from Teresa et al., Appl. Phys. Lett. 90, 252514 (2007). Copyright 2007 AIP Publishing LLC.

FIG. 18.

(a) MH hysteresis loop of the BaSrFeReO6 compound measured at 4 K and the magnetic field up to 30 T. The bottom inset shows a zoom of the magnetization around 30 T measured at 4 K and 100 K. The line marks the expected value for the saturation magnetization without the contribution of orbital magnetization. The top inset shows a comparison between the measured magnetization by a SQUID up to 5 T and the data obtained in the high-field installation (red line). Similar results for (b) Sr2FeReO6, (c) Ca2FeReO6, and (d) Sr2CrReO6. Reprinted with permission from Teresa et al., Appl. Phys. Lett. 90, 252514 (2007). Copyright 2007 AIP Publishing LLC.

Close modal
FIG. 19.

XMCD signal of the Sr2CrReO6 ceramics measured as a function of the magnetic field recorded at the Re L2 edge and 10 K. The inset shows the MH loop of the same sample measured by a SQUID at 5 K. Reprinted with permission from Majewski et al., Appl. Phys. Lett. 87, 202503 (2005). Copyright 2005 AIP Publishing LLC.

FIG. 19.

XMCD signal of the Sr2CrReO6 ceramics measured as a function of the magnetic field recorded at the Re L2 edge and 10 K. The inset shows the MH loop of the same sample measured by a SQUID at 5 K. Reprinted with permission from Majewski et al., Appl. Phys. Lett. 87, 202503 (2005). Copyright 2005 AIP Publishing LLC.

Close modal

The physical pressure also has an impact on the magnetic properties of Re-based DP bulk oxides such as A2FeReO6 (A = Ca, Ba). For example, at room temperature, the Ba2FeReO6 compound with a larger Ba cation at the A-site has a cubic crystal structure and exhibits both metallicity and magnetism, whereas the Ca2FeReO6 compound with a small Ca cation at the A-site has a monoclinic structure and exhibits insulating behavior.80–82 The higher Tc (=540 K) and magnetic coercive field (Hc = 9 kOe) are observed in the Ca2FeReO6 compound, indicating the co-existence of stronger exchange interactions and larger magnetic anisotropy in this compound. In Ca2FeReO6 with a monoclinic structure, the rotations of FeO6 and ReO6 octahedra around b and c axes are favorable, resulting in a smaller Fe–O–Re bonding angle of 156°.83,84 Thus, the pdπ hopping that controls the transport behavior in the un-rotated structure is decreased, resulting in the observed insulating behavior.85 The distorted monoclinic structure with a lower symmetry makes the ReO6 octahedra be a tetragonal distortion and leads to an enhanced magnetic anisotropy as compared with the Ba compound. Escanhoela et al.86 investigated the physical pressure dependence of the magnetic properties of ferrimagnetic A2FeReO6 DP oxides with A = Ca or Ba by using XAS at the Re L2,3 edge and powder diffraction technique. Figures 20(a) and 20(b) show the XMCD hysteresis loops of A2FeReO6 (A = Ba, Ca) DP oxides measured at 10 K with different pressures. It was found that the Ba2FeReO6 and Ca2FeReO6 compounds became magnetically harder when increasing the pressure. The coercive field of Ba2FeReO6 increased from 0.2 T to 1.55 T as the pressure was changed from 1.5 GPa to 22 GPa, and that of Ca2FeReO6 also increased about threefold under the pressure of 25 GPa. The pressure-dependent coercive field (Hc) is shown in Fig. 20(c) for both compounds. It is noticed that Ca2FeReO6 is under a pre-compressed state owing to the chemical pressure. Therefore, its coercive field change is smaller than that of the Ba2FeReO6 compound for a given physical pressure. It was also found that the volume compression dramatically increased the magnetic coercive field of these two polycrystalline samples at a rate of ΔHcV ∼ 150 Oe/Å3–200 Oe/Å3.

FIG. 20.

XMCD hysteresis loops of the Re element measured under different pressures with the x-ray energy adjusted near the Re L2 edge for (a) Ba2FeReO6 and (b) Ca2FeReO6 samples. Loops are normalized to the saturation values. (c) Pressure-dependent coercive field Hc shown comparatively for both samples. Reprinted with permission from Escanhoela et al., Phys. Rev. B 98, 054402 (2018). Copyright 2018 American Physical Society.

FIG. 20.

XMCD hysteresis loops of the Re element measured under different pressures with the x-ray energy adjusted near the Re L2 edge for (a) Ba2FeReO6 and (b) Ca2FeReO6 samples. Loops are normalized to the saturation values. (c) Pressure-dependent coercive field Hc shown comparatively for both samples. Reprinted with permission from Escanhoela et al., Phys. Rev. B 98, 054402 (2018). Copyright 2018 American Physical Society.

Close modal

2. Re-based DP thin films

Hauser et al.69 grew Sr2CrReO6 films by the magnetron sputtering method and measured their magnetic properties. Figure 21(a) shows the MH loop of Sr2CrReO6 (001) films at 5 K, from which the Ms was measured to be 1.29 μB/f.u., which is larger than the theoretical value of 1.0 μB/f.u. obtained from ferrimagnetic ionic alignments of Cr3+ (3d3↑) and Re5+ (5d2↓) ions. However, the present MS value is well consistent with the calculated value of 1.28 μB/f.u. that considers the contribution of spin–orbit coupling.87 This means strong spin–orbit coupling should exist in Sr2CrReO6 thin films. The coercive field (HC) of the Sr2CrReO6 (001) film at 5 K was measured to be 1.05 T, which means the Sr2CrReO6 thin film is a hard ferrimagnet at low temperature. However, at 300 K, the HC is reduced to be 890 Oe, indicating that the Sr2CrReO6 thin film is suitable for magneto-electronic applications at room temperature. Orna et al.71 also measured the magnetic properties of Sr2CrReO6 thin films grown by the PLD method. Figure 21(b) shows the saturation magnetization (Ms) dependence of the substrate temperature and the partial oxygen pressure. The Ms had the largest value ∼1.0 μB/f.u. when the substrate temperature was between 700 °C and 800 °C and the partial oxygen pressure was 2.6 × 10−4 Torr (open red squares), which is related to the AS fraction of the films. The AS concentration in the thin films was estimated to be in the order of 14%. The inset illustrates the room temperature MH loops of the Sr2CrReO6 films grown at different temperatures. The coercive fields of all the Sr2CrReO6 films are found to be very small (HC ∼ 120 Oe) [see the inset of Fig. 21(b)]; however, the coercive field of bulk Sr2CrReO6 is as high as 3.1 kOe at 300 K.73 Under high partial oxygen pressure PO2, the Ms value of Sr2CrReO6 films can be larger than that of bulk Sr2CrReO6. This is ascribed to the CrOx oxide segregation at the film surface, and a similar case was reported in the Sr2FeMoO6 films.88 

FIG. 21.

(a) In-plane M–H hysteresis loop measured at T = 5 K for a 1220-nm-thick Sr2CrReO6 (001) film. The measured saturation magnetization (Ms) and coercive field (Hc) are 1.29 μB/f.u. and 1.05 T at 5 K, respectively. The inset shows the M vs H curve measured at T = 300 K with Ms = 1.14 μB/f.u. and Hc = 890 Oe. Reprinted with permission from Hauser et al., Phys. Rev. B 85, 161201 (2012). Copyright 2012 American Physical Society. (b) Saturation magnetization (Ms) vs substrate temperature (Tsub) (with the oxygen pressure fixed at 2.6 × 10−4 Torr, open red squares) and oxygen pressure PO2 during deposition (with the substrate temperature fixed at 800 °C, closed blue circles). The inset shows the room temperature MH loops of the (001)-oriented Sr2CrReO6 films grown at different temperatures. Reprinted with permission from Orna et al., J. Magn. Magn. Mater. 322, 1217 (2010). Copyright 2009 Elsevier B.V.

FIG. 21.

(a) In-plane M–H hysteresis loop measured at T = 5 K for a 1220-nm-thick Sr2CrReO6 (001) film. The measured saturation magnetization (Ms) and coercive field (Hc) are 1.29 μB/f.u. and 1.05 T at 5 K, respectively. The inset shows the M vs H curve measured at T = 300 K with Ms = 1.14 μB/f.u. and Hc = 890 Oe. Reprinted with permission from Hauser et al., Phys. Rev. B 85, 161201 (2012). Copyright 2012 American Physical Society. (b) Saturation magnetization (Ms) vs substrate temperature (Tsub) (with the oxygen pressure fixed at 2.6 × 10−4 Torr, open red squares) and oxygen pressure PO2 during deposition (with the substrate temperature fixed at 800 °C, closed blue circles). The inset shows the room temperature MH loops of the (001)-oriented Sr2CrReO6 films grown at different temperatures. Reprinted with permission from Orna et al., J. Magn. Magn. Mater. 322, 1217 (2010). Copyright 2009 Elsevier B.V.

Close modal

3. Re-based DP powders

The magnetic properties of Sr2FeReO6, Ba2FeReO6, and Sr2CrReO6 powders synthesized by the molten salt method under different synthesized conditions were reported by Fuoco et al.37 Their MH loops are displayed in Fig. 22, which indicate the magnetization saturation is not fully achieved for these compounds under 7 T. Previously, it was reported that these materials could only achieve the magnetic saturation under magnetic field as high as 30 T.47,89 In Fig. 22(a), the magnetizations for the flux-prepared Sr2FeReO6 powders under different flux-to-product molar ratios and reaction time are measured to be in the range of 1.57–2.17 μB under 7 T, which was lower than the theoretical value of 3.0 μB obtained with the antiparallel spin alignments for the Fe3+ and Re5+ ions. Such smaller magnetizations are also bound up with the disordering degree at B′ and B″ sites.54,90 As for the flux-prepared Ba2FeReO6 powders [Fig. 22(b)], they exhibit a lower Hc and larger Ms (from 2.10 to 2.40 μB) due to a higher ordering degree of the Fe/Re site, as compared to Sr2FeReO6. The spontaneous magnetizations of the flux-prepared Sr2CrReO6 powders are in the range of 0.74–0.80 μB, lower than the expected value of 1.0 μB under the antiparallel spin alignments of the Cr3+ and Re5+ ions. It is found that the higher the ordering degree at the Cr/Re site is, the larger the values of Ms and Mr are in the flux-prepared Sr2CrReO6 powders. As a rule of thumb, the coercive field of the Sr2CrReO6 powders is much larger than that of either Sr2FeReO6 or Ba2FeReO6.

FIG. 22.

M–H loops of (a) Sr2FeReO6, (b) Ba2FeReO6, and (c) Sr2CrReO6 powders synthesized by the MSS method. The synthesized conditions in each curve are indicated by the flux-to-product molar ratio, reaction time, and cooling time if other than radiatively cooled. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

FIG. 22.

M–H loops of (a) Sr2FeReO6, (b) Ba2FeReO6, and (c) Sr2CrReO6 powders synthesized by the MSS method. The synthesized conditions in each curve are indicated by the flux-to-product molar ratio, reaction time, and cooling time if other than radiatively cooled. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

Close modal

1. Re-based DP bulk oxides

Transport property measurements of the Re-based DP oxides were performed by the four-probe method in conjunction with a Quantum Design PPMS system. With a constant applied magnetic field, the temperature dependence of resistivity can be measured. Prellier et al.53 reported the electrical properties of ferrimagnetic A2FeReO6 DP oxides with A = Ba and Ca. Figure 23(a) demonstrates the resistivity (ρ) of the Ba2FeReO6 ceramics vs the temperature under different magnetic fields (e.g., 0 kOe, 2 kOe, and 50 kOe). A metallic behavior was observed in the Ba2FeReO6 ceramics below 300 K. The inset of Fig. 23(a) presents the resistivity measured up to 385 K under no applied magnetic field, and the metallic behavior is still kept within this temperature region. On the contrary, Ca2FeReO6 exhibited a semiconducting behavior across 5 K–300 K, as depicted in Fig. 23(b). The observed metallic behavior in the Ba2FeReO6 compound is ascribed to the direct interaction between R e t2g and Re t2g, whereas the semiconducting behavior observed in the Ca2FeReO6 compound is attributed to the disrupted interaction between R e t2g and Re t2g by the monoclinic structural distortion (the degeneracy of t2g states is lifted). To investigate the magnetoresistance (MR) effects in the Sr2FeReO6 and Ba2FeReO6 compounds, the magnetic field dependent upon the MR ratios measured at different temperatures is demonstrated in Figs. 23(c) and 23(d), respectively. Here, the MR value is defined as

(3)

where Hpeak is the magnetic field corresponding to the maximal value of resistivity (ρ). It should be noticed that both Sr2FeReO6 and Ba2FeReO6 exhibit negative MR behavior, and their MR values increase fast at low fields but slowly at high fields. This effect becomes much apparent at low temperatures, whereas it becomes less evident at room temperature (300 K); thus, the MR is much smaller. This is the typical characteristic of intergrain MR.91 It is also noticed that Sr2FeReO6 has a significantly larger MR than Ba2FeReO6 under comparable conditions, which can be ascribed to the metal–insulator boundaries in the metallic Ba-based compound.92 The appearance of metallic or insulating behavior in the isoelectronic series of A2B′B″O6 DP compounds dependent upon the A-site cation would enable them useful in magnetic devices.

FIG. 23.

(a) Resistivity dependence of temperature for the Ba2FeReO6 compound measured at different magnetic fields. The inset shows the resistivity measured from 5 K to 385 K without magnetic field. (b) Resistivity dependence of temperature under zero-field for Ca2FeReO6. Reprinted with permission from Prellier et al., J. Phys.: Condens. Matter 12, 965 (2000). Copyright 2000 IOP Publishing Ltd. (c) and (d) MR behaviors of Sr2FeReO6 and Ba2FeReO6 compounds measured at different temperatures. Reprinted with permission from Gopalakrishnan et al., Phys. Rev. B 62, 9538 (2000). Copyright 2000 American Physical Society.

FIG. 23.

(a) Resistivity dependence of temperature for the Ba2FeReO6 compound measured at different magnetic fields. The inset shows the resistivity measured from 5 K to 385 K without magnetic field. (b) Resistivity dependence of temperature under zero-field for Ca2FeReO6. Reprinted with permission from Prellier et al., J. Phys.: Condens. Matter 12, 965 (2000). Copyright 2000 IOP Publishing Ltd. (c) and (d) MR behaviors of Sr2FeReO6 and Ba2FeReO6 compounds measured at different temperatures. Reprinted with permission from Gopalakrishnan et al., Phys. Rev. B 62, 9538 (2000). Copyright 2000 American Physical Society.

Close modal

The transport properties of Re-based ordered DPs such as A2MReO6, where A is Sr, Ca and M is Mg, Sc, Cr, Mn, Fe, Co, Ni, Zn, are also reported by Kato et al.27 The temperature dependence of resistivities (ρ) is shown in Fig. 24. It is noticed that the A2MReO6 compounds with A = Sr exhibit higher conductivity than the Ca-based ones. Insulating nature was observed in all the compounds except for the two compounds of Sr2FeReO6 and Sr2CrReO6, which are fallen into the classification of half-metals.91,93 It is observed in Fig. 24(a) that at room temperature, the resistivities of a series Sr2MReO6 DPs (M being Mg, Sc, Mn, Ni, and Zn) are in the order of kΩ⋅cm, which has nothing to do with A-site ions. However, the Sr2MReO6 DPs with M = Co, Cr, and Fe have relatively low resistivity. In the case of Ca2FeReO6 compound, a temperature-driven metal–insulator transition appeared around 150 K, which was reflected by a turning point in the resistivity curve marked by a triangle in Fig. 24(b), similar to that reported previously.74Figure 25 demonstrates the temperature dependent magnetizations of A2MReO6 (A being Sr, Ca; M being Cr, Mn, Fe, and Ni) DPs, which were measured under 1 T. The Ms values of Sr2MnReO6 and Sr2NiReO6 compounds were measured to be 2.0 and 1.0 μB/f.u., respectively, which matched well with the expected values obtained from the antiferromagnetic coupling of Mn2+ (or Ni2+) and Re6+ ions. However, the measured magnetizations of Ca2MnReO6 and Ca2NiReO6 at 1 T were much smaller than the calculated values from the antiferromagnetic coupling between Mn2+ (or Ni2+) and Re6+ ions. This may be ascribed to the large coercive fields of these two compounds (e.g., 4.0 T for Ca2MnReO6 and 2.8 T for Ca2NiReO6, respectively) and the insulating nature as well as relatively low magnetic transition temperature (Tc < 150 K). On the other hand, Sr2CrReO6 and Sr2FeReO6 compounds exhibit metallic behavior and have high magnetic transition temperatures (Tc = 635 K and 400 K). This can be ascribed to that the exchange interactions between M (= Cr, Fe) and Re spins are not disturbed and there is no gap around the Fermi level. The saturation magnetizations of Sr2CrReO6 and Sr2FeReO6 compounds were measured to be 0.56 and 2.2 μB/f.u., respectively, lower than their theoretical values of 1.0 and 3.0 μB/f.u. calculated based on the electronic configurations of Cr3+ (or Fe3+) and Re5+ ions within a simplest ionic model. Similarly, the Ca2MReO6 (M being Cr, Mn, Fe, and Ni) DPs possess comparable saturation magnetization and higher magnetic TC, as compared with the corresponding Sr-based compounds. A metal–insulator transition is observed in the ordered A2MReO6 DPs (A being Sr, Ca; M being Cr, Mn, Fe, and Ni). This phenomenon can be ascribed to the formation of the half-metallic conduction band due to the hybridization between the down-spin t2g bands of Re and M (=Cr, Fe) Cr3+ around the Fermi level. This intuitive interpretation is proven by theoretical electronic calculations of Sr2FeReO6.91 

FIG. 24.

Temperature dependence of resistivity (ρ) of the ordered double perovskites (a) Sr2MReO6 (M = Mg, Sc, Cr, Mn, Fe, Co, Ni, and Zn) and (b) Ca2MReO6 (M = Cr, Mn, Fe, Co, and Ni). Reprinted with permission from Kato et al., Phys. Rev. B 69, 184412 (2004). Copyright 2004 American Physical Society.

FIG. 24.

Temperature dependence of resistivity (ρ) of the ordered double perovskites (a) Sr2MReO6 (M = Mg, Sc, Cr, Mn, Fe, Co, Ni, and Zn) and (b) Ca2MReO6 (M = Cr, Mn, Fe, Co, and Ni). Reprinted with permission from Kato et al., Phys. Rev. B 69, 184412 (2004). Copyright 2004 American Physical Society.

Close modal
FIG. 25.

Magnetization dependence of temperature under 1 T for Sr2MReO6 and Ca2MReO6 compounds (M = Cr, Mn, Fe, and Ni). Reprinted with permission from Kato et al., Phys. Rev. B 69, 184412 (2004). Copyright 2004 American Physical Society.

FIG. 25.

Magnetization dependence of temperature under 1 T for Sr2MReO6 and Ca2MReO6 compounds (M = Cr, Mn, Fe, and Ni). Reprinted with permission from Kato et al., Phys. Rev. B 69, 184412 (2004). Copyright 2004 American Physical Society.

Close modal

In the cases of Ca2FeReO6 and Ca2CrReO6 compounds, their small structural distortion drives the one-electron bandwidth to be narrower than that in their Sr-based analogs, leading a Mott transition in these compounds. For the Sr2MnReO6 and Sr2NiReO6 compounds, their highest occupied states are the up-spin eg states for Mn2+ and Ni2+ ions, respectively. To virtually excite the M3+–Re5+ state, the up-spin eg electrons of M2+ ions are required to move toward the down-spin t2g states of Re6+ ions. Due to the lack of mixing states between t2g and eg at the adjacent sites in a perfect cubic perovskite, the above electronic transferring process is nearly prohibited in ordered DPs, which makes the Sr2MnReO6 and Sr2NiReO6 compounds exhibit insulating behavior and used as Mott insulators.

2. Re-based DP thin films

The transport properties of the Sr2CrReO6 films were reported. Figure 26(a) shows the magnetization ratio of M/M(5 K) vs temperature, which is measured under 8 kOe. The inverse magnetic susceptibility (χ−1) vs temperature is demonstrated in the inset, showing two magnetic transition temperatures (TC = 508 K for the majority phase and TC = 595 K for a secondary phase). In comparison, the TC for bulk Sr2CrReO6 is reported to be 620 K–635 K24,91 and 481 K for Sr2CrReO6 films.33 The observed higher TC may be contributed from the local regions around a small number of APBs, where there exists a stronger exchange interaction, enhancing the TC value. Figure 26(b) demonstrates the semi-log plots of the electrical resistivity of bulk Sr2CrReO6 and thin films vs temperature. The room temperature bulk resistivity was measured to be 2.1 mΩ cm. With the decreasing temperature, the electrical resistivity increases, as reported by Kato et al.93 In the Sr2CrReO6 film, the resistivity was increased more than two orders of magnitude (from 16.2 mΩ cm to 5.05 Ω cm) as the temperature was decreased from 300 K to 2 K. Such a behavior is the feature of a semiconductor (or insulator) with a gap at the Fermi level. The ln ρ vs 1000/T plot across the temperature range of 90 K–200 K is shown in the inset of Fig. 26(b), which exhibits nearly perfect linear fitting. Such temperature dependence of resistivity is frequently observed in semiconductors owing to their thermal activation,94,95 which is given by

(4)

where ρ0 is the prefactor and Ea is the activation energy. The Ea value was determined to be 9.4 meV from the linear fitting of the ln ρ vs 1000/T plot across the temperature range of 90 K–200 K. However, at temperatures below 90 K, the ln ρ vs 1000/T plot does not exhibit a linear behavior; instead, a good linear fitting is found for ln ρ vs (1000/T)0.25 across a temperature range of 55 K–90 K, and for ln ρ vs (1000/T)0.20 from 6 K to 25 K. This indicates a variable-range hopping model is appropriate in the whole temperature range (6 K–200 K), similar to those reported in the doped-semiconductors at low temperatures.96 Sohn et al.25 reported the transport properties of cation-ordered Sr2Fe1+xRe1−xO6 DP films. Figure 27(a) displays the magnetization (M) vs T curves measured under 1000 Oe for the Sr2Fe1+xRe1−xO6 films with different stoichiometric ratios, and Fig. 27(b) shows the corresponding temperature-dependent sheet resistance RS. It was found that the ferromagnetic ground state existed in the Fe-rich/ordered films, but it fully vanished in the Re-rich/disordered films. In addition, the metallic ground state was more vulnerable in the films with excess Fe ions than that with excess Re ions. The Fe-rich/ordered films (blue lines) underwent a magnetic transition around 400 K, but it completely disappeared in the Re-rich/disordered films (red lines). This was confirmed by the magnetization exhibiting no temperature dependence. The disappearance of long-range magnetic ordering in the Re-rich films was attributed to the cation disordering at the B-site, which substantially influenced the magnetic properties of DP oxides. In contrast to the magnetism, the Re-rich/disordered films exhibit the metallic ground state, as evident by the RS(T) curve in Fig. 27(b). However, the Fe-rich films behave like an insulator, and their room-T resistivity is about three orders of magnitude larger than that of the stoichiometric or Re-rich samples. Therefore, in the present Sr2Fe1+xRe1−xO6 single-crystalline films, a metal–insulator transition is observed, which exhibits much more dramatic phenomena than that reported for polycrystalline Sr2Fe1+xRe1−xO6 ceramics.97 The reason can be ascribed to the existence of many metallic grain boundaries in polycrystalline films but absence in the present films. Figure 27(c) summarizes the measured results for the Ms and Rs values of the Sr2Fe1+xRe1−xO6 DP films; based on them, the Sr2Fe1+xRe1−xO6 single-crystalline films can be classified as paramagnetic metal (PMM, x < 0), ferromagnetic metal (FMM, x ≈ 0), and room-temperature ferromagnetic insulator (FMI, x > 0) catalogs.

FIG. 26.

(a) Magnetization (M) dependence of temperature for a 1220-nm-thick Sr2CrReO6 (001) film measured at H = 8 kOe. The inset in (a) shows the temperature dependent inverse magnetic susceptibility (χ−1). (b) Semi-log plots of ρ vs T of bulk Sr2CrReO6 (black) and a 200-nm-thick Sr2CrReO6 (001) film (blue) measured by the four-probe measured method. The inset in (b) shows a plot of ln ρ vs 1000/T, showing a good linear fitting with a linear correlation coefficient R equal to 0.999 99. The activation energy (Ea) is deduced to be 9.4 meV. Reprinted with permission from Hauser et al., Phys. Rev. B 85, 161201 (2012). Copyright 2012 American Physical Society.

FIG. 26.

(a) Magnetization (M) dependence of temperature for a 1220-nm-thick Sr2CrReO6 (001) film measured at H = 8 kOe. The inset in (a) shows the temperature dependent inverse magnetic susceptibility (χ−1). (b) Semi-log plots of ρ vs T of bulk Sr2CrReO6 (black) and a 200-nm-thick Sr2CrReO6 (001) film (blue) measured by the four-probe measured method. The inset in (b) shows a plot of ln ρ vs 1000/T, showing a good linear fitting with a linear correlation coefficient R equal to 0.999 99. The activation energy (Ea) is deduced to be 9.4 meV. Reprinted with permission from Hauser et al., Phys. Rev. B 85, 161201 (2012). Copyright 2012 American Physical Society.

Close modal
FIG. 27.

(a) MT curves of the Sr2Fe1+xRe1−xO6 films measured under 1000 Oe, where the red, green, and blue lines indicate the MT curves of Re-rich, stoichiometric, and Fe-rich films, respectively. (b) Semi-log plot of the sheet resistance RS as a function of temperature. Re-rich and stoichiometric films exhibit metallic behaviors, whereas the Fe-rich film exhibits an insulating behavior. (c) RS (300 K) and MS dependence of the x value in the Sr2Fe1+xRe1−xO6 films, from which the Sr2Fe1+xRe1−xO6 single-crystalline films are classified as paramagnetic metal (PMM, x < 0), ferromagnetic metal (FMM, x ≈ 0), and room-temperature ferromagnetic insulator (FMI, x > 0) catalogs, respectively. Reprinted with permission from Sohn et al., Adv. Mater. 31, 1805389 (2019). Copyright 2019 WILEY-VCH Verlag GmbH & Co. KGaA.

FIG. 27.

(a) MT curves of the Sr2Fe1+xRe1−xO6 films measured under 1000 Oe, where the red, green, and blue lines indicate the MT curves of Re-rich, stoichiometric, and Fe-rich films, respectively. (b) Semi-log plot of the sheet resistance RS as a function of temperature. Re-rich and stoichiometric films exhibit metallic behaviors, whereas the Fe-rich film exhibits an insulating behavior. (c) RS (300 K) and MS dependence of the x value in the Sr2Fe1+xRe1−xO6 films, from which the Sr2Fe1+xRe1−xO6 single-crystalline films are classified as paramagnetic metal (PMM, x < 0), ferromagnetic metal (FMM, x ≈ 0), and room-temperature ferromagnetic insulator (FMI, x > 0) catalogs, respectively. Reprinted with permission from Sohn et al., Adv. Mater. 31, 1805389 (2019). Copyright 2019 WILEY-VCH Verlag GmbH & Co. KGaA.

Close modal

3. Re-based DP oxide powders

Fuoco et al.37 reported the transport properties of Sr2FeReO6, Ba2FeReO6, and Sr2CrReO6 powders synthesized by the MSS method, which were measured by polycrystalline pellets formed by pressing the powders but no annealing. Figure 28(a) shows their temperature dependent resistivity. A semiconducting-type behavior was observed in all the samples, which is consistent with previous reports. A stronger temperature dependence of resistivity is observed in the flux-prepared Sr2FeReO6 samples, which have resistivity about three times higher than that of the sample synthesized by the solid-state reaction. The flux-prepared Ba2FeReO6 powders had a resistivity in the range of ∼0.07 Ω cm to 0.1 Ω cm across 300 K–550 K, and the Sr2CrReO6 powders had a higher resistivity (∼13 Ω cm to 3 Ω cm). Figure 28(b) demonstrates the temperature dependent resistivity of Sr2FeReO6 powders synthesized by the solid-state method and flux-synthetic route, which was measured under magnetic fields of 0 T and 0.3 T across temperatures of 300 K–550 K. It was found that the flux-prepared Sr2FeReO6 powders had larger resistivity, which also exhibited much evident magnetic field-dependence of resistivity. This phenomenon can be attributed to the high concentrations of insulating grain boundaries in the flux-prepared Sr2FeReO6 powders. At room temperature, both Sr2CrReO6 and Ba2FeReO6 powders exhibit a similarly large intergrain tunneling magnetoresistance (ITMR) of up to ∼70% and ∼65%, respectively, whose details are shown in Figs. 29(a) and 29(b), respectively. Such polycrystalline MR at low magnetic field is larger than that reported previously for Sr2FeMoO6 (ITMR of ∼20% to 30% at 0.4 T) prepared from nanoscale particles of ∼29 nm to 45 nm.34 

FIG. 28.

(a) Resistivity dependence of temperature for Sr2FeReO6 powders synthesized by the solid-state reaction and MSS method under a 1:1 flux-to-product molar ratio for 3 h, Ba2FeReO6 synthesized at a 1:1 flux and radiatively cooled, and Sr2CrReO6 synthesized at a 1:1 flux for 6 h and radiatively cooled. (b) Resistivity dependence of temperature for the Sr2FeReO6 particles prepared by the solid-state reaction and flux-synthetic routes (1:1 M ratio, 3 h) under different magnetic fields. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

FIG. 28.

(a) Resistivity dependence of temperature for Sr2FeReO6 powders synthesized by the solid-state reaction and MSS method under a 1:1 flux-to-product molar ratio for 3 h, Ba2FeReO6 synthesized at a 1:1 flux and radiatively cooled, and Sr2CrReO6 synthesized at a 1:1 flux for 6 h and radiatively cooled. (b) Resistivity dependence of temperature for the Sr2FeReO6 particles prepared by the solid-state reaction and flux-synthetic routes (1:1 M ratio, 3 h) under different magnetic fields. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

Close modal
FIG. 29.

Temperature dependent resistivity measurements at 0 T and 0.3 T for (a) Sr2CrReO6 and (b) Ba2FeReO6 particles prepared by a NaCl/KCl flux synthesis at 800 °C for 6 h using a 1:1 flux:product molar ratio and radiatively cooled. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

FIG. 29.

Temperature dependent resistivity measurements at 0 T and 0.3 T for (a) Sr2CrReO6 and (b) Ba2FeReO6 particles prepared by a NaCl/KCl flux synthesis at 800 °C for 6 h using a 1:1 flux:product molar ratio and radiatively cooled. Reprinted with permission from Fuoco et al., Chem. Mater. 23, 5409 (2011). Copyright 2011 American Chemical Society.

Close modal

The optical behaviors of the Re-based DP oxides have promising applications in various optical devices. It is reported that the optical properties of Re-based DP oxides are controlled by their inter-band transitions between the B-site (B′ or B″) transition metal cations and the oxygen anions, which reflect the electronic states’ energy distribution in valence/conduction bands.98–100 Jeon et al.101 measured the reflectance spectra of Ba2FeReO6 and Ca2FeReO6 polycrystalline samples at 300 K across the energy range of 5.0 meV–6.0 eV, as shown in Fig. 30(a), from which the optical conductivity spectra σ(ω) of the Ba2FeReO6 and Ca2FeReO6 samples can be obtained by the Kramers–Kronig (KK) transformation, which is shown in Fig. 30(b). Two spectral peaks are clearly observed in these two samples, which are marked by α and β for Ba2FeReO6 and A, B for Ca2FeReO6. It is noticed that the intensities of the peaks (β and B) appearing on the higher energy side are much larger than those of the peaks (α and A) located on the lower energy side. It is reported that the higher energy (β and B) peaks resulted from the electric dipole allowed charge transitions from the O 2p to the Re 5d or Fe 3d orbital states, whereas the lower energy (α and A) peaks are contributed from the dd electronic transition between the Re 5d and/or Fe 3d states.102 It is also noticed that the position of the B peak in Ca2FeReO6 has a chemical shift (∼0.5 eV) toward higher energy as compared with that of Ba2FeReO6, as marked by the arrow in Fig. 30(b). This indicates the projected density of states (DOS) of the unoccupied Re t2g bands in Ca2FeReO6 is higher than that in Ba2FeReO6. The dc conductivity data of Ba2FeReO6 and Ca2FeReO6 polycrystalline samples are also given in Fig. 30(b), as marked by black and red squares. At low-frequency, σ(ω) exhibits much differences between the two samples. However, as the photon energy was beyond 2.0 eV, the σ(ω) of Ca2FeReO6 became much suppressed as compared with that of the metallic Ba2FeReO6, but the σ(ω) still had a finite Drude component at room temperature. To solve this issue, low temperature (10 K) conductivity spectra of Ca2FeReO6 were measured in comparison with those at 300 K, as shown in Fig. 30(c). The low-frequency σ(ω) became more suppressed, the spectral peak was much narrow, and a gap opened at 10 K.103 Such gap opening indicates the existence of an insulating ground state in the present Ca2FeReO6 sample.

FIG. 30.

(a) Experimentally measured reflectance spectra of Ba2FeReO6 (black line) and Ca2FeReO6 (red line) at 300 K. (b) Optical conductivity spectra σ(ω) of Ba2FeReO6 (black line) and Ca2FeReO6 (red line), where the black and red squares represent the dc conductivities of Ba2FeReO6 and Ca2FeReO6, respectively. (c) Optical conductivity spectra σ(ω) of Ca2FeReO6 measured at 300 K and 10 K are denoted by red solid and blue dashed lines, respectively. Reprinted with permission from Jeon et al., J. Phys.: Condens. Matter 22, 345602 (2010). Copyright 2010 IOP Publishing Ltd.

FIG. 30.

(a) Experimentally measured reflectance spectra of Ba2FeReO6 (black line) and Ca2FeReO6 (red line) at 300 K. (b) Optical conductivity spectra σ(ω) of Ba2FeReO6 (black line) and Ca2FeReO6 (red line), where the black and red squares represent the dc conductivities of Ba2FeReO6 and Ca2FeReO6, respectively. (c) Optical conductivity spectra σ(ω) of Ca2FeReO6 measured at 300 K and 10 K are denoted by red solid and blue dashed lines, respectively. Reprinted with permission from Jeon et al., J. Phys.: Condens. Matter 22, 345602 (2010). Copyright 2010 IOP Publishing Ltd.

Close modal

Sohn et al.25 measured the real part of optical conductivities σ1(ω) of the Sr2Fe1+xRe1−xO6 films by a spectroscopic ellipsometer across the energy of 1.2 eV–5 eV, as shown in Fig. 31(a). The inset shows the local density of states (DOS) for three bonding types (Fe3+–Re5+, Fe3+–Fe3+, Fe3+–Re6+) in Fe-rich films based on the density functional theoretical calculations.104–106 For the original Re5+–Fe3+ bonding of stoichiometric Sr2FeReO6, two types of optical transitions can be expected: (i) dd optical transition from Re 5d to Fe 3d orbitals [indicated by A in the inset of Fig. 31(a)] and (ii) p–d optical transition from O 2p to Re/Fe d orbitals [indicated by B in the inset of Fig. 31(a)]. As expected, two apparent spectral features are observed, one is a broad spectral band located below 2.5 eV (denoted A) and another one is a strong spectra peak located near 4.0 eV (denoted B) in the σ1(ω) of a stoichiometric film (green line, FMM film), which is consistent with the previous bulk data.101 However, in the Fe-rich film (blue line, FMI film), a new spectral peak C appeared at the expense of weights of A and B peaks. Peak C could be originated from the optical transition from O 2p to Fe 3d orbitals in the Fe3+–Fe3+ bonding, similar to the case of La3+Fe3+O3.107 However, any absorption peak expected from the Re6+–Fe3+ bonding, i.e., optical transition from O 2p to Re/Fe d orbitals, is not observed in the present energy regime. This is attributed to that the O 2p states of the Fe3+–Re6+ bonding are much far away from EF than those of other two bonding states with strong O 2p–Re 5d1 hybridization; thus, the corresponding optical transition requires higher energy than the present energy regime. Hauser et al.32 measured the FTIR transmission spectra of a 200-nm thick Sr2CrReO6 (001) film deposited on the SrTiO3 substrate. Its Tauc’s plot of (αE)2vs E is shown in Fig. 31(b), where α is the absorption coefficient and E is the incident photon energy.108 A linear fitting of the Tauc plot and its intercept of the extension of the linear part on the x axis clearly indicate that the direct bandgap (Eg) of the Sr2CrReO6 film is 0.21 eV. These features observed below 0.17 eV likely resulted from phonon modes. The noise appearing around 0.29 eV and in the energy range of 0.44 eV–0.50 eV contributes to water vapor and CO2 present in the FTIR chamber. The present FTIR spectrum indicates that Sr2CrReO6 is a semiconductor rather than a metal. Therefore, high temperature ferrimagnetism and semi-conductivity of the Sr2CrReO6 thin films enable them to find promising applications in the fields of spin filters, magnetic proximity switches, and energy efficient quantum electronic devices.

FIG. 31.

(a) Optical conductivity spectra of the Sr2Fe1+xRe1−xO6 films: Re-rich (x < 0, PMM), stoichiometric (x ≈ 0, FMM), and Fe-rich (x > 0, FMI) films. The inset schematically shows the local density of states (DOS) in the Fe-rich film with the Re5+–Fe3+, Fe3+–Fe3+, and Re6+–Fe3+ bondings calculated by density functional theoretical calculations. Three expected optical transitions in the Re5+–Fe3+ and Fe3+–Fe3+ bondings are marked A, B, and C, respectively. Reprinted with permission from Sohn et al., Adv. Mater. 31, 1805389 (2019). Copyright 2019 WILEY-VCH Verlag GmbH & Co. KGaA. (b) Tauc plot of (αE)2 vs E of a 200-nm-thick Sr2CrReO6 (001) film, which gives a clear bandgap Eg = 0.21 eV. Reprinted with permission from Hauser et al., Phys. Rev. B 85, 161201 (2012). Copyright 2012 American Physical Society.

FIG. 31.

(a) Optical conductivity spectra of the Sr2Fe1+xRe1−xO6 films: Re-rich (x < 0, PMM), stoichiometric (x ≈ 0, FMM), and Fe-rich (x > 0, FMI) films. The inset schematically shows the local density of states (DOS) in the Fe-rich film with the Re5+–Fe3+, Fe3+–Fe3+, and Re6+–Fe3+ bondings calculated by density functional theoretical calculations. Three expected optical transitions in the Re5+–Fe3+ and Fe3+–Fe3+ bondings are marked A, B, and C, respectively. Reprinted with permission from Sohn et al., Adv. Mater. 31, 1805389 (2019). Copyright 2019 WILEY-VCH Verlag GmbH & Co. KGaA. (b) Tauc plot of (αE)2 vs E of a 200-nm-thick Sr2CrReO6 (001) film, which gives a clear bandgap Eg = 0.21 eV. Reprinted with permission from Hauser et al., Phys. Rev. B 85, 161201 (2012). Copyright 2012 American Physical Society.

Close modal

In parallel with experimental investigations of Re-based DP oxides, theoretical calculations have also been carried out, allowing one to study electronic structures and structural aspects of Re-based DP oxides in theory. So far, several ab initio calculations such as the full potential linearized augmented plane-wave (FP-LAPW) method that is based on generalized gradient approximation (GGA),91,109–111 local density approximation (LDA), LDA + U (the Hubbard U term), and LAD + U + SOC (spin–orbital coupling) calculations,101 local spin density approximation (LSDA) and LSDA + U methods,112 and density functional theory (DFT) + U calculations113 are developed to investigate electronic, optical, mechanical, structural, and thermodynamic properties of Re-based DP oxides. As an example, the DOS of Sr2FeReO6 calculated by the FP-LAPW method based on GGA is shown in Fig. 32.91 It is found that at the Fermi level (EF), the DOS exists only in the down-spin band, whereas there is no DOS in the up-spin band. In another word, there exists an energy gap (∼1 eV) between the occupied Fe eg band and the unoccupied Re t2g band in the up-spin state [see the middle part of Fig. 32]. The nature of half-metallicity means the carriers in half-metals are completely spin-polarized charge carriers in the ground state. In an ideal ferromagnetic half-metal, at the Fermi level between the two spin (up and down) bands, only one spin band is partially occupied, while the other does not have DOS across the Fermi level. Thus, the electrical conduction is controlled only by one spin direction of the carriers. In the half-metal Sr2FeReO6 compound, theoretical calculations demonstrated that in the occupied up-spin band just below the EF, there mainly exist the Fe 3d electrons with the localized spins on the Fe sites, whereas, in the down-spin band around EF, there mainly exist the hybridized Re 5d t2g and Fe 3d t2g states. In addition, at the Fermi level, the partial DOS of Re (t2g) is 3–4 times larger than that of Fe (t2g). This implies the electrical conduction of the Sr2FeReO6 compound is dominated by the 5d conduction electrons (Re5+ = 5d2). In order to understand the importance of the on-site Coulomb interaction energy U (4.0 eV–5.0 eV), the crystal-field splitting energy, Δ (2.0 eV–3.0 eV), and the SOC (spin–orbit coupling, 0.3 eV–0.4 eV) in the electronic structures of the A2FeReO6 DP oxides with A = Ca and Ba, the total energy and electronic band structures of the A2FeReO6 DP oxides with A = Ca and Ba were calculated by LDA, LDA + U, and LAD + U + SOC methods.101Figures 33(a) and 33(b) show the calculated results for Ba2FeReO6 by LDA and LDA + U methods, respectively. Different electronic structures are observed due to the different spin directions. It is found that a finite DOS exists at the Fermi energy in the down-spin bands (marked by red dashed lines), whereas, in the up-spin bands, a gap near the Fermi level is observed (marked by black solid lines). This indicates the characteristic electronic feature of a half-metal.5 Similar results calculated by the LDA + U method were reported for Sr2FeMoO6.114,115 However, when the SOC becomes much strong in the Re-based DP oxides, the mixture of spin states and the half-metallicity can be neglected.116,117Figure 33(c) shows the calculated electronic structure of Ba2FeReO6 by using the LDA + U + SOC method. It is noticed that the SOC term leads to much changes in the electronic structure. In comparison with Fig. 33(b), great changes in the electronic states are observed in Fig. 33(c), especially that two bands are nearly parallel to each other around the Fermi energy, as indicated by blue and magenta lines. Therefore, the SOC plays a crucial role in determining the electronic structures of Ba2FeReO6. Similarly, band structures of Ca2FeReO6 with lattice distortion are also calculated by LDA, LDA + U, and LAD + U + SOC methods, which are shown in Figs. 34(a)–34(c).101 In Fig. 34(b), the LDA + U calculation predicts a semi-metallic ground state in Ca2FeReO6, in which the conduction band and valence band just get in touch with the Fermi level. Figure 34(c) shows the calculated electronic structure of Ca2FeReO6 by using the LDA + U + SOC method. However, no significant change is observed in the band dispersion as the SOC term is included. This means the U term is crucial, whereas the SOC term is less important in determining the electronic structure of Ca2FeReO6. To understand the role of lattice distortion in the electronic structure of Ca2FeReO6, theoretical calculations are also carried out by assuming that the bond angle of the Fe–O-Re bond in the Ca2FeReO6 compound is equal to 180° rather than the experimental value of 153°. Figures 34(d)–34(f) show the calculated band structures of Ca2FeReO6 by LDA, LDA + U, and LAD + U + SOC methods without considering lattice distortion. It is noticed that the band dispersions of Ca2FeReO6 are almost the same as those of Ba2FeReO6 [see Figs. 33(a)–33(c)]. Such a similarity indicates that the lattice distortion has a great influence on the electronic structure of a real Ca2FeReO6 compound. Therefore, in the real Ca2FeReO6 compound, the lattice distortion plays an important role in determining the electronic structure. In contrast, there is very little lattice distortion in Ba2FeReO6, and its electronic structure is mainly controlled by electron correlation and SOC. Similar conclusions are also drawn in the A2FeReO6 (A = Ca, Sr, and Ba) oxides based on the calculations by the LSDA and LSDA + U methods.112 To understand the correlation effects on the Re site, the partial DOSs of the Re 5d orbital in the A2FeReO6 (A = Ca, Sr, and Ba) oxides are calculated based on fully relativistic Dirac approximation, as shown in Fig. 35. It is found that Ca2FeReO6 has a much smaller Re bandwidth than the two compounds, A2FeReO6 (A = Sr and Ba). Therefore, a strong electron correlation should exist in the Re 5d bands in the Ca2FeReO6 oxide, leading to much narrow bands near the Fermi level with pseudo-gaps just above it. These bands separate from almost disconnect states that are dominated by Fe contributions above the Fermi level. A similar case happens for the Fe 3d energy bands. It is found that the Fe t2g bandwidth is decreased by about 25% from Sr2FeReO6 to Ca2FeReO6 compounds. The Fe and Re d bandwidths are different among the A2FeReO6 (A being Ca, Sr, and Ba) compounds, which can be ascribed to their different crystal distortions. The John–Teller distortions are increased in the series of A2FeReO6 (A = Ba, Sr, and Ca) oxides. The bond angle of the Fe–O–Re bond in the monoclinic Ca2FeReO6 is about 156°,92 which is much smaller than 180° in the cubic Ba2FeReO6 and tetragonal Sr2FeReO6. Such low bond angle makes the Fe–Re bond overlapping decrease and the t2g bandwidths become narrowed. In addition, the monoclinic structural distortion in the Ca2FeReO6 oxide lifts the degeneracy of the t2g levels on the Fe and Re sites. The above two factors result in more narrow energy bands for Fe and Re t2g in the Ca2FeReO6 oxides, as compared with the A2FeReO6 (A = Sr and Ba) oxides. Thus, the Fe and Re t2g electrons become much more localized and Ca2FeReO6 undergoes a Mott transition. To understand the electronic, magnetic, and structural properties of the A2FeReO6 (A being Ba, Sr, and Ca) oxides, x-ray absorption spectra (XAS) and x-ray magnetic circular dichroism (XMCD) spectra are theoretically investigated at the Re, Fe, and Ba L2,3 and Fe, Ca, and O K edges by the LSDA + U method. Figure 36 shows the calculated XAS and XMCD spectra at the Re L2,3 edges from the A2FeReO6 (A = Ba, Sr, and Ca) oxides and their corresponding experimental spectra.118,119 It is observed that the Re L3 spectra (contributed from 2p3/2 → 5d3/2,5/2 transition) have almost the same similarity and exhibit two peaks with the same edge energy position in A2FeReO6 (A being Ba, Sr, and Ca) oxides. The first peak appeared at about 10 540.1 eV, and the stronger one appeared at 10 543.6 eV. The split energy is attributed to the crystal-field splitting of d orbitals into t2g and eg states.120 The Re L2 XAS are composed of two peaks, and the low-energy peak exhibits relatively stronger intensity than the high-energy one. Theoretical calculations demonstrate an inverse peak relative intensity for the L3 and L2 XAS. The experimental dichroic L2 lines of the A2FeReO6 (A being Ba, Sr, and Ca) oxides exhibit an asymmetric negative peak with a shoulder appearing on the higher energy side. Three peaks are found in the dichroic lines at the L3 edge, which are two positive peaks with high energy and a negative one with low energy. This negative peak observed in Ba2FeReO6 and Sr2FeReO6 compounds embodies as a shoulder at low energy, but it exhibits a much large intensity in Ca2FeReO6. The dichroism at the L2 edge exhibits much stronger intensity than that at the L3 edge in all three compounds. The XMCD spectra can be qualitatively interpreted by analyzing the corresponding selection rules, orbital character, and occupation numbers of individual 5d orbitals.112 The theoretical calculations match relatively well with the experimental XAS and XMCD spectra in the A2FeReO6 (A = Ba, Sr, and Ca) oxides in both the spectral shape and relative intensities. Indeed, theoretical calculated techniques now allow one to fundamentally understand the physical properties and structural aspects of Re-based DP oxides, and now, they become important and complementary methods to the experimental probes; in particular in some cases, the physical properties of Re-based DP oxides are only directly obtained by some theoretical calculations. In parallel with much advances in theoretical computational models, the experimental scientists can grow Re-based DP oxide films by the laser-MBE method with the film growth controlled at the atomic scale and measure local physical properties at the nanoscale. It is expected, in the near future, the two trends have more chance to perform a lively theoretical–experimental dialog.

FIG. 32.

The density of states of Sr2FeReO6 and its electronic structure are calculated by the full potential linearized augmented plane-wave method based on the generalized gradient approximation. The inset shows the schematic structure of ordered A2B′B″O6, where the transition metal atoms (B′ and B″) occupy the perovskite B site and form B′O6 and B″O6 octahedra alternatively along the [111] direction. The Fermi level lies at the formed band exclusively by the Fe (t2g↓)-O(2p)-Re (t2g↓) sub-band. Reprinted with permission from Kobayashi et al., Phys. Rev. B 59, 11159 (1999). Copyright 1999 American Physical Society.

FIG. 32.

The density of states of Sr2FeReO6 and its electronic structure are calculated by the full potential linearized augmented plane-wave method based on the generalized gradient approximation. The inset shows the schematic structure of ordered A2B′B″O6, where the transition metal atoms (B′ and B″) occupy the perovskite B site and form B′O6 and B″O6 octahedra alternatively along the [111] direction. The Fermi level lies at the formed band exclusively by the Fe (t2g↓)-O(2p)-Re (t2g↓) sub-band. Reprinted with permission from Kobayashi et al., Phys. Rev. B 59, 11159 (1999). Copyright 1999 American Physical Society.

Close modal
FIG. 33.

Theoretical band structures of the Ba2FeReO6 compound calculated by using (a) LDA, (b) LDA + U, and (c) LAD + U + SOC methods. The band dispersions of up-spin and down-spin are denoted by black solid and red dashed lines, respectively. The black solid line in (c) denotes the band dispersions formed by mixing up-spin and down-spin. EF indicates the Fermi energy. The two band dispersions near the Fermi level are marked by the blue and magenta lines, which are nearly parallel to each other. Reprinted with permission from Jeon et al., J. Phys.: Condens. Matter 22, 345602 (2010). Copyright 2010 IOP Publishing Ltd.

FIG. 33.

Theoretical band structures of the Ba2FeReO6 compound calculated by using (a) LDA, (b) LDA + U, and (c) LAD + U + SOC methods. The band dispersions of up-spin and down-spin are denoted by black solid and red dashed lines, respectively. The black solid line in (c) denotes the band dispersions formed by mixing up-spin and down-spin. EF indicates the Fermi energy. The two band dispersions near the Fermi level are marked by the blue and magenta lines, which are nearly parallel to each other. Reprinted with permission from Jeon et al., J. Phys.: Condens. Matter 22, 345602 (2010). Copyright 2010 IOP Publishing Ltd.

Close modal
FIG. 34.

Theoretical band structures of Ca2FeReO6 with lattice distortion calculated by using (a) LDA, (b) LDA + U, and (c) LAD + U + SOC methods. The same band structures without lattice distortion are calculated by using (d) LDA, (e) LDA + U, and (f) LAD + U + SOC methods. The black solid and red dashed lines in (a), (b), (d), and (e) indicate the band dispersions of up-spin and down-spin, respectively. The black solid lines in (e) and (f) indicate the band dispersions of mixed up-spin and down-spin. EF indicates the Fermi energy. Reprinted with permission from Jeon et al., J. Phys.: Condens. Matter 22, 345602 (2010). Copyright 2010 IOP Publishing Ltd.

FIG. 34.

Theoretical band structures of Ca2FeReO6 with lattice distortion calculated by using (a) LDA, (b) LDA + U, and (c) LAD + U + SOC methods. The same band structures without lattice distortion are calculated by using (d) LDA, (e) LDA + U, and (f) LAD + U + SOC methods. The black solid and red dashed lines in (a), (b), (d), and (e) indicate the band dispersions of up-spin and down-spin, respectively. The black solid lines in (e) and (f) indicate the band dispersions of mixed up-spin and down-spin. EF indicates the Fermi energy. Reprinted with permission from Jeon et al., J. Phys.: Condens. Matter 22, 345602 (2010). Copyright 2010 IOP Publishing Ltd.

Close modal
FIG. 35.

Re 5d partial DOSs in A2FeReO6 oxides with A = Ba, Sr, and Ca, calculated by the LSDA relativistic Dirac approximation. Reprinted with permission from Antonov et al., Phys. Rev. B 94, 035122 (2016). Copyright 2016 American Physical Society.

FIG. 35.

Re 5d partial DOSs in A2FeReO6 oxides with A = Ba, Sr, and Ca, calculated by the LSDA relativistic Dirac approximation. Reprinted with permission from Antonov et al., Phys. Rev. B 94, 035122 (2016). Copyright 2016 American Physical Society.

Close modal
FIG. 36.

Experimental x-ray absorption spectroscopy (XAS) and x-ray magnetic circular dichroism (XMCD) spectra (open circles) recorded at the Re L2,3 edges in the A2FeReO6 (A = Ba, Sr, and Ca) oxides as compared with theoretical calculations. The experimental dichroism taken at the Re L3 edge and its theoretically calculated dichroism are enlarged by two times. Reprinted with permission from Antonov et al., Phys. Rev. B 94, 035122 (2016). Copyright 2016 American Physical Society.

FIG. 36.

Experimental x-ray absorption spectroscopy (XAS) and x-ray magnetic circular dichroism (XMCD) spectra (open circles) recorded at the Re L2,3 edges in the A2FeReO6 (A = Ba, Sr, and Ca) oxides as compared with theoretical calculations. The experimental dichroism taken at the Re L3 edge and its theoretically calculated dichroism are enlarged by two times. Reprinted with permission from Antonov et al., Phys. Rev. B 94, 035122 (2016). Copyright 2016 American Physical Society.

Close modal

Owing to the diverse electrical behaviors and huge magnetic anisotropy, Re-based DP oxides have promising applications in oxide spintronics, nonvolatile data memory devices, and multiferroic devices. For example, at room temperature, some A2FeReO6 oxides exhibit half-metallic ferromagnetism, which can be utilized as electrodes for fabricating the magnetic tunnel junctions (MTJs). In the MTJ structure, the tunneling current is dependent upon the relative orientation of the magnetization of two ferromagnetic electrodes, and it is modified by applying a magnetic field.121,122 Therefore, the MTJs constructed with Sr2FeReO6 and Sr2FeMoO6 as ferromagnetic electrodes would exhibit high TMR ratios due to the larger magneto-crystalline anisotropy because of the Re ion. Furthermore, the Re-based and Mo-based DP oxides such as Sr2FeReO6 and Sr2FeMoO6 exhibit much different coercive fields; thus, in the MTJs, the Sr2FeReO6 electrode with high coercive field may act as a pinned layer, while the Sr2FeMoO6 electrode with low coercive field acts as a free layer. The magnetization of Sr2FeMoO6 (free layer) can be rotated concerning the pinned layer (Sr2FeReO6) through a barrier in MTJs. As a consequence, it is possible to obtain different states of resistance, and each state can be functioned as a specific degree of spin, while, at the same time, the number of states can be applied in the storage or data processing system. Due to the two ferromagnetic electrodes (Sr2FeReO6 and Sr2FeMoO6) having almost the same lattice parameters, epitaxial growth of such heterostructured MTJs can be easily achieved in different crystallographic directions. This allows one to investigate electron tunneling as functions of not only the spin direction but also the orbital symmetry. It is expected that this direction will become an exciting topic in oxide spintronics. In an epitaxy bilayer of Ca2FeMoO6/Ca2FeReO6, as the insulating phase of Ca2FeReO6 appears at low temperature, the tunneling electrons across the insulator Ca2FeMoO6 layer and the counter-electrode have strongly exchanged split sub-bands. Based on this phenomenon, spin-filtering devices can be developed. Furthermore, such an epitaxial heterostructure provides easier manipulating of the spin filtering than the other systems such as Al/EuS/Al or La2/3Sr1/3MnO3/NiFe2O4 reported previously.123,124 As the important components used for dissipationless electronic and spintronic devices, ferromagnetic insulators (FMIs) can filter the electronic charges to produce pure spin currents, manipulating spins within nonmagnetic layers via the magnetic proximity effect,125–127 creating the quantum anomalous Hall effect in conjunction with topological insulators.128,129 For example, the cation-ordered DP Sr2Fe1+xRe1−xO6 thin films exhibit a FMI state and have high Curie temperature (400 K) and large Ms value (1.8 µB/f.u.). This enables them to be used in spin filters, magnetic proximity, and quantum anomalous Hall effects. Such new FMIs also find promising applications in spintronic, electronic, and energy efficient quantum devices. Another promising application of Re-based compounds is the development of multiferroic devices based on the magnetoelastic coupling. The resistance modulations in the Re-based thin films can be induced by the strain, which is generated by a piezoelectric layer beneath the DP Re-based thin film. At present, the advanced applications of Re-based DP oxides are still at their embryonic stage, and many issues need to be resolved and some technical challenges lie ahead. As a consequence, there is a long journey before accomplishing the commercialization of Re-based DP oxides.

Here, we present a comprehensive overview of the recent advances in the Re-based DP oxides, focusing on their syntheses, structural characterizations, physical properties, advanced applications, and theoretical studies on their electronic and structural aspects. The characteristic features of Re-based DP oxides such as metallic-like, half-metallic, or insulating behavior, high Curie temperature, and large carrier spin polarization enable them to find promising applications in oxide spintronic devices, multiferroic devices, and energy efficient quantum electronic devices. From the viewpoint of applications, the epitaxial growth of Re-based DP oxide thin films with high quality is highly required although the epitaxial growth of Re-based DP oxide thin films has become an emerging research field. It is expected that in the near future, remarkable advances will be achieved in the epitaxial growth of Re-based DP oxide thin films by the laser-MBE method with growth control at the atomic scale and enhanced properties. With the development of the aberration corrector (or Cs-corrector), the new generation HRTEM/STEM facility equipped with the Cs-corrector allows one to achieve a spatial resolution better than sub-Å and an energy resolution better than sub-eV, thereby making the structural characterizations of Re-based DP compounds at sub-Å available. Although the advanced applications of Re-based DP oxides are promising, they still remain in their early stage; a long journey lies ahead before the commercialization of Re-based DP oxides. However, the scientific and technical potentials of Re-based DP oxides are, for sure, great, and the future research in this field is very bright. It is hoped that this Review could attract renewed interest in the fundamental research of Re-based DP compounds, encouraging the scientific community to enter this impressive area.

K.L. collected the references and prepared the manuscript. Q.T., Y.W., L.Y., and Y.X. participated in sequence alignment. X.Z. designed the structure and modified the manuscript. All authors contributed to data interpretation and discussion and read and approved this manuscript.

The authors are grateful for financial support from the National Natural Science Foundation of China (Grant Nos. 11674161 and 11174122), the Natural Science Foundation of Jiangsu Province (Grant No. BK20181250), and undergraduate teaching reform projects from Nanjing University (Grant Nos. X20191028402 and 202010284036X).

The authors declare that they have no conflict of interest.

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
A. S.
Bhalla
,
R. Y.
Guo
, and
R.
Roy
,
Mater. Res. Innovations
4
(
1
),
3
26
(
2000
).
2.
A. R.
Chakhmouradian
and
P. M.
Woodward
,
Phys. Chem. Miner.
41
(
6
),
387
391
(
2014
).
3.
J. P.
Attfield
,
P.
Lightfoot
, and
R. E.
Morris
,
Dalton Trans.
44
(
23
),
10541
10542
(
2015
).
4.
W. E.
Pickett
and
D. J.
Singh
,
Phys. Rev. B
53
(
3
),
1146
1160
(
1996
).
5.
K.-I.
Kobayashi
,
T.
Kimura
,
H.
Sawada
,
K.
Terakura
, and
Y.
Tokura
,
Nature
395
(
6703
),
677
680
(
1998
).
6.
J. B.
Goodenough
and
J. M.
Longo
,
Landolt-Bornstein: Numerical Data and Functional Relationships in Science and Technology
, New Series, Group III: Crystal and Solid State Physics, Magnetic and Other Properties of Oxides and Related Compounds, Part A (
Springer
,
1970
).
7.
B. N.
Parida
,
N.
Panda
,
R.
Padhee
, and
R. K.
Parida
,
Phase Transitions
91
(
6
),
638
648
(
2018
).
8.
P. N.
Lekshmi
,
G. R.
Raji
,
M.
Vasundhara
,
M. R.
Varma
,
S. S.
Pillai
, and
M.
Valant
,
J. Mater. Chem. C
1
(
40
),
6565
6574
(
2013
).
9.
R.
Nechache
,
C.
Harnagea
,
L.-P.
Carignan
,
O.
Gautreau
,
L.
Pintilie
,
M. P.
Singh
,
D.
Ménard
,
P.
Fournier
,
M.
Alexe
, and
A.
Pignolet
,
J. Appl. Phys.
105
(
6
),
061621
(
2009
).
10.
S.
Wurmehl
and
J. T.
Kohlhepp
,
J. Phys. D: Appl. Phys.
41
(
17
),
173002
(
2008
).
11.
Y.-H.
Huang
,
R. I.
Dass
,
Z. L.
Xing
, and
J. B.
Goodenough
,
Science
312
(
5771
),
254
257
(
2006
).
12.
M. A.
Green
,
A.
Ho-Baillie
, and
H. J.
Snaith
,
Nat. Photonics
8
(
7
),
506
514
(
2014
).
13.
A.
Hossain
,
A. K. M.
Atique Ullah
,
P.
Sarathi Guin
, and
S.
Roy
,
J. Sol-Gel Sci. Technol.
93
(
3
),
479
494
(
2020
).
14.
Z. P.
Pei
,
Y.
Lu
,
H.
Wu
,
W. R.
Xia
, and
X. H.
Zhu
,
J. Chin. Ceram. Soc.
47
,
518
529
(
2019
).
15.
S.
Vasala
and
M.
Karppinen
,
Prog. Solid State Chem.
43
(
1-2
),
1
36
(
2015
).
16.
P. M.
Woodward
,
Acta Crystallogr., Sect. B: Struct. Sci.
53
,
32
43
(
1997
).
17.
R.
Roy
,
J. Am. Ceram. Soc.
37
(
12
),
581
588
(
1954
).
18.
F.
Galasso
,
L.
Katz
, and
R.
Ward
,
J. Am. Chem. Soc.
81
(
4
),
820
823
(
1959
).
19.
D.
Serrate
,
J. M.
De Teresa
, and
M. R.
Ibarra
,
J. Phys.: Condens. Matter
19
(
2
),
023201
(
2007
).
20.
A.
Hossain
,
P.
Bandyopadhyay
, and
S.
Roy
,
J. Alloys Compd.
740
,
414
427
(
2018
).
21.
X.
Chen
,
J.
Xu
,
Y.
Xu
,
F.
Luo
, and
Y.
Du
,
Inorg. Chem. Front.
6
(
9
),
2226
2238
(
2019
).
22.
A. V.
Flores
,
A. E.
Krueger
,
A. J.
Stiner
,
H. M.
Albert
,
T.
Mansur
,
V.
Willis
,
C. C.
Lee
,
L. J.
Garay
,
L. T.
Nguyen
,
M. A.
Frank
,
P. W.
Barnes
, and
A. M.
Fry-Petit
,
Prog. Solid State Chem.
56
,
100251
(
2019
).
23.
A. T.
Lee
and
C. A.
Marianetti
,
Phys. Rev. B
97
(
4
),
045102
(
2018
).
24.
J. M.
De Teresa
,
D.
Serrate
,
C.
Ritter
,
J.
Blasco
,
M. R.
Ibarra
,
L.
Morellon
, and
W.
Tokarz
,
Phys. Rev. B
71
(
9
),
092408
(
2005
).
25.
C.
Sohn
,
E.
Skoropata
,
Y.
Choi
,
X.
Gao
,
A.
Rastogi
,
A.
Huon
,
M. A.
McGuire
,
L.
Nuckols
,
Y. W.
Zhang
,
J. W.
Freeland
,
D.
Haskel
, and
H. N.
Lee
,
Adv. Mater.
31
(
4
),
1805389
(
2019
).
26.
M.
Karppinen
and
H.
Yamauchi
,
Frontiers in Magnetic Materials
, edited by
A. V.
Narlikar
(
Springer
,
Berlin
,
2005
), pp.
153
184
.
27.
H.
Kato
,
T.
Okuda
,
Y.
Okimoto
,
Y.
Tomioka
,
K.
Oikawa
,
T.
Kamiyama
, and
Y.
Tokura
,
Phys. Rev. B
69
(
18
),
184412
(
2004
).
28.
J.
Blasco
,
J.
Sesé
,
J. A.
Rodríguez-Velamazán
,
C.
Ritter
, and
J.
Herrero-Martín
,
J. Phys. Conden. Matter
21
(
21
),
216008
(
2009
).
29.
T. W.
Lim
,
S. D.
Kim
,
K. D.
Sung
,
Y. M.
Rhyim
,
H.
Jeen
,
J.
Yun
,
K. H.
Kim
,
K. M.
Song
,
S.
Lee
,
S. Y.
Chung
,
M.
Choi
, and
S. Y.
Choi
,
Sci. Rep.
6
,
19746
(
2016
).
30.
S. W.
Wang
,
L. D.
Chen
,
T.
Hirai
, and
J. K.
Guo
,
J. Inorg. Mater.
16
(
6
),
1055
1061
(
2001
).
31.
S. J.
Wu
,
L. C.
DeJonghe
, and
M. N.
Rahaman
,
J. Am. Ceram. Soc.
79
(
8
),
2207
2211
(
1996
).
32.
A. J.
Hauser
,
J. M.
Lucy
,
H. L.
Wang
,
J. R.
Soliz
,
A.
Holcomb
,
P.
Morris
,
P. M.
Woodward
, and
F. Y.
Yang
,
Appl. Phys. Lett.
102
(
3
),
032403
(
2013
).
33.
J.
Orna
,
L.
Morellon
,
P. A.
Algarabel
,
J. A.
Pardo
,
C.
Magen
,
M.
Varela
,
S. J.
Pennycook
,
J. M.
De Teresa
, and
M. R.
Ibarra
,
J. Magn. Magn. Mater.
322
(
9-12
),
1217
1220
(
2010
).
34.
Y.
Tomioka
,
T.
Okuda
,
Y.
Okimoto
,
R.
Kumai
,
K.-I.
Kobayashi
, and
Y.
Tokura
,
Phys. Rev. B
61
(
1
),
422
427
(
2000
).
35.
C. L.
Yuan
,
S. G.
Wang
,
W. H.
Song
,
T.
Yu
,
J. M.
Dai
,
S. L.
Ye
, and
Y. P.
Sun
,
Appl. Phys. Lett.
75
(
24
),
3853
3855
(
1999
).
36.
X. H.
Li
,
Y.
Sun
,
W. J.
Lu
,
R.
Ang
,
S. B.
Zhang
,
X. B.
Zhu
,
W. H.
Song
, and
J. M.
Dai
,
Solid State Commun.
145
(
3
),
98
102
(
2008
).
37.
L.
Fuoco
,
D.
Rodriguez
,
T.
Peppel
, and
P. A.
Maggard
,
Chem. Mater.
23
(
24
),
5409
5414
(
2011
).
38.
H.
Dery
,
P.
Dalal
,
Ł.
Cywiński
, and
L. J.
Sham
,
Nature
447
(
7144
),
573
576
(
2007
).
39.
S. A.
Wolf
,
D. D.
Awschalom
,
R. A.
Buhrman
,
J. M.
Daughton
,
S.
von Molnar
,
M. L.
Roukes
,
A. Y.
Chtchelkanova
, and
D. M.
Treger
,
Science
294
(
5546
),
1488
1495
(
2001
).
40.
K.
Ando
,
Science
312
(
5782
),
1883
1885
(
2006
).
41.
A. H.
Macdonald
,
P.
Schiffer
, and
N.
Samarth
,
Nat. Mater.
4
(
3
),
195
202
(
2005
).
42.
R. A.
de Groot
,
F. M.
Mueller
,
P. G. V.
Engen
, and
K. H. J.
Buschow
,
Phys. Rev. Lett.
50
(
25
),
2024
2027
(
1983
).
43.
Y.
Yamada
,
K.
Ueno
,
T.
Fukumura
,
H. T.
Yuan
,
H.
Shimotani
,
Y.
Iwasa
,
L.
Gu
,
S.
Tsukimoto
,
Y.
Ikuhara
, and
M.
Kawasaki
,
Science
332
(
6033
),
1065
1067
(
2011
).
44.
Q. U.
Ain
,
S.
Naseem
, and
S.
Nazir
,
Sci. Rep.
10
,
13778
(
2020
).
45.
S.
Chakraverty
,
A.
Ohtomo
, and
M.
Kawasaki
,
Appl. Phys. Lett.
97
(
24
),
243107
(
2010
).
46.
S.
Nazir
and
Q. U.
Ain
,
J. Alloys Compd.
847
,
155906
(
2020
).
47.
J. M.
De Teresa
,
J. M.
Michalik
,
J.
Blasco
,
P. A.
Algarabel
,
M. R.
Ibarra
,
C.
Kapusta
, and
U.
Zeitler
,
Appl. Phys. Lett.
90
(
25
),
252514
(
2007
).
48.
T. K.
Mandal
,
C.
Felser
,
M.
Greenblatt
, and
J.
Kuebler
,
Phys. Rev. B
78
(
13
),
134431
(
2008
).
49.
J. M.
Michalik
,
J. M.
De Teresa
,
J.
Blasco
,
C.
Ritter
,
P. A.
Algarabel
,
M. R.
Ibarra
, and
C.
Kapusta
,
Solid State Sci.
12
(
7
),
1121
1130
(
2010
).
50.
J.
Blasco
,
J. M.
Michalik
,
J.
Garcia
,
G.
Subias
, and
J. M.
De Teresa
,
Phys. Rev. B
76
(
14
),
144402
(
2007
).
51.
J.
Wang
,
N. N.
Zu
,
Y.
Wang
, and
Z. J.
Wu
,
J. Magn. Magn. Mater.
339
,
163
167
(
2013
).
52.
A.
Nishiyama
,
Y.
Doi
, and
Y.
Hinatsu
,
J. Solid State Chem.
248
,
134
141
(
2017
).
53.
W.
Prellier
,
V.
Smolyaninova
,
A.
Biswas
,
C.
Galley
,
R. L.
Greene
,
K.
Ramesha
, and
J.
Gopalakrishnan
,
J. Phys.: Condens. Matter
12
(
6
),
965
973
(
2000
).
54.
J. M.
De Teresa
,
D.
Serrate
,
J.
Blasco
,
M. R.
Ibarra
, and
L.
Morellon
,
Phys. Rev. B
69
(
14
),
144401
(
2004
).
55.
D.
Serrate
,
J. M.
De Teresa
,
J.
Blasco
,
M. R.
Ibarra
,
L.
Morellon
, and
C.
Ritter
,
Eur. Phys. J. B
39
(
1
),
35
40
(
2004
).
56.
S.-Y.
Choi
,
J. B.
Lim
,
Y.
Ikuhara
,
D.
Suvorov
, and
J.-H.
Jeon
,
Microsc. Microanal.
19
,
25
28
(
2013
).
57.
J.
Navarro
,
L.
Balcells
,
F.
Sandiumenge
,
M.
Bibes
,
A.
Roig
,
B.
Martínez
, and
J.
Fontcuberta
,
J. Phys.: Condens. Matter
13
(
37
),
8481
8488
(
2001
).
58.
S. Y.
Chung
,
S. Y.
Choi
,
T.
Yamamoto
, and
Y.
Ikuhara
,
Phys. Rev. Lett.
100
(
12
),
125502
(
2008
).
59.
S. Y.
Chung
,
S. Y.
Choi
,
S.
Lee
, and
Y.
Ikuhara
,
Phys. Rev. Lett.
108
(
19
),
195501
(
2012
).
60.
P.-L.
Ho
,
C.-P.
Yu
,
Q.
Zhang
,
K.
Song
,
J. P.
Buban
,
S.-Y.
Choi
,
R. E.
Dunin-Borkowski
,
J.
Mayer
,
N.-H.
Tai
,
J.
Zhu
,
L.
Jin
, and
X.
Zhong
,
Ultramicroscopy
193
,
137
142
(
2018
).
61.
J. S.
Stephens
and
D. W.
Cruickshank
,
Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem.
26
,
222
226
(
1970
).
62.
Z. Y.
Wu
,
D. C.
Xian
,
T. D.
Hu
,
Y. N.
Xie
,
Y.
Tao
,
C. R.
Natoli
,
E.
Paris
, and
A.
Marcelli
,
Phys. Rev. B
70
(
3
),
033104
(
2004
).
63.
J. M.
Michalik
,
J. M.
De Teresa
,
D.
Serrate
,
J.
Blasco
, and
M. R.
Ibarra
,
J. Magn. Magn. Mater.
316
(
2
),
413
416
(
2007
).
64.
T.
Manako
,
M.
Izumi
,
Y.
Konishi
,
K.-I.
Kobayashi
,
M.
Kawasaki
, and
Y.
Tokura
,
Appl. Phys. Lett.
74
(
15
),
2215
2217
(
1999
).
65.
H. Q.
Yin
,
J.-S.
Zhou
,
J.-P.
Zhou
,
R.
Dass
,
J. T.
McDevitt
, and
J. B.
Goodenough
,
Appl. Phys. Lett.
75
(
18
),
2812
2814
(
1999
).
66.
A.
Masuno
,
M.
Haruta
,
M.
Azuma
,
H.
Kurata
,
S.
Isoda
,
M.
Takano
, and
Y.
Shimakawa
,
Appl. Phys. Lett.
89
(
21
),
211913
(
2006
).
67.
M.
Sakai
,
A.
Masuno
,
D.
Kan
,
M.
Hashisaka
,
K.
Takata
,
M.
Azuma
,
M.
Takano
, and
Y.
Shimakawa
,
Appl. Phys. Lett.
90
(
7
),
072903
(
2007
).
68.
P.
Padhan
,
H. Z.
Guo
,
P.
LeClair
, and
A.
Gupta
,
Appl. Phys. Lett.
92
(
2
),
022909
(
2008
).
69.
A. J.
Hauser
,
J. R.
Soliz
,
M.
Dixit
,
R. E. A.
Williams
,
M. A.
Susner
,
B.
Peters
,
L. M.
Mier
,
T. L.
Gustafson
,
M. D.
Sumption
,
H. L.
Fraser
,
P. M.
Woodward
, and
F. Y.
Yang
,
Phys. Rev. B
85
(
16
),
161201
(
2012
).
70.
L.
Liu
and
J. H.
Edgar
,
Mater. Sci. Eng., R
37
(
3
),
61
127
(
2002
).
71.
H.
Asano
,
N.
Koduka
,
A.
Tsuzuki
, and
M.
Matsui
,
Appl. Phys. Lett.
85
(
2
),
263
265
(
2004
).
72.
L.
Balcells
,
J.
Navarro
,
M.
Bibes
,
A.
Roig
,
B.
Martínez
, and
J.
Fontcuberta
,
Appl. Phys. Lett.
78
(
6
),
781
783
(
2001
).
73.
J. M.
Michalik
,
J. M.
De Teresa
,
C.
Ritter
,
J.
Blasco
,
D.
Serrate
,
M. R.
Ibarra
,
C.
Kapusta
,
J.
Freudenberger
, and
N.
Kozlova
,
Europhys. Lett.
78
(
1
),
17006
(
2007
).
74.
H.
Kato
,
T.
Okuda
,
Y.
Okimoto
,
Y.
Tomioka
,
K.
Oikawa
,
T.
Kamiyama
, and
Y.
Tokura
,
Phys. Rev. B
65
(
14
),
144404
(
2002
).
75.
T.
Alamelu
,
U. V.
Varadaraju
,
M.
Venkatesan
,
A. P.
Douvalis
, and
J. M. D.
Coey
,
J. Appl. Phys.
91
(
10
),
8909
8911
(
2002
).
76.
B.
Yuan
,
J. P.
Clancy
,
J. A.
Sears
,
A. I.
Kolesnikov
,
M. B.
Stone
,
Z.
Yamani
,
C.
Won
,
N.
Hur
,
B. C.
Jeon
,
T. W.
Noh
,
A.
Paramekanti
, and
Y. J.
Kim
,
Phys. Rev. B
98
(
14
),
214433
(
2018
).
77.
B. T.
Thole
,
P.
Carra
,
F.
Sette
, and
G.
van der Laan
,
Phys. Rev. Lett.
68
(
12
),
1943
1946
(
1992
).
78.
P.
Carra
,
B. T.
Thole
,
M.
Altarelli
, and
X.
Wang
,
Phys. Rev. Lett.
70
(
5
),
694
697
(
1993
).
79.
P.
Majewski
,
S.
Geprägs
,
O.
Sanganas
,
M.
Opel
,
R.
Gross
,
F.
Wilhelm
,
A.
Rogalev
, and
L.
Alff
,
Appl. Phys. Lett.
87
(
20
),
202503
(
2005
).
80.
S.
Gong
,
S.-D.
Guo
,
P.
Chen
, and
B.-G.
Liu
,
RSC Adv.
5
(
78
),
63165
63174
(
2015
).
81.
M.
Sikora
,
O.
Mathon
,
P.
van der Linden
,
J. M.
Michalik
,
J. M.
de Teresa
,
C.
Kapusta
, and
S.
Pascarelli
,
Phys. Rev. B
79
(
22
),
220402
(
2009
).
82.
W.
Westerburg
,
O.
Lang
,
C.
Ritter
,
C.
Felser
,
W.
Tremel
, and
G.
Jakob
,
Solid State Commun.
122
(
3-4
),
201
206
(
2002
).
83.
K.
Oikawa
,
T.
Kamiyama
,
H.
Kato
, and
Y.
Tokura
,
J. Phys. Soc. Jpn.
72
(
6
),
1411
1417
(
2003
).
84.
N.
Auth
,
G.
Jakob
,
W.
Westerburg
,
C.
Ritter
,
I.
Bonn
,
C.
Felser
, and
W.
Tremel
,
J. Magn. Magn. Mater.
272-276
,
E607
E608
(
2004
).
85.
S.
Nakamura
,
M.
Tanaka
,
H.
Kato
, and
Y.
Tokura
,
J. Phys. Soc. Jpn.
72
(
2
),
424
428
(
2003
).
86.
C. A.
Escanhoela
, Jr.
,
G.
Fabbris
,
F.
Sun
,
C.
Park
,
J.
Gopalakrishnan
,
K.
Ramesha
,
E.
Granado
,
N. M.
Souza-Neto
,
M.
van Veenendaal
, and
D.
Haskel
,
Phys. Rev. B
98
(
5
),
054402
(
2018
).
87.
D.
Serrate
,
J. M.
De Teresa
,
P. A.
Algarabel
,
C.
Marquina
,
J.
Blasco
,
M. R.
Ibarra
, and
J.
Galibert
,
J. Phys.: Condens. Matter
19
(
43
),
436226
(
2007
).
88.
M.
Besse
,
F.
Pailloux
,
A.
Barthélémy
,
K.
Bouzehouane
,
A.
Fert
,
J.
Olivier
,
O.
Durand
,
F.
Wyczisk
,
R.
Bisaro
, and
J.-P.
Contour
,
J. Cryst. Growth
241
(
4
),
448
454
(
2002
).
89.
J. M.
Michalik
,
J. M.
De Teresa
,
J.
Blasco
,
P. A.
Algarabel
,
M. R.
Ibarra
,
C.
Kapusta
, and
U.
Zeitler
,
J. Phys.: Condens. Matter
19
(
50
),
506206
(
2007
).
90.
C.-R.
Lin
,
Y.-M.
Chu
, and
S.-C.
Wang
,
Mater. Lett.
60
(
17-18
),
2270
2273
(
2006
).
91.
K.-I.
Kobayashi
,
T.
Kimura
,
Y.
Tomioka
,
H.
Sawada
,
K.
Terakura
, and
Y.
Tokura
,
Phys. Rev. B
59
(
17
),
11159
11162
(
1999
).
92.
J.
Gopalakrishnan
,
A.
Chattopadhyay
,
S. B.
Ogale
,
T.
Venkatesan
,
R. L.
Greene
,
A. J.
Millis
,
K.
Ramesha
,
B.
Hannoyer
, and
G.
Marest
,
Phys. Rev. B
62
(
14
),
9538
9542
(
2000
).
93.
H.
Kato
,
T.
Okuda
,
Y.
Okimoto
,
Y.
Tomioka
,
Y.
Takenoya
,
A.
Ohkubo
,
M.
Kawasaki
, and
Y.
Tokura
,
Appl. Phys. Lett.
81
(
2
),
328
330
(
2002
).
94.
J. M.
Lucy
,
A. J.
Hauser
,
H. L.
Wang
,
J. R.
Soliz
,
M.
Dixit
,
R. E. A.
Williams
,
A.
Holcombe
,
P.
Morris
,
H. L.
Fraser
,
D. W.
McComb
,
P. M.
Woodward
, and
F. Y.
Yang
,
Appl. Phys. Lett.
103
(
4
),
042414
(
2013
).
95.
J. F.
Shackelford
,
Introduction to Materials Science for Engineers
, 7th ed. (
Prentice-Hall
,
New York
,
2008
).
96.
C.
Michel
,
S. D.
Baranovskii
,
P. J.
Klar
,
P.
Thomas
, and
B.
Goldlücke
,
Appl. Phys. Lett.
89
(
11
),
112116
(
2006
).
97.
M.
Abe
,
T.
Nakagawa
, and
S.
Nomura
,
J. Phys. Soc. Jpn.
35
(
5
),
1360
1365
(
1973
).
98.
J. B.
Philipp
,
P.
Majewski
,
L.
Alff
,
A.
Erb
,
R.
Gross
,
T.
Graf
,
M. S.
Brandt
,
J.
Simon
,
T.
Walther
,
W.
Mader
,
D.
Topwal
, and
D. D.
Sarma
,
Phys. Rev. B
68
(
14
),
144431
(
2003
).
99.
R.
Vidya
,
P.
Ravindran
,
A.
Kjekshus
, and
H.
Fjellvåg
,
Phys. Rev. B
70
(
18
),
184414
(
2004
).
100.
S.
Baidya
and
T.
Saha-Dasgupta
,
Phys. Rev. B
84
(
3
),
035131
(
2011
).
101.
B. C.
Jeon
,
C. H.
Kim
,
S. J.
Moon
,
W. S.
Choi
,
H.
Jeong
,
Y. S.
Lee
,
J.
Yu
,
C. J.
Won
,
J. H.
Jung
,
N.
Hur
, and
T. W.
Noh
,
J. Phys.: Condens. Matter
22
(
34
),
345602
(
2010
).
102.
J. H.
Jung
,
S. J.
Oh
,
M. W.
Kim
,
T. W.
Noh
,
J. Y.
Kim
,
J. H.
Park
,
H. J.
Lin
,
C. T.
Chen
, and
Y.
Moritomo
,
Phys. Rev. B
66
(
10
),
104415
(
2002
).
103.
K. H.
Kim
,
J. Y.
Gu
,
H. S.
Choi
,
D. J.
Eom
,
J. H.
Jung
, and
T. W.
Noh
,
Phys. Rev. B
55
(
7
),
4023
4026
(
1997
).
104.
Z.
Fang
,
K.
Terakura
, and
J.
Kanamori
,
Phys. Rev. B
63
(
18
),
180407
(
2001
).
105.
106.
M. D.
Scafetta
,
A. M.
Cordi
,
J. M.
Rondinelli
, and
S. J.
May
,
J. Phys.: Condens. Matter
26
(
50
),
505502
(
2014
).
107.
T.
Arima
and
Y.
Tokura
,
J. Phys. Soc. Jpn.
64
(
7
),
2488
2501
(
1995
).
108.
A. J.
Hauser
,
J.
Zhang
,
L.
Mier
,
R. A.
Ricciardo
,
P. M.
Woodward
,
T. L.
Gustafson
,
L. J.
Brillson
, and
F. Y.
Yang
,
Appl. Phys. Lett.
92
(
22
),
222901
(
2008
).
109.
S.
Azam
,
M. A.
Kamran
,
M. W.
Iqbal
,
M.
Irfan
,
S.
Abdullah
,
A.
Mahmood
,
M. S.
Khan
,
T.
Alharbi
,
A.
Majid
,
S.
Goumri-Said
,
R.
Khenata
, and
X.
Wang
,
J. Mol. Model.
26
(
6
),
158
(
2020
).
110.
S. A.
Dar
,
V.
Srivastava
,
U. K.
Sakalle
, and
G.
Pagare
,
Comput. Condens. Matter
14
,
137
143
(
2018
).
111.
X.
Chen
,
D.
Parker
,
K. P.
Ong
,
M.-H.
Du
, and
D. J.
Singh
,
Appl. Phys. Lett.
102
(
10
),
102403
(
2013
).
112.
V. N.
Antonov
,
L. V.
Bekenov
, and
A.
Ernst
,
Phys. Rev. B
94
(
3
),
035122
(
2016
).
113.
A. T.
Lee
and
C. A.
Marianetti
,
Phys. Rev. B
97
(
4
),
045102
(
2018
).
114.
H. J.
Xiang
and
M. H.
Whangbo
,
Phys. Rev. B
75
(
5
),
052407
(
2007
).
115.
K.
Kuepper
,
M.
Kadiroglu
,
V.
Postnikov
,
K. C.
Prince
,
M.
Matteucci
,
V. R.
Galakhov
,
H.
Hesse
,
G.
Borstel
, and
M.
Neumann
,
J. Phys.: Condens. Matter
17
(
27
),
4309
4317
(
2005
).
116.
G.
Vaitheeswaran
,
V.
Kanchana
, and
A.
Delin
,
Appl. Phys. Lett.
86
(
3
),
032513
(
2005
).
117.
K. W.
Lee
and
W. E.
Pickett
,
Phys. Rev. B
77
(
11
),
115101
(
2008
).
118.
J.
Herrero-Martín
,
G.
Subías
,
J.
Blasco
,
J.
García
, and
M. C.
Sánchez
,
J. Phys.: Condens. Matter
17
(
33
),
4963
4976
(
2005
).
119.
A.
Winkler
,
N.
Narayanan
,
D.
Mikhailova
,
K. G.
Bramnik
,
H.
Ehrenberg
,
H.
Fuess
,
G.
Vaitheeswaran
,
V.
Kanchana
,
F.
Wilhelm
,
A.
Rogalev
,
A.
Kolchinskaya
, and
L.
Alff
,
New J. Phys.
11
(
7
),
073047
(
2009
).
120.
F. M. F.
de Groot
,
Physica B
208-209
(
1-4
),
15
18
(
1995
).
121.
T.
Miyazaki
and
N.
Tezuka
,
J. Magn. Magn. Mater.
139
,
L231
L234
(
1995
).
122.
J. S.
Moodera
,
L. R.
Kinder
,
T. M.
Wong
, and
R.
Meservey
,
Phys. Rev. Lett.
74
,
3273
3276
(
1995
).
123.
J. S.
Moodera
,
X.
Hao
,
G. A.
Gibson
, and
R.
Meservey
,
Phys. Rev. Lett.
61
(
5
),
637
640
(
1988
).
124.
U.
Lüders
,
M.
Bibes
,
K.
Bouzehouane
,
E.
Jacquet
,
J.-P.
Contour
,
S.
Fusil
,
J.-F.
Bobo
,
J.
Fontcuberta
,
A.
Barthélémy
, and
A.
Fert
,
Appl. Phys. Lett.
88
(
8
),
082505
(
2006
).
125.
D. P.
DiVincenzo
,
J. Appl. Phys.
85
(
8
),
4785
4787
(
1999
).
126.
H. X.
Yang
,
A.
Hallal
,
D.
Terrade
,
X.
Waintal
,
S.
Roche
, and
M.
Chshiev
,
Phys. Rev. Lett.
110
(
4
),
046603
(
2013
).
127.
S.
Sakai
,
S. V.
Erohin
,
Z. I.
Popov
,
S.
Haku
,
T.
Watanabe
,
Y.
Yamada
,
S.
Entani
,
S.
Li
,
P. V.
Avramov
,
H.
Naramoto
,
K.
Ando
,
P. B.
Sorokin
, and
Y.
Yamauchi
,
Adv. Funct. Mater.
28
(
20
),
1800462
(
2018
).
128.
I.
Vobornik
,
U.
Manju
,
J.
Fujii
,
F.
Borgatti
,
P.
Torelli
,
D.
Krizmancic
,
Y. S.
Hor
,
R. J.
Cava
, and
G.
Panaccione
,
Nano Lett.
11
(
10
),
4079
4082
(
2011
).
129.
C.
Tang
,
C.-Z.
Chang
,
G.
Zhao
,
Y.
Liu
,
Z.
Jiang
,
C.-X.
Liu
,
M. R.
McCartney
,
D. J.
Smith
,
T.
Chen
,
J. S.
Moodera
, and
J.
Shi
,
Sci. Adv.
3
(
6
),
e1700307
(
2017
).