Next Article in Journal
Correction: Dhatwalia et al. Rubus ellipticus Sm. Fruit Extract Mediated Zinc Oxide Nanoparticles: A Green Approach for Dye Degradation and Biomedical Applications. Materials 2022, 15, 3470
Next Article in Special Issue
Ferrimagnetic Ordering and Spin-Glass State in Diluted GdFeO3-Type Perovskites (Lu0.5Mn0.5)(Mn1−xTix)O3 with x = 0.25, 0.50, and 0.75
Previous Article in Journal
Transient Magnetic Properties of Non-Grain Oriented Silicon Steel under Multi-Physics Field
Previous Article in Special Issue
VOx Phase Mixture of Reduced Single Crystalline V2O5: VO2 Resistive Switching
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dielectric and Spin-Glass Magnetic Properties of the A-Site Columnar-Ordered Quadruple Perovskite Sm2CuMn(MnTi3)O12

1
International Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS), Namiki 1-1, Tsukuba 305-0044, Ibaraki, Japan
2
Graduate School of Chemical Sciences and Engineering, Hokkaido University, North 10 West 8, Kita-ku, Sapporo 060-0810, Hokkaido, Japan
3
Institute of Scientific and Industrial Research, Osaka University, Mihogaoka 8-1, Ibaraki 567-0047, Osaka, Japan
*
Author to whom correspondence should be addressed.
Materials 2022, 15(23), 8306; https://doi.org/10.3390/ma15238306
Submission received: 11 October 2022 / Revised: 18 November 2022 / Accepted: 21 November 2022 / Published: 23 November 2022
(This article belongs to the Special Issue Synthesis, Structure and Properties of Metal Oxides)

Abstract

:
Perovskite-type ABO3 oxides show a number of cation-ordered structures, which have significant effects on their properties. The rock-salt-type order is dominant for B cations, and the layered order for A cations. In this work, we prepared a new perovskite-type oxide, Sm2CuMn(MnTi3)O12, with a rare columnar A-site order using a high-pressure, high-temperature method at about 6 GPa and about 1700 K. Its crystal structure was studied with synchrotron powder X-ray diffraction. The compound crystallizes in space group P42/nmc (No. 137) at room temperature with a = 7.53477 Å and c = 7.69788 Å. The magnetic properties of the compound were studied with dc and ac magnetic susceptibility measurements and specific heat. Spin-glass (SG) magnetic properties were found with TSG = 7 K, while specific heat, in the form of Cp/T, showed a strong, very broad anomaly developing below 20 K and peaking at 4 K. The dielectric constant of Sm2CuMn(MnTi3)O12 was nearly frequency and temperature independent between 8 K and 200 K, with a value of about 50. Cu2+ doping drastically modified the magnetic and dielectric properties of Sm2CuMn(MnTi3)O12 in comparison with the parent compound Sm2MnMn(MnTi3)O12, which showed a long-range ferrimagnetic order at 34–40 K. The antisite disorder of Cu2+ and Mn2+ cations between square-planar and octahedral sites was responsible for the SG magnetic properties of Sm2CuMn(MnTi3)O12.

Graphical Abstract

1. Introduction

The properties of the perovskite-structure oxide material, ABO3, are controlled by their chemical compositions and degrees of cation orderings [1,2]. There are perovskites with B-site cation orderings, A-site cation orderings, and both types of orderings. In the case of B-site ordering, the rock-salt-type order is dominant [3]. In the case of A-site ordering in ABO3, the layered-type order is dominant [1,2,4], but there are other types of ordering [5,6]. There are also two special families of perovskites with A-site orderings: A-site-ordered quadruple perovskites, AA′3B4O12 [7,8,9], and A-site columnar-ordered quadruple perovskites, A2A′A″B4O12 [10]. Quadruple perovskites can have ordered arrangements of 3d transition metals at the A (in general) perovskite sites in addition to the B sites. The resulting B–B, A–B, and A–A exchange interactions can produce complex interaction patterns and frustration networks and result in competing magnetic ground states, a large number of magnetic transitions and unexpected magnetism [11].
With A = R = rare earth elements and Bi and A′ = A″ = B = Mn, interesting classes of perovskite manganites are formed, namely RMn7O12 [9] and RMn3O6 (in a short formula) [12]. They show several magnetic transitions with spin reorientations [9], and some members have incommensurate magnetic structures [13]. The Cu2+ doping of RMn7O12 and RMn3O6 has beneficial effects on their magnetic properties. in a sense that magnetic transition temperatures significantly increase, e.g., from about 80–87 K in RMn7O12 [13] to 360–400 K in RCu3Mn3O12 [14], and from about 60–77 K in RMn3O6 [12] to 160–180 K in R2CuMnMn4O12 [11,15]. In case of BiMn7O12 [9], Cu2+ doping results in complex structural behavior [16,17,18], complex magnetic behavior, and almost a linear rise of magnetic transition temperatures in BiCuxMn7–xO12 [18] for 0.8 ≤ x ≤ 3 from about 30 K (for 0 < x < 0.8) to 360 K for x = 3 [19]. The beneficial effects of Cu2+ doping also took place in Y2MnGaMn4O12, which shows spin-glass magnetic properties at 26 K [20], as Y2CuGaMn4O12 exhibits long-range ferrimagnetic ordering at 115 K [21].
Sm2MnMn(MnTi3)O12 is a member of the A-site columnar-ordered quadruple perovskites. The magnetic properties of Sm2MnMn(MnTi3)O12 [22,23] were somewhat unexpected, as it shows a long-range ferrimagnetic ordering with TC = 34–40 K and a well-defined M–H hysteresis loop with remnant magnetization of 2.3–2.4 μB/f.u. at 5 K. The concentration of 3d transition metals (Mn2+) at the B sites (25%) was below the percolation limit for the corner-shared octahedral network. Nevertheless, Mn2+ cations at the B sites were involved in the long-range ordering with a noticeable ordered magnetic moment [23]. Similar compounds without magnetic cations at the B sites (e.g., Ca2MnMnTi4O12 [24] and NaRMnMnTi4O12 [25]) show antiferromagnetic (AFM) transitions at lower temperatures of about 12 K. Therefore, there should be a noticeable involvement of the A–A and A–B exchange interactions in Sm2MnMn(MnTi3)O12 to stabilize the ferrimagnetic structure out of AFM one and to increase the magnetic transition temperature. In addition, Sm2MnMn(MnTi3)O12 was the first example among A-site columnar-ordered quadruple perovskites, demonstrating relaxor-type dielectric properties with broad maxima on the temperature dependence of a dielectric constant near 220 K [22].
In this work, we investigated the effects of Cu2+ doping on the magnetic and dielectric properties of the parent Sm2MnMn(MnTi3)O12 compound and prepared Sm2CuMn(MnTi3)O12 using a high-pressure, high-temperature method. However, in this case, the magnetic properties of the parent compound were “degraded” by Cu2+ doping, as only spin-glass (SG) magnetic properties were observed below TSG = 7 K. We attributed this degradation to antisite disorder. The relaxor-type dielectric properties of the parent compound disappeared, and Sm2CuMn(MnTi3)O12 demonstrated a frequency and temperature independent dielectric constant between 10 K and 200 K, with a value of about 50.

2. Experimental

Sm2CuMn(MnTi3)O12 was prepared using a high-pressure, high-temperature method using a belt-type high-pressure machine at 6 GPa and about 1700 K for 2 h in a Pt capsule. After annealing at 1700 K, the samples were quenched to room temperature (RT) by turning off the heating current, and the pressure was slowly released. Stoichiometric amounts of Sm2O3 (99.9%), CuO (99.9%), MnO (99.99%), and TiO2 (99.9%) were used as an initial oxide mixture with the 1:1:2:3 ratio, respectively. Commercial Sm2O3, CuO, and TiO2 chemicals were used. A single-phase MnO oxide was prepared from a commercial MnO2 chemical by annealing at 1273 K for 4 h in a 20% H2 + 80% Ar gas flow.
X-ray powder diffraction (XRPD) data were collected at RT with a RIGAKU MiniFlex600 diffractometer (CuKα radiation; a 2θ range of 8–100°; a step of 0.02°, and scan speed of 1°/min). The synchrotron XRPD data were collected at RT on the BL15XU beamline (the former NIMS beamline) of SPring-8 [26] between 2.04° and 60.23° at 0.003° intervals in 2θ with a wavelength of λ = 0.65298 Å. The data between 6° and 60.23° were used in the refinements as no reflections were observed and expected below 6°. The sample was inserted into a Lindemann glass capillary tube (inner diameter: 0.1 mm), which was rotated during the measurements. The Rietveld analysis of all XRPD data was performed using the RIETAN-2000 program [27].
Scanning electron microscopy (SEM) images and energy-dispersive X-ray (EDX) spectra were obtained on a Hitachi Miniscope TM3000 (operating at 15 kV).
SQUID magnetometers (Quantum Design, MPMS-XL-7T and MPMS3) were used for the magnetic measurements. Temperature dependence was measured between 2 and 400 K in applied fields of 100 Oe and 10 kOe under both zero-field-cooled (ZFC) and field-cooled on cooling (FCC) conditions on an MPMS-XL-7T. Magnetic-field dependence was measured at T = 2 K and 5 K between −70 and 70 kOe on MPMS3. Frequency dependent alternating current (ac) susceptibility measurements were performed on cooling with a Quantum Design MPMS3 instrument at different frequencies (f), different applied oscillating magnetic fields (Hac), and different static dc field (Hdc). Relaxation curves were measured on MPMS3 using the following procedure: the sample was cooled down from 50 K to a measurement temperature at zero magnetic field, then a magnetic field of 100 Oe was applied, and magnetization was measured (as one scan within 2 s) as a function of time every 5 s.
Specific heat, Cp, was measured by cooling from 270 K to 2 K at zero magnetic field and from 150 K to 2 K at magnetic field of 90 kOe by a pulse relaxation method using a commercial calorimeter (Quantum Design PPMS).
The dielectric constant and dielectric loss were measured on a NOVOCONTROL Alpha-A High Performance Frequency Analyzer in a frequency range from 100 Hz to 665 kHz in a temperature range from 8 K to 330 K (on heating) at zero magnetic field.

3. Results and Discussion

The as-synthesized Sm2CuMn(MnTi3)O12 contained a small amount of CuO impurity. In addition, the synchrotron XRPD pattern showed the presence of Pt impurity. However, Pt appeared from Pt capsules used in the synthesis and can be considered as an extrinsic impurity. The presence of a CuO impurity suggests that the main phase should be slightly Cu-deficient in comparison with the target composition. The morphology of the sample is shown in Figure 1. The particle sizes varied between about 10 and 50 μm. The Ti:Mn:Sm:Cu cation ratio determined with EDX was 3.24(8):2.02(5):1.96(4):0.77(7), respectively. These values were close to the nominal values within 3σ.
All of the reflections on the laboratory and synchrotron XRPD patterns (except CuO and Pt) could be indexed in a tetragonal system in space group P42/nmc (No. 137) (Figure 2). Sm2CuMnMnTi3O12 was found to crystallize in the parent structure of the A-site columnar-ordered quadruple perovskites, A2A′A″B4O12 [10]. Therefore, the structural data for the parent compound Sm2MnMn(MnTi3)O12 [22,23] were taken as an initial starting model.
In the structural analysis, we first assumed ideal cation distributions (that is, Sm at the A site, Cu at the square-planar A′ site, Mn at the tetrahedral A″ site, and 0.75Ti + 0.25Mn at the octahedral B site) and refined the occupation factors (g) together with all of the other structural and nonstructural parameters (except g(B): one cation occupation factor should always be fixed to avoid significant correlations among the refined g parameters). In addition, in the structural analysis, we always assumed that Ti4+ cations were located at the B site, as Ti4+ cations strongly prefer octahedral sites [28].
The refined g values were as follows: g(Sm–A) = 0.9167(16), g(Cu–A′) = 0.914(7), and g(Mn–A″) = 1.047(7). These values suggest that the ideal cation distribution was not realized, and there were some antisite disorders. The g(Cu–A′) value suggested that this site should contain lighter elements that could only be Mn (with the above assumption on Ti). When only Mn was placed at the square-planar A′ site, the occupation factor was g(Mn–A′) = 1.112(8), meaning that heavier elements should also be at this site. Because it was difficult to precisely refine the distribution of Mn and Cu with X-ray diffraction, we introduced a virtual atom: MC = 0.5Mn + 0.5Cu. The precise distribution of Mn and Cu could only be determined with neutron diffraction. The disordering of cations at the Cu site was also observed as in many cases of such perovskites [21,22,25,29].
The refined structural parameters and primary bond lengths and angles in Sm2CuMn(MnTi3)O12 are listed in Table 1 and Table 2. The experimental, calculated, and difference synchrotron patterns are shown in Figure 3. The crystal structure of Sm2CuMn(MnTi3)O12 is illustrated in the inset of Figure 3.
Our model suggested that a small fraction of Cu2+ cations should be located at the B site. Indirect evidence for the location of Cu2+ cations at the octahedral site can be seen from the resulting Ti/MC–O bond lengths. In the parent and related compounds, R2MnMn(MnTi3)O12 (R = Nd, Sm, Eu, and Gd), the Ti/Mn–O bond lengths were about 1.99, 2.01, and 2.01 Å (from both the synchrotron [22,29] and neutron [23] diffraction data), resulting in an octahedral distortion parameter, Δ, of about 0.2 × 10–4. On the other hand, in Sm2CuMn(MnTi3)O12, the Ti/MC–O bond lengths were about 1.95, 1.99, and 2.02 Å resulting in Δ of about 2.0 × 10–4. This rise in the octahedral distortion could be caused by the presence of a small amount of Jahn–Teller active Cu2+ cations at this site.
Magnetic susceptibility curves, χ versus T, of Sm2CuMn(MnTi3)O12 under applied magnetic fields of 0.1 kOe and 10 kOe are shown on Figure 4. There was a divergence between the 100 Oe ZFC and FCC curves at 7 K and a relatively sharp maximum on the 100 Oe ZFC curve at 7 K. A divergence between the ZFC and FCC curves almost disappeared under 10 kOe. These features are typical for spin-glass transitions [30,31,32]. Isothermal magnetization, M versus H, curves demonstrated an extended S-type shape with very weak and narrow hysteresis (Figure 5). Almost no hysteresis was observed at 5 K because 5 K was close to its TSG = 7 K; on the other hand, the hysteresis was noticeably wider at a lower temperature of 2 K. Such M versus H curves are also typical for spin glasses [30,31,32].
The inverse magnetic susceptibilities (χ−1 versus T) followed the Curie–Weiss law at high temperatures (Figure 4). To obtain the effective magnetic moment and the Curie–Weiss temperature, we performed fits between 250 and 345 K using the 10 kOe FCC curves (the fit and fitting parameters are summarized on Figure 4). The experimental effective magnetic moment was close to the expected one (8.803 μB; in the calculations we used 1.5 μB for Sm3+ [33]). The negative Curie–Weiss temperature shows that the main magnetic interactions were antiferromagnetic in nature. The ratio between the Curie–Weiss temperature (−81.5 K) and TSG (the so-called frustration ratio) was about 11, indicating a strong degree of magnetic frustration. We note that CuO was in an antiferromagnet with transition temperatures of 213 K and 230 K. Therefore, CuO impurity should not affect the reported magnetic properties at low temperatures.
To confirm the spin-glass nature of the sample, we measured ac magnetic susceptibility curves (Figure 6 and Figure 7). We note that no dependence of the χ′ and χ″ values on the applied Hac field was observed (inset of Figure 6). We indeed observed typical features of spin-glasses: peak positions were frequency-dependent and shifted to higher temperatures with increasing frequency; in addition, peak intensity was suppressed on the χ′ versus T curves and enhanced on the χ″ versus T curves with increasing frequency. All of these features are typical for spin glasses [30,31,32]. In addition, the shape of the χ′ versus T and the χ″ versus T curves was also typical for spin glasses. The criterion, which quantifies the relative change of the spin-glass temperature per frequency decade and is defined as ΔTSG/[TSGΔlog(f)], was about 0.023 for Sm2CuMn(MnTi3)O12 (with TSG = 7.2 K at f = 2 Hz and TSG = 7.6 K at f = 500 Hz). This value is often observed in different spin-glass materials [30,31,32].
Sm2CuMn(MnTi3)O12 shows time-dependent magnetic properties below TSG, namely magnetization relaxation (Figure 8). Above TSG, no noticeable relaxation of magnetization was detected. Time-dependent magnetic properties, such as relaxation, are typical features of spin-glass systems. Relaxation below TSG was fitted by the stretched exponential function, f(t) = M0MSG × exp[−(t/tr)β] [30], and the resultant parameters are listed on Table 3. The most important parameter is the mean relaxation time, tr, and it decreases monotonically with increasing temperature.
The specific heat data showed a noticeable magnetic contribution to the total specific heat below about 20 K, where it could be clearly seen as a rise in Cp/T values below 20 K (Figure 9). No λ-type anomaly was detected in the Cp versus T curve (inset of Figure 9, a green curve). Instead, a broad anomaly was seen in the Cp versus T curve, which gave a broad peak centered at 4 K in the Cp/T versus T curve. Therefore, specific heat measurements confirmed the absence of long-range magnetic ordering. A magnetic field of 90 kOe slightly suppressed the peak near 4 K and moved the magnetic entropy into the 14–40 K range.
The temperature dependence of the dielectric constant and dielectric loss is shown in Figure 10. The dielectric constant was nearly temperature and frequency-independent between 8 and 200 K. Above about 200 K, a sharp rise in the dielectric constant was observed, where the magnitude of the rise depended on frequency. This behavior typically originates from the Maxwell–Wagner contribution due to increased conductivity. No broad anomalies were observed in Sm2CuMn(MnTi3)O12 in comparison with the parent compound Sm2MnMn(MnTi3)O12. This fact shows that Cu2+ doping drastically modified the dielectric properties as well, in addition to the magnetic properties. We note that Pt impurity was only observed in a powder sample, which could contain parts from the surface. The surfaces of a pellet used for dielectric measurements were polished. Therefore, Pt impurity should not present in a pellet and affect dielectric measurements.
Spin-glass magnetic properties were also observed in Sm2MnZn(MnTi3)O12 at TSG = 6.5 K, with a significant antisite disorder [34]. This fact shows that antisite structural disorder should play a major role in the modification of magnetic properties of the parent Sm2MnMn(MnTi3)O12 compound, not the nature of dopant cations (magnetic as Cu2+ or non-magnetic as Zn2+). Both Sm2CuMn(MnTi3)O12 and Sm2MnZn(MnTi3)O12 demonstrated similar low-temperature specific heat features (Figure 9b).
The beneficial effects of Cu2+ doping in RMn7O12 [13,14], RMn3O6 [11,12,15], and Y2MnGaMn4O12 [20,21] originate from the fact that Cu2+ doping is aliovalent doping, which produces Mn4+ cations. A mixture of Mn3+ and Mn4+ at the B sites of perovskites significantly enhanced the exchange interactions and magnetic transition temperatures. On the other hand, Cu2+ doping in the parent Sm2MnMn(MnTi3)O12 compound was isovalent doping. Such doping did not change the oxidation state of Mn, while the antisite disordering “degraded” the magnetic properties.

4. Conclusions

A new member of the A-site columnar-ordered quadruple perovskite family, Sm2CuMn(MnTi3)O12, was prepared using a high-pressure, high-temperature method. Cu2+ doping significantly modified the properties of the parent Sm2MnMn(MnTi3)O12 compound, as spin-glass magnetic properties at TSG = 7 K were observed in Sm2CuMn(MnTi3)O12 in comparison with the long-range ferrimagnetic order at TC = 34–40 K in Sm2MnMn(MnTi3)O12. In addition, relaxor-like dielectric properties of Sm2MnMn(MnTi3)O12 disappeared in Sm2CuMn(MnTi3)O12, which showed a nearly temperature and frequency-independent dielectric constant between 8 and 200 K with a value of about 50.

Author Contributions

A.A.B.: investigation, data analysis, conceptualization, supervision, writing—original draft, and writing—review and editing. R.L.: investigation. K.Y.: investigation, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This study was partly supported by JSPS KAKENHI grant numbers JP20H05276 and JP22H04601.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available from A.A.B. upon reasonable request.

Acknowledgments

This study was partly supported by JSPS KAKENHI grant numbers JP20H05276 and JP22H04601. The synchrotron radiation experiments were conducted at the former NIMS beamline (BL15XU) of SPring-8 with the approval of the former NIMS Synchrotron X-ray Station (proposal numbers: 2019A4501 and 2019B4500). We thank Y. Katsuya and M. Tanaka for their help at SPring-8.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Abakumov, A.M.; Tsirlin, A.A.; Antipov, E.V. Transition-Metal Perovskites. In Comprehensive Inorganic Chemistry II (Second Edition): From Elements to Applications; Reedijk, J., Poeppelmeier, K.R., Eds.; Elsevier: Amsterdam, The Netherlands, 2013; Volume 2, pp. 1–40. [Google Scholar]
  2. King, G.; Woodward, P.M. Cation ordering in perovskites. J. Mater. Chem. 2010, 20, 5785–5796. [Google Scholar] [CrossRef]
  3. Vasala, S.; Karppinen, M. A2B′B′′O6 perovskites: A review. Prog. Solid State Chem. 2015, 43, 1–36. [Google Scholar] [CrossRef]
  4. Knapp, M.C.; Woodward, P.M. A-site cation ordering in AA′BB′O6 perovskites. J. Solid State Chem. 2006, 179, 1076–1085. [Google Scholar] [CrossRef]
  5. Rubel, M.H.K.; Miura, A.; Takei, T.; Kumada, N.; Mozahar Ali, M.; Nagao, M.; Watauchi, S.; Tanaka, I.; Oka, K.; Azuma, M.; et al. Superconducting double perovskite bismuth oxide prepared by a low-temperature hydrothermal reaction. Angew. Chem. Int. Ed. 2014, 126, 3673–3677. [Google Scholar]
  6. Rubel, M.H.K.; Takei, T.; Kumada, N.; Mozahar Ali, M.; Miura, A.; Tadanaga, K.; Oka, K.; Azuma, M.; Yashima, M.; Fujii, K.; et al. Hydrothermal synthesis, crystal structure, and superconductivity of a double-perovskite Bi oxide. Chem. Mater. 2016, 28, 459–465. [Google Scholar]
  7. Vasil’ev, A.N.; Volkova, O.S. New functional materials AC3B4O12 (review). Low Temp. Phys. 2007, 33, 895–914. [Google Scholar] [CrossRef]
  8. Long, Y.W. A-site ordered quadruple perovskite oxides AA′3B4O12. Chin. Phys. B 2016, 25, 078108. [Google Scholar] [CrossRef]
  9. Belik, A.A.; Johnson, R.D.; Khalyavin, D.D. The rich physics of A-site-ordered quadruple perovskite manganites AMn7O12. Dalton Trans. 2021, 50, 15458–15472. [Google Scholar] [CrossRef]
  10. Belik, A.A. Rise of A-site columnar-ordered A2A′A″B4O12 quadruple perovskites with intrinsic triple order. Dalton Trans. 2018, 47, 3209–3217. [Google Scholar] [CrossRef] [Green Version]
  11. Vibhakar, A.M.; Khalyavin, D.D.; Manuel, P.; Liu, J.; Belik, A.A.; Johnson, R.D. Spontaneous rotation of ferrimagnetism driven by antiferromagnetic spin canting. Phys. Rev. Lett. 2020, 124, 127201. [Google Scholar]
  12. Zhang, L.; Matsushita, Y.; Yamaura, K.; Belik, A.A. Five-fold ordering in high-pressure perovskites RMn3O6 (R = Gd-Tm and Y). Inorg. Chem. 2017, 56, 5210–5218. [Google Scholar] [CrossRef] [PubMed]
  13. Johnson, R.D.; Khalyavin, D.D.; Manuel, P.; Zhang, L.; Yamaura, K.; Belik, A.A. Magnetic structures of the rare-earth quadruple perovskite manganites RMn7O12. Phys. Rev. B Condens. Matter Mater. Phys. 2018, 98, 104423. [Google Scholar] [CrossRef] [Green Version]
  14. Sanchez-Benitez, J.; Alonso, J.A.; Martinez-Lope, M.J.; de Andres, A.; Fernandez-Diaz, M.T. Enhancement of the Curie temperature along the perovskite series RCu3Mn4O12 driven by chemical pressure of R3+ cations (R = rare earths). Inorg. Chem. 2010, 49, 5679–5685. [Google Scholar] [CrossRef] [PubMed]
  15. Belik, A.A.; Khalyavin, D.D.; Zhang, L.; Matsushita, Y.; Katsuya, Y.; Tanaka, M.; Johnson, R.D.; Yamaura, K. Intrinsic triple order in A-site columnar-ordered quadruple perovskites: Proof of concept. ChemPhysChem 2018, 19, 2449–2452. [Google Scholar] [CrossRef]
  16. Belik, A.A.; Matsushita, Y.; Khalyavin, D.D. Reentrant structural transitions and collapse of charge and orbital orders in quadruple perovskites. Angew. Chem. Int. Ed. 2017, 56, 10423–10427. [Google Scholar] [CrossRef]
  17. Khalyavin, D.D.; Johnson, R.D.; Orlandi, F.; Radaelli, P.G.; Manuel, P.; Belik, A.A. Emergent helical texture of electric dipoles. Science 2020, 369, 680–684. [Google Scholar] [CrossRef]
  18. Belik, A.A.; Matsushita, Y.; Tanaka, M.; Johnson, R.D.; Khalyavin, D.D. A plethora of structural transitions, distortions and modulations in Cu-doped BiMn7O12 quadruple perovskites. J. Mater. Chem. C 2021, 9, 10232–10242. [Google Scholar] [CrossRef]
  19. Kayser, P.; Martinez-Lope, M.J.; Alonso, J.A.; Sanchez-Benitez, J.; Fernandez, M.T. High-pressure synthesis and characterization of BiCu3(Mn4−xFex)O12 (x = 0, 1.0, 2.0) complex perovskites. J. Solid State Chem. 2013, 204, 78–85. [Google Scholar] [CrossRef]
  20. Liu, R.; Khalyavin, D.D.; Tsunoda, N.; Kumagai, Y.; Oba, F.; Katsuya, Y.; Tanaka, M.; Yamaura, K.; Belik, A.A. Spin-glass magnetic properties of A-site columnar-ordered quadruple perovskites Y2MnGa(Mn4–xGax)O12 with 0 ≤ x ≤ 3. Inorg. Chem. 2019, 58, 14830–14841. [Google Scholar]
  21. Belik, A.A.; Khalyavin, D.D.; Matsushita, Y.; Yamaura, K. Triple A-site cation ordering in the ferrimagnetic Y2CuGaMn4O12 perovskite. Inorg. Chem. 2022, 61, 14428–14435. [Google Scholar] [CrossRef]
  22. Belik, A.A.; Zhang, L.; Liu, R.; Khalyavin, D.D.; Katsuya, Y.; Tanaka, M.; Yamaura, K. Valence variations by B-site doping in A-site columnar-ordered quadruple perovskites Sm2MnMn(Mn4−xTix)O12 with 1 ≤ x ≤ 3. Inorg. Chem. 2019, 58, 3492–3501. [Google Scholar] [CrossRef]
  23. Vibhakar, A.M.; Khalyavin, D.D.; Manuel, P.; Liu, R.; Yamaura, K.; Belik, A.A.; Johnson, R.D. Effects of magnetic dilution in the ferrimagnetic columnar ordered Sm2MnMnMn4−xTixO12 perovskites. Phys. Rev. B Condens. Matter Mater. Phys. 2020, 102, 214428. [Google Scholar] [CrossRef]
  24. Aimi, A.; Mori, D.; Hiraki, K.; Takahashi, T.; Shan, Y.J.; Shirako, Y.; Zhou, J.S.; Inaguma, Y. High-pressure synthesis of A-site ordered double perovskite CaMnTi2O6 and ferroelectricity driven by coupling of A-site ordering and the second-order Jahn–Teller effect. Chem. Mater. 2014, 26, 2601–2608. [Google Scholar] [CrossRef]
  25. Liu, R.; Scatena, R.; Khalyavin, D.D.; Johnson, R.D.; Inaguma, Y.; Tanaka, M.; Matsushita, Y.; Yamaura, K.; Belik, A.A. High-pressure synthesis, crystal structures, and properties of A-site columnar-ordered quadruple perovskites NaRMn2Ti4O12 with R = Sm, Eu, Gd, Dy, Ho, Y. Inorg. Chem. 2020, 59, 9065–9076. [Google Scholar] [CrossRef]
  26. Tanaka, M.; Katsuya, Y.; Matsushita, Y.; Sakata, O. Development of a synchrotron powder diffractometer with a one-dimensional X-ray detector for analysis of advanced materials. J. Ceram. Soc. Jpn. 2013, 121, 287–290. [Google Scholar] [CrossRef] [Green Version]
  27. Izumi, F.; Ikeda, T. A Rietveld-analysis program RIETAN-98 and its applications to zeolites. Mater. Sci. Forum 2000, 321, 198–205. [Google Scholar] [CrossRef]
  28. Waroquiers, D.; Gonze, X.; Rignanese, G.-M.; Welker-Nieuwoudt, C.; Rosowski, F.; Gobel, M.; Schenk, S.; Degelmann, P.; André, R.; Glaum, R.; et al. Statistical analysis of coordination environments in oxides. Chem. Mater. 2017, 29, 8346–8360. [Google Scholar] [CrossRef]
  29. Liu, R.; Tanaka, M.; Mori, H.; Inaguma, Y.; Yamaura, K.; Belik, A.A. Ferrimagnetic and relaxor ferroelectric properties of R2MnMn(MnTi3)O12 perovskites with R = Nd, Eu, and Gd. J. Mater. Chem. C 2021, 9, 947–956. [Google Scholar] [CrossRef]
  30. Mydosh, J.A. Spin Glass: An Experimental Introduction; Taylor & Francis: London, UK, 1993. [Google Scholar]
  31. von Lohneysen, H. Low energy excitations in amorphous metals. Phys. Rep. 1981, 79, 161–212. [Google Scholar] [CrossRef]
  32. Binder, K.; Young, A.P. Spin-glasses: Experimental facts, theoretical concepts, and open questions. Rev. Mod. Phys. 1986, 58, 801–976. [Google Scholar] [CrossRef]
  33. Kittel, C. Introduction to Solid State Physics, 8th ed.; John Wiley and Sons, Inc.: New York, NY, USA, 2005. [Google Scholar]
  34. Belik, A.A.; Liu, R.; Yamaura, K. Spin-glass to long-range order to spin-glass evolution of magnetic properties with the composition in Sm2MnZnMn4−xTixO12 (x = 1, 2, and 3) perovskites. J. Alloys Compd. 2023, 932, 166297. [Google Scholar] [CrossRef]
Figure 1. Scanning electron microscopy (SEM) images of the fractured surface of the as-synthesized Sm2CuMn(MnTi3)O12 sample. The scale bars are 100 µm (left) and 30 µm (right); magnification is 1000 (left) and 2000 (right). The surface is partially polished in the right panel. No polishing has been done in the left panel.
Figure 1. Scanning electron microscopy (SEM) images of the fractured surface of the as-synthesized Sm2CuMn(MnTi3)O12 sample. The scale bars are 100 µm (left) and 30 µm (right); magnification is 1000 (left) and 2000 (right). The surface is partially polished in the right panel. No polishing has been done in the left panel.
Materials 15 08306 g001
Figure 2. Laboratory powder X-ray diffraction pattern of Sm2CuMn(MnTi3)O12 in a 2θ range from 14° to 60.5°. Possible Bragg reflection positions for Sm2CuMn(MnTi3)O12 (the first row) and CuO impurities (the second row) are shown. The (hkl) indices of all of the observed reflections of Sm2CuMn(MnTi3)O12 are given.
Figure 2. Laboratory powder X-ray diffraction pattern of Sm2CuMn(MnTi3)O12 in a 2θ range from 14° to 60.5°. Possible Bragg reflection positions for Sm2CuMn(MnTi3)O12 (the first row) and CuO impurities (the second row) are shown. The (hkl) indices of all of the observed reflections of Sm2CuMn(MnTi3)O12 are given.
Materials 15 08306 g002
Figure 3. Full experimental (black crosses), calculated (red line), and difference (blue line at the bottom) room-temperature synchrotron powder X-ray diffraction patterns of Sm2CuMn(MnTi3)O12 in a 2θ range of 6° and 60°. The brown tick marks show possible Bragg reflection positions for the main phase, the blue tick marks are for CuO impurity (2.0 wt.%), and the green ones are for Pt impurity (0.3 wt.%). The inset shows a tetragonal crystal structure of Sm2CuMn(MnTi3)O12; TiO6 octahedra (gray), MnO4 tetrahedra (green), and ideal CuO4 square-planar units (red) are plotted; Sm atoms are given by black circles; split Cu sites are shown by yellow circles.
Figure 3. Full experimental (black crosses), calculated (red line), and difference (blue line at the bottom) room-temperature synchrotron powder X-ray diffraction patterns of Sm2CuMn(MnTi3)O12 in a 2θ range of 6° and 60°. The brown tick marks show possible Bragg reflection positions for the main phase, the blue tick marks are for CuO impurity (2.0 wt.%), and the green ones are for Pt impurity (0.3 wt.%). The inset shows a tetragonal crystal structure of Sm2CuMn(MnTi3)O12; TiO6 octahedra (gray), MnO4 tetrahedra (green), and ideal CuO4 square-planar units (red) are plotted; Sm atoms are given by black circles; split Cu sites are shown by yellow circles.
Materials 15 08306 g003
Figure 4. The left-hand axis shows the ZFC (filled symbols) and FCC (empty symbols) dc magnetic susceptibility curves (χ = M/H) of Sm2CuMn(MnTi3)O12 at 100 Oe (black) and 10 kOe (red). The right-hand axis gives the 10 kOe FCC χ−1 versus T curve with the Curie–Weiss fit between 250 K and 345 K (black line). Parameters of the fits are shown on the figure. The inset shows details below 30 K.
Figure 4. The left-hand axis shows the ZFC (filled symbols) and FCC (empty symbols) dc magnetic susceptibility curves (χ = M/H) of Sm2CuMn(MnTi3)O12 at 100 Oe (black) and 10 kOe (red). The right-hand axis gives the 10 kOe FCC χ−1 versus T curve with the Curie–Weiss fit between 250 K and 345 K (black line). Parameters of the fits are shown on the figure. The inset shows details below 30 K.
Materials 15 08306 g004
Figure 5. M versus H curves of Sm2CuMn(MnTi3)O12 at T = 2 K (black) and T = 5 K (red) (f.u.: formula unit). The inset shows details near the origin. Parameters of the M versus H curve at T = 2 K are given: MS is the magnetization value at H = 70 kOe, MR is the remnant magnetization, and HC is the coercive field.
Figure 5. M versus H curves of Sm2CuMn(MnTi3)O12 at T = 2 K (black) and T = 5 K (red) (f.u.: formula unit). The inset shows details near the origin. Parameters of the M versus H curve at T = 2 K are given: MS is the magnetization value at H = 70 kOe, MR is the remnant magnetization, and HC is the coercive field.
Materials 15 08306 g005
Figure 6. (a) Real (χ′) and (b) imaginary (χ″) parts of the ac magnetic susceptibility curves of Sm2CuMn(MnTi3)O12 at different frequencies. The insets in (a,b) show the χ′ versus T and χ″ versus T curves at different Hac = 0.05, 0.5, and 5 Oe and one frequency (f = 300 Hz) (the χ″ data at Hac = 0.05 Oe are not shown because they were too noisy).
Figure 6. (a) Real (χ′) and (b) imaginary (χ″) parts of the ac magnetic susceptibility curves of Sm2CuMn(MnTi3)O12 at different frequencies. The insets in (a,b) show the χ′ versus T and χ″ versus T curves at different Hac = 0.05, 0.5, and 5 Oe and one frequency (f = 300 Hz) (the χ″ data at Hac = 0.05 Oe are not shown because they were too noisy).
Materials 15 08306 g006
Figure 7. (ad) χ′ versus T curves of Sm2CuMn(MnTi3)O12 at f = 2 Hz and 500 Hz and different bias dc fields: (a) Hdc = 0 Oe, (b) 100 Oe, (c) 1 kOe, and (d) 10 kOe. Insets show χ″ versus T curves (the χ″ data at f = 2 Hz and Hdc = 10 kOe were too noisy and not shown). (e) All χ′ versus T curves at f = 500 Hz are shown in one figure. (f) All χ″ versus T curves at f = 500 Hz are shown in one figure.
Figure 7. (ad) χ′ versus T curves of Sm2CuMn(MnTi3)O12 at f = 2 Hz and 500 Hz and different bias dc fields: (a) Hdc = 0 Oe, (b) 100 Oe, (c) 1 kOe, and (d) 10 kOe. Insets show χ″ versus T curves (the χ″ data at f = 2 Hz and Hdc = 10 kOe were too noisy and not shown). (e) All χ′ versus T curves at f = 500 Hz are shown in one figure. (f) All χ″ versus T curves at f = 500 Hz are shown in one figure.
Materials 15 08306 g007
Figure 8. Relaxation curves defined as 100 × [M(t) − M(0)]/M(0) versus time (t) for Sm2CuMn(MnTi3)O12 at temperatures of 2, 3, 4, 5, 6, 7, and 10 K. Experimental points are given by symbols, and the red line shows the fit at 3 K as an example. The equation used for fitting and the resultant parameters are listed in Table 3.
Figure 8. Relaxation curves defined as 100 × [M(t) − M(0)]/M(0) versus time (t) for Sm2CuMn(MnTi3)O12 at temperatures of 2, 3, 4, 5, 6, 7, and 10 K. Experimental points are given by symbols, and the red line shows the fit at 3 K as an example. The equation used for fitting and the resultant parameters are listed in Table 3.
Materials 15 08306 g008
Figure 9. (a) Cp/T versus T curves of Sm2CuMn(MnTi3)O12 at H = 0 Oe (red triangles) and 90 kOe (blue triangles) in comparison with the parent compound Sm2MnMn(MnTi3)O12 at H = 0 Oe (black circles) and 90 kOe (brown circles). Cp is the total specific heat. The arrows show the positions of the magnetic anomalies. The inset shows the Cp/T versus T curve at H = 0 Oe below 30 K, and the Cp versus T curve at H = 0 Oe (green circles). For the Cp versus T curve, the Cp values were divided by 20, and the Cp unit is J K−1 mol−1. (b) Comparison of Cp/T versus T data for Sm2CuMn(MnTi3)O12 (red triangles) and Sm2MnZn(MnTi3)O12 (black circles) [34] at H = 0 Oe.
Figure 9. (a) Cp/T versus T curves of Sm2CuMn(MnTi3)O12 at H = 0 Oe (red triangles) and 90 kOe (blue triangles) in comparison with the parent compound Sm2MnMn(MnTi3)O12 at H = 0 Oe (black circles) and 90 kOe (brown circles). Cp is the total specific heat. The arrows show the positions of the magnetic anomalies. The inset shows the Cp/T versus T curve at H = 0 Oe below 30 K, and the Cp versus T curve at H = 0 Oe (green circles). For the Cp versus T curve, the Cp values were divided by 20, and the Cp unit is J K−1 mol−1. (b) Comparison of Cp/T versus T data for Sm2CuMn(MnTi3)O12 (red triangles) and Sm2MnZn(MnTi3)O12 (black circles) [34] at H = 0 Oe.
Materials 15 08306 g009
Figure 10. Temperature dependence of the dielectric constant at different frequencies in Sm2CuMn(MnTi3)O12. The inset shows temperature dependence of dielectric loss in the logarithmic scale.
Figure 10. Temperature dependence of the dielectric constant at different frequencies in Sm2CuMn(MnTi3)O12. The inset shows temperature dependence of dielectric loss in the logarithmic scale.
Materials 15 08306 g010
Table 1. Structure parameters of Sm2CuMn(MnTi3)O12 from synchrotron powder diffraction data (λ = 0.65298 Å) at room temperature.
Table 1. Structure parameters of Sm2CuMn(MnTi3)O12 from synchrotron powder diffraction data (λ = 0.65298 Å) at room temperature.
Crystal systemTetragonal
Space groupP42/nmc (No. 137, cell choice 2)
Z2
Caclulated density (g/cm3)6.15
Formula weight (g/cm3)809.737
Used d range (Å)0.6507–6.238
a (Å)7.53477(1)
c (Å)7.69788(1)
V3)437.0301(8)
g(Sm)0.9413(9)Sm + 0.0587Mn
z(Sm)0.22194(4)
B(Sm) (Å2)0.796(7)
g(Cu)0.4597MC + 0.0403Sm
z(Cu)0.7804(3)
B(Cu) (Å2)1.25(7)
g(Mn)0.963(3)Mn + 0.037Sm
B(Mn) (Å2)0.38(6)
g(Ti)0.25MC + 0.75Ti
B(Ti) (Å2)0.400(9)
y(O1)0.0571(3)
z(O1)−0.0379(3)
B(O1) (Å2)0.33(5)
y(O2)0.5363(3)
z(O2)0.5745(3)
B(O2) (Å2)0.36(5)
x(O3)0.44161(23)
B(O3) (Å2)1.48(7)
Rwp (%)3.16
Rp (%)2.17
RI (%)2.63
RF (%)1.69
The Sm site is in the 4d site (0.25, 0.25, z); Cu is in the 4c site (0.75, 0.25, z); Mn is in the 2b site (0.75, 0.25, 0.25); Ti is in the 8e site (0, 0, 0); O1 and O2 are in the 8g site (0.25, y, z), and O3 is in the 8f site (x, −x, 0.25). g is the occupation factor. g(O1) = 1, g(O2) = 1, and g(O3) = 1. MC is a virtual atom: 0.5Mn + 0.5Cu.
Table 2. Bond lengths (in Å), bond angles (in deg), and distortion parameters of TiO6 (Δ) in Sm2CuMn(MnTi3)O12 at room temperature.
Table 2. Bond lengths (in Å), bond angles (in deg), and distortion parameters of TiO6 (Δ) in Sm2CuMn(MnTi3)O12 at room temperature.
Sm–O1 × 22.352(3)
Sm–O1 × 22.472(2)
Sm–O2 × 22.437(2)
Sm–O3 × 42.744(1)
Cu–O3 × 42.055(2)
Mn–O2 × 42.102(3)
Ti–O1 × 21.954(1)
Ti–O2 × 21.988(1)
Ti–O3 × 22.023(1)
Δ(TiO6)2.0 × 10−4
Ti–O1–Ti × 2149.17(9)
Ti–O2–Ti × 2142.75(9)
Ti–O3–Ti × 2144.17(9)
Table 3. Results of the fittings of the relaxation curves of Sm2CuMn(MnTi3)O12 at different temperatures.
Table 3. Results of the fittings of the relaxation curves of Sm2CuMn(MnTi3)O12 at different temperatures.
T (K)M0MSGtr (s)β
222.80(13)23.67(18)1270(24)0.4362(4)
326.01(12)27.20(20)979(16)0.4461(4)
422.81(10)23.94(18)800(11)0.4495(4)
517.74(6)18.62(13)671(8)0.4558(4)
612.34(14)13.01(9)554(6)0.4654(4)
The fitting equation is f(t) = M0MSG × exp[−(t/tr)β] [30] applied to the 100 × [M(t) − M(0)]/M(0) versus time (t) curves.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Belik, A.A.; Liu, R.; Yamaura, K. Dielectric and Spin-Glass Magnetic Properties of the A-Site Columnar-Ordered Quadruple Perovskite Sm2CuMn(MnTi3)O12. Materials 2022, 15, 8306. https://doi.org/10.3390/ma15238306

AMA Style

Belik AA, Liu R, Yamaura K. Dielectric and Spin-Glass Magnetic Properties of the A-Site Columnar-Ordered Quadruple Perovskite Sm2CuMn(MnTi3)O12. Materials. 2022; 15(23):8306. https://doi.org/10.3390/ma15238306

Chicago/Turabian Style

Belik, Alexei A., Ran Liu, and Kazunari Yamaura. 2022. "Dielectric and Spin-Glass Magnetic Properties of the A-Site Columnar-Ordered Quadruple Perovskite Sm2CuMn(MnTi3)O12" Materials 15, no. 23: 8306. https://doi.org/10.3390/ma15238306

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop