2016 Volume 63 Issue 7 Pages 605-608
A new double perovskite Ca2FeMnO6 with a layered arrangement of Mn4+ and the unusually high valence Fe4+ was obtained by oxidizing the brownmillerite Ca2FeMnO5 with ozone at 200 °C. The low-temperature topotactic reaction kept the layered cation arrangement of the brownmillerite but oxidized Mn3+ to Mn4+ and Fe3+ to Fe4+. The crystal structure with a P21/c space group was analyzed with the synchrotron X-ray diffraction data. The changes in the valence states of Fe and Mn were confirmed by bond valence sum calculations, X-ray absorption spectroscopy, and Mössbauer spectra measurements.
The stable valence states of Fe ions in oxides are 2+ and 3+. Unusually high valence states of Fe, however, can be stabilized by synthesis under strong oxidizing atmosphere. CaFeO3 is one of the oxides containing such unusually high valence Fe ions1–3). The compound shows a characteristic behavior, i.e. charge disproportionation (2Fe4+ → Fe3+ + Fe5+) at 290 K with the structural and metal-to-insulator transitions. In the charge disproportionated state below the temperature Fe3+ and Fe5+ ions are arranged in a rock-salt type manner4). LaCu3Fe4O12 is another compound containing the unusually high valence Fe5). The compound with high valence state Fe3.75+ shows intersite charge transfer (3Cu2+ + 4Fe3.75+ → 3Cu3+ + 4Fe3+) at 393 K with paramagnetism-to-antiferromagnetism and metal-to-insulator transitions. Both charge disproportionation and intersite charge transfer can be considered as the results of localization of oxygen p holes, which are produced by strong hybridization of Fe-d and O-p orbitals, to relieve the instability of the unusually high valence states. In CaFeO3 and LaCu3Fe4O12 the Fe ions are arranged in the three-dimensional perovskite-type structures, and the localization of p holes at the Fe sites results in the charge disproportionation whereas that at the Cu sites gives the intersite charge transfer.
There have been so far no reports on compounds with unusually high valence Fe ions arranged in a two-dimensional way. We thus attempt to make such a new compound with Fe4+ and another transition-metal ion in a layered arranged manner. Because highly oxidizing atmosphere at high temperature is needed to stabilize the unusually high valence state of Fe in the oxides1,5), it is usually difficult to keep the layered arrangement of the two kinds of transition-metal ions. In this study, we focus on a brownmillerite-type structure compound, Ca2FeMnO5 with the layered arrangement of Fe3+ and Mn3+ 6,7), and modify it to a layered double perovskite Ca2FeMnO6 with Fe4+ and Mn4+ (see the crystal structures in Fig. 1) by low temperature topochemical oxidation with ozone. An advantage by using an ozone strong oxidizing agent is that we can increase the Fe valence state in the oxide at a relatively low temperature with keeping the ion ordered arrangement.
Crystal structures of brownmillerite Ca2FeMnO5 (left) and double perovskite Ca2FeMnO6 (right). Ca2FeMnO6 is obtained from Ca2FeMnO5 by low-temperature topotactic oxidation by ozone.
The precursor material Ca2FeMnO5 was prepared by solid-phase reaction. Stoichiometric amount of Ca2CO3, Fe2O3, and MnCO3 were mixed, calcined at 1000 °C in air, ground, and fired at 1250 °C in air. The obtained powder sample was then pressed into a pellet and annealed at 1100 °C in vacuum. The prepared Ca2FeMnO5 sample was oxidized to Ca2FeMnO6 at 200 °C for 6 hours in ozone containing oxygen flow.
The crystal structures of the compounds were analyzed with powder X-ray diffraction (XRD) data taken at room temperature at the beamline BL02B2 in SPring-8 with a wave length 0.775 Å. Rietveld refinement was performed using the RIETAN-FP program8). The valence states of Fe and Mn were estimated by the bond valence sum (BVS) calculations from the refined bond lengths9,10), 57Fe Mössbauer spectroscopy, and X-ray absorption spectroscopy (XAS). The 57Fe Mössbauer spectra were obtained in transmission geometry in combination with a constant-acceleration spectrometer using 57Co/Rh as a radiation source. α-Fe was used as a control for velocity calibration and isomer shift (IS). The obtained spectra were fitted by a least-squares method with Lorentzian functions. XAS spectra near K-edges of Fe and Mn were measured at the beam line BL01B1 in SPring-8. The spectra at 300 K were obtained by a transmission method.
Fig. 2 (a) shows the XRD pattern of the precursor Ca2FeMnO5 and the result of Rietveld refinement. The observed pattern is well fitted with the brownmillerite structure with the Pnma space group. The refinement results listed in Table 1 are essentially the same as those reported previously11,12). The Fe and Mn ions respectively are located at the tetrahedral and octahedral sites in a layered manner along the b axis. The refined Fe-O bond lengths, 1.805(5), 1.805(5), 1.967(1), and 1.869(1) Å give the BVS value of 3.1 (Table 2), indicating Fe3+ at the tetrahedral site. The Mn-O bond lengths, 2.244(5) Å, along the b axis and 1.922(1) and 1.903(1) Å along the a and c axes, form distorted octahedra and give the BVS value of 3.2 (Table 2), suggesting Mn3+. This highly anisotropic distortion of the octahedra is thus due to the Jahn-Teller active Mn3+ ions. The structure refinement result confirms that Ca2FeMnO5 crystallizes in the brownmillerite structure with the layered arrangement of the Fe3+O4 tetrahedra and the Mn3+O6 octahedra.
Synchrotron X-ray diffraction patterns at room temperature and the results of Rietveld analyses for brownmillerite Ca2FeMnO5 and double perovskite Ca2FeMnO6. The dots and solid line represent observed and calculated patterns, respectively. The plots below the diffraction patterns are the differences between the observed and calculated intensities. Vertical marks below the profiles are Bragg reflection positions of main phase (green) and a small amount of Ca2MnO4 impurity (purple).
Atom | x | y | z | B (Å2) | Occupancy |
---|---|---|---|---|---|
Ca1 | 0.9862(5) | 0.1095(1) | 0.4779(4) | 0.79(5) | 1.0 |
Fe1 | 0.9547(5) | 0.25 | 0.9399(5) | 0.86(6) | 1.0 |
Mn1 | 0.0 | 0.0 | 0.0 | 0.35(5) | 1.0 |
O1 | 0.260(3) | 0.9872(4) | 0.244(1) | 0.9(1) | 1.0 |
O2 | 0.019(1) | 0.14411(3) | 0.068(1) | 0.9(1) | 1.0 |
O3 | 0.090 (2) | 0.25 | 0.623(2) | 0.9(1) | 1.0 |
Compound | Ion | BVS |
---|---|---|
Ca2FeMnO5 | Fe | 3.1 |
Mn | 3.2 | |
Ca2FeMnO6 | Fe | 4.0 |
Mn | 4.1 |
The valence state of Fe in Ca2FeMnO5 is also examined by Mössbauer spectroscopy. The spectrum at room temperature shown in Fig. 3 (a) is fitted with a magnetically ordered sextet component, and its IS and hyperfine field respectively are 0.18 mm/sec 35.0 T. These values are typical for magnetic Fe3+ and they are also close to the values reported in the previous reports13). The result thus also confirms that the Fe ion in the brownmillerite Ca2FeMnO5 has 3+ valence state at the single tetrahedral site.
Mössbauer spectra of Ca2FeMnO5 (a) and Ca2FeMnO6 (b) at room temperature. The dots are observed data and solid lines are fittings.
Fig. 2 (b) shows the XRD pattern of the ozone oxidized Ca2FeMnO6. The pattern is well reproduced with a perovskite structure model with the space group of P21/c and √2aPC × 2aPC × √2aPC unit cell (aPC represents a perovskite unit cell). The details of refined parameters are listed in Table 3. All the oxygen sites are fully occupied making the FeO6 and MnO6 octahedra, and the crystal structure of Ca2FeMnO6 is the double perovskite. Although the refined crystal structure has the monoclinic space group P21/c symmetry, the refined β angle converges to 90.00(1) °. The BVS values are 4.0 for Fe and 4.1 for Mn (Table 2), confirming that both Fe and Mn are oxidized to the high valence states. By the low-temperature topotactic reaction, oxygen ions are introduced into the Ca2FeMnO5, changing the Fe3+O4 tetrahedra to the Fe4+O6 octahedra in Ca2FeMnO6 with keeping the two-dimensional layered arrangement of Fe and Mn ions. Since the scattering factors of X-ray for Fe and Mn are close each other, it was difficult to confirm the layered ordering by X-ray diffraction. We thus analyzed the detailed structure with neutron diffraction, and the refined B-site occupancies of Fe/Mn were 92.2(4)/7.8(4) and 14.4(2)/85.6(2) for each layer confirming that the layered ordering remains in Ca2FeMnO6 6).
Atom | x | y | z | B (Å2) | Occupancy |
---|---|---|---|---|---|
Ca1 | 0.2470(7) | 0.9918(8) | 0.0307(3) | 0.46(3) | 1.0 |
Fe1 | 0.0 | 0.5 | 0.0 | 0.4(1) | 1.0 |
Mn1 | 0.5 | 0.5 | 0.0 | 0.1(1) | 1.0 |
O1 | 0.253(2) | 0.081(2) | 0.4879(9) | 0.1(1) | 1.0 |
O2 | 0.16(2) | 0.722(4) | 0.272(4) | 0.3(2) | 1.0 |
O3 | 0.531(2) | 0.303(4) | 0.726(4) | 0.5 | 1.0 |
The Mössbauer spectroscopy for Ca2FeMnO6 also confirms the unusually high valence state of Fe. The spectrum at room temperature, shown in Fig. 3 (b), consists of a singlet component with IS of 0.02 mm/sec, which is close to that observed for Fe4+ in CaFeO3 1). Because no other spectra observed, all Fe3+ ions in the precursor brownmillerite are fully oxidized to Fe4+ in the double perovskite Ca2FeMnO6.
The changes in the valence states of both Fe and Mn are also clearly seen in the XAS results. Fig. 4 (a) shows the XAS spectra near the Fe K-edge for Ca2FeMnO5 and Ca2FeMnO6. The figure also includes the spectrum for Fe2O3 as a reference of Fe3+. The Fe absorption edge of Ca2FeMnO5 at about 7120 eV is quite similar to that of Fe2O3, confirming Fe in Ca2FeMnO5 is in the Fe3+ valence state. The absorption edge of Ca2FeMnO6 is shifted to a higher energy than those of Fe2O3 and Ca2FeMnO5, indicating that Fe in Ca2FeMnO6 becomes in the unusually high valence state. It is also interesting to see the change in the peak shape of the pre-edge at about 7109 eV. This pre-edge peak mainly reflects the Fe-3d electronic states through the 1s→3d quadrupole transition, but the 1s→4p dipole transition also contributes the intensity when the Fe is located at the tetrahedral site. The observed decrease in the intensity of the pre-edge peaks clearly indicates that the oxygen coordination of Fe changes from the tetrahedral one in Ca2FeMnO5 to the octahedral one in Ca2FeMnO6. The Mn K-XAS spectra also show the clear shift by the low-temperature oxidation, as shown in Fig. 4 (b). The absorption edge of Ca2FeMnO6, which is located at about 6550 eV, is higher than that of Ca2FeMnO5 at 6547 eV, indicating the increase in the valence state of Mn.
X-ray absorption spectra of Ca2FeMnO5 and Ca2FeMnO6 near Fe-K (a) and Mn-K (b) edges The spectrum of Fe2O3 is also shown in (a) as a reference of Fe3+.
All the results of the crystal structure analysis, the Mössbauer spectroscopy, and the XAS measurements, show that the precursor brownmillerite Ca2FeMnO5 changes to the fully oxygenated double perovskite Ca2FeMnO6 by low-temperature topotactic reaction. The Fe3+ in the tetrahedra and Mn3+ in the octahedra in Ca2FeMnO5 change to the unusually high valence Fe4+ and Mn4+ both in the octahedra in Ca2FeMnO6. An important finding in the present study is that the two-dimensional layered arrangement of Fe and Mn are kept, and the obtained Ca2FeMnO6 has the two-dimensional arranged unusual Fe4+ ions in the crystal structure, which is the first compound made. Interestingly, we recently found that Fe4+ in Ca2FeMnO6 shows charge disproportionation in the two-dimensional layers6,7).
We prepared the brownmillerite-structure oxide Ca2FeMnO5 and oxidized it to a layered double perovskite Ca2FeMnO6 by low-temperature topotactic oxidation with ozone. The precursor Ca2FeMnO5 had the layered arrangement of Fe3+O4 tetrahedra and highly distorted Mn3+O6 octahedra, while the oxidized Ca2FeMnO6 had the layered arrangement of Fe4+O6 and Mn4+O6 octahedra. The changes in the valence states of Fe and Mn ions were confirmed by BVS calculations, XAS measurement and Mössbauer spectroscopy. The obtained Ca2FeMnO6 had the two-dimensional layered arrangement of the unusually high valence Fe4+.
We thank D. Kan and H. Seki in ICR Kyoto Univ. (Japan) for useful discussions. The synchrotron radiation experiments at SPring-8 were performed with the approval of the Japan Synchrotron Radiation Research Institute (proposal Nos: 2014B1770, 2013A1008, 2013A1009, 2013B1017, and 2012B1978). This work was supported by Grants-in-Aid for Scientific Research (Nos. 19GS0207, 22740227, and 24540346) and by a grant for the Joint Project of Chemical Synthesis Core Research Institutions from the Ministry of Education, Culture, Sports, Science and Technology of Japan. The work was also supported by Japan Science and Technology Agency, CREST.