The mixed valence multiferroic LnFe2+Fe3+O4 (where Ln = Y, Lu, and Yb) can reversibly uptake oxygen into its lattice, which is evidenced by a crystallographic phase transition along with the appearance of structural modulations. In this study, we show that the Mn-substituted version of this multiferroic can also be readily oxidized to LnFe3+Mn3+O4.5 revealing similar oxygen storage behavior. Through neutron, electron, and synchrotron x-ray diffraction studies, we observe a structural modulation that we attribute to a displacement wave in the fully oxidized compound. This wave exhibits commensurability with a wavevector q = (−2/7, 1/7, 0). Bond valence summation analysis of plausible interstitial oxygen positions suggests that oxygen insertion likely occurs at the middle of the Fe/Mn–O bipyramid layers. The structural modulation of LnFeMnO4.5 is two-dimensional, propagates along the ab-plane, and is highly symmetric as 12 identical modulation vectors are observed in the diffraction patterns. The nature of the lanthanide, Ln3+, does not seem to influence such modulations since we observe identical satellite reflections for all three samples of Ln = Y, Lu, and Yb. Both LnFeMnO4 and LnFeMnO4.5 display spin glassy behavior with 2D short-range magnetic ordering being observed in LnFeMnO4. Analysis of the neutron diffraction data reveals a correlation length of ∼10 nm. Upon oxidation to LnFeMnO4.5, the short-range magnetic order is significantly suppressed.
I. INTRODUCTION
Hexagonal-layered LnM2+M3+O4 (Ln = lanthanide, M = transitional metal) has been attracting great interest in materials chemistry and physics due to its unique crystal structure and high tunability in compositions.1–4 The structure of LnM2+M3+O4 consists of alternating layers of edge-sharing LnO octahedra with a double layer of edge-sharing MO trigonal bipyramids. The M2+ and M3+ cations occupy the same crystallographic sites. In the LnO octahedra layer, the Ln3+ cations arrange in a triangular lattice, whereas in the MO layer, the M2+/M3+ cations are found in a non-corrugated honeycomb lattice. Both the triangular5 and the honeycomb6 lattices can display geometric frustration relevant to magnetism. This potential spin frustration has attracted attention from the condensed matter community as this family of oxides could host quantum spin-liquid behavior. Such candidates include those where S = ½ as found in YbMgGaO4,7–10 YbZnGaO4,11,12 and LuCuGaO4.13 The fact that M2+ and M3+ share the same crystallographic site also leads to a great variety of interesting chemical and physical phenomena. In the famous case of mixed-valence LnFe2O4, where Fe2+ and Fe3+ coexist at the same sites, ferrimagnetic ordering coincides with charge ordering between Fe2+ and Fe3+, leading to multiferroicity.14–19
Oxygen non-stoichiometry in LnM2+M3+O4 has also attracted great interest since physical properties of LnM2+M3+O4, especially LnFe2O4, are extremely sensitive to oxygen stoichiometry.20–25 However, the oxygen-abundant form of LnFe2O4+δ was not realized until Hervieu et al. observed that LuFe2O4 could be oxidized to the hyper-stoichiometric LuFe2O4.5 at a relatively low temperature of 200 °C in air.26 The oxygen uptake behavior of other LnFe2O4 (Ln = Yb, Y, and In) was later reported following this discovery.27–31 The structural transition from LnFe2O4 to LnFe2O4.5 is also found to be highly reversible, implying a reversible oxygen uptake/release behavior in the LnFe2O4 system.29,30,32 The oxygen uptake behavior of LnFe2O4 suggests its potential as an oxygen storage material and can also lead to interesting structure transitions and physical property changes.26,29
Inspired by the interesting oxygen uptake behavior in LnFe2O4, we hypothesize that the substitution of Fe2+ by Mn2+ will control the oxygen uptake behavior and the nature of associated structural transitions and magnetism. In this study, we systematically explore the change before and after the oxidation of LnFe3+Mn2+O4. We focus on the difference in the average structure, local structure, structural modulation, and magnetic properties between the reduced phase and corresponding oxidized phase. We investigate the possible influence of the Ln cations on oxygen uptake and compare them with their LnFe2O4 analogs.
II. EXPERIMENT
A. Materials synthesis
LnMnFe2O4 (Ln = Yb, Lu, and Y) series, labeled as the reduced phases, were prepared using solid-state reactions.2 Ln2O3 (Ln = Yb, Lu, and Y), Fe2O3, MnO, and Fe powder (99.9% pure) were ground in a 1:1:2 ratio, respectively, to prepare around 2 g of the target compound. The powder mixtures were pressed into pellets of 13 mm in diameter. The pellets were placed in a 2 ml alumina crucible and sealed inside evacuated 8 mm diameter quartz ampoules. The ampoules were then heated to 1180 °C (heating rate 10 °C/min) for at least 12 h and subsequently quenched in an ice water bath.
The AB2O4 phase is metastable; hence, quenching is required to kinetically lock it in at lower temperatures. Quenching limits the number of oxide impurities. The elemental ratio of the product was verified with ICP and scanning electron microscopy (SEM, Hitachi SU-70) in conjunction with energy dispersive spectroscopy (EDS, Bruker XFlash 6/60), SEM-EDS. To prepare the oxidized phases, LnMnFe2O4.5 (Ln = Yb, Lu, and Y), for ex situ studies, the as-synthesized samples were oxidized at 600 °C (heating rate 10 °C/min) under air for 12 h and then cooled to room temperature. These samples are labeled as oxidized phases.
B. Thermogravimetric analysis (TGA)
TGA was conducted on reduced phase LnMnFe2O4 (Ln = Yb, Lu, Y). In total, 5–10 mg of sample was used for each measurement. TGA measurement with isotherm heating was performed. Samples of ∼10 mg were heated to 350 °C (ramping rate 10 °C/min) in air, and the temperature was held for 6 h.
C. High-resolution synchrotron x-ray and neutron diffraction
LnMnFe2O4 (Ln = Yb, Lu, and Y) series as well as their oxidized phases LnMnFe2O4.5 were characterized with high-resolution synchrotron x-ray powder diffraction (SXPD). The experiments were performed on the 11-BM beamline at the Advanced Photon Source (APS) at Argonne National Laboratory. X-rays were of wavelength 0.457 895 Å. Room temperature data were collected. To compare with the SXPD patterns and explore the possible magnetic structures, we also took time-of-flight (TOF) neutron diffraction patterns for LnMnFe2O4 and LnMnFe2O4.5 at different temperatures between 5 and 300 K. The TOF patterns were collected on the powder diffractometer POWGEN beamline at the Spallation Neutron Source (Oak Ridge National Laboratory). Rietveld analysis was carried out using JANA33,34 and GSASii.35 Structures of oxidized phases were solved using the charge flipping method.36,37 Structural modulation of the oxidized phase was refined with JANA.
D. Synchrotron x-ray and neutron pair distribution function analysis
The x-ray total scattering data used for x-ray (λ = 0.457 895 Å) pair distribution function (XPDF) analysis were collected on the 11-IDB beamline also at the APS. Data were collected at room temperature (∼300 K) and were reduced using GSASii.35 Neutron total scattering data were obtained at the POWGEN beamline at the SNS. Data were collected at 5 and 300 K.
E. Raman spectroscopy
Raman spectroscopy was obtained with Yvon Jobin LabRam ARAMIS using a 532 nm laser source. Powders were used for the measurement. Because the samples strongly absorb the laser light and can be oxidized, a long exposure time (>5 min) and a small aperture were used to obtain the Raman scattering signal.
F. Transmission electron microscopy (TEM) and electron energy loss spectroscopy (EELS)
Transmission electron microscopy (TEM) images, as well as electron diffraction patterns, were taken using a JEOL 2100F field emission gun (scanning) TEM, (S)TEM, equipped with Gatan image filter (GIF, model of Tridiem 863), operated at 200 kV accelerating voltage.
G. Magnetic susceptibility measurement
Temperature-dependent magnetic susceptibility measurements were performed on both reduced and oxidized phases with a SQUID magnetic property measurement system (MPMS, Quantum Design). For each sample, a portion of powder with known mass was loaded into a gel capsule with cotton suspended inside a plastic straw. Field-cooled (FC) and zero field-cooled (ZFC) magnetization measurements were recorded in direct current mode from room temperature (300 K) to base temperature under an applied field of 100 Oe. Zero field AC susceptibility at 1, 10, 100, and 1000 Hz was also performed at the temperature range of interest.
III. RESULTS
A. Average structure
Since crystal structure of LnMnFeO4 (Ln = Lu, Yb, and Y) is isostructural to the parent phase LnFe2O4 (space group ),2,4,38 we used the latter’s structure as the starting model for our Rietveld analysis. The refinements were performed with combined neutron and synchrotron x-ray powder datasets. The neutron diffraction pattern with the Rietveld refinement fit for the YbFeMnO4 sample is displayed in Fig. 1(a), and the corresponding plots for the Ln = Lu and Y samples are presented in Figs. S4 and S8. The refined lattice parameters and atom positions of LnMnFeO4 (Ln = Lu, Yb, and Y) are shown in Tables S1, S3, and S6.
(a) Neutron powder diffraction pattern and the Rietveld refinement fit of (a) reduced phase YbMnFeO4 and (b) oxidized phase YbMnFeO4.5. Crystal structure of (c) YbMnFeO4 and (e) average crystal structure of YbMnFeO4.5 (Both and lead to the identical structure.).
(a) Neutron powder diffraction pattern and the Rietveld refinement fit of (a) reduced phase YbMnFeO4 and (b) oxidized phase YbMnFeO4.5. Crystal structure of (c) YbMnFeO4 and (e) average crystal structure of YbMnFeO4.5 (Both and lead to the identical structure.).
No significant difference was found between LnMnFeO4 and LnFe2O4. The Ln site displays significant displacement along the c-direction, which is commonly observed in the LnM2+M3+O4 phase.29,39 We noted that some studies use a structure model, where the position of Ln (0, 0, and 0) is split into two partially occupied sites (0, 0, and 0 ± δ).40 Both models suggest a static positional disorder on the Ln site.
We performed TGA measurement on LnMnFeO4 under the oxygen flow, and the result for YbFeMnO4 is displayed in Fig. 2 (Ln = Lu, Y in Figs. S4b and S8b). Similar to what was observed for LnFe2O4 (Ln = Lu, Yb, and Y),26,29,30,32 LnFeMnO4 also uptakes oxygen into its lattice at a relatively low temperature of 200 °C. The maximum weight gain was 2.51%, 2.21%, and 3.88%, corresponding to the change from LnFeMnO4 to LnFeMnO4.5. Assuming that Mn2+ is oxidized to Mn3+, the theoretical mass gain should be 2.29, 2.3, and 3.03% for Ln = Lu, Yb, and Y, respectively.
To verify the Mn oxidation state, we took electron energy-loss (EEL) spectra of Mn before and after oxidation (Fig. S13). By examining the electron energy-loss near-edge structure (ELNES) on the Mn L2,3-edge, which corresponds to the transitions from the Mn 2p core orbitals to the unoccupied Mn 3d state (dipole selection rule ∆ℓ = ±1, where ℓ is an angular quantum number),41 one can obtain information on the chemical coordination of Mn. Our EELS results confirm that the Mn valence state transits from Mn2+ into Mn3+,42 and more details on the ELNES analysis can be found in the supplementary material.
The powder diffraction patterns of LnFeMnO4.5 can be indexed with the space group, the same that was found in YbFe2O4.529 and LuFe2O4.5.26 However, we also found space group to provide an identical fit. We then, therefore, applied the charge flipping method to the SXRD patterns to elucidate the crystal structure of the oxidized phase, LnFeMnO4.5. Such a methodology helped determine the details of the Ln (Ln = Yb, Lu, and Yb), Fe/Mn, and one of the oxygen atom positions(O1). Afterward, we employed a Fourier difference map to locate the average positions of other oxygen atoms (O2 and O3) with both the NPD and SXRD datasets. Combined refinement was then performed based on the solved structure.
The neutron diffraction pattern with the Rietveld refinement fit of LnFeMnO4.5 is shown in Fig. 1(b) for Ln = Yb and Figs. S4c and S8c for Ln = Lu and Y. The refined lattice parameters and atom positions for LuFeMnO4.5, YbFeMnO4.5, and YFeMnO4.5 are included in Tables S2, S4, and S6. Not surprisingly, the solved average structure of LnFeMnO4.5 (Ln = Lu, Yb, and Y) resembles what is reported on YbFe2O4.529 and is highly correlated with the original reduced phase LnFeMnO4. Figures 1(c) and 1(d) show the structure of LnMnFeO4 and LnFeMnO4.5, respectively. Both and space groups can lead to the same structure for LnFeMnO4.5 because the atoms only occupy special locations.
From LnMnFeO4 to LnFeMnO4.5, the abundant oxygen atoms are inserted in the middle of Fe/Mn oxide bipyramid double layers, which leads to the shift of the layers along the ab-directions.29 Due to the layer shift, the newly generated unit cell is only 1/3 of the original unit cell (ao = ar, co = 1/3cr). Among all three LnFeMnO4.5 (Ln = Lu, Yb, and Y) samples, we noted that the diffraction peaks for LnFeMnO4.5 show a broader and more asymmetric nature than LnFeMnO4, which is an indication of stacking faults upon oxidation.43,44 The appearance of the stacking faults also implies the layer-shift mechanism upon oxygen uptake.
We note that the refinement on the LnFeMnO4.5 with the solved average structure gives relatively high Rwp values (16%). The misfit mainly arises from the unaccounted satellite peaks [Figs. 1(b), S4c, and S8c], which are similar in all three samples of Ln. The appearance of the satellite peaks indicates a structural modulation,45,46 which has also been observed in YbFe2O4.529 and LuFe2O4.5.26 Given the similarity of the average structures between LuFe2O4.5/YbFe2O4.5, structural modulation in LnFeMnO4.5 is expected. However, the positions and the distributions of the satellite peaks in LnFeMnO4.5 (Ln = Lu, Yb, and Y) appear to be completely different from those reported in YbFe2O4.5 and LuFe2O4.5. We will discuss the structural modulation of LnFeMnO4.5 in detail in Sec. III C. For both LnMnFeO4 and LnMnFeO4.5 (Ln = Lu, Yb, and Y), no cation ordering was observed for the Mn and Fe positions.
B. Local structure
The Rietveld refinement results indicate that the oxidation of LnFeMnO4 to LnFeMnO4.5 leads to a considerable contraction in the c-direction and a slight expansion along the a-direction when the unit cells are normalized to the same chemical formula (ar = 1/3cr, ao = co). This trend is not influenced by the nature of Ln, as shown in Tables S1–S6. This c-direction contraction is also evidenced by the Raman measurements (Figs. 3, S4g, and S7g), which afford information on the local structure. The Raman spectrum of LnFe2O4 is well studied,47–52 and the spectrum we recorded of YbFeMnO4 (blue curve in Fig. 3) resembles that of LnFe2O4. The major phonon modes of YbFeMnO4 are assigned based on the Raman spectra studies of LnFe2O4.51
Raman spectroscopy comparison between YbMnFeO4 and its oxidized phase YbMnFeO4.5.
Raman spectroscopy comparison between YbMnFeO4 and its oxidized phase YbMnFeO4.5.
For the Raman spectra of the oxidized phase YbFeMnO4.5 (red curve), we were not able to assign the phonon modes exactly due to the new space group symmetry and structural modulations. However, the structures of YbFeMnO4 and YbFeMnO4.5 are highly correlated; hence, there must be some resemblance between their phonon modes. The most intense peak (assigned to at 610 cm−1) for the reduced phase YbFeMnO4 corresponds to the stretching vibration of the lattice along the c-direction. After oxidation to YbFeMnO4.5, the layered framework remains; thus, this mode should still exist in the oxidized phase. We argue that the most intense peak blueshifts to 720 cm−1 for YbFeMnO4.5. Based on such an assignment, oxidation of these materials leads to stronger covalency and shorter bond distances along the c-direction.
Neutron PDF analysis also affords information on the local structure. The PDF curves before and after oxidation are presented in Fig. 4 for Ln = Yb (Fig. S4f for Ln = Lu). The comparison of neutron PDFs indicates that oxidation mainly changes the local environment of the Fe/Mn and O atoms. The Fe–O1 distance (along the c-direction) is shortened, and the Fe–O2 distance is elongated (along the horizontal direction) after oxygen uptake, which is also confirmed by the refined average structure. O–O distances around 2.5–3.0 Å are shortened after oxidation, implying that the crowding among the oxygens is due to oxygen insertion. It is also noted that the relative O–O pair intensities of LnFeMnO4.5 are enhanced compared with LnFeMnO4, confirming more oxygen atoms are accommodated into the structure after oxygen uptake. The x-ray PDFs of LnFeMnO4 and LnFeMnO4.5 (Ln = Yb and Lu) are shown in Figs. S3 and S7. The M-O and O–O pairs are much less obvious compared with the neutron PDFs due to the insensitivity of oxygen compared with neutron scattering. However, we do confirm from x-ray PDFs that the observable atomic distances along the c-direction are shortened after the oxygen uptake, consistent with the Raman and Rietveld results.
Neutron pair distribution function (PDF) comparison between YbMnFeO4 (top) and YbMnFeO4.5 (bottom).
Neutron pair distribution function (PDF) comparison between YbMnFeO4 (top) and YbMnFeO4.5 (bottom).
C. Satellite reflections from structural modulations
To better understand the structural differences between LnFeMnO4 and LnFeMnO4.5, high-resolution TEM (HRTEM) imaging and electron diffraction (ED) were performed on powder grains (Fig. 5 for Ln = Yb). The orientations of the ED patterns were determined based on the simulated electron diffraction (ED) patterns of the refined crystal structures (Figs. S11 and S12). For the reduced phase YbFeMnO4, we observed that most grains prefer to orient either along the [001] direction [Figs. 5(a) and 5(b)] or the [010] direction [Figs. 5(c) and 5(d)], suggesting that (001) and (010) are the preferred exposed facets for YbFeMnO4.
(a) HRTEM image and (b) ED pattern of YbFeMnO4 along the incident electron beam of [001]. (c) HRTEM image and (d) ED pattern of YbFeMnO4 along the incident electron beam of [010]. (e) HRTEM image and (f) ED pattern of YbFeMnO4.5 along the incident electron beam of [001]. (g) HRTEM image and (h) ED pattern of YbFeMnO4.5 along the incident electron beam of [110].
(a) HRTEM image and (b) ED pattern of YbFeMnO4 along the incident electron beam of [001]. (c) HRTEM image and (d) ED pattern of YbFeMnO4 along the incident electron beam of [010]. (e) HRTEM image and (f) ED pattern of YbFeMnO4.5 along the incident electron beam of [001]. (g) HRTEM image and (h) ED pattern of YbFeMnO4.5 along the incident electron beam of [110].
The fast Fourier-transform (FFT) pattern converted from the HRTEM images and ED patterns also shows a hexagonal arrangement of diffraction spots, confirming the rhombohedral cells. The HRTEM image [Fig. 5(c)] along the incident electron beam of [010] direction shows a layered stacking fringe, consistent with the refined crystal structure where LnO octahedra layers and Fe/MnO bipyramid double layers are alternatively stacked. It is reported that satellite peaks can be observed in the ED patterns for LnFe2O4 due to the charge ordering of Fe2+ and Fe3+.15,17,53,54 However, in our case of YbFeMnO4, there is no charge ordering because the Mn2+/Fe3+ pair is much more energy-favored than the Fe2+/Mn3+ pair.18,55
After oxygen uptake, the major Bragg diffraction spots along the [001] direction maintain the hexagonal arrangement, whereas those along the [110] direction remain vertically aligned, confirming the unchanged rhombohedral unit cell (a = b, α = β = 90°, and γ = 120°) after oxidation. HRTEM image [Fig. 5(g)] along the incident electron beam of [110] direction also shows a layered stacking fringe, indicating that the stacked layered framework persists after oxygen uptake. We also realized that most grains are now either oriented along the [001] direction [Figs. 5(e) and 5(f)] or the [110] direction [Figs. 5(g) and 5(h)] after the oxidation, which implies that the (110) planes observed in the oxidized phase are transformed from the (010) planes in the reduced phase. This transformation strengthens the argument that the structural transition from YbFeMnO4 to YbFeMnO4.5 is through the mechanism of layer shift along the ab directions.
Apart from the major reflections in the ED patterns for YbFeMnO4.5, satellite reflections are noted in both [001] and [110] directions [Figs. 5(f) and 5(h)]. The existence of satellite reflections is clearly an indication of structural modulations, which is consistent with the observed satellites in both synchrotron x-ray and neutron diffraction. In the [001] direction [Fig. 5(f)], each main Bragg diffraction spot was surrounded by 12 satellite spots. We interpret this number of satellites as a pair of reflections at each hexagonal edge with some angle of splitting. The symmetric arrangement of the satellites implies a modulation vector q of high symmetry. In the [110] direction [Fig. 5(h)], the additional superlattice spots were observed within the two main Bragg diffraction spots and were horizontal to the main Bragg diffraction spot with a modulation wave vectors q = (0, 1/3, and 0), suggesting that the modulation is 2D and propagates within the ab-plane [q= (a, b, and 0)]. Interestingly, the structural modulation seems so strong that it can be rendered in real-space HRTEM images. Both Figs. 5(e) and 5(g) show aperiodic structures with a nanometer scale apart from the atomic arrangement. The satellites are also visible in the FFT patterns [Figs. 5(e) and 5(g) inserts]. The observed HRTEM images and ED patterns are highly identical, regardless of the Ln site elements. Figures S14 and S15 show HRTEM images and ED patterns of LuFeMnO4 and YFeMnO4 before and after oxidation. The satellite Bragg diffraction spots appearing in the ED patterns of oxidized phases along [001] and [110] directions are highly identical to what is observed on YbFeMnO4.5, showing a 12-direction and highly symmetric structural modulation that propagates within the ab-plane.
Structural modulation is reported on LuFe2O4.5 and YbFe2O4.5 as well. However, the observed propagation of structural modulation in LnFe2O4.5 shows a great deviation from what we observed on LnFeMnO4.5. In the LnFe2O4.5 system (Ln = Yb and Lu), the structural modulation is reported to propagate along ac and bc planes [modulation wave vectors q = (a, 0, c) ].26,27,29 In addition, only 6 instead of 12 propagation directions of the modulation are observed in their ED patterns.26,27,29
D. Magnetic properties
The magnetic properties of reduced phase LnFeMnO4 and oxidized phase LnFeMnO4.5 are investigated through magnetic susceptibility and neutron diffraction measurements. Susceptibility measurement of YbFeMnO4 has been previously reported, and our results on YbFeMnO4 display similar behavior.18 The DC susceptibility from 2 to 300 K (Fig. S16a) shows that the reduced phase YbFeMnO4 displays an antiferromagnetic transition around 65 K. The Currie-Weiss fit (250–300 K, Fig. S16b) provides an effective moment size of 4.02 uB, per formula unit, which deviates significantly from the theoretical value of 11.8 uB based on the addition of all three magnetic atoms (one Mn2+, one Fe3+, and one Yb3+). This deviation suggests that there are still strong magnetic correlations among the magnetic ions of YbFeMnO4 even at room temperature.
The AC susceptibility measurements [Figs. 6(a) and 6(b)] show that the antiferromagnetic transition temperature varies when different frequencies are applied. We calculated the value of ΔT/T/Δ log f to be 0.007 (where ΔT = peak shift, T = peak temperature, and Δ log f = difference of the logarithm of frequency). This calculated value is close to the minimum one for spin glassy systems (∼0.004).56 In the imaginary part (χ″) for YbFeMnO4, a broad peak or shoulder was also observed below Tf (around 30 K) and was found to shift with changing the frequency, which may be related to the anisotropic spin freezing.57,58
AC susceptibility (a) real component and (b) imaginary component of YbFeMnO4. (c) Temperature dependent neutron diffraction of direction of YbFeMnO4 and (d) Zoom-in magnetic peaks. AC susceptibility (e) real component and (f) imaginary component of YbFeMnO4.5. (g) Temperature dependent neutron diffraction of direction of YbFeMnO4.5 and (h) zoom-in magnetic peaks.
AC susceptibility (a) real component and (b) imaginary component of YbFeMnO4. (c) Temperature dependent neutron diffraction of direction of YbFeMnO4 and (d) Zoom-in magnetic peaks. AC susceptibility (e) real component and (f) imaginary component of YbFeMnO4.5. (g) Temperature dependent neutron diffraction of direction of YbFeMnO4.5 and (h) zoom-in magnetic peaks.
Time of flight powder neutron diffractions at 300, 100, 40, and 5 K on YbFeMnO4 are presented in Fig. 6(c). We did not observe any structural transition from 300 to 5 K as the diffraction peaks from the nuclear component remain nearly identical across the whole temperature range. PDFs from 5 to 300 K do not show any major deviations either (Figs. S20 and S21). However, broad magnetic peaks at Q = 1.25 Å−1 and Q = 2.5 Å−1 grow upon cooling. The magnetic peaks appear even at room temperature, implying a strong magnetic correlation, which is consistent with the DC susceptibility analysis. We noted the magnetic peaks are considerably broad and highly asymmetric, suggesting the magnetic ordering is short range and two-dimensional.59–61 We indexed the magnetic peaks as the (1/3, 1/3, l) and (2/3, 2/3, l).
After oxidation to YbFeMnO4.5, the magnetic properties significantly altered. DC susceptibility measurements from 2 to 300 K (Fig. S16) indicate YbFeMnO4.5 also displays an antiferromagnetic transition, however, at a much lower temperature (around 30.5 K compared with 62 K for YbFeMnO4). The moment gained from the Currie-Weiss fit is 15.61 uB per formula, still deviating from the theoretical value (15.3 uB: one Mn3+, one Fe3+, one Yb3+). AC susceptibility measurements [Figs. 6(d) and 6(e)] show that this antiferromagnetic transition temperature varies when different frequencies are applied. We calculated ΔT/T/Δ log f to be 0.0066, still very close to the minimum value for spin glass systems. In the imaginary part (χ″) for YbFeMnO4.5, a relatively broad peak except for the peak for Tf was also observed around 10 K and was found to shift with changing the frequency [Fig. 6(e)].
Time of flight powder neutron diffraction at 300, 10, 40, and 5 K on LuFeMnO4.5 are presented in Fig. 6(f). Similar to YbFeMnO4, no nuclear structural transition is observed from 300 to 5 K (also true for the PDFs at 5 and 300 K). A magnetic peak is still visible at Q = 1.25 Å−1 even at room temperature, but more diffuse compared with YbFeMnO4.
We found that the Ln site element has minimal influence on the magnetic properties of the reduced phase LnFeMnO4 and the oxidized phase LnFeMnO4.5. The same magnetic susceptibility and temperature-dependent neutron diffraction measurements were performed on the Ln = Lu and Y samples, and the results are provided in Figs. S16–S18. Important parameters gained from the magnetic measurements are listed in Table S7. All LnFeMnO4 and LnFeMnO4.5 show spin-glass behavior. The effective moments from Curie-Weiss fits never agree with the theoretical values based on isolated moments. In the neutron datasets, the reduced phases of LnFeMnO4 all show broad and highly asymmetric magnetic peaks upon cooling. The positions of the magnetic peaks can be indexed into (1/3, 1/3, l) and (2/3, 2/3, l), regardless of the Ln cations. Upon oxidation, magnetic peaks become much less visible, and there is a decrease in temperature for the antiferromagnetic transition.
IV. DISCUSSION
A. Lattice contraction upon oxidation
Table I compares the normalized lattice parameters and some of the special distances within the crystal structure of LnFeMnO4 before and after oxygen uptake. The refined lattice parameters indicate that for all LnFeMnO4 (Ln = Y, Lu, and Yb) samples, oxygen uptake leads to contraction along the c-direction. Measuring the thickness of the LnO octahedra layer and the Fe/MnO bipyramid double layer, we found out the Fe/Mn bipyramid double layer undergoes contraction, whereas the LnO octahedra layer expands after the oxygen uptake, as shown in Table I. Thus, the contraction of the c-lattice parameter of the oxidized phase is mainly a result of the contraction of Fe/MnO bipyramid double layers. Previous studies show the oxidation of LnFe2O4 (Ln = Y, Lu, and Yb) leads to the expansion along c-directions,26,29,30 which is opposite to what we have observed on LnFeMnO4, as presented in Fig. 7.
Sublayer thickness and special bond distance of LnFeMnO4 and LnFeMnO4.5 (*O2–O3 and O3–O3 distance only exist in the oxidized phase LnFeMnO4.5. ** parameters for LnFeMnO4.5 are based on the average structure).
. | Sublayer thickness (Å) . | Bond distance (Å) . | ||||
---|---|---|---|---|---|---|
Material . | MO polyhedral double layer . | LnO octahedra layer . | M-M vertical . | M-M horizontal . | O2–O3 * . | O3–O3 * . |
LuMnFeO4 | 6.226 | 2.346 | 3.276 | 3.444 | ||
**LuMnFeO4.5 | 5.815 | 2.436 | 3.173 | 3.478 | 2.152 | 2.529 |
YbMnFeO4 | 6.231 | 2.304 | 3.25 | 3.459 | ||
**YbMnFeO4.5 | 5.756 | 2.488 | 3.175 | 3.488 | 2.169 | 2.488 |
YMnFeO4 | 6.167 | 2.32 | 3.194 | 3.496 | ||
**YMnFeO4.5 | 5.722 | 2.523 | 3.168 | 3.519 | 2.059 | 2.561 |
. | Sublayer thickness (Å) . | Bond distance (Å) . | ||||
---|---|---|---|---|---|---|
Material . | MO polyhedral double layer . | LnO octahedra layer . | M-M vertical . | M-M horizontal . | O2–O3 * . | O3–O3 * . |
LuMnFeO4 | 6.226 | 2.346 | 3.276 | 3.444 | ||
**LuMnFeO4.5 | 5.815 | 2.436 | 3.173 | 3.478 | 2.152 | 2.529 |
YbMnFeO4 | 6.231 | 2.304 | 3.25 | 3.459 | ||
**YbMnFeO4.5 | 5.756 | 2.488 | 3.175 | 3.488 | 2.169 | 2.488 |
YMnFeO4 | 6.167 | 2.32 | 3.194 | 3.496 | ||
**YMnFeO4.5 | 5.722 | 2.523 | 3.168 | 3.519 | 2.059 | 2.561 |
(a) c lattice parameter comparison of LnFeMnO4 before and after oxygen uptake. (b) c lattice parameter comparison of LnFe2O4 before and after oxygen uptake.
(a) c lattice parameter comparison of LnFeMnO4 before and after oxygen uptake. (b) c lattice parameter comparison of LnFe2O4 before and after oxygen uptake.
The oxygen uptake of LnFeMnO4 correlates with the oxidation of Mn2+ (d5) into Mn3+ (d4) within the structure, whereas the oxygen uptake of LnFe2O4 correlates with the oxidation of Fe2+(d6) into Fe3+(d5). The different electron configuration changes upon oxidation (d5 to d4 for oxidation of LnFeMnO4 and d6 to d5 for oxidation of LnFe2O4) can influence the crystal field effect in the Mn/FeO5 bipyramid,32,62 causing opposite lattice expansion behaviors for LnFeMnO4 and LnFe2O4.
B. Oxygen insertion mechanism
The oxygen uptake of LnFeMnO4 into LnFeMnO4.5 leads to the insertion of extra oxygen into the crystal lattice. The average structure of LnFeMnO4.5 [Fig. 1(d)] indicates that the extra oxygens (O3) are located in between the Fe/Mn–O double layer. Based on such an observation, intuitively, we can infer that the oxygen insertion process on LnFeMnO4.5 should happen in between the Fe/Mn–O bipyramid layers [in the middle of the Fe/Mn–O bipyramid double layer, close to the lattice plane (002)].29 If this proposed mechanism is correct, at the early stage of oxygen uptake of LnFeMnO4, where the oxygen uptake is far smaller than 0.5, the inserted oxygen atoms should exist between the Fe/Mn–O bipyramid layers as interstitial oxygen before the phase transition occurs. To further verify this proposed insertion mechanism, the bond valence summation approach63,64 was applied to the reduced phase LnFeMnO4 to estimate the relative energy of the interstitial oxygen ions at different locations. The program softBV is commonly used to estimate the migration pathways for ions based on the energy of the sites estimated from the bond valence summation.65,66 The calculated energy of sites can also be used to estimate the preferred occupying sites of certain ions. Figure 8 displays the sites for interstitial oxygen (position coordinates are included in Table S8), which are calculated to have the lowest energy. We clearly observe that the possible low-energy interstitial oxygen sites are all concentrated in between the Fe/Mn–O bipyramid layers, suggesting that the oxygen insertion in between the Fe/Mn–O bipyramid layers is the most energy favorable.
Energy favored oxygen interstitial sites in reduced phase LnFeMnO4 calculated by softBV.65,66 In the structure on the left, the darker the yellow color, the lower energy for interstitial oxygen to occupy and the higher chances for the interstitial oxygen to appear at the early stage of oxygen uptake for LnFeMnO4.
Energy favored oxygen interstitial sites in reduced phase LnFeMnO4 calculated by softBV.65,66 In the structure on the left, the darker the yellow color, the lower energy for interstitial oxygen to occupy and the higher chances for the interstitial oxygen to appear at the early stage of oxygen uptake for LnFeMnO4.
Bond valence summation analysis also indicates that the Mn/Fe sites are under-bonded in LnFeMnO4. The bond valence sum of the Mn/Fe site is calculated to be around +2.25 compared with the theoretic value of +2.5. This under-bonded environment in LnFeMnO4 potentially allows the Mn/Fe site to bond extra oxygen atoms under oxidation conditions.
C. Structural modulation of the oxidized phase LnFeMnO4.5
Structural modulation upon oxidation of LnFe2O4 into LnFe2O4.5 has been recently reported.26,27,29 Also, it was shown that the propagation of modulation changes in LuFe2O4, depending on how much oxygen the material can uptake.26,28 Our study only focuses on the fully oxidized phase LnFeMnO4.5. As mentioned previously, YFeMnO4.5, LuFeMnO4.5, and YbFeMnO4.5 possess similar modulation structures as they all show a similar arrangement of satellite peaks in the diffraction patterns. Figures 9(a) and 9(b) show the enlarged ED patterns of YbFeMnO4.5 from Figs. 5(f) and 5(h), respectively. The 12 modulation vectors can be clearly seen in Fig. 9(a). Since direction does not change between the real space and reciprocal space, the direction of modulations can be visualized in the real crystal structure based on the direction of satellites in the electron diffraction. Figure 9(c) presents the top view (from the c-direction) of the YbFeMnO4.5 average structure lattice, with the direction of modulation vectors indicated. We note that the directions of the 12 modulations are indeed identical in the average structure.
Analysis of modulation vector from electron diffraction patterns in (a) [001] direction and (b) [110] direction (diffraction peak profile shown in inset). (c) Display of the modulation vector in the real space structure in [001] direction. Refinement of modulation wave vector with (d) synchrotron x-ray and (e) neutron diffractions of YbFeMnO4.5. Subsets show the zoom in satellite peaks.
Analysis of modulation vector from electron diffraction patterns in (a) [001] direction and (b) [110] direction (diffraction peak profile shown in inset). (c) Display of the modulation vector in the real space structure in [001] direction. Refinement of modulation wave vector with (d) synchrotron x-ray and (e) neutron diffractions of YbFeMnO4.5. Subsets show the zoom in satellite peaks.
The angle between the wave vector q and a-axis is measured to be around 19°, aligned with the diagonal line of the two-unit cells [shown in Fig. 9(c)]. The modulation wave vector q is measured to be (−0.28, 0.14, and 0), close to the commensurate vector (−2/7, 1/7, 0). Assuming that it is indeed commensurate, the 12 modulation vectors are measured to be (−2/7, 1/7, 0), (1/7, −2/7, 0), (3/7, 1/7, 0), (3/7, 2/7, 0), (2/7, 3/7, 0), (1/7, 3/7, 0), (−3/7, −1/7, 0), (−3/7, −2/7, 0), (−2/7, −3/7, 0), (−1/7, −3/7, 0), (2/7, −1/7, 0), and (1/7, −2/7, 0).
Our observed ED patterns are homogeneous in the samples, as all tested crystalline grains show the same results. From these observations, we quickly realized that all 12 modulation wave vectors can be generated through symmetry operations from the space group (symmetry operations of are listed in Table S9)67 on wave vector q. However, applying symmetry operations from space group (symmetry operations of are listed in Table S9) on wave vector q can only generate six identical modulation wave vectors, in which case, two independent modulation wave vectors q = (−2/7, 17, 0) and q′ = (−17, 2/7, 0) are necessary to describe the structural modulations. From the symmetry of the modulation vectors, seems to be a better choice for describing the average structure of LnFeMnO4.5. However, it is rare but might be possible that the modulation wave vectors q = (−2/7, 1/7, and 0) and q′ = (−1/7, 2/7, and 0) are twined, in which case, will be a better choice for describing the average structure of LnFeMnO4.5.
Although LnFe2O4.5 also shows structural modulations,26,27,29 the observed modulation wave vectors of LnFe2O4.5 (Ln = Lu and Fe) are in great difference with our LnFeMnO4.5 system. LuFe2O4.5 show a modulation wave vector (0.31, 0, and 0.33)26 and YbFe2O4.5 show a modulation wave vector (0.31, 0, and 0.33),29 both of which propagate within bc- and ac-planes rather than the ab-plane. The multiplicity of the modulation wave vector observed in LuFe2O4.5 and YbFe2O4.5 is less than that of the LnFeMnO4.5 system, indicating a lower symmetry for the structural modulations. Since LnFe2O4.5 and LnFeMnO4.5 show nearly identical average structures, the ultimate reason for their very different modulation vectors is not clear. We previously noted that the oxidation in LnFe2O4 leads to c-direction expansion,26,29 whereas oxidation in LnFeMnO4 to LnFeMnO4.5 leads to c-direction contraction. Likely that different anisotropic lattice expansion behavior upon oxidation is correlated with the different structural modulations.
We further refined the structural modulation of LnFeMnO4.5 with the powder diffraction data. Although the measured modulation wave vector from ED results seems to be a commensurate value, treating the structure as a commensurate case would have required building a 7 × 7 supercell, which can be extremely complex. Thus, here, we approximate the structural modulation as an incommensurate case and apply the superspace approach.45,46 The modulation wave vectors are indexed and refined on both synchrotron and neutron diffractions. Figures 9(d) and 9(e) show the refinement of YbFeMnO4.5 with SXRD and NPD data. The refinements of LuFeMnO4.5 and YFeMnO4.5 are displayed in Figs. S21–S24. Here, we only applied one modulation wave vector q= (1/7, 3/7, and 0) with group, and we can see that is already enough to account for all the positions of satellite peaks, confirming the symmetric nature of 12 modulation vectors. The indexed and refined modulation vectors from powder diffractions are listed in Table II, which are in good agreement with ED. We also note that those satellite peaks in the powder diffraction patterns display extreme Lorentzian broadening features [Figs. 9(d) and 9(e)], suggesting that the modulation is not continuous across whole crystal grains but has nano-size domains.
Measured/indexed modulation vector of LnFeMnO4.5 from electron, x-ray, and Neutron diffractions.
. | Electron diffraction . | X-ray . | Neutron . |
---|---|---|---|
LuFeMnO4.5 | (0.14, 0.42, 0) | (0.1429, 0.4287, 0) | (0.1437, 0.4311, 0) |
YbFeMnO4.5 | (0.14, 0.42, 0) | (0.1433, 0.4299, 0) | (0.1431, 0.4293, 0) |
YFeMnO4.5 | (0.14, 0.42, 0) | (0.1445, 0.4335, 0) | (0.1435, 0.4305, 0) |
. | Electron diffraction . | X-ray . | Neutron . |
---|---|---|---|
LuFeMnO4.5 | (0.14, 0.42, 0) | (0.1429, 0.4287, 0) | (0.1437, 0.4311, 0) |
YbFeMnO4.5 | (0.14, 0.42, 0) | (0.1433, 0.4299, 0) | (0.1431, 0.4293, 0) |
YFeMnO4.5 | (0.14, 0.42, 0) | (0.1445, 0.4335, 0) | (0.1435, 0.4305, 0) |
Due to the existence of stacking faults and small domains of structural modulation, solving the modulation structure of LnFeMnO4.5 from only powder diffraction is extremely challenging. Thus, we aimed to qualitatively understand the modulation rather than fully solve the modulations. To alleviate the complexity modulation refinement, group was used, and only modulation along one q vector (1/7, 3/7, and 0) was refined. We also assumed the modulation along other modulation vectors should be symmetric or centrosymmetric to the modulation along one q vector (1/7, 3/7, and 0). We mainly used YbFeMnO4.5 to perform analysis for understanding, as the quality of diffraction patterns of LuFeMnO4.5 and YFeMnO4.5 are lower due to much more severe stacking faults. For SXRD, we only refined the modulation of Ln and Fe/Mn sites as x-ray is not sensitive to oxygen. For NPD of YbFeMnO4.5, the Yb, Fe/Mn, and all O sites were refined. The refinement on LuFeMnO4.5 was also attempted. The displacive modulation along a, b, and c for Yb, Fe, and oxygen atoms are shown in Figs. S25–S39, which is refined from neutron diffraction of YbFeMnO4.5. The refined modulation parameters are listed in Tables S10–S13.
Figure 10 shows a displacive modulation of O1, Fe/Mn, O3, and O2 sites along the c-direction. The Yb, Fe/Mn, and O1 sites all show strong periodic displacement along the modulation vector. Displacement of Fe and O1 along the c direction is relatively strong. We noted that the modulation curves for Fe/Mn and O1 along the c-direction have a phase difference of t/2, causing periodic breathing of Fe–O1 bonds, whereas O2 and O3 are partially occupied. Their occupancy periodicities also show a phase difference of roughly t/2, suggesting that O2 and O3 do not appear in the same unit cell. In the average structure of LnFeMnO4.5, the distance between O2 and O3 is around 2–2.2 A (Table I), which is far smaller than two times the radius of the O2− anion (2.4–2.8 A).68 The alternate occupancy of O2 and O3 makes physical sense as the occupancy at the same unit cell of O2 and O3 would cause a covalent overlap. Since structure has an inversion center, inversion of the modulation of O3 shows that the periodicity of the occupancy of O2 and O3 inversion sites will have a t/2 phase shift. The phase shift of the inversion sites O3′ (or O2′) indicates that O3′ and O3 should probably not appear in the same unit cell either. Since modulation appears after oxygen insertion, it is likely that the ordering of oxygen vacancy ordering is one of the major drivers for the structural modulations.
(a) Displacive modulation of O1, Fe/Mn, and O3 site along c direction. O1, Fe/Mn, and O3 are shown together since they have the same x and y coordinates (1/3, 2/3, and z). (b) Displacive modulation of O2 site along c direction. t represents modulation period.
(a) Displacive modulation of O1, Fe/Mn, and O3 site along c direction. O1, Fe/Mn, and O3 are shown together since they have the same x and y coordinates (1/3, 2/3, and z). (b) Displacive modulation of O2 site along c direction. t represents modulation period.
D. Short-range magnetic correlations
The magnetic properties of YFeMnO4 have been previously studied.60,61,69,70 The AC susceptibility suggests that LnFeMnO4 shows a weak spin glass-like state with the spin freezing transition. Our temperature-dependent neutron diffraction indicates that 2D short-range magnetic ordering exists in LnFeMnO4, regardless of Ln. Following the order of Ln = Y, Yb, and Lu, as the c-lattice parameter increases, the spin-frozen transition temperature decreases slightly increases (Table S7), suggesting the dominant magnetic interaction might be along the ab-plane rather than the c-direction.
Magnetization measurement on single crystal YFeMnO4 at 4.2 K shows an anisotropic ferrimagnetic behavior of YFeMnO4. A larger magnetic hysteresis loop was observed on YFeMnO4 parallel to c-direction than vertical to it. Two-dimensional ferrimagnetic ordering is also observed in the LnFe2O4 systems.14,71,72 In the LnFe2O4 or LnFeMnO4 systems, the Fe/Mn–O bipyramid double layers are separated by the Ln-O octahedral layers [shown in Fig. 1(c)], leading to extremely weak magnetic interactions between Fe/Mn–O bipyramid double layers in the c-direction.
Warren Line shape fitting of magnetic diffraction peak of (a)–(c) LnFeMnO4 and (d)–(f) LnFeMnO4.5.
Warren Line shape fitting of magnetic diffraction peak of (a)–(c) LnFeMnO4 and (d)–(f) LnFeMnO4.5.
Fitted magnetic correlation length of LnFeMnO4 before and after oxygen uptake.
Correlation length (Å) . | Lu . | Yb . | Y . |
---|---|---|---|
LnFeMnO4 | 75.6 | 115.6 | 89.9 |
LnFeMnO4.5 | 12.1 | 16.1 | 10.5 |
Correlation length (Å) . | Lu . | Yb . | Y . |
---|---|---|---|
LnFeMnO4 | 75.6 | 115.6 | 89.9 |
LnFeMnO4.5 | 12.1 | 16.1 | 10.5 |
Different from LnFeMnO4, long-range 2D ferrimagnetic magnetic ordering develops upon cooling in LnFe2O4,14,24,71,72,76 although the crystal structure and the location of the magnetic moments in the structure are identical for LnFeMnO4 and LnFe2O4. The disordering of Fe3+ and Mn2+ is believed to contribute to the lack of long-range magnetic ordering in the LnFeMnO4 system.60,61 Charge ordering exists in the LnFe2O4 system, where Fe2+(d6) and Fe3+(d5) are located at different layers within the Fe–O bipyramid double layer.15,49,71 For LnFeMnO4, Mn2+(d5) and Fe3+(d5) are randomly distributed Fe/Mn–O bipyramid double layers due to the lack of charge order. However, we must note that both Fe3+ and Mn2+ in LnFeMnO4 have the same electron configuration d5 despite carrying different charges. Fe3+ and Mn2+ should possess the same magnetic dipole moment as they are in the same trigonal bipyramidal crystal field environment in the structure. The magnetic superexchange interaction thus should be the same on each Mn2+ or Fe3+ site despite charge disordering.
Figure 12 displays a qualitative analysis of the superexchange interaction in the LnFeMnO4 system. Only the top and side views of the Fe/Mn–O bipyramid double layers are shown for simplicity. Two major superexchange pathways are indicated, J1 and J2. We write J1 between two Fe3+/Mn2+ ions that are connected through two O2− anions (M-O-M angle = 97.803°). Fe3+/Mn2+ ions form the top and bottom bipyramid layers, respectively. We designate J2 between two Fe3+/Mn2+ ions within the same bipyramid layer; hence, the cations are connected by only one O2− ion (M-O-M angle = 118.2°). It is worth noting that J1 exactly aligns with the (1/3, 1/3, l) lattice plane, which is where the magnetic peak appears in the neutron diffraction patterns. However, we are unclear how the charge disorder influences other magnetic interactions other than through M-O-M superexchange.
Display of two major superexchange interaction between Fe/Mn–Fe/Mn ions in the LnFeMnO4 structure in (a) [001] and (b) [100] directions. Green and yellow colors represent the same atom but in different layers.
Display of two major superexchange interaction between Fe/Mn–Fe/Mn ions in the LnFeMnO4 structure in (a) [001] and (b) [100] directions. Green and yellow colors represent the same atom but in different layers.
Oxidation of LnFeMnO4 into LnFeMnO4.5 causes the decrease of the Tf from around 60–30 K, implying a weakening of the magnetic interactions between moments. We note that there is a a-lattice parameter expansion upon oxygen uptake of LnFeMnO4 into LnFeMnO4.5. The weakening of the superexchange interaction along the ab-plane may contribute to the decrease of Tf upon oxidation. It is also evident that oxygen uptake greatly suppresses the short-range magnetic ordering. Figures 11(d)–11(f) present the Warren line shape fitting on the magnetic peak (located at Q ∼ 1.2 Å−1) of YFeMnO4.5, LuFeMnO4.5, and YbFeMnO4.5. The fitted magnetic correlation lengths are displayed in Table III. The magnetic correlation lengths of LnFeMnO4.5 (Ln = Y, Lu, and Yb) are 1–2 nm, much shorter than that of LnFeMnO4 (∼10 nm). We consider several factors that cause the suppression of short-range magnetic ordering upon oxygen uptake of LnFeMnO4. First, oxygen uptake leads to Mn2+(d5) oxidation into Mn3+(d4). The disordering of Mn3+(d4) and Fe3+(d5) cause the disordering of the different magnetic dipole moments. Second, the insertion of the extra O2− changes the existing magnetic superexchange interactions. The new O2− sites will undoubtedly connect some magnetic ions, which disturb the original ordered superexchange interactions. Finally, the displacement of Mn/Fe sites due to strong structural modulation in the oxidized phase can also disturb the originally ordered superexchange geometry.
Interestingly, in both reduced phase YbMnFeO4 and oxidized phase YbFeMnO4.5, their magnetic behavior does not differ from LnMnFeO4 and LnFeMnO4.5 (Ln = Y and Lu), although Yb3+ carries magnetic moment (4.5 μB in octahedra field),77 but Y3+ and Lu3+ do not. It seems the YbO octahedra layer does not influence the spin-frozen states or the nature of the short-range magnetic order. This “non-involvement” of Yb3+ might be the result of magnetic frustration of the YbO octahedra layer. The severe displacement of Yb3+ in the structure and triangular arrangement of Yb3+ leads to strong magnetic frustration. Indeed, the isostructural YbZnGaO4 and YbMgGaO4 (where only Yb3+ carries moment) do not display any magnetic transition even at 2 K due to the magnetic frustration in the YbO octahedra layer.12,78
V. CONCLUSION
In summary, like LnFe2O4, LnFeMnO4 displays an astonishing oxygen uptake behavior at low temperatures (∼200 °C). Despite the similarity of the average crystal structure of fully oxidized phases, LnFe2O4 and LnFeMnO4, the anisotropic lattice expansion behavior and structural modulations show significant variations. The change of the electron configuration of the transition metal upon oxidation (d6 to d5 for LnFe2O4, d5 to d6 for LnFeMnO4) seems to play an important role in those differences. We also found that the oxidation of LnFeMnO4 greatly suppresses its magnetic ordering. Our studies expand the LnM2+M3+O4 family to oxygen storage materials. The modifiable nature of LnM2+M3+O4 shows its potential to be tuned for many applications involving oxygen carriers such as chemical looping, catalytic oxidation/reduction, and oxygen ion conductors. The dramatic change of the physical property after the oxidation of LnFeMnO4 suggests it can be used for oxygen gas sensing as well.
SUPPLEMENTARY MATERIAL
See the supplementary material for additional characterizations of LnFe2+Fe3+O4 and LnFe2+Fe3+O4.5 (where Ln = Y, Lu, and Yb) including refinement of average structure with Synchrotron x-ray and Neutron diffraction, TGA, Pair distribution function, Raman, TEM and magnetic susceptibility measurement. Crystal structure Refinement parameters are included. Crystal structure files (cif files) are provided.
ACKNOWLEDGMENTS
We acknowledge the NIST Cooperative Agreement Nos. 70NANB20H139 and 70NANB17H301 for support. Use of the Advanced Photon Source at Argonne National Laboratory was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences, under Contract No. DE-AC02-06CH11357. This research used resources at Spallation Neutron Source, a DOE Office of Science User Facility operated by the Oak Ridge National Laboratory. We thank M. Kirkham at POWGEN, SNS, and ORNL for her help with the powder neutron diffraction experiment.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Tianyu Li: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Investigation (lead); Methodology (lead); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (equal). Sz-Chian Liou: Formal analysis (supporting); Investigation (supporting); Writing – review & editing (supporting). Stephanie J. Hong: Writing – review & editing (supporting). Qiang Zhang: Investigation (supporting); Writing – review & editing (supporting). H. Cein Mandujano: Investigation (supporting); Writing – review & editing (supporting). Efrain E. Rodriguez: Conceptualization (supporting); Funding acquisition (lead); Methodology (supporting); Resources (lead); Supervision (lead); Validation (lead); Writing – review & editing (lead).
DATA AVAILABILITY
The data that support the findings of this study are available from the corresponding author upon reasonable request.