Next Article in Journal
Model of Production System Evaluation with the Influence of FDM Machine Reliability and Process-Dependent Product Quality
Next Article in Special Issue
Mode Coupling at around M-Point in PZT
Previous Article in Journal
Bone Stress Evaluation with and without Cortical Bone Using Several Dental Restorative Materials Subjected to Impact Load: A Fully 3D Transient Finite-Element Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Crystal and Magnetic Structure Transitions in BiMnO3+δ Ceramics Driven by Cation Vacancies and Temperature

by
Dmitry V. Karpinsky
1,2,3,*,
Maxim V. Silibin
2,4,5,
Dmitry V. Zhaludkevich
1,2,
Siarhei I. Latushka
1,2,
Vadim V. Sikolenko
2,6,7,
Daniel M. Többens
8,
Denis Sheptyakov
9,
Vladimir A. Khomchenko
10 and
Alexei A. Belik
11
1
Laboratory of Technology and Physics of Crystals Growth, Scientific-Practical Materials Research Centre of NAS of Belarus, 220072 Minsk, Belarus
2
Institute for Advanced Materials and Technologies, National Research University of Electronic Technology “MIET”, 124498 Zelenograd, Russia
3
Department of Materials Science and Physico-Chemistry of Materials, South Ural State University, av. Lenina, 76, 454080 Chelyabinsk, Russia
4
Scientific-Manufacturing Complex “Technological Centre”, 124498 Zelenograd, Russia
5
Institute for Bionic Technologies and Engineering, I.M. Sechenov First Moscow State Medical University, 119991 Moscow, Russia
6
Frank Laboratory of Neutron Physics, Joint Institute for Nuclear Research, 141980 Dubna, Russia
7
Institute of Applied Geosciences, Karlsruhe Institute of Technology, 76131 Karlsruhe, Germany
8
Department Structure and Dynamics of Energy Materials, Helmholtz-Zentrum Berlin für Materialien und Energie, 14109 Berlin, Germany
9
Laboratory for Neutron Scattering and Imaging, Paul Scherrer Institut, 5232 Villigen, Switzerland
10
CFisUC, Department of Physics, University of Coimbra, 3004-516 Coimbra, Portugal
11
International Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS), Namiki 1-1, Tsukuba, Ibaraki 305-0044, Japan
*
Author to whom correspondence should be addressed.
Materials 2021, 14(19), 5805; https://doi.org/10.3390/ma14195805
Submission received: 31 August 2021 / Revised: 28 September 2021 / Accepted: 29 September 2021 / Published: 4 October 2021

Abstract

:
The crystal structure of BiMnO3+δ ceramics has been studied as a function of nominal oxygen excess and temperature using synchrotron and neutron powder diffraction, magnetometry and differential scanning calorimetry. Increase in oxygen excess leads to the structural transformations from the monoclinic structure (C2/c) to another monoclinic (P21/c), and then to the orthorhombic (Pnma) structure through the two-phase regions. The sequence of the structural transformations is accompanied by a modification of the orbital ordering followed by its disruption. Modification of the orbital order leads to a rearrangement of the magnetic structure of the compounds from the long-range ferromagnetic to a mixed magnetic state with antiferromagnetic clusters coexistent in a ferromagnetic matrix followed by a frustration of the long-range magnetic order. Temperature increase causes the structural transition to the nonpolar orthorhombic phase regardless of the structural state at room temperature; the orbital order is destroyed in compounds BiMnO3+δ (δ ≤ 0.14) at temperatures above 470 °C.

1. Introduction

In the last two decades, complex oxides of transition metals possessing both magnetic and electric dipole orders (so called multiferroics) have attracted increased interest of researchers [1,2]. The most studied single-phase multiferroic is bismuth ferrite, which has high transition temperatures to magnetic (650 K) and ferroelectric (1100 K) phases [1,3]. Bismuth manganite (BiMnO3) is also a well-known magnetoelectric material with perovskite structure with transition temperature to the magnetically ordered state of TC ~ 102 K [4]. The ferromagnetic (FM) order in BiMnO3 is caused by the presence of orbital ordering, which is destroyed at a temperature of ~475 K; while the type of structural distortion does not change, the crystal structure remains to be monoclinic; at temperatures above T ~ 770 K, the crystal structure changes to orthorhombic (space group (SG) Pnma) [5]. The magnetic structure of BiMnO3 is determined by positive exchange interactions formed between Mn3+ ions, while the type of the exchange interactions strongly depends on the mutual orientation of the 3 d orbitals of Mn ions as well as the geometry of the Mn-O-Mn chemical bonds [6].
It is known that the crystal structure and magnetic properties of BiMnO3-based compounds can be significantly modified via chemical doping [7,8]. Magnetic properties of BiMnO3-based compounds can also be controllably changed via doping on the A- and/or B- perovskite positions [9,10]. It is known that isovalent substitution in the A-position with rare-earth elements causes a stabilization of long-range antiferromagnetic (AFM) order accompanied by an increase in the magnetic transition temperature [11], whereas the nominal oxidation state of the Mn ions remains unchanged. Chemical substitution of Mn ions for other transition-metal ions also causes a disruption of ferromagnetic order [4,12], thus confirming a dominant role of the orbital order in magnetic properties of the BiMnO3-based compounds [13].
The chemical doping schemes assume an influence of the dopant ions on the crystal structure and electronic configuration of the B-site cations and their chemical bond character, thus hampering an analysis of the exchange interactions formed between manganese ions. On the other hand, a self-doping approach [14] can be used as an entirely internal stimulus which causes a modification of the crystal structure and thus changes the magnetic state of the compounds, thus facilitating an understanding of the structure-properties relationship. In the present study, we have investigated the crystal structure of ceramics BiMnO3+δ depending on the intrinsic (cation vacancies) and extrinsic (temperature) stimuli—the factors affecting magnetic state of the compounds. The obtained results have allowed us to clarify a correlation between the cation content, crystal structure and orbital ordering in the ceramic compounds BiMnO3+δ at different temperatures, thus showing that cation vacancies can be used to control orbital order and thus magnetic properties of the manganites.

2. Experimental

Ceramic compounds BiMnO3+δ with nominal oxygen excess δ = 0.02, 0.08, 0.14 were prepared from simple oxides Bi2O3 (99.9999%, Rare Metallic Co. Ltd.), MnO2 (99.997%, Alfa Aesar), and Mn2O3 taken in accordance with the chemical formulas Bi0.993Mn0.993O3, Bi0.974Mn0.974O3 and Bi0.955Mn0.955O3 respectively. Mn2O3 was prepared from MnO2 by heating in air at 923 K in 5 h. The oxygen content and phase purity of MnO2 were confirmed by X-ray powder diffraction and thermogravimetric analysis (TGA) before its use. The synthesis was performed using a belt-type high-pressure apparatus at a pressure of 6 GPa and temperature of about 1600 K for 40 min in sealed Pt capsules. After synthesis, the pressure was slowly released, and the samples were quenched at room temperature. Chemical composition of the compounds and the cation ratio was confirmed by electron probe microanalysis and density analysis as described in a previous study [15] as well as precise structural data presented below. The crystal and magnetic structure of the compounds were analyzed using X-ray diffraction data obtained with a PanAlytical X’pert Pro diffractometer (Panalytical B.V., Almelo, Netherlands), synchrotron powder diffraction (SPD) data obtained at the KMC-2 beamline (BESSY-II electron storage ring [16]), and neutron diffraction measurements performed with the HRPT diffractometer (PSI) [17]. Temperature dependent SPD measurements were run under ambient atmosphere in the range 10° < 2ϴ < 100° and temperatures 290–970 K with a step of 0.014° at λ ~ 1.77 Å; neutron diffraction data were recorded in the temperature range 290–750 K (λ ~1.494 Å). The X-ray and neutron diffraction data were analyzed by the Rietveld method using the FullProf software [18,19]. Magnetization measurements were performed using Cryogen Free Measurement System (CFMS) (Cryogenic ltd, London, UK). Differential scanning calorimetry (DSC) measurements were carried out with a Phoenix 204 F1 setup (Netzsch GmbH, Selb, Germany) in nitrogen flow at a heating and cooling rate of 10 °C per min in the range 300–700 K (nitrogen is considered as a neutral purge gas having high heat conductivity in the mentioned temperature range).

3. Results and Discussion

Analysis of the laboratory and synchrotron X-ray diffraction patterns of the compounds BiMnO3+δ confirms the changes of crystal structures from the monoclinic structure (C2/c) to another monoclinic (P21/c) and then to the orthorhombic structure (Pnma) with nominal increase in the oxygen content which is consistent with available data [15]. The mentioned sequence of the structural transformations is accompanied by a gradual decrease in the unit cell volume, while the crystal symmetry of the compounds changes nonmonotonously. Along with the information about the structural transformation, the diffraction measurements have also provided detailed information about the structural parameters, ionic coordinates and occupations. The occupation values calculated for the bismuth and manganese ions confirm nearly equal amount of the cation vacancies of about 3% and 5% for the compounds with nominal oxygen excess δ = 0.08 and 0.14 which correspond to their chemical formulas. Limited accuracy of the diffraction data could not exactly approve the chemical composition of the compound BiMnO3.02, while the reliability factors can be notably improved assuming a small deficit in the cation occupations during the refinement procedure. Further decrease in the cation/anion ratio specific to the BiMnO3.08 sample leads to a stabilization of the monoclinic structure having lower symmetry, viz. P21/c (SG #14) as compared to the stoichiometric compound—C2/c (SG #15)—while further decrease in the cation content leads to a stabilization of the nonpolar orthorhombic structure Pnma (SG #62).
It should be noted that the space group P21/c is the only common maximal subgroup for both C2/c and Pnma space groups. The symmetry of the crystal structure of BiMnO3.08 is lower than that attributed to BiMnO3.02 as determined by the extinction rule (Figure 1) [20]. Indeed, the diffraction pattern obtained for BiMnO3.08 reveals a splitting of the reflection located at 2ϴ ≈ 21.7 deg., and the extra reflection appears at 2ϴ ≈ 21.4 deg. (indexed as 1 ¯ 02 in the space group P21/c (Figure 1)), which points to the symmetry lowering. An appearance of the extra reflections associated with a lowering of the crystal symmetry is caused by an occupation of the new less-symmetrical positions of the manganese and oxygen ions over the lattice. The space group P21/c (without centering) allows a splitting of the reflection 110, thus leading to an appearance of the satellite 011 forbidden in C2/c; the extra reflection 1 ¯ 02 is associated with Mn ions in the Wyckoff position 2c (0; 0; ½) formed as a result of splitting of the Wyckoff position 4d (¼; ¼; ½) specific for C2/c into 2c (0;0; ½) and 2b (½; 0; 0) positions of the space group P21/c (Table 1). Further decrease in the amount of the cation content down to ~95.5% (BiMnO3.14) from the stoichiometric composition leads to a decrease in the unit cell volume resulting in a stabilization of orthorhombic symmetry (Pnma). The orthorhombic lattice is characterized by one structural position for Mn ions—4b (0; 0; ½) (Table 1) and two different positions occupied by oxygen ions—4c and 8d.
Decrease in the cation content leads to a reduction of magnetization of the compounds as declared in the previous works [15,21] and confirmed by our neutron diffraction and magnetization data (Figure 2). The changes in the magnetic structure are mainly caused by two factors—modification in the orbital ordering of the manganese ions and a dilution of the magnetic sublattice by the cation vacancies. The first factor is determined by the geometry of the bond lengths Mn-O and angles Mn-O-Mn, the second factor is associated with a character of the vacancies distribution over the A- and B- sublattices of the perovskite structure. Both mentioned factors have different impact on the magnetic structure of the compounds.
A decrease in the cation content even up to ~ 5% (in case of the compound BiMnO3.14) should not necessarily lead to a suppression of the long-range magnetic order observed experimentally [15,21] (Figure 2), e.g., other manganites as LaMnO3 and LuMnO3, keep stable long-range magnetic structure up to 6–10% of the Mn vacancies [14,22,23,24].
The orbital ordering and the magnetic state of the compounds BiMnO3+δ are strongly dependent on the spatial orientation of the chemical bond lengths and thus the type of the exchange interaction between Mn ions which can be of negative or positive character. Careful analysis of the diffraction data has allowed us to clarify that in the compound BiMnO3.02, the manganese vacancies are mainly located in the Wyckoff position 4d (¼; ¼; ½) occupied by Mn(2) ions, while the orbital ordering remains the same as it was observed in the stoichiometric compound [13]. Thus, in the compound BiMnO3.02 with crystal structure characterized by two independent positions of Mn ions, viz. 4d (¼; ¼; ½) and 4e (0; y; ¾), the exchange interactions Mn(1)-O(1)-Mn(2) and Mn(1)-O(3)-Mn(2) have positive character while that between Mn(1)-O(2)-Mn(2) is of negative character (Table 1, Figure 3). The positive character of the exchange interactions Mn(1)-O(1, 3)-Mn(2) is determined by the Goodenough–Kanamori rules [25,26], as the mentioned couplings are formed by the half-filled dz2 orbital of Mn(1) ions and the empty dx2-y2 orbital of Mn(2) ions; in turn, negative character of the exchange interactions Mn(1)-O(2)-Mn(2) is associated with a coupling between two empty dx2-y2 orbitals.
Lowering of the structural symmetry observed for the compound BiMnO3.08 is caused by an ordering of the cation vacancies. It results in a presence of three different structural positions for the Mn ions. Based on the refined structural data, we conclude that in the compound BiMnO3.08, the manganese vacancies are mainly located in the Mn(2) position 2b and partly in the Mn(1) position 4e. It should be noted that the space groups C2/c and P21/c follow the group–subgroup relation and the position 4d in the SG C2/c is split into the positions 2b and 2c in the SG P21/c. Modification of the crystal structure occurring in the BiMnO3.08 compound as compared to BiMnO3.02 results in a modification of the geometry of the orbital ordering. The orbital ordering specific to the BiMnO3.08 compound is characterized by four positive and two negative exchange interactions (Figure 3b, Mn(1) ion is located in the center). While in the BiMnO3.02 compound (Figure 3a) four positive exchange interactions form two pathways with long-range ordered ferromagnetic coupling, in the BiMnO3.08 compound, there is only one pathway associated with long-range ferromagnetic order; the next one has frustrated character where positive and negative interactions are alternated and the last pathway is characterized by negative interactions thus fostering a long-range antiferromagnetic order (Table 1, Figure 3b).
In the BiMnO3.14 compound, the orbital ordering is strongly suppressed as confirmed by the structural analysis, viz. lesser difference between the Mn-O bond lengths (Figure 3c, Table 1), wherein the manganese vacancies drastically frustrate positive superexchange interactions thus leading to a further decrease in the magnetization. It should also be noted that the charge neutrality in the nonstoichiometric compounds BiMnO3+δ is achieved through the formation of Mn4+ ions [14]. To keep the electric neutrality in the BiMnO3.14 compound, the concentration of Mn4+ ions should be about 28%; a similar value is derived from the BVS (bond valence sum) parameters calculated during the structure refinement procedure. The estimated amount of the Mn ions in the 4+ oxidation state should dramatically affect the magnetic structure through a suppression of the orbital order in the compound as well as providing strongly positive exchange interactions Mn4+-O-Mn3+. Based on the available data, one can conclude that the magnetic properties of the BiMnO3.14 compound are mainly determined by partly suppressed orbital ordering as positive exchange interactions Mn4+-O-Mn3+ have only a local character and do not contribute to the formation of long-range ferromagnetic order [27].
Based on the obtained structural data, we conclude that the cation (manganese) vacancies and the character of their distribution over the lattice provide a change in the geometry of the orbital arrangement, thus causing a suppression of the ferromagnetic properties. One should also take into account a formation of Mn4+ ions that ensure the charge neutrality of the nonstoichiometric compounds, while even in the sample with the maximal amount of the cation vacancies (BiMnO3.14), the Mn4+ ions did not notably provide a revival of the ferromagnetic order.

Temperature Dependent Structural Investigations

Analysis of the temperature-dependent synchrotron X-ray and neutron diffraction data has allowed us to clarify an evolution of the structural state and to evaluate the modification of the orbital ordering in the compounds having different structures at room temperature. Temperature-dependent structural data permitted an estimation of the impact of the cation vacancies on the temperature stability of the different structural states and orbital ordering which determines the magnetic structure and properties.
In case of the BiMnO3.02 compound, a temperature increase leads to a gradual reduction in the structural distortion as confirmed by a decrease in the intensity of the reflections specific to the monoclinic structure, e.g., 110, −112, 112, 312 (Figure 1). In particular, the unit cell parameters gradually increase with temperature showing ~1% increment in the unit cell volume over the range 320–670 K associated with monotonous increase of the a- and b- parameters (Figure 4), whereas the c-parameter slowly decreases, thus leading to more symmetric structure. Temperature increase above 670 K leads to a drastic modification in the unit cell parameters associated with a structural transition to the nonpolar orthorhombic phase with Pnma symmetry. The temperature range of the orthorhombic phase stability is quite narrow (770–820 K); further temperature increase leads to a chemical decomposition of the sample as confirmed by the diffraction data and the results of the DSC measurements. It should be noted that the structural transition into the orthorhombic phase is accompanied by a decrease in the primitive cell volume of about 0.7% pointing at more dense packing of the oxygen octahedra in the lattice. Analysis of the chemical bond lengths allowed us to conclude that the orbital ordering found in the BiMnO3.02 compound is destroyed at temperatures above 500 K, the disruption occurring in the narrow temperature range of about 50 K. The data obtained show that the orbital ordering collapses simultaneously along three pathways in contrast to the concentration induced modification of the orbital ordering.
In the BiMnO3.08 compound, the structural phase transition to the nonpolar orthorhombic phase occurs at lower temperature as compared to the BiMnO3.02 compound. At a temperature of ~470 K, the structural state of the compound BiMnO3.08 is characterized by two coexisting phases—monoclinic P21/c and the nonpolar orthorhombic with the space group Pnma. At temperatures above 500 K, the crystal structure of the compound BiMnO3.08 is single phase orthorhombic up to 670 K. Above this temperature, the compound begins to decompose. Whereas the difference in the cell volumes calculated for the monoclinic and orthorhombic phases (0.4%) is notably smaller as compared to the data obtained for the BiMnO3.02 compound, the diffraction measurements confirm the absence of any intermediate structural phase in this phase transition.
The structural parameters calculated for the BiMnO3.08 compound within the monoclinic phase demonstrate the tendency previously observed for the BiMnO3.02 compound with the temperature increase, assuming a collapse of the orbital ordering at a temperature of ~470 K. The compound BiMnO3.14 maintains the orthorhombic structure up to ~820 K, where the chemical decomposition starts. An evolution of the calculated lattice parameters points to a nonmonotonous change of the structural distortions. The derived chemical bond lengths indicate a complete disruption of the orbital order at temperatures above 470 K, which might also be indicated by a modification of the lattice parameters about this temperature (Figure 4).

4. Conclusions

Synchrotron X-ray and neutron diffraction measurements allowed clarification of the details of the evolution of the crystal and magnetic structures of the BiMnO3+δ compounds as a function of temperature and the cation content. One can conclude that the structural stability of the monoclinic phases reduces with an increase in the number of the cation vacancies. The temperature of the structural transitions to the nonpolar orthorhombic phase reduces with an increase in the value of nominal oxygen excess; the orbital ordering specific to the compounds is destroyed at temperatures about 500 K. Increase in the number of the cation vacancies causes a rearrangement of the orbital order (which is accompanied by a change of the dominant superexchange interactions) followed by a disruption of the orbital ordering and removal of the long-range spin order.

Author Contributions

D.V.K.: Supervision, Conceptualization, Investigation, Writing—Original Draft. M.V.S., D.V.Z., S.I.L., V.V.S., D.M.T., D.S.: Investigation, Data analysis. V.A.K., D.S., A.A.B.: Validation, Conceptualization and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by RSF (project 21-19-00386) and the APC was funded by RSF (project 21-19-00386).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by the RSF (project 21-19-00386). The authors acknowledge HZB/BESSY for the allocation of synchrotron diffraction measurements and HZB/BESSY staff for the assistance with diffraction experiments. The authors thank P. Hüttenlloch and O. Ivanshina for DSC measurements and I. Radulov for magnetization measurements. V.A.K. acknowledges the support provided by national funds from FCT—Fundaçaão para a Ciência e a Tecnologia, I.P., within the projects UIDB/04564/2020 and UIDP/04564/2020. This work is partly based on experiments performed at the Swiss spallation neutron source SINQ, Paul Scherrer Institute, Villigen, Switzerland.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Spaldin, N.A.; Ramesh, R. Advances in magnetoelectric multiferroics. Nat. Mater. 2019, 18, 203–212. [Google Scholar] [CrossRef] [PubMed]
  2. Fernández-Posada, C.M.; Castro, A.; Kiat, J.-M.; Porcher, F.; Peña, O.; Algueró, M.; Amorín, H. A novel perovskite oxide chemically designed to show multiferroic phase boundary with room-temperature magnetoelectricity. Nat. Commun. 2016, 7, 12772. [Google Scholar] [CrossRef]
  3. Ederer, C.; Spaldin, N.A. Weak ferromagnetism and magnetoelectric coupling in bismuth ferrite. Phys. Rev. B 2005, 71, 060401. [Google Scholar] [CrossRef] [Green Version]
  4. Belik, A.A. Local distortions in multiferroic BiMnO3 as a function of doping. Sci. Technol. Adv. Mater. 2011, 12, 044610. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kimura, T.; Kawamoto, S.; Yamada, I.; Azuma, M.; Takano, M.; Tokura, Y. Magnetocapacitance effect in multiferroic BiMnO3. Phys. Rev. B 2003, 67, 180401. [Google Scholar] [CrossRef]
  6. Coey, J.M.D.; Viret, M.; von Molnár, S. Mixed-valence manganites. Adv. Phys. 1999, 48, 167–293. [Google Scholar] [CrossRef]
  7. Karoblis, D.; Diliautas, R.; Raudonyte-Svirbutaviciene, E.; Mazeika, K.; Baltrunas, D.; Beganskiene, A.; Zarkov, A.; Kareiva, A. The Synthesis and Characterization of Sol-Gel-Derived SrTiO3-BiMnO3 Solid Solutions. Crystals 2020, 10, 1125. [Google Scholar] [CrossRef]
  8. Revathi, B.; Balakrishnan, L.; Chandar, N.K. Structural, morphological, optical, dielectric and magnetic field sensing characteristics of Bi1-xKxMnO3 and BiMn1-yCoyO3 nanopowders: A comparative study. Mater. Lett. 2019, 256, 126655. [Google Scholar] [CrossRef]
  9. Belik, A.A. Structural, magnetic, and dielectric properties of solid solutions between BiMnO3 and YMnO3. J. Solid State Chem. 2017, 246, 8–15. [Google Scholar] [CrossRef]
  10. Thakur, S.; Singh, K.; Pandey, O.P. Sr doped BiMO3 (M = Mn, Fe, Y) perovskites: Structure correlated thermal and electrical properties. Mater. Chem. Phys. 2017, 187, 96–103. [Google Scholar] [CrossRef]
  11. Kumar, P.; Dayal, V. Critical behavior and non-universal low-field magnetic scaling in La1-xBixMnO3 (x=0.4 & 0.6) perovskite manganite oxide. AIP Conf. Proc. 2015, 1665, 030014. [Google Scholar]
  12. Azuma, M.; Kanda, H.; Belik, A.A.; Shimakawa, Y.; Takano, M. Magnetic and structural properties of BiFe1−xMnxO3. J. Magn. Magn. Mater. 2007, 310, 1177–1179. [Google Scholar] [CrossRef]
  13. Moreira dos Santos, A.; Cheetham, A.K.; Atou, T.; Syono, Y.; Yamaguchi, Y.; Ohoyama, K.; Chiba, H.; Rao, C.N.R. Orbital ordering as the determinant for ferromagnetism in biferroic BiMnO3. Phys. Rev. B 2002, 66, 064425. [Google Scholar] [CrossRef]
  14. Figueiras, F.G.; Karpinsky, D.; Tavares, P.B.; Gonçalves, J.N.; Yañez-Vilar, S.; Moreira Dos Santos, A.F.; Franz, A.; Tovar, M.; Agostinho Moreira, J.; Amaral, V.S. Novel multiferroic state and ME enhancement by breaking the AFM frustration in LuMn1−xO3. Phys. Chem. Chem. Phys. 2017, 19, 1335–1341. [Google Scholar] [CrossRef] [PubMed]
  15. Belik, A.A.; Kodama, K.; Igawa, N.; Shamoto, S.-i.; Kosuda, K.; Takayama-Muromachi, E. Crystal and Magnetic Structures and Properties of BiMnO3+δ. J. Am. Chem. Soc. 2010, 132, 8137–8144. [Google Scholar] [CrossRef] [PubMed]
  16. Többens, D.M.; Zander, S. KMC-2: An X-ray beamline with dedicated diffraction and XAS endstations at BESSY II. JLSRF 2016, 2, A49. [Google Scholar] [CrossRef] [Green Version]
  17. Fischer, P.; Frey, G.; Koch, M.; Könnecke, M.; Pomjakushin, V.; Schefer, J.; Thut, R.; Schlumpf, N.; Bürge, R.; Greuter, U.; et al. High-resolution powder diffractometer HRPT for thermal neutrons at SINQ. Physica B 2000, 276–278, 146–147. [Google Scholar] [CrossRef]
  18. Rodríguez-Carvajal, J. Recent advances in magnetic structure determination by neutron powder diffraction. Physica B 1993, 192, 55–69. [Google Scholar] [CrossRef]
  19. Rietveld, H.M. A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr. 1969, 2, 65–71. [Google Scholar] [CrossRef]
  20. Cowley, J.M.; Moodie, A.F.; Miyake, S.; Takagi, S.; Fujimoto, F. The extinction rule for reflexions in symmetrical electron-diffraction spot patterns. Acta Crystallogr. Sect. A 1961, 14, 87–88. [Google Scholar] [CrossRef]
  21. Sundaresan, A.; Mangalam, R.V.K.; Iyo, A.; Tanaka, Y.; Rao, C.N.R. Crucial role of oxygen stoichiometry in determining the structure and properties of BiMnO3. J. Mater. Chem. C 2008, 18, 2191–2193. [Google Scholar] [CrossRef]
  22. Figueiras, F.G.; Karpinsky, D.; Tavares, P.B.; Das, S.; Leitão, J.V.; Brück, E.H.; Moreira, J.A.; Amaral, V.S. Breaking the geometric magnetic frustration in controlled off-stoichiometric LuMn1+zO3+δ compounds. Phys. Chem. Chem. Phys. 2016, 18, 13519–13523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Markovich, V.; Rozenberg, E.; Gorodetsky, G.; Jung, G.; Fita, I.; Puzniak, R.; Wisniewski, A.; Martin, C.; Hebert, S.; Raveau, B. Vacancies at Mn-sites in LaMn1-xO3 manganites: Interplay between ferromagnetic interactions and hydrostatic pressure. J. Appl. Phys. 2004, 95, 7112–7114. [Google Scholar] [CrossRef]
  24. Shames, A.I.; Auslender, M.; Rozenberg, E.; Gorodetsky, G.; Hébert, S.; Martin, C. Ferromagnetic ordering in LaMn1−xO3 manganites: EMR probing. J. Magn. Magn. Mater. 2007, 316, e640–e643. [Google Scholar]
  25. Goodenough, J.B. An interpretation of the magnetic properties of the perovskite-type mixed crystals La1-xSrxCoO3−λ. J. Phys. Chem. Solids 1958, 6, 287–297. [Google Scholar] [CrossRef]
  26. Kanamori, J. Superexchange interaction and symmetry properties of electron orbitals. J. Phys. Chem. Solids 1959, 10, 87–98. [Google Scholar] [CrossRef]
  27. Paraskevopoulos, M.; Mayr, F.; Hemberger, J.; Loidl, A.; Heichele, R.; Maurer, D.; Müller, V.; Mukhin, A.A.; Balbashov, A.M. Magnetic properties and the phase diagram of La1-xSrxMnO3 for x <=0.2. J. Phys. Condens. Matter. 2000, 12, 3993–4011. [Google Scholar] [CrossRef]
Figure 1. SPD pattern of the compound BiMnO3.02 refined using the space group C2/c. Inset shows the reflections specific to the compounds BiMnO3+δ (δ = 0.02, 0.08, 0.14). Images on the right show the temperature evolution of the selected SPD reflections of the compound BiMnO3.02.
Figure 1. SPD pattern of the compound BiMnO3.02 refined using the space group C2/c. Inset shows the reflections specific to the compounds BiMnO3+δ (δ = 0.02, 0.08, 0.14). Images on the right show the temperature evolution of the selected SPD reflections of the compound BiMnO3.02.
Materials 14 05805 g001
Figure 2. Refined NPD pattern of the compound BiMnO3.08 using the space group P21/c. Inset shows M(T) magnetization curves for the compounds BiMnO3+δ (δ = 0.02, 0.08, 0.14) measured in a field-cooled mode (H~100 Oe). Images on the right show the temperature evolution of selected NPD reflections of the compound BiMnO3.08.
Figure 2. Refined NPD pattern of the compound BiMnO3.08 using the space group P21/c. Inset shows M(T) magnetization curves for the compounds BiMnO3+δ (δ = 0.02, 0.08, 0.14) measured in a field-cooled mode (H~100 Oe). Images on the right show the temperature evolution of selected NPD reflections of the compound BiMnO3.08.
Materials 14 05805 g002
Figure 3. Schemes of the oxygen octahedra of the compounds BiMnO3+δ; (a) denotes the scheme specific to the compound with δ = 0.02; (b) —δ = 0.08; (c) —δ = 0.14); the dz2 orbitals are highlighted in red and yellow colors (yellow color of the dz2 orbitals in case of BiMnO3.14 compound denotes partial suppression of the orbital ordering which leads to an uncertainty in the type of the exchange interactions Mn-O(2)’-Mn, marked by “?”).
Figure 3. Schemes of the oxygen octahedra of the compounds BiMnO3+δ; (a) denotes the scheme specific to the compound with δ = 0.02; (b) —δ = 0.08; (c) —δ = 0.14); the dz2 orbitals are highlighted in red and yellow colors (yellow color of the dz2 orbitals in case of BiMnO3.14 compound denotes partial suppression of the orbital ordering which leads to an uncertainty in the type of the exchange interactions Mn-O(2)’-Mn, marked by “?”).
Materials 14 05805 g003
Figure 4. The unit cell parameters derived from the SPD and NPD data of the compounds BiMnO3+δ with δ = 0.02, 0.08, 0.14 - (ac), respectively. The DSC curves obtained for the respective compounds are presented at the bottom. The unit cell parameters are presented in normalized form, viz. ap = a/√6, bp = b/√2, cp = c/√6·(for C2/c and P21/c phases) and ap = a/√2, bp = b/4, cp = c/(2√2)·(for Pnma phase).
Figure 4. The unit cell parameters derived from the SPD and NPD data of the compounds BiMnO3+δ with δ = 0.02, 0.08, 0.14 - (ac), respectively. The DSC curves obtained for the respective compounds are presented at the bottom. The unit cell parameters are presented in normalized form, viz. ap = a/√6, bp = b/√2, cp = c/√6·(for C2/c and P21/c phases) and ap = a/√2, bp = b/4, cp = c/(2√2)·(for Pnma phase).
Materials 14 05805 g004
Table 1. Selected structural parameters of the BiMnO3+δ compounds refined based on the synchrotron and neutron diffraction data.
Table 1. Selected structural parameters of the BiMnO3+δ compounds refined based on the synchrotron and neutron diffraction data.
SampleSGMn-O-Mn/deg.Magn.
Order
Mn-O/Å
BiMnO3.02C2/cMn(1)-O(1)-Mn(2) (148.05)FMMn(1)-O(1): 1.955(7); Mn(1)-O(2): 2.037(5);
Mn(1)-O(3): 2.076(4); Mn(2)-O(1): 2.125(8);
Mn(2)-O(2): 2.081(5); Mn(2)-O(3): 1.899(7);
Mn(1)-O(2)-Mn(2) (158.96)AFM
Mn(1)-O(3)-Mn(2) (152.32)FM
BiMnO3.08P21/cMn(1)-O(1)-Mn(2) (144.08)AFMMn(1)-O(1): 2.045(7); Mn(1)-O(2): 1.983(5);
Mn(1)-O(3): 1.922(6); Mn(1)-O(4): 2.098(6);
Mn(1)-O(5): 1.993(6); Mn(1)-O(6): 1.944(5);
Mn(2)-O(1): 2.008(8); Mn(2)-O(3): 2.121(2);
Mn(2)-O(5): 1.900(4); Mn(3)-O(2): 2.091(6);
Mn(3)-O(4): 2.029(2); Mn(3)-O(6): 1.965(3);
Mn(1)-O(2)-Mn(3) (160.10)FM
Mn(1)-O(3)-Mn(2) (154.28)FM
Mn(1)-O(4)-Mn(3) (157.91)FM
Mn(1)-O(5)-Mn(2) (153.26)FM
Mn(1)-O(6)-Mn(3) (163.87)AFM
BiMnO3.14PnmaMn-O(1)-Mn (155.87)AFMMn-O(1): 1.995(3);
Mn-O(2): 1.935(2);
Mn-O(2): 2.027(7).
Mn-O(2)-Mn (172.38)FM
Coordinates of the Mn ions at 300 K calculated for BiMnO3.02 (SG #15, C2/c): Mn(1)—4e (0, 0.216, 0.75); Mn(2)—4d (0.25, 0.25, 0.5); for BiMnO3.08 (SG #14, P21/c): Mn(1)—4e (0.271, 0, 0.22); Mn(2)—2b (0.5, 0, 0); Mn(3)—2c (0, 0, 0.5); for BiMnO3.14 (SG #62, Pnma): Mn—4b (0; 0; 0.5). The type of exchange interaction (FM or AFM) is noted for each ternary Mn-O-Mn.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Karpinsky, D.V.; Silibin, M.V.; Zhaludkevich, D.V.; Latushka, S.I.; Sikolenko, V.V.; Többens, D.M.; Sheptyakov, D.; Khomchenko, V.A.; Belik, A.A. Crystal and Magnetic Structure Transitions in BiMnO3+δ Ceramics Driven by Cation Vacancies and Temperature. Materials 2021, 14, 5805. https://doi.org/10.3390/ma14195805

AMA Style

Karpinsky DV, Silibin MV, Zhaludkevich DV, Latushka SI, Sikolenko VV, Többens DM, Sheptyakov D, Khomchenko VA, Belik AA. Crystal and Magnetic Structure Transitions in BiMnO3+δ Ceramics Driven by Cation Vacancies and Temperature. Materials. 2021; 14(19):5805. https://doi.org/10.3390/ma14195805

Chicago/Turabian Style

Karpinsky, Dmitry V., Maxim V. Silibin, Dmitry V. Zhaludkevich, Siarhei I. Latushka, Vadim V. Sikolenko, Daniel M. Többens, Denis Sheptyakov, Vladimir A. Khomchenko, and Alexei A. Belik. 2021. "Crystal and Magnetic Structure Transitions in BiMnO3+δ Ceramics Driven by Cation Vacancies and Temperature" Materials 14, no. 19: 5805. https://doi.org/10.3390/ma14195805

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