Next Article in Journal
Perspectives on Molecular Materials—A Tribute to Professor Peter Day
Next Article in Special Issue
Simple Realistic Model of Spin Reorientation in 4f-3d Compounds
Previous Article in Journal
Counterintuitive Single-Molecule Magnet Behaviour in Two Polymorphs of One-Dimensional Compounds Involving Chiral BINOL-Derived Bisphosphate Ligands
Previous Article in Special Issue
A Quantum-Mechanical Study of Antiphase Boundaries in Ferromagnetic B2-Phase Fe2CoAl Alloy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnetic Behaviour of Perovskite Compositions Derived from BiFeO3

by
Andrei N. Salak
1,*,
João Pedro V. Cardoso
1,
Joaquim M. Vieira
1,
Vladimir V. Shvartsman
2,
Dmitry D. Khalyavin
3,
Elena L. Fertman
4,
Alexey V. Fedorchenko
4,
Anatoli V. Pushkarev
5,
Yury V. Radyush
5,
Nikolai M. Olekhnovich
5,
Róbert Tarasenko
6,
Alexander Feher
6 and
Erik Čižmár
6,*
1
Department of Materials and Ceramics Engineering/CICECO-Aveiro Institute of Materials, University of Aveiro, 3810-193 Aveiro, Portugal
2
Institute for Materials Science and CENIDE-Center for Nanointegration Duisburg-Essen, University of Duisburg-Essen, 45141 Essen, Germany
3
ISIS Facility, Rutherford Appleton Laboratory, Chilton, Didcot, Oxfordshire OX11 0QX, UK
4
B. Verkin Institute for Low Temperature Physics and Engineering, National Academy of Sciences of Ukraine, Nauky 47, 61103 Kharkiv, Ukraine
5
Scientific-Practical Materials Research Centre, National Academy of Sciences of Belarus, P. Brovka 19, 220072 Minsk, Belarus
6
Institute of Physics, Faculty of Science, Pavol Jozef Šafárik University, Park Angelinum 9, 041 54 Košice, Slovakia
*
Authors to whom correspondence should be addressed.
Magnetochemistry 2021, 7(11), 151; https://doi.org/10.3390/magnetochemistry7110151
Submission received: 5 October 2021 / Revised: 10 November 2021 / Accepted: 14 November 2021 / Published: 16 November 2021
(This article belongs to the Special Issue Ferromagnetism)

Abstract

:
The phase content and sequence, the crystal structure, and the magnetic properties of perovskite solid solutions of the (1−y)BiFeO3yBiZn0.5Ti0.5O3 series (0.05 ≤ y ≤ 0.90) synthesized under high pressure have been studied. Two perovskite phases, namely the rhombohedral R3c and the tetragonal P4mm, which correspond to the structural types of the end members, BiFeO3 and BiZn0.5Ti0.5O3, respectively, were revealed in the as-synthesized samples. The rhombohedral and the tetragonal phases were found to coexist in the compositional range of 0.30 ≤ y ≤ 0.90. Magnetic properties of the BiFe1−y[Zn0.5Ti0.5]yO3 ceramics with y < 0.30 were measured as a function of temperature. The obtained compositional variations of the normalized unit-cell volume and the Néel temperature of the BiFe1−y[Zn0.5Ti0.5]yO3 perovskites in the range of their rhombohedral phase were compared with the respective dependences for the BiFe1−yB3+yO3 perovskites (where B3+ = Ga, Co, Mn, Cr, and Sc). The role of the high-pressure synthesis in the formation of the antiferromagnetic states different from the modulated cycloidal one characteristic of the parent BiFeO3 is discussed.

1. Introduction

Bismuth ferrite is one of few type-I multiferroics, namely those solids in which the coexisting (anti)ferroelectric order and the (anti)ferromagnetic order are caused by independent mechanisms [1]. Ferroelectricity in BiFeO3 is induced by the electronic instability of the lone-pair Bi3+ cations, while the antiferromagnetism in this material with the Néel temperature, TN, as high as 643 K results from the superexchange interactions between Fe3+ cations [2]. Besides, the distorted crystal structure of this perovskite, involving large polar atomic displacements and octahedral tilting, gives rise to competing antisymmetric Dzyaloshinskii-Moriya (DM) interactions [3]. The part of antisymmetric exchange associated with the polar distortions favours a spatially modulated ground state in the form of a long-period incommensurate cycloid. Contrary, the antisymmetric exchange imposed by the octahedral tilting requires a non-modulated canted weak ferromagnetic (FM) state. In the undoped BiFeO3, the former contribution wins promoting the long-period modulated spin ordering that averages the net magnetization to zero [4]. This modulation, however, can be suppressed via chemical modification or thin-film strain engineering, resulting in a state where both the ferroelectric polarization and spontaneous magnetization coexist.
The bismuth site substitutions in BiFeO3 are the most studied. The compositional dependent structural transitions and variations of the magnetic ordering of the Bi1−xA3+xFeO3 perovskites series (0 ≤ x < 1) have been revealed [5,6,7,8]. In particular, the minimum substitution rates for rare earth cations sufficient to destroy the cycloidal modulation were estimated [9]. Besides, the correlations between the size of the substituting cation and the transition temperature were found [5,10].
Chemical modifications in the iron site of BiFeO3 appear to be the direct approach to tune the magnetic behaviour of this material. However, using the conventional synthesis routes, it is possible to achieve the substitution rates of a few at.% only. Most of the reported single-phase BiFe1−yB3+yO3 perovskite compositions with x > 0.10 were prepared using the high-pressure synthesis technique [11,12,13,14,15,16,17,18,19,20,21]. The only exception appears to be the BiFe1−yMnyO3 system, in which up to about 30 at.% of the iron-to-manganese substitution is possible via the conventional ceramic route [22]. It should be noticed here that the B-site substituted compositions derived from bismuth ferrite using the conventional route belong to the same space group, R3c, as that of the parent BiFeO3, while high-pressure synthesis can result in the formation of other structural phases. For instance, the same perovskite composition, BiFe0.75Mn0.25O3, prepared by solid-state synthesis at ambient pressure or via high-pressure synthesis is rhombohedral or orthorhombic, respectively [22]. The full-range range substitutions of Fe3+ with trivalent cations whose ionic radii are considerably smaller (gallium [16]) or considerably larger (scandium [19]) than the iron one were successfully performed using high-pressure synthesis. In high-pressure stabilized perovskite solid solutions of the BiFe1−yScyO3 system a series of structural transitions with increasing y was found. Moreover, it was revealed that annealing the as-prepared BiFe1−yScyO3 perovskites (y ≥ 0.3) results in irreversible transformations into new perovskite phases with interesting combinations of ferroic orders [23]. It was demonstrated that the observed effect is a manifestation of conversion polymorphism, which is a general phenomenon in the high-pressure stabilized oxygen-octahedral structural phases [23]. Structure, dielectric response, and magnetic behaviour of the as-prepared and the converted polymorphs of the (1-y)BiFeO3yBiScO3 perovskites have been considered in great detail [24,25,26,27]. Magnetic ordering was detected in the BiFe1−yScyO3 compositions with up to 60 at.% of scandium with a near-linear TN(y) dependence. In the 0.1 ≤ y < 0.3 range of this solid solution system, some peculiarities of the temperature-dependent magnetic moment below TN were observed and associated with possible transitions between three different antiferromagnetic (AFM) structures, namely those corresponding to collinear, canted, and cycloidal spin arrangements [26]. Similar temperature anomalies of the magnetic behaviour below TN were then revealed in the Fe-rich compositional range of the BiFe1−y[Zn0.5Ti0.5]yO3 perovskites phases prepared using high-pressure synthesis [28]. The (1-y)BiFeO3yBiZn0.5Ti0.5O3 series is of interest as the promising lead-free system in which the compositional range of coexistence of two polar phases (the morphotropic phase boundary/region, MPB) occurs.
Although a number of the BiFe1−yB3+yO3 perovskite series has already been prepared and characterized [11,12,13,14,15,16,17,18,19,20,21,28,29], to the best of our knowledge, the obtained structural and magnetic data have not been generalized in respect of the ionic size of the substituting cation. This is certainly worthy of consideration as the comparative studies of the variation of structural characteristics and transition temperatures in solid solutions and series the isomorphous substitutions of are known to be very convenient to understand some features and predict properties of new compositions [30].
In this paper, we considered the compositional behaviours of the crystal structure and the magnetic properties of the BiFe1−y[Zn0.5Ti0.5]yO3 perovskite phases and compared them with the respective dependences of BiFe1−yB3+yO3 perovskites (where B3+ = Ga, Co, Mn, Cr, and Sc) in the vicinity of parent bismuth ferrite. Among these, Cr3+, Mn3+, and Co3+ are magnetic cations of transition metals from the same 3d series to which iron belongs, while Ga3+, Sc3+ and [Zn0.5Ti0.5]3+ are non-magnetic. Besides, as compared with iron, Ga3+ is smaller, Sc3+ is considerably bigger, and [Zn0.5Ti0.5]3+ is slightly bigger than Fe3+ in octahedral coordination. Although no simple model based on microscopic magneto-structural correlations can be applied and predicting the behaviour of the doped perovskites requires extensive DFT calculations that include detailed information about the structural modifications [7], the compared cases appear to be various enough to conclude on possible correlations and trends.

2. Results

Analysis of the XRD data of the as-synthesized (unannealed) samples of the (1−y)BiFeO3yBiZn0.5Ti0.5O3 series has revealed no crystalline phase apart from the perovskite ones. It was found from the comparison of the XRD patterns of the compositions with increasing y that the samples with y < 0.30 are single-phase perovskites with the rhombohedral R3c structure. An increase of the Zn-Ti content results in the appearance and growth of new diffraction peaks (Figure 1). These peaks were associated with the tetragonal perovskite phase similar to that of the parent BiZn0.5Ti0.5O3 [31]. In the range of 0.30 ≤ y ≤ 0.90, the rhombohedral and the tetragonal phases coexist. These two were the only phases detected in the whole compositional range, and no other perovskite phase has been revealed. This is in contradiction with the results of Pan et al. [18]) who observed an intermediate monoclinic phase in BiFe1−y[Zn0.5Ti0.5]yO3 between y = 0.40 and 0.50. It should be noted, however, that Pan et al. studied the annealed samples while no thermal treatment was performed in this work. As mentioned in the Introduction, annealing of the high-pressure stabilized materials can lead to irreversible polymorph transformations [23].
The compositional range of coexistence of the rhombohedral and the tetragonal structural phases (MPB) in the BiFeO3–BiZn0.5Ti0.5O3 system is essentially broader than that observed in the BiMg0.5Ti0.5O3–BiZn0.5Ti0.5O3 solid solutions [32], in which the bismuth magnesium titanate is a structural analogue of PbZrO3 [33]. In the latter system, a coexistence of the perovskite phases was observed in the compositional range narrower than 5 at.%. A wide-range coexistence of the perovskite phases is very typical of the compositions derived from bismuth ferrite since the energy landscape of BiFeO3 is rather flat [34].
The crystal structure refinement was successful considering the two perovskite phases in the as-prepared the BiFe1−y[Zn0.5Ti0.5]yO3 samples, namely the rhombohedral R3c and the tetragonal P4mm, which correspond to the structural types of the end members, BiFeO3 and BiZn0.5Ti0.5O3, respectively.
The compositional variations of the primitive perovskite unit–cell parameters (ap, cp, and αp) and the normalized unit-cell volume (Vp = V/Z) are shown in Figure 2. The parameters were calculated from the refinement data using the relations for the basis vectors of the rhombohedral R3c structure and the parent cubic cell [28]. One can see no significant increment of any of the parameters with y over the whole range. The maximum relative variations were observed for the cp value (~0.6%, the P4mm phase) and the Vp value (~1.4%, R3c phase). As a result, the difference between the normalized unit-cell values of the phases is almost constant over their coexistence range (Figure 2b).
The most representative results of magnetic measurements of the BiFe1−y[Zn0.5Ti0.5]yO3 samples are shown in Figure 3 and Figure 4. It was earlier found [28] that the Néel temperature is more pronounced in magnetic data for the heat-treated samples. Therefore, the annealed samples were used for the estimation of the TN values. The temperature dependence of the magnetic moment measured in the field-cooled (FC) regime in an applied magnetic field of 500 Oe from the temperature high enough above the transition temperature down to 330 K is shown in Figure 3.
Two anomalies in the temperature-dependent magnetic moment considered as indications of magnetic transformations (assigned as Tm and Ta) were observed in the AFM state of the BiFe1−yScyO3 ceramics in the compositional range of 0.1 ≤ y < 0.25 [26]. The signature of similar behaviour was also observed in the BiFe1−y[Zn0.5Ti0.5]yO3 phase with y = 0.25 [28]. Therefore, the low-temperature magnetic moment of BiFe1−y[Zn0.5Ti0.5]yO3 with 0.1 ≤ y < 0.2 was measured in the temperature range of 5–400 K in a small applied field of 50 Oe. Similar behaviour was observed for as-synthesized and annealed samples. As can be seen from the data measured using as-synthesized samples in zero-field-cooled (ZFC) regime in Figure 4, for low y values, there is no clear signature of the transformations as mentioned above in contrast to BiFe1−yScyO3 [26]. Possible transformations of the magnetic structure shown by arrows in Figure 4 were revealed for y = 0.2 only from the derivative of the M/H(T) curve.
An interesting behaviour was observed in the magnetization loops. The shape of the magnetization loops of the as-synthesized BiFe1−y[Zn0.5Ti0.5]yO3 samples resembles those of BiFe1−yScyO3 [26], which can be described as a superposition of linear AFM and hysteretic FM contribution. The annealing leads to an increase in coercivity, remnant magnetization, and the total magnetization at the maximum applied field. Example data are depicted for y = 0.15 in the inset of Figure 5 with HC = 1.43 kOe and HC = 4.63 kOe for the as-synthesized and the annealed sample, respectively. The observed change in the shape of the magnetization loops after the annealing is of particular interest. The magnetization loops clearly indicate the presence of metamagnetic behaviour particularly pronounced in the compositions with y = 0.05 and y = 0.1. Apparently, the magnetic state of the studied samples is nonhomogeneous, and they consist of at least two phases. One of them is weak ferromagnet (i.e., canted antiferromagnet), and another is either a modulated cycloid, similar to the undoped BiFeO3, or collinear antiferromagnet (no spin canting) as in the ground state of polar BiFe0.7Sc0.3O3 [23]. The metamagnetic behaviour can be attributed to the latter phase, where the magnetic field switches the spin ordering from the modulated to collinear. The field-induced transition is reversible at room temperature resulting in the unusual shape of the magnetization loops. It has to be pointed out that such transition is also well-known in BiFeO3 [36,37], however, with the critical field significantly higher than in the present case. The phase fraction of the modulated/collinear metamagnetic phase decreases with y, and it vanishes in the compositions with y ≥ 0.2. These experimental observations can be interpreted as a composition-induced first-order phase transition with an extremely large phase coexisting region. In this scenario, the hysteretic region might also depend on temperature, resulting in a very complex composition-temperature-field phase diagram. On the other hand, the weak ferromagnet phase itself can exhibit metamagnetic re-orientation of the magnetic moments at low magnetic fields, as reported in the case of high-pressure synthesized BiFe0.75Mn0.25O3 [22].

3. Discussion

It follows from the available data on the bulk perovskite BiFe1−yB3+yO3 solid solutions that the BiFeO3-type rhombohedral phase remains in the compositions with up to about 30 at.% substitution rate regardless of the preparation method. In particular, the R3c range is y ≤ 0.3 for Ga [16], y ≤ 0.25 for Co [14], y ≤ 0.30 for Mn [12], y ≤ 0.25 for Sc [19], and y ≤ 0.30 for Zn0.5Ti0.5 (see Results). The only known exception is the iron-to-chromium substitution, at which BiFe0.50Cr0.50O3 is still rhombohedral [13,20]. In the cases when the perovskite BiFe1−yB3+yO3 phase is prepared via high-pressure synthesis, annealing may extend (by about 5 at.%) the compositional range of the rhombohedral structure [22,23].
Figure 6 shows the value of the normalized unit-cell volume for the BiFe1−yB3+yO3 perovskites as a function of y in the range of their R3c phase. The Vp(y) dependences in this range are roughly linear with the slopes, which correlate well with the ionic radii of these B3+ cations in octahedral coordination.
In the BiFe1−yB3+yO3 systems, in which iron is substituted by Mn, Cr, or Sc, the rhombohedral R3c phase borders with the antipolar orthorhombic Pnma phase [12,19,20]. When the substituting element is Co or Ga, the R3c phase is followed by the monoclinic Cm one [14,16]. According to results reported by Pan et al. [18], increasing y in the annealed (1−y)BiFeO3yBiZn0.5Ti0.5O3 ceramics leads to a crossover from the rhombohedral to the monoclinic Cc structure, while the data obtained in this work indicate that the next perovskite phase in the ceramics as-synthesized under high-pressure is the tetragonal P4mm.
Generally, the structure sequence (starting from the rhombohedral one at y = 0) in the BiFe1−yB3+yO3 perovskites is determined by the structural type of the BiB3+O3 end member. In particular, in spite of the considerable size difference in the B3+ cations, the as-synthesized metastable perovskites BiMnO3, BiCrO3, and BiScO3 all are monoclinic C2/c [35]. As a result, the sequence of the structural phases in the BiFeO3 derived solid solutions with these perovskites is the same, namely R3cPnmaC2/c as y is increased. Analogously, in the solid solutions with BiCoO3 and BiZn0.5Ti0.5O3, which are both tetragonal P4mm [31,38], the sequence is R3cCm (or Cc)–P4mm. The crystal structure of BiGaO3 is of a pyroxene-type [16]. Therefore, the structure sequence observed in the BiFe1−yGayO3 series is essentially different from the aforementioned ones.
The BiFe1−yB3+yO3 perovskites are all antiferromagnets, at least in the range of the rhombohedral phase. Their TN values are shown in Figure 7 as functions of the substitution rate. One can see that regardless of the nature of the substituting B3+ cation, the Néel temperature decreases with y. Moreover, the negative increment, ΔTN/Δy is roughly the same for the solid solutions, in which B3+ = Mn, Cr, Sc, and Zn0.5Ti0.5 (as obtained in this work from Figure 3). To the best of our knowledge, no data on the TN(y) dependence has been reported regarding the iron-to-cobalt substitution. Taking into account the TN(y) behaviour for the systems with manganese and chromium, a possible compositional variation of the Néel temperature for the (1-y)BiFeO3yBiCoO3 system has been suggested (dashed line in Figure 7). Surprisingly, in the case of the iron-to-scandium substitution [26], magnetic ordering still remains at y = 0.60.
In the annealed samples of the BiFe1−yScyO3 series with y = 0.30, a reversible transition between the AFM state with the cycloidal incommensurate modulation and the collinear AFM ground state was observed at Tm = 230 K [23]. The as-synthesized BiFe0.70Sc0.30O3 phase is the orthorhombic Pnma, while the annealed polymorph of this composition is the rhombohedral R3c. No such AFM-AFM transition has been detected in the as-prepared material [23]. That is why it was suggested that conversion polymorphysm is responsible for the formation of the collinear AFM ground state in BiFe0.7Sc0.3O3. However, Rusakov et al. [21] have reported on the reversible transition between the AFM states with the cycloidal and the collinear spin arrangements in the high-pressure stabilized BiFe0.8Cr0.2O3 at Tm= 260 K. This composition demonstrates no conversion polymorphism and is rhombohedral before and after annealing. Therefore, the formation of the AFM states different from the modulated cycloidal one can be rather associated with the features caused by the high-pressure synthesis of the aforementioned perovskites.
In addition to the transition at Tm, the magnetic measurements of the BiFe1−yScyO3 samples with compositions 0.10 ≤ y ≤ 0.30 revealed some anomalies in the M(T) dependences at Ta (Tm < Ta < TN) that were also associated with the transitions between different AFM states [26]. As seen from Figure 4, no clear evidence of the AFM-AFM phase transformation (at Tm and/or Ta) in BiFe1−y[Zn0.5Ti0.5]yO3 can be found for the compositions with y < 0.20, which may point to the differences in the lattice distortions induced by the substituting atoms of different sizes. Relatively large scandium (with the C.N. 6 ionic radius of 0.89 Å versus 0.69 Å for iron) may cause significant destruction of cycloidal spin arrangement in the parent BiFeO3 at much lower substitution levels as for B3+ = Cr and Zn0.5Ti0.5 that are similar in size to Fe3+.
A substitution usually induces short-range (deformation of Fe3+ coordination octahedra) and long-range structural distortions with the Fe3+–O2−–Fe3+ bond angle change. This may lead to the change in magnetic exchange and the spin canting due to the induced DM interaction, thus enhancing the FM component of the magnetization [10,40,41], as seen in Figure 5. On the other hand, the bismuth site substitution can also cause FM contribution originating from the created oxygen vacancies due to charge compensation achieved by oxygen deficiency after introducing alkali earth ions (e.g., Ca2+) in the BiFeO3 crystal structure [42]. An enhancement of ferromagnetism in BiFeO3 can also be achieved by reducing crystallite size into the nanometre scale when the size becomes comparable with the AFM cycloid period of ~62 nm [43,44]. However, the FM contribution may be enhanced by mechanically induced distortions even in larger BiFeO3 crystallites [45] (as revealed from significant hysteresis in the magnetization loops similar to those shown in Figure 5) when the material is prepared using the mechanochemical synthesis. Such an observation would point to the critical role of the high-pressure synthesis procedure of BiFe1−yB3+yO3 perovskite solid solutions introducing mechanical strain and affecting the magnitude of uncompensated magnetic moments in initial cycloidal AFM arrangement. The enhancement of the ferromagnetism and possible spin canting, suggested from the observed shape of the hysteresis loops in Figure 5, may not be solely the result of chemically induced distortions of the iron substitution but a combined effect.
The dependence of the Néel temperature on the B3+ substitution rate in BiFe1−yB3+yO3 perovskites shown in Figure 7 obviously follows the same trend for different types of iron substitution. Presented results also suggest that the phenomenon of reversible transitions between magnetic states with different types of AFM ordering (collinear, canted, and cycloidal spin arrangements) shares the same features and deserves a particular study. It is very likely that this phenomenon is rather general and was overlooked in the systems with B3+ = Co and Ga.

4. Materials and Methods

Ceramics of the BiFe1−y[Zn0.5Ti0.5]yO3 series (0.05 ≤ y ≤ 0.90) were synthesized under high pressure from the precursors prepared via a solid-state reaction from the stoichiometric oxide mixtures. Details of the precursor preparation and the high-pressure synthesis can be found in Ref. [28].
Phase analysis of the samples before and after annealing was performed using a PANalytical X’Pert Powder X-ray diffractometer (XRD, Ni-filtered Cu Kα radiation) at room temperature. Before the XRD measurements, the samples were reduced into powders. The crystal structure and the magnetic structure of the samples were refined using the FULLPROF package [46].
Magnetic properties of the ceramic samples were measured in the range of 5–300 K using a commercial Quantum Design MPMS3 magnetometer in applied fields up to 70 kOe in both ZFC and FC regimes. For the ZFC measurements, the samples were heated to 400 K, demagnetized from the applied field of 10 kOe to zero field in the Oscillate Mode, and the residual field was removed by the built-in Magnet Reset quench procedure. High temperature (over the range of 300–800 K) measurements were done using a commercial Quantum Design MPMS-XL5 magnetometer equipped with an oven insert. Some of the ceramic samples were annealed prior to the magnetic measurements. Annealings were done in air at 720 K for 1 h.

5. Conclusions

In the as-synthesized (unannealed) ceramics of the BiFe1−y[Zn0.5Ti0.5]yO3 series (0.05 ≤ y ≤ 0.90) prepared using high-pressure synthesis, two perovskite crystalline phases were detected, namely the rhombohedral R3c, which is similar to that in the parent BiFeO3, and the tetragonal P4mm as that in the high-pressure stabilized BiZn0.5Ti0.5O3. No other crystalline phases have been revealed in the obtained samples. The rhombohedral and the tetragonal phases coexist in a wide compositional range (morphotropic phase region) of 0.30 ≤ y ≤ 0.90. In this region, the relative difference between the normalized unit-cell values (Vp) of the phases is almost constant, ΔVp/Vp ≈ 7%.
The magnetic behaviour of the BiFe1−y[Zn0.5Ti0.5]yO3 solid solutions with y < 0.30 is typical of antiferromagnets whose Néel temperature (TN) linearly decreases with y. Ferromagnetic contribution to their magnetic moment was revealed. This contribution was found to be more substantial in the annealed samples.
The Vp(y) dependences of the BiFe1−yB3+yO3 perovskites (B3+ = Ga, Co, Mn, Cr, Sc, and Zn0.5Ti0.5) in the compositional range of their rhombohedral phase are approximately linear with the slopes, which correlate well with the ionic radii of these B3+ cations in octahedral coordination. In particular, the biggest positive slope and the biggest negative slope are observed in the series with B3+ = Co and Sc, respectively.
In contrast to the B3+ ionic size dependent Vp(y) behaviours of the BiFe1−yB3+yO3 perovskites, the compositional dependences of the Néel temperature in the range of their rhombohedral crystalline phases are essentially similar regardless of the nature (magnetic or non-magnetic) of the B3+ cation.
The anomalies in the temperature behaviour of the magnetic moment below TN observed in the BiFe1−yB3+yO3 perovskites with B3+ = Cr, Sc, and Zn0.5Ti0.5 are assumed to indicate to the reversible transitions between magnetic states with different types of antiferromagnetic ordering (collinear, canted, and cycloidal spin arrangements). Occurrence of such transitions is likely to be characteristic of the high-pressure stabilized nature of the BiFe1−yB3+yO3 perovskites and deserves a particular study.

Author Contributions

Conceptualization, A.N.S., V.V.S., D.D.K., and E.Č.; data curation, E.L.F., A.V.F., E.Č., R.T., Y.V.R., A.V.P., and J.P.V.C.; formal analysis, E.Č., R.T. and J.P.V.C.; investigation, E.Č., Y.V.R., A.V.P., J.P.V.C., and A.N.S.; methodology, D.D.K. and A.N.S.; project administration, E.Č. and N.M.O.; resources, A.F. and J.M.V.; supervision, A.F., and J.M.V.; validation, N.M.O. and J.M.V.; visualization, E.Č., J.P.V.C., and A.N.S.; writing—original draft, A.N.S.; writing—review and editing, E.Č., D.D.K., V.V.S., A.F., J.M.V., and A.N.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was done in the frame of the bilateral Slovakia-Belarus project APVV SK-BY-RD-19-0008/T20SLKG-001 funded by the Slovak Research and Development Agency and the Belarusian Republican Foundation for Fundamental Research, respectively. J.P.V.C acknowledges the financial support of FCT—the Portuguese Foundation for Science and Technology for the Ph.D. grant SFRH/BD/145281/2019. A.N.S. acknowledges the financial support of national funds (OE) through FCT—Portugal in the scope of the framework contract foreseen in the numbers 4, 5, and 6 of the article 23, of the Decree-Law 57/2016, of 29 August, changed by Law 57/2017, of 19 July. The research done in University of Aveiro was supported by the project CICECO-Aveiro Institute of Materials, UIDB/50011/2020 and UIDP/50011/2020, financed by national funds through the Portuguese Foundation for Science and Technology (FCT)/MCTES.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khomskii, D. Classifying multiferroics: Mechanisms and effects. Physics 2009, 2, 20. [Google Scholar] [CrossRef] [Green Version]
  2. Catalan, G.; Scott, J.F. Physics and applications of bismuth ferrite. Adv. Mater. 2009, 21, 2463–2485. [Google Scholar] [CrossRef]
  3. Pyatakov, A.P.; Zvezdin, A.K. Magnetoelectric and multiferroic media. Phys.-Uspekhi 2012, 55, 557–581. [Google Scholar] [CrossRef]
  4. Sosnowska, I.; Peterlin-Neumaier, T.; Steichele, E. Spiral magnetic ordering in bismuth ferrite. J. Phys. C Solid State Phys. 1982, 15, 4835–4846. [Google Scholar] [CrossRef]
  5. Arnold, D.C. Composition-driven structural phase transitions in rare-earth-doped BiFeO3 ceramics: A review. IEEE Trans. Ultrason. Ferroelectr. Freq. Control. 2015, 62, 62–82. [Google Scholar] [CrossRef]
  6. Song, G.; Song, Y.; Su, J.; Song, X.; Zhang, N.; Wang, T.; Chang, F. Crystal structure refinement, ferroelectric and ferromagnetic properties of Ho3+ modified BiFeO3 multiferroic material. J. Alloys Compd. 2017, 696, 503–509. [Google Scholar] [CrossRef]
  7. Gebhardt, J.; Rappe, A.M. Doping of BiFeO3: A comprehensive study on substitutional doping. Phys. Rev. B 2018, 98, 125202. [Google Scholar] [CrossRef] [Green Version]
  8. Surdu, V.A.; Truşcă, R.; Vasile, B.S.; Oprea, O.; Tanasa, E.; Diamandescu, L.; Andronescu, E.; Ianculescu, A. Bi1-xEuxFeO3 powders: Synthesis, characterization, magnetic and photoluminescence properties. Nanomaterials 2019, 9, 1465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Mumtaza, F.; Nasira, S.; Hassnain Jaffaria, G.; Ismat Shah, S. Chemical pressure exerted by rare earth substitution in BiFeO3: Effect on crystal symmetry, band structure and magnetism. J. Alloys Compd. 2021, 876, 160178. [Google Scholar] [CrossRef]
  10. Yang, C.-H.; Kan, D.; Takeuchi, I.; Nagarajan, V.; Seidel, J. Doping BiFeO3: Approaches and enhanced functionality. Phys. Chem. Chem. Phys. 2012, 14, 15953–15962. [Google Scholar] [CrossRef]
  11. Sosnowska, I.; Schäfer, W.; Kockelmann, W.; Andersen, K.H.; Troyanchuk, I.O. Crystal structure and spiral magnetic ordering of BiFeO3 doped with manganese. Appl. Phys. A 2002, 74, s1040–s1042. [Google Scholar] [CrossRef]
  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. Suchomel, M.R.; Thomas, C.I.; Allix, M.; Rosseinsky, M.J.; Fogg, A.M.; Thomas, M.F. High pressure bulk synthesis and characterization of the predicted multiferroic Bi(Fe1/2Cr1/2)O3. Appl. Phys. Lett. 2007, 90, 112909. [Google Scholar] [CrossRef]
  14. Azuma, M.; Niitaka, S.; Hayashi, N.; Oka, K.; Takano, M.; Funakubo, H.; Shimakawa, Y. Rhombohedral–tetragonal phase boundary with high Curie temperature in (1-x)BiCoO3–xBiFeO3 solid solution. Jpn. J. Appl. Phys. 2008, 47, 7579–7581. [Google Scholar] [CrossRef]
  15. Oka, K.; Koyama, T.; Ozaaki, T.; Mori, S.; Shimakawa, Y.; Azuma, M. Polarization rotation in the monoclinic perovskite BiCo1−xFexO3. Angew. Chem. Int. Ed. 2012, 51, 7977–7980. [Google Scholar] [CrossRef]
  16. Belik, A.A.; Rusakov, D.A.; Furubayashi, T.; Takayama-Muromachi, E. BiGaO3-based perovskites: A large family of polar materials. Chem. Mater. 2012, 24, 3056–3064. [Google Scholar] [CrossRef]
  17. Khalyavin, D.D.; Salak, A.N.; Olekhnovich, N.M.; Pushkarev, A.V.; Radyush, Y.V.; Manuel, P.; Raevski, I.P.; Zheludkevich, M.L.; Ferreira, M.G.S. Polar and antipolar polymorphs of metastable perovskite BiFe0.5Sc0.5O3. Phys. Rev. B 2014, 89, 174414. [Google Scholar] [CrossRef]
  18. Pan, Z.; Chen, J.; Yu, R.; Yamamoto, H.; Rong, Y.; Hu, L.; Li, Q.; Lin, K.; You, L.; Zhao, K.; et al. Giant polarization and high temperature monoclinic phase in a lead-free perovskite of Bi(Zn0.5Ti0.5)O3-BiFeO3. Inorg. Chem. 2016, 55, 9513–9516. [Google Scholar] [CrossRef]
  19. Salak, A.N.; Khalyavin, D.D.; Pushkarev, A.V.; Radyush, Y.V.; Olekhnovich, N.M.; Shilin, A.D.; Rubanik, V.V. Phase formation in the (1-y)BiFeO3-yBiScO3 system under ambient and high pressure. J. Solid State Chem. 2017, 247, 90–96. [Google Scholar] [CrossRef]
  20. Raevski, I.P.; Kubrin, S.P.; Pushkarev, A.V.; Olekhnovich, N.M.; Radyush, Y.V.; Titov, V.V.; Malitskaya, M.A.; Raevskaya, S.I.; Chen, H. The effect of Cr substitution for Fe on the structure and magnetic properties of BiFeO3 multiferroic. Ferroelectrics 2018, 525, 1–10. [Google Scholar] [CrossRef]
  21. Rusakov, V.S.; Pokatilov, V.S.; Sigov, A.S.; Belik, A.A.; Matsnev, M.E. Changes in the magnetic structure of multiferroic BiFe0.80Cr0.20O3 with temperature. Phys. Solid State 2019, 61, 1030–1036. [Google Scholar] [CrossRef]
  22. Belik, A.A.; Abakumov, A.M.; Tsirlin, A.A.; Hadermann, J.; Kim, J.; Van Tandeloo, G.; Takayama-Muromachi, E. Structure and magnetic properties of BiFe0.75Mn0.25O3 perovskite prepared at ambient and high pressure. Chem. Mater. 2011, 42, 4505–4514. [Google Scholar] [CrossRef]
  23. Khalyavin, D.D.; Salak, A.N.; Fertman, E.L.; Kotlyar, O.V.; Eardley, E.; Olekhnovich, N.M.; Pushkarev, A.V.; Radyush, Y.V.; Fedorchenko, A.; Desnenko, V.A.; et al. The phenomenon of conversion polymorphism in Bi-containing metastable perovskites. Chem. Commun. 2019, 55, 4683–4686. [Google Scholar] [CrossRef]
  24. Fedorchenko, A.; Fertman, E.L.; Salak, A.N.; Desnenko, V.A.; Čižmár, E.; Feher, A.; Vaisburd, A.I.; Olekhnovich, N.M.; Pushkarev, A.V.; Radyush, Y.V.; et al. Unusual magnetic properties of the polar orthorhombic BiFe0.5Sc0.5O3 perovskite. J. Magn. Magn. Mater. 2018, 465, 328–332. [Google Scholar] [CrossRef]
  25. Fertman, E.L.; Fedorchenko, A.; Desnenko, V.A.; Shvartsman, V.V.; Lupascu, D.C.; Salamon, S.; Wende, H.; Vaisburd, A.I.; Stanulis, A.; Ramanauskas, R.; et al. Exchange bias effect in bulk multiferroic BiFe0.5Sc0.5O3. AIP Adv. 2020, 10, 045102. [Google Scholar] [CrossRef] [Green Version]
  26. Fertman, E.L.; Fedorchenko, A.V.; Čižmár, E.; Vorobiov, S.; Feher, A.; Radyush, Y.V.; Pushkarev, A.V.; Olekhnovich, N.M.; Stanulis, A.; Barron, A.R.; et al. Magnetic diagram of the high-pressure stabilized multiferroic perovskites of the BiFe1-yScyO3 series. Crystals 2020, 10, 950. [Google Scholar] [CrossRef]
  27. Shvartsman, V.V.; Khalyavin, D.D.; Olekhnovich, N.M.; Pushkarev, A.V.; Radyush, Y.V.; Salak, A.N. Spontaneous and induced ferroelectricity in the BiFe1-xScxO3 perovskite ceramics. Phys. Status Solidi A 2021, 218, 2100173. [Google Scholar] [CrossRef]
  28. Čižmár, E.; Vorobiov, S.; Kliuikov, A.; Radyush, Y.V.; Pushkarev, A.V.; Olekhnovich, N.M.; Cardoso, J.P.; Salak, A.N.; Feher, A. Structural and magnetic phase transitions in the Fe-rich compositional ranges of the multiferroic BiFe1-x[Zn0.5Ti0.5]xO3 perovskites. Integr. Ferroelectr. 2021. [Google Scholar] [CrossRef]
  29. Arafat, S.S. Structural transition and magnetic properties of high Cr-doped BiFeO3 ceramic. Cerâmica 2020, 66, 114–118. [Google Scholar] [CrossRef]
  30. Salak, A.N.; Shilin, A.D.; Bushinski, M.V.; Olekhnovich, N.M.; Vyshatko, N.P. Structural regularities and dielectric phenomena in the compound series PbB3+1/2Nb1/2O3. Mater. Res. Bull. 2000, 35, 1429–1438. [Google Scholar] [CrossRef]
  31. Suchomel, M.R.; Fogg, A.M.; Allix, M.; Niu, H.; Claridge, J.B.; Rosseinsky, M.J. Bi2ZnTiO6:  a lead-free closed-shell polar perovskite with a calculated ionic polarization of 150 μC cm−2. Chem. Mater. 2006, 18, 4987–4989. [Google Scholar] [CrossRef]
  32. Salak, A.N.; Shvartsman, V.V.; Cardoso, J.P.; Pushkarev, A.V.; Radyush, Y.V.; Olekhnovich, N.M.; Khalyavin, D.D.; Vieira, J.M.; Čižmár, E.; Feher, A. The orthorhombic-tetragonal morphotropic phase boundary in high-pressure synthesized BiMg0.5Ti0.5O3-BiZn0.5Ti0.5O3 perovskite solid solutions. J. Phys. Chem. Solids 2022, 161, 110392. [Google Scholar] [CrossRef]
  33. Khalyavin, D.D.; Salak, A.N.; Vyshatko, N.P.; Lopes, A.B.; Olekhnovich, N.M.; Pushkarev, A.V.; Maroz, I.I.; Radyush, Y.V. Crystal structure of metastable perovskite Bi(Mg1/2Ti1/2)O3: Bi-based structural analogue of antiferroelectric PbZrO3. Chem. Mater. 2006, 18, 5104–5110. [Google Scholar] [CrossRef]
  34. Prosandeev, S.; Wang, D.; Ren, W.; Íñiguez, J.; Bellaiche, L. Novel nanoscale twinned phases in perovskite oxides. Adv. Funct. Mater. 2013, 23, 234–240. [Google Scholar] [CrossRef]
  35. Kubel, F.; Schmid, H. Structure of a ferroelectric and ferroelastic monodomain crystal of the perovskite BiFeO3. Acta Cryst. B 1990, 46, 698–702. [Google Scholar] [CrossRef] [Green Version]
  36. Popov, Y.F.; Zvezdin, A.K.; Vorob’ev, G.P.; Kadomtseva, A.M.; Murashev, V.A.; Rakov, D.N. Linear magnetoelectric effect and phase transitions in bismuth ferrite, BiFeO3. JETP Lett. 1993, 57, 69–73. [Google Scholar]
  37. Ruette, B.; Zvyagin, S.; Pyatakov, A.P.; Bush, A.; Li, J.F.; Belotelov, V.I.; Zvezdin, A.K.; Viehland, D. Magnetic-field-induced phase transition in BiFeO3 observed by high-field electron spin resonance: Cycloidal to homogeneous spin order. Phys. Rev. B. 2004, 69, 064114. [Google Scholar] [CrossRef] [Green Version]
  38. Belik, A.A. Polar and nonpolar phases of BiMO3: A review. J. Solid State Chem. 2012, 195, 32–40. [Google Scholar] [CrossRef]
  39. Belik, A.A.; Iikubo, S.; Kodama, K.; Igawa, N.; Shamoto, S.; Niitaka, S.; Azuma, M.; Shimakawa, Y.; Takano, M.; Izumi, F.; et al. Neutron powder diffraction study on the crystal and magnetic structures of BiCoO3. Chem. Mater. 2006, 18, 798–803. [Google Scholar] [CrossRef]
  40. Sharma, V.; Ghosh, R.K.; Kuanr, B.K. Investigation of room temperature ferromagnetism in transition metal doped BiFeO3. J. Phys. Condens. Matter 2019, 31, 395802. [Google Scholar] [CrossRef]
  41. Freitas, V.F.; Bonadio, T.G.M.; Dias, G.S.; Protzek, O.A.; Medina, A.N.; Cótica, L.F.; Santos, I.A.; Garcia, D.; Eiras, J.A. On the microscopic mechanism for the stabilization of structural and ferroic states in displacive multiferroics. J. Appl. Phys. 2013, 113, 114105. [Google Scholar] [CrossRef]
  42. Sánchez-De Jesús, F.; Bolarín-Miró, A.M.; Cortés-Escobedo, C.A.; Barba-Pingarrón, A.; Pedro-García, F. Enhanced ferromagnetic and electric properties of multiferroic BiFeO3 by doping with Ca. J. Alloys Compd. 2020, 824, 153944. [Google Scholar] [CrossRef]
  43. Huang, F.; Wang, Z.; Lu, X.; Zhang, J.; Min, K.; Lin, W.; Ti, R.; Xu, T.; He, J.; Yue, C.; et al. Peculiar magnetism of BiFeO3 nanoparticles with size approaching the period of the spiral spin structure. Sci. Rep. 2013, 3, 2907. [Google Scholar] [CrossRef] [Green Version]
  44. Da Silva, K.L.; Menzel, D.; Feldhoff, A.; Kubel, C.; Bruns, M.; Paesano, A., Jr.; Duvel, A.; Wilkening, M.; Ghafari, M.; Hahn, H.; et al. Mechanosynthesized BiFeO3 nanoparticles with highly reactive surface and enhanced magnetization. J. Phys. Chem. C 2011, 115, 7209–7217. [Google Scholar] [CrossRef]
  45. Da Silva, K.L.; Trautwein, R.S.; Da Silva, R.B.; Fabián, M.; Čižmár, E.; Holub, M.; Skurikhina, O.; Harničárová, M.; Girman, V.; Menzel, D.; et al. Suppression of the cycloidal spin arrangement in BiFeO3 caused by the mechanically induced structural distortion and its effect on magnetism. Front. Matter. 2021. [Google Scholar] [CrossRef]
  46. Rodriguez-Carvajal, J. Recent advances in magnetic structure determination by neutron powder diffraction. Phys. B Phys. Condens. Matter 1993, 192, 55–69. [Google Scholar] [CrossRef]
Figure 1. The XRD patterns of the (1−y)BiFeO3yBiZn0.5Ti0.5O3 samples as-synthesized under high pressure. The numbers at the diffractograms denote the y values. The shadow areas indicate the angular ranges of (001)p, (011)p and (111)p reflection families of the primitive perovskite lattice. The dotted lines point out the 2Theta positions of the reflections corresponding to the rhombohedral R3c phase (red lines) and the tetragonal P4mm phase (blue lines).
Figure 1. The XRD patterns of the (1−y)BiFeO3yBiZn0.5Ti0.5O3 samples as-synthesized under high pressure. The numbers at the diffractograms denote the y values. The shadow areas indicate the angular ranges of (001)p, (011)p and (111)p reflection families of the primitive perovskite lattice. The dotted lines point out the 2Theta positions of the reflections corresponding to the rhombohedral R3c phase (red lines) and the tetragonal P4mm phase (blue lines).
Magnetochemistry 07 00151 g001
Figure 2. (a) The primitive perovskite cell parameters and (b) the normalized unit-cell volume of the BiFe1−y[Zn0.5Ti0.5]yO3 perovskite phases as a function of y with the ranges of phase coexistence (0.30 ≤ y ≤ 0.90) indicated. The data for the end members (y = 0 and y = 1) were taken from Refs. [31,35], respectively. The polyhedral representations of the respective structures are shown.
Figure 2. (a) The primitive perovskite cell parameters and (b) the normalized unit-cell volume of the BiFe1−y[Zn0.5Ti0.5]yO3 perovskite phases as a function of y with the ranges of phase coexistence (0.30 ≤ y ≤ 0.90) indicated. The data for the end members (y = 0 and y = 1) were taken from Refs. [31,35], respectively. The polyhedral representations of the respective structures are shown.
Magnetochemistry 07 00151 g002
Figure 3. The normalized temperature-dependent magnetic moment of the annealed BiFe1−y[Zn0.5Ti0.5]yO3 samples measured in the FC regime in the temperature range above 330 K.
Figure 3. The normalized temperature-dependent magnetic moment of the annealed BiFe1−y[Zn0.5Ti0.5]yO3 samples measured in the FC regime in the temperature range above 330 K.
Magnetochemistry 07 00151 g003
Figure 4. Temperature dependence of M/H of the as-synthesized BiFe1−y[Zn0.5Ti0.5]yO3 samples with 0.1 ≤ y < 0.2 measured in the ZFC regime in the temperature range of 5–400 K. The Tm and Ta were estimated from the derivative of the M/H(T) curve.
Figure 4. Temperature dependence of M/H of the as-synthesized BiFe1−y[Zn0.5Ti0.5]yO3 samples with 0.1 ≤ y < 0.2 measured in the ZFC regime in the temperature range of 5–400 K. The Tm and Ta were estimated from the derivative of the M/H(T) curve.
Magnetochemistry 07 00151 g004
Figure 5. Magnetization loops of the annealed BiFe1−y[Zn0.5Ti0.5]yO3 samples measured at 300 K normalized to the magnetization value Mmax at the maximum applied field. The inset shows the magnetization loop for the composition with y = 0.15 before and after annealing.
Figure 5. Magnetization loops of the annealed BiFe1−y[Zn0.5Ti0.5]yO3 samples measured at 300 K normalized to the magnetization value Mmax at the maximum applied field. The inset shows the magnetization loop for the composition with y = 0.15 before and after annealing.
Magnetochemistry 07 00151 g005
Figure 6. The compositional behaviour of the normalized unit-cell volume for the BiFe1−yB3+yO3 perovskites with B3+ = Ga [16], Co [14], Mn [11], Cr [20], Sc [19], and Zn0.5Ti0.5 (this work). The data corresponding to the rhombohedral R3c phase range are only shown. The Vp value of the BiFe0.70Sc0.30O3 perovskite (open symbol) was determined by refinement of neutron diffraction data collected at room temperature on the annealed sample with the R3c symmetry (see Ref. [23] for details).
Figure 6. The compositional behaviour of the normalized unit-cell volume for the BiFe1−yB3+yO3 perovskites with B3+ = Ga [16], Co [14], Mn [11], Cr [20], Sc [19], and Zn0.5Ti0.5 (this work). The data corresponding to the rhombohedral R3c phase range are only shown. The Vp value of the BiFe0.70Sc0.30O3 perovskite (open symbol) was determined by refinement of neutron diffraction data collected at room temperature on the annealed sample with the R3c symmetry (see Ref. [23] for details).
Magnetochemistry 07 00151 g006
Figure 7. The Néel temperature as a function of y for the BiFe1−yB3+yO3 perovskites with B3+ = Co [39], Mn [12], Cr [29], Sc [26] and Zn0.5Ti0.5 (this work). A suggested TN(y) behaviour for the (1−y)BiFeO3yBiCoO3 perovskites is shown with the dashed line.
Figure 7. The Néel temperature as a function of y for the BiFe1−yB3+yO3 perovskites with B3+ = Co [39], Mn [12], Cr [29], Sc [26] and Zn0.5Ti0.5 (this work). A suggested TN(y) behaviour for the (1−y)BiFeO3yBiCoO3 perovskites is shown with the dashed line.
Magnetochemistry 07 00151 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Salak, A.N.; Cardoso, J.P.V.; Vieira, J.M.; Shvartsman, V.V.; Khalyavin, D.D.; Fertman, E.L.; Fedorchenko, A.V.; Pushkarev, A.V.; Radyush, Y.V.; Olekhnovich, N.M.; et al. Magnetic Behaviour of Perovskite Compositions Derived from BiFeO3. Magnetochemistry 2021, 7, 151. https://doi.org/10.3390/magnetochemistry7110151

AMA Style

Salak AN, Cardoso JPV, Vieira JM, Shvartsman VV, Khalyavin DD, Fertman EL, Fedorchenko AV, Pushkarev AV, Radyush YV, Olekhnovich NM, et al. Magnetic Behaviour of Perovskite Compositions Derived from BiFeO3. Magnetochemistry. 2021; 7(11):151. https://doi.org/10.3390/magnetochemistry7110151

Chicago/Turabian Style

Salak, Andrei N., João Pedro V. Cardoso, Joaquim M. Vieira, Vladimir V. Shvartsman, Dmitry D. Khalyavin, Elena L. Fertman, Alexey V. Fedorchenko, Anatoli V. Pushkarev, Yury V. Radyush, Nikolai M. Olekhnovich, and et al. 2021. "Magnetic Behaviour of Perovskite Compositions Derived from BiFeO3" Magnetochemistry 7, no. 11: 151. https://doi.org/10.3390/magnetochemistry7110151

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