You are on page 1of 8

Journal of Magnetism and Magnetic Materials 585 (2023) 171142

Contents lists available at ScienceDirect

Journal of Magnetism and Magnetic Materials


journal homepage: www.elsevier.com/locate/jmmm

Research article

Unusual magnetic properties of (Sr, Mn)-substituted BiFeO3 near the polar/


antipolar phase boundary
V.A. Khomchenko *, M. Das , M.S.C. Henriques , J.A. Paixão
CFisUC, Department of Physics, University of Coimbra, P-3004-516 Coimbra, Portugal

A R T I C L E I N F O A B S T R A C T

Keywords: Investigations of the crystal structure and magnetic properties of the Bi0.85Sr0.15Fe1-xMnxO3 (0.1 ≤ x ≤ 0.5)
BiFeO3 ferromanganites were carried out. The materials maintain the single-phase polar R3c structure of the pure bis­
Magnetic properties muth ferrite up to x = 0.35 (a further increase in the Mn content stabilizes the antipolar Pnam phase), thus
Multiferroics
exhibiting the coexistence of spontaneous polarization and magnetic order over a wide range of Mn concen­
Perovskites
trations, sufficient to influence the cycloidal antiferromagnetic arrangement typical of the parent compound and
contribute to the canted antiferromagnetic phase formation (at room temperature). A complex magnetic
behaviour with irregular changes in remanent magnetization and coercive field suggesting the temperature-
driven modification of the canted antiferromagnetic structure near the polar/antipolar phase boundary has
also been found.

1. Introduction [7,8] resulting in zero net magnetization. The cycloidal modulation


inhibiting the observation of the linear magnetoelectric coupling can be
Magnetoelectric multiferroics [1,2] have attracted considerable modified/removed by applying a strong (above 10 T at room tempera­
attention due to the coupling between the ferroelectric and magnetic ture) magnetic field [9]. Chemical substitution at the Bi or/and Fe sites
order parameters, which makes them promising for a variety of appli­ [10–12] can reduce the threshold field of the cycloidal structure to
cations including data storage and energy harvesting [3,4]. While large canted structure transformation or even stabilize the canted antiferro­
linear magnetoelectric effects should be seen in ferroelectric and, magnetic order permitting the linear magnetoelectric response [13]. The
simultaneously, ferromagnetic materials [5], no reliable data indicating known examples of chemical substitution-related strategies to achieve
the possibility to combine the electric polarization and ferromagnetic spontaneous magnetization in the multiferroic BiFeO3 thus exploit the
spin ordering in a single phase has yet been reported. The applicability possibility to affect the magnetic anisotropy to suppress its cycloidal
of traditional multiferroic materials is largely limited by their specific antiferromagnetic structure in favour of the canted antiferromagnetic
crystal and magnetic structures, which often restrict the range of one. The value of spontaneous magnetization associated with the canted
achievable properties. One potential approach in engineering multi­ antiferromagnetic arrangement in the polar phase depends on the
ferroic materials is through chemical substitution. By introducing mul­ structural characteristics [14] and does not typically exceed 0.3 emu/g
tiple elements into the material’s crystal lattice, one can stabilize [11,12].
different crystal structures, modify the magnetic and ferroelectric An alternative approach towards attaining a large switchable
behaviour, and tune the coupling between the order parameters. magnetization in the BiFeO3-based multiferroics has been recently
Bismuth ferrite (BiFeO3) is an antiferromagnetic and ferroelectric proposed [15,16]. This approach exploits the doping scheme in which
perovskite-type compound which, due to its exceptionally high Néel and Fe3+ ions are partially replaced by manganese in the mixed (Mn3+/
Curie temperatures (TN≈ 640 K and TC≈ 1100 K, respectively) and large Mn4+) oxidation state to provide the ferromagnetic exchange coupling
spontaneous polarization (Ps ~ 100 µC/cm2), is often considered as a Mn3+: t32ge1g –O– Mn4+: t32ge0g (the coexistence of Mn3+ and Mn4+ in the
model room-temperature multiferroic with great potential for practical crystal lattice is achieved by a partial heterovalent substitution of Bi3+
use [6]. The material possesses quite complex magnetic structure with for an alkali-earth ion AE2+). The previous investigation of the crystal
the cycloidal modulation of the canted G-type antiferromagnetic order and magnetic structures as well as magnetic properties of the

* Corresponding author.
E-mail address: uladzimir@uc.pt (V.A. Khomchenko).

https://doi.org/10.1016/j.jmmm.2023.171142
Received 2 June 2023; Received in revised form 31 July 2023; Accepted 12 August 2023
Available online 14 August 2023
0304-8853/© 2023 The Author(s). Published by Elsevier B.V. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
V.A. Khomchenko et al. Journal of Magnetism and Magnetic Materials 585 (2023) 171142

Fig. 2. (a) Normalized lattice parameters for the polar rhombohedral [an = a/
Fig. 1. Observed, calculated and difference X-ray powder diffraction patterns √2, cn = c/(2√3); space group R3c] (red symbols) and antipolar orthorhombic
taken for the Bi0.85Sr0.15Fe1-xMnxO3 [x = 0.3 (a), 0.4 (b) and 0.5 (c)] samples at [an = a/√2, bn = b/(2√2), cn = c/4; space group Pnam] (blue symbols) sam­
room temperature. ples. (b) Primitive cell volume for the Bi0.85Sr0.15Fe1-xMnxO3 samples. The lines
are guides to the eye.
Bi0.85Ca0.15Fe1-xMnxO3 (0.1 ≤ x ≤ 0.5) compounds (the Ca2+/Bi3+ ratio
was chosen as close to the largest one allowing to keep the spontaneous pestle, and pressed into pellets. The pellets were further placed into
electric polarization over a wide range of Mn concentration) [17,18] has closed alumina crucibles, heated up to 950 ◦ C with a rate of 300 ◦ C/h,
shown that the increase in Mn content gives rise to the sequence of phase annealed for 20 h, and then cooled down to room temperature at a rate
transitions from the canted antiferromagnetic polar state to the collinear of ~ 100 ◦ C/h. Phase purity of the samples was examined by scanning
antiferromagnetic polar state, and then to the paramagnetic nonpolar electron microscopy (SEM) and X-ray diffraction (XRD) techniques.
state at room temperature [15,16]. The appearance of spontaneous Secondary electron microscopy and backscattered electron microscopy
magnetization associated with a modification of the collinear antifer­ images of fractured surfaces of gold–palladium coated Bi0.85Sr0.15
romagnetic spin arrangement specific to the ferroelectric Bi0.85Ca0.15 Fe1-xMnxO3 ceramics were obtained with a VEGA-3 SB scanning electron
Fe1-xMnxO3 samples near the polar / nonpolar phase boundary has been microscope (TESCAN). XRD patterns were collected using a D8 Advance
observed with decreasing temperature. A metamagnetic transition diffractometer (Bruker) with Cu Kα radiation. To identify the phase
accompanied by the large (up to ~ 4 emu/g at low temperature) jumps composition and crystal structure of the materials under study, the
of magnetization taking place in a small (a few kOe) magnetic field has diffraction patterns were analyzed with the FullProf software suite [19].
also been found. The observation of the unusual magnetic behaviour in Contact atomic force microscopy (AFM) and piezoresponse force mi­
the Bi0.85Ca0.15Fe1-xMnxO3 series has motivated us to perform the croscopy (PFM) measurements of the mechanically-polished ceramics
comparative study of the crystal structure and magnetic state of the were conducted using a NTEGRA Prima scanning probe microscope (NT-
Bi0.85Sr0.15Fe1-xMnxO3 (0.1 ≤ x ≤ 0.5) compounds to investigate the MDT Spectrum Instruments). The PPP-NCHR-50 probes (Nanosensors)
effect of the cationic size mismatch on the multiferroic properties of the with a resonant frequency of ~260 kHz were used. Magnetization
Bi1-yAEyFe1-xMnxO3 perovskites. measurements were carried out in magnetic fields up to 9 T applied over
the temperature range from 2 K to 400 K. A cryogen-free Physical
2. Experimental Properties Measurement System (PPMS DynaCool, Quantum Design)
was used.
Ceramic samples of the Bi0.85Sr0.15Fe1-xMnxO3 (x = 0.1, 0.2, 0.3,
0.35, 0.4, 0.5) ferromanganites were obtained by a solid-state reaction
between Bi2O3, SrCO3, Fe2O3, and Mn2O3. The powdered reagents were
taken in stoichiometric cation ratios, mixed using an agate mortar and a

2
­
­
V.A. Khomchenko et al. Journal of Magnetism and Magnetic Materials 585 (2023) 171142

3. Results and discussion

3.1. XRD, PFM, SEM

Results of the Rietveld refinement of the X-ray diffraction data


collected for the Bi0.85Sr0.15Fe1-xMnxO3 (x ≤ 0.35) compounds at room
temperature (Fig. 1 (a)) indicate that the materials possess the non-
centrosymmetric rhombohedral structure (space group R3c) character­
istic of the pure BiFeO3 [20] and low-doped Bi1-ySryFeO3-y/2 [21–24].
The related structural distortions (as described with respect to the
reference cubic perovskite ABO3) can be presented in terms of the su­
perposition of an alternating clockwise / counterclockwise tilting of the
BO6 octahedra along the threefold pseudo-cubic axis (tilt system
a− a− a− ) [25] and polar displacements of ions along the same [1 1 1]cubic
lattice direction. A further increase in the Mn concentration gives rise to
the appearance of the antipolar orthorhombic phase (space group Pnam)
specific to the BiFe1-xMnxO3 (0.25 < x < 0.65) perovskites [26] (Fig. 1
(b, c)). Comparison of the phase boundaries in the Bi0.85Sr0.15
Fe1-xMnxO3 and BiFe1-xMnxO3 systems implies that the Sr substitution
favours the expansion of the Mn concentration range compatible with
the ferroelectric phase existence.
The Pnam structure, which combines the modulated in-phase /
antiphase (a− a− c+/a− a− c− ) oxygen octahedra tilting with the PbZrO3-
like [27] antipolar displacements of the A-site ions along the [110/11
0]cubic directions, is known to be stabilized via rare-earth (RE) substi­
tution of the parent bismuth ferrite [28–30]. The composition-driven
polar-to-antipolar structural phase transition has also been reported
for the (Bi, RE)Fe1-xMnxO3 series with invariable concentrations of RE
ions in the A-sublattice [31,32]. The Mn substitution-induced R3c →
Pnam transformation can also be observed in the Bi0.9Ca0.1Fe1-xMnxO3
system (at x = 0.45) [33]. On the contrary, the crystal structure of the
Bi0.85Ca0.15Fe1-xMnxO3 solid solution [16] seems not to follow the trend
toward the formation of the intermediate antipolar phase: this system
Fig. 3. Topography AFM (right column) and vertical PFM (left column) images
undergoes the polar-to-nonpolar (R3c → Pnma) phase transition at 0.4 < of the mechanically-polished ceramics of Bi0.85Sr0.15Fe1-xMnxO3 (x = 0.1, 0.3
x < 0.45 (the same nonpolar Pnma structure is characteristic of the end and 0.5). Image size: 5 × 5 μm. The bright and dark regions in the PFM images
member of this series, CaMnO3) [34]. correspond to the ferroelectric domains with the polarization directed to the
Compositional dependence of the normalized lattice parameters of free surface and to the bulk of a sample, respectively.
the Bi0.85Sr0.15Fe1-xMnxO3 (x ≤ 0.5) samples is shown in Fig. 2. One can
see the change in the rate of the primitive cell volume decrease taking force microscopy data (Fig. 3). In the PFM measurements conducted on
place for the rhombohedral compounds with increasing Mn concentra­ the samples with the rhombohedral structure, an AC voltage applied
tion above x = 0.1 (primitive cell volume of the x = 0 compound is between the sample and the conductive cantilever introduces a periodic
reported to be within the range from 61.63 Å3 to 62.51 Å3 [24,35,36]). variation in the A⋅cos ϕ signal (A is the amplitude of the measured
The same behaviour reflecting a change in mechanism of defect for­ mechanical oscillation and ϕ is the phase of the piezoresponse) reflect­
mation in the aliovalently-doped samples with increasing x has been ing the morphology of ferroelectric domains. The absence of cross-
previously observed for the Bi0.85Ca0.15Fe1-xMnxO3 system [16]. It has correlation between the sample topography and the piezoresponse ex­
been shown that the deviation from the linear decrease in the lattice cludes a topographic origin of the PFM signals. The orthorhombic
parameters of the isostructural solid solutions suggested by the Vegard’s sample with x = 0.5 shows no apparent piezoelectric response, as ex­
law can be explained by considering the charge compensation routes pected for a material with a centrosymmetric crystal structure.
characteristic of the Bi1-yAEyFe1-xMnxO3 series [16]. In the low-doped Magnetic properties of BiFeO3-based multiferroics can be strongly
Bi1-yAEyFeO3-y/2 perovskites, the aliovalent substitution of Bi3+ with affected by presence of ferro- or ferrimagnetic impurities [45,46].
alkali-earth cations AE2+ results in the oxygen vacancies formation, iron Although the XRD measurements detected no traces of secondary phases
keeping its highly-stable oxidation state +3 [37–39]. A partial replace­ in the samples under study, their phase purity was additionally verified
ment of Fe by Mn is accompanied by the filling of the anion vacancies by scanning electron microscopy technique. Indeed, the comparison of
indicating that the valence state of manganese at x ≤ y is +4 [40,41]. secondary electron microscopy images (providing the information on
Above the threshold concentration (i.e. at x > y), the requirement of the topography of the sample’s surface) and backscattered electron
charge compensation suggests that the extra manganese is present in microscopy images (which are heavily dependent on the atomic number
trivalent state [42] and, thus, the coexistence of Mn4+ with Mn3+ should of the elements composing the phase) would allow us to identify
be observed [43]. The slowing in the rate of the lattice parameters different phases at the submicrometre level [46]. Examples of such
decrease which takes place for the Bi1-yAEyFe1-xMnxO3 systems after measurements revealing no composition-dependent variations in the
attaining the threshold Mn concentration (x = y) is then determined by image contrast are shown in Fig. 4. Similar to the trend previously
the difference in the ionic radii of Mn4+, Mn3+ and Fe3+ cations (rMn4+ = described for the Bi0.9Ca0.1Fe1-xMnxO3 [33] and Bi0.85Ca0.15Fe1-xMnxO3
0.53 Å, rMn3+ ≈ rFe3+ = 0.645 Å) [44]. series [16], the increase (from ~ 1 μm to ~ 10 μm) in the typical grain
Validity of the proposed structural model which assumes that the size of the Bi0.85Sr0.15Fe1-xMnxO3 ceramics was observed with increasing
Bi0.85Sr0.15Fe1-xMnxO3 system undergoes the polar / antipolar structural Mn content. The subsequent study of magnetic properties of the Sr/Mn-
phase transition at x = 0.4 (Fig. 2) is confirmed by the piezoresponse doped compounds didn’t find any clear correlation between the

3
­
V.A. Khomchenko et al. Journal of Magnetism and Magnetic Materials 585 (2023) 171142

Fig. 5. (a) and (b): Field dependencies of magnetization measured for the
Bi0.85Sr0.15Fe1-xMnxO3 (0.1 ≤ x ≤ 0.5) compounds at T = 300 K. The inset in (b)
shows the enlarged part of the main figure for a higher clarity.

the presence of weak ferromagnetic phase with the canted antiferro­


magnetic ordering) and a metamagnetic behaviour at high fields.
Spontaneous magnetization characteristic of this sample (as evaluated
from the linear extrapolation of the high-field M (H) dependence to H =
0) is practically coincident with the remanent one. The high-field
anomaly (the increase in magnetic susceptibility with increasing mag­
netic field) is very similar to that reflecting the onset of the field-induced
Fig. 4. First row: SEM images obtained for the Bi0.85Sr0.15Fe1-xMnxO3 (x = 0.1 antiferromagnetic to weak ferromagnetic transformation characteristic
and 0.2) ceramics in secondary electron mode. Second, third and fourth rows:
of the phase with the cycloidal antiferromagnetic spin arrangement [9].
SEM images obtained for the Bi0.85Sr0.15Fe1-xMnxO3 (x = 0.3, 0.4 and 0.5) ce­
The chemical substitution-driven partial elimination of the cycloidal
ramics in secondary electron (left column) and backscattered electron (right
column) modes. The scale bar (highlighted in yellow) is 5 μm.
modulation resulting in the co-existence of the cycloidal and canted
antiferromagnetic phases is commonly observed in the low-doped oxy­
gen-deficient Bi1-yAEyFeO3-y/2 multiferroics [47,48]. It is assumed that
microstructure and the magnetic state of the samples.
the local strain fields generated by oxygen vacancies tend to suppress the
coupling between gradients of the magnetic order parameter to stabilize
3.2. Magnetic properties the G-type antiferromagnetic structure with the moments lying in the
basal hexagonal plane [49,50] (such an orientation gives rise to spin
Analysis of the field dependences of magnetization obtained for the canting (weak ferromagnetism) due to the Dzyaloshinskii-Moriya
Bi0.85Sr0.15Fe1-xMnxO3 (x ≤ 0.5) compounds at room temperature (Fig. 5 interaction) [51].
(a, b)) has found rather complex evolution of magnetic behaviour of the The magnetic hysteresis loop measurement conducted for the sample
solid solutions with increasing Mn concentration. The magnetic hys­ with x = 0.2 (Fig. 5(a), S2) demonstrates a one and a half-fold increase in
teresis loop measured on the sample with x = 0.1 exhibits the features the spontaneous magnetization implying an increase in the volume of
indicative of its inhomogeneous magnetic state (see Fig. S1 in the sup­ the weak ferromagnetic phase. Another effect of the 20% Mn substitu­
plementary materials file for more details). Indeed, the measurement tion is a decrease in the threshold field of the magnetic field-induced
reveals both a small (0.043 emu/g) remanent magnetization (testifying cycloidal structure → canted structure transition. Such a decrease in

4
V.A. Khomchenko et al. Journal of Magnetism and Magnetic Materials 585 (2023) 171142

Fig. 7. Temperature dependencies of magnetization measured for the


Bi0.85Sr0.15Fe1-xMnxO3 (x = 0.3, 0.2 and 0.1) samples with an applied field H =
2 kOe in zero field cooling (ZFC), field-cooled cooling (FCC) and field-cooled
warming (FCW) modes.

octahedra rotation-related components of the Dzyaloshinskii-Moriya


interaction [53].
In contrast to the magnetic properties of the rare-earth-substituted
compounds implying that the complete suppression of the cycloidal
order in the ferroelectric phase is never reached [54], the M(H)
dependence obtained for the Bi0.85Sr0.15Fe0.7Mn0.3O3 ferrite suggests the
collapse of the cycloidal spin modulation (Fig. 5 (a), S3). Indeed, the
sample demonstrates a weak ferromagnetic behaviour without any
metamagnetic anomalies which could be attributed to the cycloidal
phase existence. Its spontaneous magnetization, Ms = 0.08 emu/g, is
comparable with those characteristic of the weak ferromagnetic solid
solutions of Bi1-ySryFeO3-y/2 [47] and Bi1-xCaxFe1-xMnxO3 [18] near the
polar/nonpolar phase boundary. It is worth noting that the preceding
investigations of the magnetic properties of the BiFe1-xMnxO3 [55] and
Bi0.85Ca0.15Fe1-xMnxO3 [16] multiferroics didn’t reveal any increase in
the magnetization with increasing Mn content that would assume the
stabilization of the weak ferromagnetic state at room temperature.
A further increase in the Mn concentration dramatically changes
magnetic state of the Bi0.85Sr0.15Fe1-xMnxO3 compounds at room tem­
perature (Fig. 5 (b)). One can see that the samples with 0.35 ≤ x ≤ 0.5
are characterized by the linear-like field dependences of magnetization
expected for either collinear antiferromagnetic or paramagnetic mate­
rials. However, a more careful examination of the M (H) dependences
obtained for the samples with x = 0.35 and x = 0.4 confirms the exis­
tence of a small (0.0065 emu/g and 0.0057 emu/g, respectively)
remanent magnetization (inset in Fig. 5 (b)) calling the preliminary
conclusion into question.
Fig. 6. Temperature dependencies of magnetization measured for the To clarify the character of magnetic state of the Bi0.85Sr0.15
Bi0.85Sr0.15Fe1-xMnxO3 [x = 0.5 (a), 0.4 (b) and 0.35 (c)] samples with an Fe1-xMnxO3 samples near the polar/antipolar phase boundary, the
applied field H = 2 kOe in zero field cooling (ZFC), field-cooled cooling (FCC) measurements of the temperature dependences of magnetization using
and field-cooled warming (FCW) modes. zero field cooling (ZFC), field-cooled cooling (FCC) and field-cooled
warming (FCW) protocols have been carried out (Fig. 6). Analysis of
the critical field accompanied by the appearance of a remanent the M (T) dependences allows us to conclude that the samples with x =
magnetization is very typical of the isovalent substituted Bi1-yREyFeO3 0.5, x = 0.4 and x = 0.35 undergo the paramagnetic state / magnetically
multiferroics [10,52] demonstrating the instability of the cycloidal ordered state transition at 214(1) K, 325(1) K and 375(1) K, respec­
ordering with respect to the cooperative chemical substitution-induced tively. The Néel temperatures (determined from the inflection points of
structural distortions affecting the polarization-related and oxygen the M(T) curves) are very close to those specific to the

5
­
V.A. Khomchenko et al. Journal of Magnetism and Magnetic Materials 585 (2023) 171142

Fig. 9. Temperature dependencies of magnetization measured for the


Bi0.85Sr0.15Fe0.65Mn0.35O3 sample with applied fields H = 1 kOe and H = 5 kOe
Fig. 8. Temperature dependencies of magnetization measured for the in zero field cooling (ZFC), field-cooled cooling (FCC) and field-cooled warming
Bi0.85Sr0.15Fe0.6Mn0.4O3 sample with applied fields H = 1 kOe and H = 5 kOe in (FCW) modes.
zero field cooling (ZFC), field-cooled cooling (FCC) and field-cooled warming
(FCW) modes.
of ordered magnetic moment per Fe/Mn atom (reflecting frustration of
the exchange interactions with increasing Mn concentration) and
Bi0.85Ca0.15Fe1-xMnxO3 compounds with the corresponding Mn con­
frequency-dependent magnetic behaviour supporting the scenario of the
centrations [16].
antiferromagnetic (weak ferromagnetic) and glassy phase coexistence
The ZFC and FC curves obtained for the sample with x = 0.5 shows a
[15,16].
large irreversibility below 200 K (i. e. close to the Néel point) with a cusp
A very unusual magnetization vs. temperature dependences with a
observed in the ZFC dependence at much lower temperature of 13 K
large hysteresis between the FCC and FCW curves assuming a first-order
(Fig. 6 (a)). A similar behaviour has previously been described for the
magnetic phase transition have been obtained for the samples with x =
Bi0.5Sr0.5Fe0.5Mn0.5O3 compound [41]. The magnetic properties study
0.4 and x = 0.35 (Fig. 6 (b, c)). A strong magnetic field makes the
supplemented by the neutron diffraction and Mössbauer spectroscopy
hysteresis less pronounced (Figs. 8, 9). To clarify the origin of the
measurements [41] has shown that the material orders antiferromag­
temperature-driven phase transformation, a series of measurements of
netically at 226 K. Below the Néel point, the random and competing
the field dependences of magnetization at different temperatures has
superexchange interactions between the iron and manganese ions result
been carried out (Figs. S4, S5). Analysis of the remanent magnetization
in formation of a cluster glass state coexisting with the long-range
and coercive field of the Bi0.85Sr0.15Fe0.65Mn0.35O3 sample as functions
antiferromagnetism over a broad temperature range [41]. The obser­
of temperature reveals the anomalous decrease in the mentioned pa­
vation of a bifurcation between ZFC and FC dependences with a low-
rameters with decreasing temperature (Fig. 10 (a, b)). The anomaly
temperature maximum in ZFC curves (which, in contrast to the
appears below 350 K (where the sample exhibits a weak ferromagnetic
Bi0.85Sr0.15Fe1-xMnxO3 (0.1 ≤ x ≤ 0.5) samples (Figs. 6, 7), is seen near
behaviour) and persists down to 250 K (where the sample is character­
an irreversibility temperature) has recently been reported for the
ized by a close to linear field dependence of magnetization with a
Bi0.85Ca0.15Fe1-xMnxO3 (0.1 ≤ x ≤ 0.5) compounds [16]. The neutron
vanishingly small remanent magnetization expected for a collinear
diffraction and magnetization measurements indicate a strong reduction

6
V.A. Khomchenko et al. Journal of Magnetism and Magnetic Materials 585 (2023) 171142

Fig. 10. Temperature dependencies of remanent magnetization (upper row) and coercive field (lower row) obtained for the Bi0.85Sr0.15Fe1-xMnxO3 [x = 0.35 (a, b)
and x = 0.4 (c, d)] samples.

The temperature dependences of the remanent magnetization and


coercive field obtained for the sample with x = 0.4 (Fig. 10 (c, d)) are in
many aspects similar to those specific to the compound with x = 0.35
(Fig. 10 (a, b)). The main difference is the shift of the anomalous region
to a lower temperature (the decrease in the remanent magnetization
appears below 150 K) and the existence of second anomaly (a small
minimum seen at 275 K). Taking into account the two-phase character of
this sample and the non-monotonic compositional variation of the Néel
temperature in the 0.35 ≤ x ≤ 0.5 series (Fig. 6), the latter anomaly can
be attributed to the onset of the magnetic ordering in the orthorhombic
phase.
The recent neutron diffraction study of the Bi0.85Ca0.15Fe0.6Mn0.4O3
compound has found that the material undergoes the magnetic phase
transition associated with reorientation of antiferromagnetically
coupled magnetic moments from the c to a axis with decreasing tem­
perature [15]. Such a reorientation is accompanied by the appearance of
a small net magnetization due to spin canting [51]. The magnetization
measurements conducted for the Bi0.85Sr0.15Fe1-xMnxO3 samples with x
= 0.35 and x = 0.4 reveal the features implying the opposite scenario of
the temperature-driven changes in the magnetic anisotropy resulting in
a spin reorientation. In contrast to the Bi0.85Ca0.15Fe1-xMnxO3 com­
pounds, in which the Mn substitution modifies the cycloidal spin mod­
ulation towards a collinear antiferromagnetic order with spins aligned
Fig. 11. Field dependencies of magnetization measured for the Bi0.85Sr0.15 along the polar c axis at room temperature [16], the Sr/Mn-doping
Fe1-xMnxO3 (0.1 ≤ x ≤ 0.5) compounds at T = 2 K. The inset shows the enlarged seems to favour the composition-driven cycloidal to canted antiferro­
part of the main figure for a higher clarity. magnetic structure (in which spins should be oriented perpendicular to
the c axis) [51] transformation in the R3c phase. The complex
antiferromagnet). A further temperature decrease results in a progress­ temperature-dependent behaviour of magnetization found for the Sr/
ing increase in the remanent magnetization, which is likely related to the Mn-substituted samples near the polar/antipolar phase boundary
increasing contribution of a glassy phase responsible for the unsaturated (Figs. 6 (b, c), 10) suggests that the canted antiferromagnetic order is
character of the M vs. T and M vs. H dependences at low temperatures modified as the temperature decreases. Contrary to the reorientational
(Figs. 6, 7, 11) and the large magnetization observed for the Mn- transition found in the Bi0.85Ca0.15Fe0.6Mn0.4O3 compound [15], the
enriched samples in high magnetic fields (Fig. 11) [15,16]. temperature-induced magnetic structure transformation specific to the

7
­
V.A. Khomchenko et al. Journal of Magnetism and Magnetic Materials 585 (2023) 171142

Sr/Mn-doped samples is observed at relatively high temperatures [7] I. Sosnowska, T. Peterlin-Neumaier, E. Steichele, J. Phys. C Solid State Phys. 15
(1982) 4835.
(where contribution to the magnetic behaviour from a glassy phase is
[8] R.D. Johnson, P. Barone, A. Bombardi, R.J. Bean, S. Picozzi, P.G. Radaelli, Y.S. Oh,
small) and, therefore, is not accompanied by any pronounced meta­ S.-W. Cheong, L.C. Chapon, Phys. Rev. Lett. 110 (2013), 217206.
magnetic anomalies (Figs. S4, S5, 11). [9] S. Kawachi, A. Miyake, T. Ito, S.E. Dissanayake, M. Matsuda, W. Ratcliff II, Z. Xu,
Y. Zhao, S. Miyahara, N. Furukawa, M. Tokunaga, Phys. Rev. Mater. 1 (2017),
024408.
4. Conclusions [10] I.O. Troyanchuk, D.V. Karpinsky, M.V. Bushinsky, V.A. Khomchenko, G.N. Kakazei,
J.P. Araujo, M. Tovar, V. Sikolenko, V. Efimov, A.L. Kholkin, Phys. Rev. B 83
(2011), 054109.
A study of the crystal structure, microstructure and local ferroelectric
[11] V.A. Khomchenko, M.S. Ivanov, D.V. Karpinsky, S.V. Dubkov, M.V. Silibin, J.
and magnetic properties of the Bi0.85Sr0.15Fe1-xMnxO3 (0.1 ≤ x ≤ 0.5) A. Paixão, Mater. Lett. 235 (2019) 46.
compounds has been performed. The XRD, SEM and PFM measurements [12] V.A. Khomchenko, D.V. Karpinsky, S.I. Latushka, A. Franz, V.V. Sikolenko, S.
confirm that the materials are impurity-free and indicate that the in­ V. Dubkov, M.V. Silibin, J.A. Paixão, J. Mater. Chem. C 7 (2019) 6085.
[13] A.M. Kadomtseva, A.K. Zvezdin, Y.F. Popov, A.P. Pyatakov, G.P. Vorob’ev, JETP
crease in Mn concentration gives rise to the polar to antipolar (R3c → Lett. 79 (2004) 571.
Pnam) structural phase transition at x≈ 0.4. The magnetization mea­ [14] C. Weingart, N. Spaldin, E. Bousquet, Phys. Rev. B 86 (2012), 094413.
surements reveal the decrease in the Néel temperature with increasing [15] V.A. Khomchenko, M.V. Silibin, M.V. Bushinsky, S.I. Latushka, D.V. Karpinsky,
J. Phys. Chem. Solid 146 (2020), 109612.
Mn content and demonstrate that the combined Sr/Mn substitution [16] V.A. Khomchenko, M. Das, J.A. Paixão, M.V. Silibin, D.V. Karpinsky, J. Alloy.
tends to supress the cycloidal antiferromagnetic arrangement specific to Compd. 901 (2022), 163682.
the parent BiFeO3 in favour of the canted antiferromagnetic structure [17] L.H. Yin, Y.P. Sun, F.H. Zhang, W.B. Wu, X. Luo, X.B. Zhu, Z.R. Yang, J.M. Dai, W.
H. Song, R.L. Zhang, J. Alloy. Compd. 488 (2009) 254.
typical neither of the BiFe1-xMnxO3 nor of the Bi0.85Ca0.15Fe1-xMnxO3 [18] D.V. Karpinsky, M.V. Silibin, A.V. Trukhanov, A.L. Zhaludkevich, T. Maniecki,
multiferroics near their polar/antipolar (nonpolar) phase boundaries (at W. Maniukiewicz, V. Sikolenko, J.A. Paixão, V.A. Khomchenko, J. Phys. Chem.
room temperature). A temperature lowering can modify this structure, Solid 126 (2019) 164.
[19] J. Rodríguez-Carvajal, Phys. B 192 (1993) 55.
as evidenced by the observation of the anomalous decrease in the
[20] F. Kubel, H. Schmid, Acta Cryst. B 46 (1990) 698.
remanent magnetization and coercive field for the samples with x = 0.35 [21] S. Hussain, S.K. Hasanain, G.H. Jaffari, N.Z. Ali, M. Siddique, S.I. Shah, J. Alloy.
and x = 0.4 implying the existence of the temperature decrease-driven Compd. 622 (2015) 8.
[22] S. Chauhan, M. Kumar, P. Pal, RSC Adv. 6 (2016) 68028.
reorientation of the G-type-ordered magnetic moments from the a to c
[23] F. Pedro-García, A.M. Bolarín-Miró, F. Sánchez-De Jesús, C.A. Cortés-Escobedo,
axis. Z. Valdez-Nava, G. Torres-Villaseñor, Ceram. Int. 44 (2018) 8087.
[24] S.K. Mandal, P. Kiran, P.S. Rao, A. Chandra, J. Alloy. Compd. 922 (2022), 166107.
[25] A.M. Glazer, Acta Cryst. A 31 (1975) 756.
CRediT authorship contribution statement
[26] D.V. Karpinsky, M.V. Silibin, S.I. Latushka, D.V. Zhaludkevich, V.V. Sikolenko,
H. Al-Ghamdi, A.H. Almuqrin, M.I. Sayyed, A.A. Belik, Nanomaterials 12 (2022)
V.A. Khomchenko: Conceptualization, Investigation, Formal anal­ 1565.
[27] A.M. Glazer, K. Roleder, J. Dec, Acta Cryst. B 49 (1993) 846.
ysis, Visualization, Writing – original draft, Writing – review & editing.
[28] S. Karimi, I.M. Reaney, Y. Han, J. Pokorny, I. Sterianou, J. Mater. Sci. 44 (2009)
M. Das: Investigation, Formal analysis. M.S.C. Henriques: Methodol­ 5102.
ogy, Investigation. J.A. Paixão: Methodology, Investigation, Validation, [29] V.A. Khomchenko, I.O. Troyanchuk, M.V. Bushinsky, O.S. Mantytskaya,
Resources, Funding acquisition. V. Sikolenko, J.A. Paixão, Mater. Lett. 65 (2011) 1970.
[30] M. Kubota, K. Oka, Y. Nakamura, H. Yabuta, K. Miura, Y. Shimakawa, M. Azuma,
Jpn. J. Appl. Phys. 50 (2011) 09NE08.
[31] V.A. Khomchenko, I.O. Troyanchuk, M.I. Kovetskaya, J.A. Paixa͂ o, J. Appl. Phys.
Declaration of Competing Interest
111 (2012), 014110.
[32] V. A. Khomchenko, I. O. Troyanchuk, M. I. Kovetskaya, M. Kopcewicz, and J. A.
The authors declare that they have no known competing financial Paixa͂ o, J. Phys. D: Appl. Phys. 45, 045302 (2012).
interests or personal relationships that could have appeared to influence [33] V.A. Khomchenko, L.C.J. Pereira, J.A. Paixão, J. Mater. Sci. 50 (2015) 1740.
[34] K.R. Poeppelmeier, M.E. Leonowicz, J.C. Scanlon, J.M. Longo, W.B. Yelon, J. Solid
the work reported in this paper. State Chem. 45 (1982) 71.
[35] W. Ma, Q. Sun, M. Sun, L. Bai, Y. Liu, J. Zhang, J. Yang, Appl. Surf. Sci. 571 (2022),
Data availability 151130.
[36] M. Salah, M.M. El-Desoky, H.E. Ali, I. Morad, J. Magn. Magn. Mater. 557 (2022),
169511.
Data will be made available on request. [37] V.V. Pokatilov, V.S. Pokatilov, A.S. Sigov, V.M. Cherepanov, Inorg. Mater. 45
(2009) 683.
[38] K. Sardar, J. Hong, G. Catalan, P.K. Biswas, M.R. Lees, R.I. Walton, J.F. Scott, S.A.
Acknowledgments T. Redfern, J. Phys. Condens. Matter 24 (2012), 045905.
[39] A.T. Kozakov, A.G. Kochur, V.I. Torgashev, K.A. Googlev, S.P. Kubrin, V.
This work was supported by national funds from FCT – Fundação G. Trotsenko, A.A. Bush, A.V. Nikolskii, J. Alloy. Compd. 664 (2016) 392.
[40] I.O. Troyanchuk, D.V. Karpinskii, M.V. Bushinskii, A.N. Chobot, N.V. Pushkarev,
para a Ciência e a Tecnologia, I.P., within the projects UIDB/04564/ O. Prohnenko, M. Kopcewicz, R. Szymczak, Crystallogr. Rep. 54 (2009) 1172.
2020 and UIDP/04564/2020. Access to TAIL-UC facility funded under [41] P. Mandal, S.S. Bhat, Y. Sundarayya, A. Sundaresan, C.N.R. Rao, V. Caignaert,
QREN-Mais Centro project ICT_2009_02_012_1890 is gratefully B. Raveau, E. Suard, RSC Adv. 2 (2012) 292.
[42] D. Kothari, V.R. Reddy, A. Gupta, D.M. Phase, N. Lakshmi, S.K. Deshpande, A.
acknowledged. M. Awasthi, J. Phys. Condens. Matter 19 (2007), 136202.
[43] J. Wei, D. Xue, Electrochem. Solid St. 10 (2007) G85.
Appendix A. Supplementary data [44] R.D. Shannon, Acta Cryst. A 32 (1976) 751.
[45] F.E.N. Ramirez, G.A.C. Pasca, J.A. Souza, Phys. Lett. A 379 (2015) 1579.
[46] V.A. Khomchenko, D.V. Karpinsky, J.A. Paixão, J. Mater. Chem. C 5 (2017) 3623.
Supplementary data to this article can be found online at https://doi. [47] I.O. Troyanchuk, M.V. Bushinsky, D.V. Karpinsky, V. Sirenko, V. Sikolenko,
org/10.1016/j.jmmm.2023.171142. V. Efimov, Eur. Phys. J. B 73 (2010) 375.
[48] V.A. Khomchenko, J.A. Paixão, J. Appl. Phys. 116 (2014), 214105.
[49] J.A. Schiemer, R.L. Withers, Y. Liu, M.A. Carpenter, Chem. Mater. 25 (2013) 4436.
References [50] V.A. Khomchenko, J.A. Paixão, J. Phys. Condens. Matter 28 (2016), 166004.
[51] C. Ederer, N.A. Spaldin, Phys. Rev. B 71 (2005), 060401.
[1] M. Fiebig, T. Lottermoser, D. Meier, M. Trassin, Nat. Rev. Mater. 1 (2016) 16046. [52] V.A. Khomchenko, I.O. Troyanchuk, D.V. Karpinsky, J.A. Paixão, J. Mater. Sci. 47
[2] N.A. Spaldin, R. Ramesh, Nat. Mater. 18 (2019) 203. (2012) 1578.
[3] B. Sun, G. Zhou, L. Sun, H. Zhao, Y. Chen, F. Yang, Y. Zhao, Q. Song, Nanoscale [53] R.D. Johnson, P.A. McClarty, D.D. Khalyavin, P. Manuel, P. Svedlindh, C.S. Knee,
Horiz. 6 (2021) 939. Phys. Rev. B 95 (2017), 054420.
[4] R. Gupta, R.K. Kotnala, J. Mater. Sci. 57 (2022) 12710. [54] I.O. Troyanchuk, D.V. Karpinsky, M.V. Bushinsky, O.S. Mantytskaya, N.
[5] W. Eerenstein, N.D. Mathur, J.F. Scott, Nature 442 (2006) 759. V. Tereshko, V.N. Shut, J. Am. Ceram. Soc. 94 (2011) 4502.
[6] N. Wang, X. Luo, L. Han, Z. Zhang, R. Zhang, H. Olin, Y. Yang, Nano-Micro Lett. 12 [55] A.A. Belik, A.M. Abakumov, A.A. Tsirlin, J. Hadermann, J. Kim, G. Van Tendeloo,
(2020) 81. E. Takayama-Muromachi, Chem. Mater. 23 (2011) 4505.

You might also like