Next Article in Journal
A Review of the Impact of Graphene Oxide on Cement Composites
Next Article in Special Issue
Impact of Ni Doping on the Microstructure and Mechanical Properties of TiB2 Films
Previous Article in Journal
Defect-Tolerant Memristor Crossbar Circuits for Local Learning Neural Networks
Previous Article in Special Issue
Interface Optimization and Thermal Conductivity of Cu/Diamond Composites by Spark Plasma Sintering Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improved Ferroelectric and Magnetic Properties of Bismuth Ferrite-Based Ceramics by Introduction of Non-Isovalent Ions and Grain Engineering

1
Guangdong Provincial Key Laboratory of Electronic Functional Materials and Devices, Huizhou University, Huizhou 516001, China
2
School of Mechanical and Electrical Engineering, Hebei Vocational College of Rail Transportation, Shijiazhuang 050801, China
3
School of Materials Science and Engineering, Harbin Institute of Technology (Shenzhen), Shenzhen 518055, China
4
School of Undergraduate Education, Shenzhen Polytechnic University, Shenzhen 518055, China
5
Department of Physics, Saveetha School of Engineering, Saveetha Institute of Medical and Technical Sciences (SIMATS), Thandalam, Chennai 602105, India
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(3), 215; https://doi.org/10.3390/nano15030215
Submission received: 31 December 2024 / Revised: 21 January 2025 / Accepted: 25 January 2025 / Published: 29 January 2025
(This article belongs to the Special Issue Design and Applications of Heterogeneous Nanostructured Materials)

Abstract

:
Single-phase multiferroics exhibiting ferroelectricity and ferromagnetism are considered pivotal for advancing next-generation multistate memories, spintronic devices, sensors, and logic devices. In this study, the magnetic and electric characteristics of bismuth ferrite (BiFeO3) ceramics were enhanced through compositional design and grain engineering. BiFeO3 ceramic was co-substituted by neodymium (Nd) and niobium (Nb), two non-isovalent elements, via the spark plasma sintering process using phase-pure powder prepared via sol-gel as the precursor. The symmetry of the sintered Nd–Nb co-doped samples changed from R3c to Pnma, accompanied by a decrease in the loss tangent, grain size, and leakage current density. The reduction in the leakage current density of the co-doped samples was ~three orders of magnitude. Moreover, ferroelectric, dielectric, and magnetic properties were substantially improved. The remanent polarization and magnetization values of the optimized Nd–Nb co-doped BiFeO3 sample were 3.12 μC cm−2 and 0.15 emu g−1, respectively. The multiferroic properties were enhanced based on multiple factors such as structural distortion caused by co-doping, grain size reduction, suppression of defect charges via donor doping, space-modulated spin structure disruption, and an increase in magnetic ions. The synergistic approach of composition design and grain engineering sets a paradigm for the advancement of multiferroic materials.

1. Introduction

Compounds with ≥2 primary ferroic orders (such as ferromagnetism, ferroelasticity, and ferroelectricity) exhibit multiferroic characteristics. These compounds have been technologically and economically important over the past two decades, and continue to be now and will be in the future [1]. The effective magnetoelectric coupling at room temperature has several applications in data storage systems, adjustable microwave components, sensors, and random-access memory devices [2,3,4]. Magnetoelectric multiferroic materials are of three types: (i) single-phase type I multiferroics with distinct magnetization and polarization origin, (ii) single-phase type II multiferroics with polarization induced by a particular magnetic configuration, and (iii) composite materials or heterostructures (piezoelectric multilayer and artificial ferromagnetic capacitors) with strain-mediated interactions modulating magnetic properties under an electric field (E) [1].
Bismuth ferrite (BiFeO3, BF) has emerged as the quintessential example of single-phase type I multiferroic materials at room temperature since 2003 owing to its substantial ferroelectric polarization and elevated phase transition temperatures (Curie temperature (TC) of ~825 °C and Néel temperature (TN) of ~370 °C) [5]. It exhibits a distorted rhombohedral perovskite structure belonging to the R3c space group with lattice parameter values of length (a) and angle (α) of 3.958 Å and 89.30°, respectively [6]. It is characterized by better ferroelectric properties than lead-free materials such as bismuth titanate (Bi4Ti3O12) [7], with simultaneous antiferromagnetic and ferroelectric characteristics at ambient temperature [6]. BF crystals exhibit ferromagnetism and ferroelectricity owing to the partially filled 3D orbits of ferric ions (Fe3+) and the stereochemically active lone-pair electrons in the 6s2 orbitals of bismuth ions (Bi3+) [8]. The theoretical ferroelectric polarization value of BF in the [111] direction is in the range of 90–100 μC cm−2 [9], while the magnetoelectric effect is prominent with the coupling energy of ~1.25 meV [10]. Although exceptional electrical properties of BF are predicted theoretically, it is practically challenging owing to the high leakage current density (J). This is attributed to oxygen vacancy, Bi evaporation, the formation of the frequent secondary phase, the presence of mixed Fe3+ and ferrous (Fe2+) ions, and weak ferromagnetism, which substantially limit its application in various sophisticated devices [10]. The central challenge for BiFeO3-based multiferroic ceramics has been to concurrently enhance ferroelectricity and ferromagnetism, thereby improving the magnetoelectric coupling.
Chemical substitution at the A-, B-, or A/B-sites of BF is an effective method to increase its ferroelectric properties by reducing leakage current [1,2,3,4,11]. Lakshmi et al. [12] reported a substantial reduction in the leakage current of BF ceramics doped with niobium (Nb) of high valency. The doping of Nb could alter the Fe-O-Fe bond angle, potentially disrupting the spin cycloid structure. Abe et al. [13] examined the effects of cation substitution with different valences in BiFeO3 ceramics and observed a marked reduction in leakage current by substituting BF with 10% titanium ion (Ti4+), which was attributed to a decrease in oxygen vacancies. Yamamoto et al. [14] identified a transition from a cycloidal spin structure to a collinear spin structure of BF polycrystalline ceramics substituted by cobalt (Co), which induced spontaneous magnetization at room temperature owing to the spin-canting phenomenon occurring in collinear structures. The doping of A-sites with rare-earth ions is a beneficial strategy. Studies indicated that the crystal structure was stabilized via the partial replacement of the Bi3+ ions of BF with rare-earth ions by enhancing ionic bond strength, which suppressed secondary phases and oxygen vacancies primarily formed by the evaporation of Bi. Moreover, magnetocrystalline anisotropy increased, which was responsible for the energetically unfavorable spin cycloid structure [15,16,17,18,19,20]. Recently, the symmetry modulation of R3c to Pna21 and eventually to Pbnm with the increasing substitution of lanthanum (La)- [19], neodymium (Nd)- [16], samarium (Sm)- [20], gadolinium (Gd)- [18], and La/lutetium (Lu) [17] was developed as an effective method to improve the multiferroic characteristics of BF ceramics. The emergence of the Pna21 polar phase under chemical pressure is a determining factor in this process. Quan et al. [15] reported the improved magnetic properties of Bi0.8Ce0.2FeO3, which were attributed to the high canting angle and repressed cycloidal spin structure of the cerium (Ce)-doped sample. Nonzero magnetization was observed in La-substituted BF [21], a phenomenon owing to the chemical compression-induced rotation of iron oxide (FeO6) octahedra. In the FeO6 octahedra, each Fe3+ is surrounded by six O2− and the two nearest neighboring Fe3+ are connected by an O2−, where the Fe-3d energy level and O-2p energy level hybridizing contributes to double-exchange interaction. The Fe-O-Fe bond angle changed with the doping La element, which was helpful for the exchange interaction by enhancing the overlap of the Fe 3d-O 2p orbitals. Thus, these modified the symmetry structure and suppressed the spin cycloid. Thus, the presence of rare-earth ions alters chemical pressure and disrupts the spin cycloid, progressively enhancing magnetic properties.
Hence, it was anticipated that co-doping the Fe and Bi sites of BF might decrease J while simultaneously enhancing magnetic and ferroelectric properties. Ke et al. [22] co-doped BF ceramics using Nd and titanium (Ti) to obtain structures with morphotropic phase boundaries (MPB) using a solid-state sintering method. The synthesized samples exhibited improved magnetic and ferroelectric properties at the boundaries with the highest remanent magnetization (Mr) and polarization (Pr) values of 0.27 emu g−1 and 29 μC cm−2, respectively. Gumiel et al. [23] reported that antiferromagnetism could be converted to weak ferromagnetism via co-doping (Nd-Ti), which also decreased electric conductivity, increased the electric polarization, and stabilized the converted state. Wang et al. [24] co-doped BF ceramics with Sm and Nb via spark plasma sintering and the sol-gel process. The doped samples exhibited reduced J with enhanced magnetic and ferroelectric properties. Gao et al. [25] reported the synthesis of polar nanoregions by co-doping BF with Sm and scandium (Sc), which increased the dielectric breakdown field. Nonetheless, reports on the improved magnetic and ferroelectric properties and the leakage current in BiFeO3 ceramics co-doped with Nd and high-valence Nb are scarce.
On the other hand, the technique used to fabricate ceramics significantly influences their macroscopic multiferroic characteristics. Several methods, such as rapid liquid-phase sintering [26], the sol-gel method [27], electric current activated sintering [28], flash sintering [29], and microwave sintering [30] were used to prepare single-phase BF. The spark plasma sintering (SPS) method is considered a more efficient technique than the aforementioned techniques for synthesizing materials with micro-/nanostructures that can be modulated. The pulsed electric current and uniaxial stress generated during SPS facilitate rapid sintering and effective heat transfer, inducing enhanced ceramic compaction within a short duration [31]. It is characterized by low sintering temperatures and less holding time, which inhibit Bi volatilization and the valence state change of Fe. SPS was used to successfully fabricate highly dense BF and BF-based ceramics with a pure phase and enhanced multiferroic and dielectric properties [24,31,32]. In this context, the synergistic impact of co-substitution in ferroelectric ceramics can be beneficial for enhancing physical properties. This inspired us to use SPS to prepare nanocrystalline BiFeO3 ceramics co-doped with rare-earth Nd3+ and high-valence Nb5+.
In this study, inspired by the concept of the grain size and ionic radius effect, the proposed strategy involves the co-substitution of non-isovalent ions (Nd and Nb) at the A/B sites via the sol-gel method followed by SPS to enhance the ferroelectric and magnetic characteristics of BF ceramics. The Bi1−xNdxFe0.95Nb0.05O3 (x = 0, 0.05, 0.10, 0.15, and 0.20) ceramic samples were sintered using phase-pure sol-gel-derived precursor powders. The impact of Nd–Nb co-doping on the ferroelectric, structural, dielectric, and magnetic characteristics of the synthesized samples was thoroughly investigated.

2. Experimental Procedure

BF and Bi1−xNdxFe0.95Nb0.05O3 (x = 0.05, 0.10, 0.15, and 0.20) nanopowder samples in a pure phase were synthesized using the sol-gel method, as documented in our earlier studies [33,34]. These as-synthesized powder samples were loaded into a 10 mm-diameter graphite mode and sintered via SPS. SPS was performed following the reported procedure [24,35]. The pellet was sintered at 700 °C, with a heating and cooling rate of 100 °C/min. The temperature was held there for 5 min under a vacuum of 10−2 Pa. A uniaxial and stable stress of 50 MPa was maintained both in the heating and soaking processes. The Bi1−xNdxFe0.95Nb0.05O3 samples with x values of 0.05, 0.10, 0.15, and 0.20 were denoted as 5NdNb, 10NdNb, 15NdNb, and 20NdNb, respectively.
The phase composition of the sintered ceramic samples via SPS was examined via X-ray diffraction (XRD). The microstructures of the fractured surfaces of the ceramic samples were analyzed using a field emission–scanning electron microscope (FE-SEM, model Hitachi S-4700). The samples were ground to obtain 1 mm thickness, coated with silver electrodes, and subsequently annealed for 2 h at 120 °C to analyze the dielectric properties using a computerized impedance analyzer (PSM1735, Newton 4th Ltd., UK). Polarization versus electric field (PE) and J were measured using a ferroelectric testing system (Premier-II by Radiant Technologies, Inc.). Magnetic properties were analyzed using a physical property measurement system (PPMS, DynaCool-9T, by QUANTUM DESIGN). All experimental data were recorded at ambient temperature.

3. Results and Discussion

Figure 1a presents the XRD patterns of pristine and Nd–Nb co-substituted BiFeO3 ceramic samples. Phase-pure BF ceramics were effectively synthesized via SPS in conjunction with the powder samples obtained via the sol-gel method [24]. The diffraction peaks of (104), (110), and (113) were indexed to a characteristic perovskite structure distorted rhombohedrally with the R3c space group [24]. Impure peaks in the doped ceramics were absent, and the sharpness of the diffraction peaks indicates high crystallinity. No additional peaks were observed for ceramics up to a co-doped level of 0.10, suggesting R3c as the primary phase structure. The multiple peaks at 2θ of ~32°, 40°, 52°, and 57° merged to form a broad peak, indicating the successful doping of Nd and Nb in BF [36]. The magnified patterns at 2θ of ~32° presented in Figure 1b show the overlapping of the split peaks (104) and (110). Moreover, the peaks shifted to high diffraction angles, indicating a reduction in the unit cell size and lattice constant owing to the smaller ionic radius of Nd3+ than that of Bi3+ and the similar ionic radii of Nb5+ and Fe3+ (Nb5+ = 0.64 Å, Nd3+ = 0.983 Å, Fe3+ = 0.645 Å, and Bi3+ = 1.173 Å) [34,37]. This observation further confirms the successful integration of Nb and Nd within BF. A few peaks of 15NdNb and 20NdNb re-split to form new peaks, particularly for 20NdNb, suggesting the presence of a mixture of orthorhombic Pnma and rhombohedral R3c phases, conforming to the reported results [37]. Therefore, it can be inferred that Bi and Fe substituted by Nd and Nb, respectively, substantially altered the crystal structure and lattice parameters of BF.
The quantitative correlation between structural transition and doping concentration was elucidated via Rietveld refinement using the General Structure Analysis System (GSAS) program coupled with the graphical user interface (EXPGUI). Figure 1c presents the Rietveld refinement of the XRD patterns of 10NdNb. The refined lattice parameters, phase constitution, and R-factors of the samples are listed in Table 1. The minimal discrepancy between the observed and calculated data, along with GOF (goodness of fit, Rwp/Rexp) values ranging from 1.3 to 1.6, confirms the success of the refinement process. The determined a and c of the pristine BF ceramics were 5.571 and 13.872 Å, respectively, which align with the R3c space group (JCPDS No. 86-1518) [24]. The refinement considered two phases for Nd–Nb co-doped BF, such as R3c + Pbnm, R3c + Pbam, or R3c + Pnma. The R3c + Pnma model yielded the most accurate fit. The R3c fraction decreased from 97.45% to 75.12% upon Nd–Nb co-doping. Lattice parameters a and c exhibited a steady increase and decrease, respectively, which can be ascribed to the different ionic radii of Fe, Bi, Nd, and Nb and the change in bonding angle. This explains the observed shift in diffraction peaks to large angles in Figure 1b (for samples: BFO, 5NdNb, 10NdNb, and 15NdNb).
To examine the surface morphology and grain size variations in the ceramics, SEM was utilized. Figure 2 illustrates the distinct grain structure of Nd–Nb co-doped BF samples. Analytical software (Nano Measurer 1.2, Department of Chemistry, Fudan University, Fudan, China) was employed to calculate the grain-size distributions of the as-prepared ceramics. We reported earlier that undoped BFO comprises large polyhedral grains with sizes of ~1–3 μm [24]. The obtained average grain size of BFO samples was about 2.16 μm (Figure 2a). A few pores existed at grain boundaries owing to oxygen vacancies and Bi volatilization [38]. In contrast, the Nd–Nb co-doped samples exhibited flatter boundary facets with no evident pores and a uniform distribution. The gradual reduction in grain size with increasing Nd–Nb co-doping is evident in Figure 2b–e. The grain size diminished with the increase in the co-doped content, reaching an average size <300 nm of 20NdNb, indicating that Nd–Nb co-doping effectively inhibits grain growth. It is known that the presence of vacancies in oxide materials is beneficial for ion transport during sintering and thus generates larger grains for ceramic. The marked decrease in grain size can be attributed to the low diffusivity of Nd3+ and the suppression of Bi volatilization and oxygen vacancy formation because of Nd–Nb co-doping [39]. Figure 2g,h illustrates the surface scan results, showing even distribution of elements within the scanned area without non-uniform aggregation or segregation. Generally, the microstructure significantly impacts the macroscopic properties, as documented in studies on ferroelectric oxides [1,2,40]. The influence of grain size on magnetic properties in BiFeO3 ceramics has been noted in single-phase BiFeO3 [41]. Consequently, with the pronounced changes in surface morphology induced by Nd–Nb co-doping, the multiferroic characteristics and electronic structure of Nd–Nb co-doped BF differ from those of the pristine BF, which is discussed in the subsequent section.
Figure 3 depicts the change in dielectric behavior (dielectric constant and dielectric loss) of pristine BF and Nd–Nb co-doped BF with temperature at 1 MHz frequency. The dielectric constant was independent of temperature below 200 °C, signifying the stability of the ferroelectric order. This stability is attributed to the high ferroelectric TC and the unique micro- and electronic structures of BF [42]. The effect of lattice vibrations on electron movement diminished at high frequencies, leading to a less pronounced variation in dielectric properties with temperature at these frequencies. The dielectric constant increased gradually with a further increase in temperature owing to thermally activated ionic conductivity caused by an increase in defects, such as oxygen vacancies, under higher temperatures [16,19]. The loss tangent of BF increased with increasing temperature, whereas the same for Nd–Nb co-doped BF remained relatively stable below 200 °C. The dielectric constant of Nd–Nb co-doped BF increased, while the loss tangent exhibited an opposite trend to those of BF, suggesting that the dielectric characteristics of Nd–Nb co-doped BF were altered. The dielectric constants of BF, 5NdNb, 10NdNb, 15NdNb, and 20NdNb at room temperature were 143, 176, 187, 211, and 132, respectively. The SEM images of Nd–Nb co-doped BF indicate that the grain sizes of Nd–Nb co-doped samples were reduced, which increased the grain boundary. Additionally, Nd–Nb co-doping could induce phase transition and hinder Bi volatilization. Thus, resistivity was increased and the oxygen vacancy concentration was decreased [26]. Consequently, the dielectric constant was enhanced, while the loss tangent was reduced. Although the loss tangent remained slightly high, the incorporation of other ABO3 materials and manganese oxide (MnO2), or by subjecting the samples to annealing under various atmospheric conditions, could potentially have reduced the dielectric loss [43].
The ferroelectric properties of the synthesized ceramics were investigated by measuring ferroelectric hysteresis (PE) loops at room temperature and 10 Hz using a ferroelectric analyzer [Figure 4a]. The Pr and coercivity field (Ec) values of the ceramics are presented in Figure 4b and Table 2. The results were directly compared based on the magnitude, as the measurements were conducted using consistent parameters. Figure 4a shows no inflation of polarization loops, although they were not completely saturated, indicating the production of high-quality ceramics and a marked reduction in the leaky behavior of BFO. BF ceramics exhibited unsaturated and substantially inflated PE loop area owing to a high J [31]. Furthermore, it is noteworthy that the ferroelectric polarization was highly dependent on the Nd–Nb concentration, with a substantial enhancement observed for the 10NdNb ceramics (see Figure 4a). Notably, Ec and Pr exhibited nonlinear behavior, as illustrated in Figure 4b. BF exhibited Pr and Ec values of ~0.41 μC cm−2 and ~4.45 kV cm−1, respectively. Nd–Nb doping substantially improved the shape of PE loops. The Pr changed with the doping content, exhibiting an optimal value of ~3.12 μC cm−2 for 10NdNb, which is considerably higher than that of BF (0.41 μC cm−2). Thus, Nd–Nb co-doping is an appropriate strategy for enhancing the ferroelectric characteristics of BF ceramics. The substitution of Bi and Fe sites of BF by Nd3+ and Nb5+ ions, respectively, altered the grain size and structural characteristics of the crystals, affecting the Ec and Pr of the materials.
In the doped samples, some Bi sites were replaced by Nd, and the lone-pair electrons of Bi3+ ions (s2 electrons) hybridized with the empty p orbitals of Bi3+ or O2− ions, forming a localized lobe. This phenomenon induces non-centrosymmetric distortion, thereby enhancing ferroelectricity. Nd–Nb co-doping is also responsible for changing lattice parameters. Consequently, the improved ferroelectricity was attributed to the decreased concentration of oxygen vacancy, and even the decrease in the amount of Fe2+, which helps to mitigate domain-pinning effects [44]. Additionally, the grain boundaries in the rhombic-phase ceramics impeded the movement of electric domains and polarization reversal, thereby enhancing ferroelectric polarization [42]. However, samples doped with highly concentrated dopants (x > 0.10) increased the Pnma phase. 10NdNb represents the optimal sample based on its position at the morphotropic-phase boundary [45]. The reduced ferroelectricity of 15NdNb might be attributed to several factors: (i) the stereochemically active 6s2 lone-pair electrons of Bi3+ and empty 6p orbits induce the ferroelectricity of BF, so when the doping content exceeds a certain value, Nd-doped Bi sites decrease the concentration of lone-pair electrons (s2 electrons) and deteriorate the stereochemical activity of the lone pair owing to the spherically distributed electron density of Nd3+ [36]; (ii) the highly centrosymmetric orthorhombic phase fails to exhibit ferroelectricity [31] and induce polarization easily; or (iii) a high Nd concentration indicates a low ability of the cations to displace within the crystal lattice, resulting in a low ion displacement and ferroelectric polarization [44]. Wu et al. [42] investigated Tb-doped BiFeO3 (x = 0.05 and 0.1) synthesized via an optimized sol-gel method and reported a maximum Pr of 6.9 μC/cm2. Islam et al. [38] prepared Bi0.8Ba0.2Fe0.9Ta0.1O3 ceramic through a solid-state reaction, achieving a Pr of approximately 0.15 μC/cm2. Zhang et al. [46] reported a Pr of 0.75 μC/cm2 for Bi0.95Dy0.05Fe0.95Mn0.05O3 ceramics prepared by a solid-state reaction. Hua et al. [47] reported that the Pr of Bi0.925Ho0.075Fe0.95Mn0.05O3 ceramic was measured to be 0.0948 μC/cm2. Our values are comparably favorable.
Figure 5a illustrates the J versus E characteristics of the as-sintered ceramics measured at ambient temperature. The plots exhibit symmetry under negative and positive biases. The J of BF, 5NdNb, 10NdNb, 15NdNb, and 20NdNb ceramics at an electric field of ±30 kV cm−1 were ~1.26 × 10−5, 3.39 × 10−7, 1.78 × 10−7, 5.82 × 10−8, and 9.51 × 10−8 A cm−2, respectively. Thus, Nd–Nb co-doped BF ceramics were remarkably lower J than BF under identical electric fields. The smallest J of Nd–Nb co-doped BF was ~three orders of magnitude lower than that of BF, suggesting superior insulating properties. This improvement can be ascribed to the reduced Bi volatility, charge defects (oxygen or Bi vacancies), and Fe2+ content [2]. The bond enthalpy values of Nb−O and Nd−O were higher than those of Fe−O and Bi−O, stabilizing the perovskite structure through the substitution of Bi and Fe by Nd and Nb, respectively, which also diminished Bi evaporation [39,48]. Moreover, the substitution of Nb5+ ions, which had a higher valence, for Fe3+ ions, helped to occupy oxygen vacancies. Consequently, the co-doping of Nd and Nb effectively diminished the leakage current density in these ceramic samples.
Five primary leakage mechanisms of perovskite ferroelectric materials have been identified: Schottky emission, space–charge limited current (SCLC), Poole–Frenkel emission, ohmic conduction, and Fowler–Nordheim (FN) tunneling [35]. Specifically, ohmic conduction and SCLC are the predominant leakage mechanisms of BF-based materials. Leakage current behavior can be modeled by applying the equation J SCLC E n . The leakage behavior follows SCLC and ohmic conduction mechanisms when exponent (n) values are 2 and 1, respectively. The leakage data were reanalyzed by plotting log J against log E, as depicted in Figure 5b. The value of n of BF was ~2, suggesting that leakage was predominantly owing to the SCLC mechanism. The slope (n) of Nd–Nb co-doped BF lay between 1 and 2, indicating combined SCLC and ohmic conduction mechanisms. Thus, the slope of BF decreased with the introduction of Nd–Nb, which implies a reduction in mobile charge carriers due to enhanced structural stability [44].
Figure 6 displays the magnetic hysteresis loops of the prepared ceramic samples at ambient temperature. The Mr and coercivity (Hc) data of each sample are presented in Figure 6b and Table 2. None of the loops reached saturation under the magnetic field of 50 kOe. The magnetic moment of BF was nearly negligible owing to its G-type antiferromagnetic character, similar to reported studies [49,50]. BF exhibited a nearly linear loop with a negligible Mr (~5.67 × 10−4 emu g−1), characteristic of antiferromagnetic materials, which is in agreement with other findings [49,50]. The loop centers of Nd–Nb co-doped BF expanded, indicating a shift towards weak ferromagnetism and a substantial increase in Mr, suggesting a transition from an antiferromagnetic state to a ferromagnetic state [24]. The Mr values of 5NdNb, 10NdNb, 15NdNb, and 20NdNb were ~0.07, 0.12, 0.15, and 0.14 emu g−1, respectively, which are over two orders of magnitude higher than those of BF. This enhancement is attributed to structural distortion. Nd–Nb co-doped BF ceramics induce a structural transformation with the coexistence of rhombohedral and orthorhombic phases. The equilibrium between antiparallel spin lattices of Fe3+ adjacent to each other is disrupted because of the partially substituted Bi3+ and Fe3+ by Nd3+ and Nb5+, respectively, with different ionic radii and valencies. This symmetry distortion typically disrupts the spatially non-uniform spin helical structure, altering bond lengths and angles, which in turn increases the spin tilt angle and results in a net macroscopic magnetization [15]. The symmetry-driven Dzyaloshinsky–Moriya (DM) interaction mechanism suggests that the tilted antiferromagnetic spin orders demonstrate enhanced magnetic characteristics [16]. Concurrently, the unique spin–helix long-range ordered structure in BiFeO3 is disrupted, liberating additional magnetic moments [51], which couples with increased polarization. The high coercivity is linked to the magnetic anisotropy induced by co-doping. As documented by Monem and colleagues [52], the peak Mr for Bi0.93Sr0.07Fe0.8Zr0.2O3 ceramic, synthesized via a tartaric acid-assisted sol-gel method, reaches approximately 3.67 emu/g. Yadav et al. [53] and Rao et al. [39] separately synthesized Bi0.9Nd0.1Fe1−xTixO3 and Bi0.85Sm0.15Fe0.9Sc0.1O3 ceramics through conventional solid-state reactions, obtaining Mr values of about 3.15 and 0.2 emu/g, respectively. Evidently, the current findings are in good agreement with these studies. Without a doubt, the current results suggest that the approach of compositional design coupled with grain engineering is a promising strategy for enhancing the multiferroic properties of BiFeO3 at room temperature.

4. Conclusions

Phase-pure multiferroic Nd–Nb co-doped BF ceramics were successfully synthesized via spark plasma sintering. The phase composition of BF ceramics was altered from an exclusive R3c phase to a mixed phase comprising R3c and Pbnm phases via Nd–Nb co-doping. The reduced-loss tangent and J were correlated to the decrease in the defect concentration and refined grain size. Nd–Nb co-doping enhanced the ferroelectric, dielectric, and magnetic properties. The primary leakage mechanism of the co-doped sample was based on combined SCLC and ohmic conduction. The Bi0.9Nd0.1Fe0.95Nb0.05O3 ceramic sample exhibited optimal values of Pr and moderate Mr at ~3.12 μC cm−2 and 0.12 emu g−1, respectively. The improvement in ferromagnetic and ferroelectric properties of the Nd–Nb co-doped BF samples was attributed to the crystal structure, decreased leakage current, and the presence of a disrupted spiral spin configuration. These findings may offer fresh insights for advancing the multiferroic properties of BiFeO3 ceramics through compositional optimization and grain size manipulation, potentially paving the way for significant practical applications.

Author Contributions

Methodology, H.Y.; investigation, Z.Z. and M.W.; data curation, X.W., Y.C., and H.M.; writing—original draft, T.W. and P.R.; writing—review and editing, T.W.; supervision, S.S. and Q.M.; project administration, Q.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Guangdong Basic and Applied Basic Research Foundation (grant No. 2022A1515140002 and 2023A1515140074), the College Students’ Innovative Entrepreneurial Training Plan Program (S202310577064), the Project of General Colleges and Universities in Guangdong Province (grant No. 2021ZDJS079), the Program for Innovative Research Team of Guangdong Province & Huizhou University (IRTHZU), the Shenzhen Science and Technology Program (No. JCYJ20220531094612029), and the National Natural Science Foundation of China (grant No. 62204162). The authors would like to express their gratitude to MJEditor (www.mjeditor.com, accessed on 30 December 2024) for the expert linguistic services.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hu, Z.; Stenning, G.B.G.; Zhang, H.; Shi, Y.; Koval, V.; Hu, W.; Zhou, Z.; Jia, C.; Abrahams, I.; Yan, H. Non-volatile voltage-controlled magnetization in single-phase multiferroic ceramics at room temperature. J. Mater. 2025, 11, 100857. [Google Scholar] [CrossRef]
  2. Sun, Y.; Liu, J.; Sun, T.; Yu, Z.; Zheng, Z.; Ge, M.; Bai, L. Structural evolution and enhanced multiferroicity in BiFeO3-based ceramics via rare earth element co-substitution. J. Eur. Ceram. Soc. 2025, 45, 117021. [Google Scholar] [CrossRef]
  3. Billah, A.; Matsuno, Y.; Anju, A.N.; Koike, K.; Kubota, S.; Hirose, F.; Ahmmad, B. Unusual behavior of magnetic coercive fields with temperature and applied field in La-doped BiFeO3 ceramics. ACS Appl. Electron. Mater. 2023, 5, 4261–4267. [Google Scholar] [CrossRef]
  4. Liu, L.; Zhu, X.L.; Chen, X.M. Improved ferroelectricity and magnetoelectric coupling effect of multielements co-substituted BiFeO3 ceramics. J. Am. Ceram. Soc. 2023, 107, 933–944. [Google Scholar] [CrossRef]
  5. Selbach, S.M.; Tybell, T.; Einarsrud, M.A.; Grande, T. The ferroic phase transitions of BiFeO3. Adv. Mater. 2008, 20, 3692–3696. [Google Scholar] [CrossRef]
  6. Catalan, G.; Scott, J.F. Physics and applications of bismuth ferrite. Adv. Mater. 2009, 21, 2463–2485. [Google Scholar] [CrossRef]
  7. Gao, X.; Zhang, L.; Wang, C.; Shen, Q. Enhanced magnetic and ferroelectric properties of textured multiferroic Bi5Ti3FeO15-CoFe2O4 ceramics prepared by plasma activated sintering. Ceram. Int. 2017, 43, 6607–6611. [Google Scholar] [CrossRef]
  8. Fiebig, M. Revival of the magnetoelectric effect. J. Phys. D Appl. Phys. 2005, 38, R123–R152. [Google Scholar] [CrossRef]
  9. Teague, J.R.; Gerson, R.; James, W.J. Dielectric hysteresis in single crystal BiFeO3. Solid. State Commun. 1970, 8, 1073–1074. [Google Scholar] [CrossRef]
  10. Shi, X.X.; Liu, X.Q.; Chen, X.M. Readdressing of magnetoelectric effect in bulk BiFeO3. Adv. Funct. Mater. 2017, 27, 1604037. [Google Scholar] [CrossRef]
  11. Boukhari, M.; Abdelkafi, Z.; Abdelmoula, N.; Khemakhem, H.; Randrianantoandro, N. Enhanced dielectric and optical properties in Zn2+ and Zr4+ co-doping BiFeO3 ceramic. J. Mater. Sci. Mater. Electron. 2023, 34, 1218. [Google Scholar] [CrossRef]
  12. Lakshmi, S.D.; Banu, I.B.S. Multiferroism and magnetoelectric coupling in single-phase Yb and X (X = Nb, Mn, Mo) co-doped BiFeO3 ceramics. J. Sol-Gel Sci. Technol. 2018, 89, 713–721. [Google Scholar] [CrossRef]
  13. Abe, K.; Sakai, N.; Takahashi, J.; Itoh, H.; Adachi, N.; Ota, T. Leakage current properties of cation-substituted BiFeO3 ceramics. Jpn. J. Appl. Phys. 2010, 49, 09MB01. [Google Scholar] [CrossRef]
  14. Yamamoto, H.; Kihara, T.; Oka, K.; Tokunaga, M.; Mibu, K.; Azuma, M. Spin structure change in co-substituted BiFeO3. J. Phys. Soc. Jpn. 2016, 85, 064704. [Google Scholar] [CrossRef]
  15. Quan, Z.; Liu, W.; Hu, H.; Xu, S.; Sebo, B.; Fang, G.; Li, M.; Zhao, X. Microstructure, electrical and magnetic properties of Ce-doped BiFeO3 thin films. J. Appl. Phys. 2008, 104, 084106. [Google Scholar] [CrossRef]
  16. Chen, J.; Xu, B.; Liu, X.Q.; Gao, T.T.; Bellaiche, L.; Chen, X.M. Symmetry modulation and enhanced multiferroic characteristics in Bi1−xNdxFeO3 ceramics. Adv. Funct. Mater. 2019, 29, 1806399. [Google Scholar] [CrossRef]
  17. Chen, J.; Liu, L.; Zhu, X.L.; Gareeva, Z.; Zvezdin, A.; Chen, X.M. The involvement of Pna21 phase in the multiferroic characteristics of La/Lu co-substituted BiFeO3 ceramics. Appl. Phys. Lett. 2021, 119, 112901. [Google Scholar] [CrossRef]
  18. Chen, J.; Su, N.; Zhu, X.L.; Liu, X.Q.; Zhang, J.-X.; Chen, X.M. Electric-field-controlled magnetism due to field-induced transition of Pna21/R3c in Bi1−xGdxFeO3 ceramics. J. Mater. 2021, 7, 967–975. [Google Scholar] [CrossRef]
  19. Liu, L.; Zhu, X.L.; Gareeva, Z.; Zvezdin, A.; Eiras, J.A.; Chen, X.M. Symmetry evolution and modulation of multiferroic characteristics in Bi1-xLaxFeO3 ceramics. Appl. Phys. Lett. 2022, 120, 13904. [Google Scholar] [CrossRef]
  20. Liu, L.; Chen, J.; Zhu, X.L.; Chen, X.M. Structure evidence of Pna21 phase and field-induced transition of Pna21/R3c in Bi1−xSmxFe0.99Ti0.01O3 ceramics. Appl. Phys. Lett. 2021, 118, 142904. [Google Scholar] [CrossRef]
  21. Petkov, V.; Zafar, A.; Kenesei, P.; Shastri, S. Chemical compression and ferroic orders in La substituted BiFeO3. Phys. Rev. Mater. 2023, 7, 054404. [Google Scholar] [CrossRef]
  22. Ke, H.; Zhang, L.; Zhang, H.; Li, F.; Luo, H.; Cao, L.; Wang, W.; Wang, F.; Jia, D.; Zhou, Y. Electric/magnetic behaviors of Nd/Ti co-doped BiFeO3 ceramics with morphotropic phase boundary. Scr. Mater. 2019, 164, 6–11. [Google Scholar] [CrossRef]
  23. Gumiel, C.; Jardiel, T.; Bernardo, M.S.; Villanueva, P.G.; Urdiroz, U.; Cebollada, F.; Aragó, C.; Caballero, A.C.; Peiteado, M. Combination of structural and microstructural effects in the multiferroic response of Nd and Ti co-doped BiFeO3 bulk ceramics. Ceram. Int. 2019, 45, 5276–5283. [Google Scholar] [CrossRef]
  24. Wang, T.; Song, S.H.; Ma, Q.; Tan, M.L.; Chen, J.J. Highly improved multiferroic properties of Sm and Nb co-doped BiFeO3 ceramics prepared by spark plasma sintering combined with sol-gel powders. J. Alloys Compd. 2019, 795, 60–68. [Google Scholar] [CrossRef]
  25. Gao, X.; Li, Y.; Chen, J.; Yuan, C.; Zeng, M.; Zhang, A.; Gao, X.; Lu, X.; Li, Q.; Liu, J.-M. High energy storage performances of Bi1−xSmxFe0.95Sc0.05O3 lead-free ceramics synthesized by rapid hot press sintering. J. Eur. Ceram. Soc. 2019, 39, 2331–2338. [Google Scholar] [CrossRef]
  26. Lin, F.; Yu, Q.; Deng, L.; Zhang, Z.; He, X.; Liu, A.; Shi, W. Effect of La/Cr codoping on structural transformation, leakage, dielectric and magnetic properties of BiFeO3 ceramics. J. Mater. Sci. 2017, 52, 7118–7129. [Google Scholar] [CrossRef]
  27. Wang, C.A.; Pang, H.Z.; Zhang, A.H.; Lu, X.B.; Gao, X.S.; Zeng, M.; Liu, J.M. Enhanced ferroelectric polarization and magnetization in BiFe1−xScxO3 ceramics. Mater. Res. Bull. 2015, 70, 595–599. [Google Scholar] [CrossRef]
  28. Perez-Maqueda, L.A.; Gil-Gonzalez, E.; Wassel, M.A.; Jha, S.K.; Perejon, A.; Charalambous, H.; Okasinski, J.; Sanchez-Jimenez, P.E.; Tsakalakos, T. Insight into the BiFeO3 flash sintering process by in-situ energy dispersive X-ray diffraction (ED-XRD). Ceram. Int. 2019, 45, 2828–2834. [Google Scholar] [CrossRef]
  29. Bernardo, M.S.; Jardiel, T.; Caballero, A.C.; Bram, M.; Gonzalez-Julian, J.; Peiteado, M. Electric current activated sintering (ECAS) of undoped and titanium-doped BiFeO3 bulk ceramics with homogeneous microstructure. J. Eur. Ceram. Soc. 2019, 39, 2042–2049. [Google Scholar] [CrossRef]
  30. Cai, W.; Gao, R.; Fu, C.; Yao, L.; Chen, G.; Deng, X.; Wang, Z.; Cao, X.; Wang, F. Microstructure, enhanced electric and magnetic properties of Bi0.9La0.1FeO3 ceramics prepared by microwave sintering. J. Alloys Compd. 2019, 774, 61–68. [Google Scholar] [CrossRef]
  31. Yu, C.; Viola, G.; Zhang, D.; Zhou, K.; Koval, V.; Mahajan, A.; Wilson, R.M.; Tarakina, N.V.; Abrahams, I.; Yan, H. Phase evolution and electrical behaviour of samarium-substituted bismuth ferrite ceramics. J. Eur. Ceram. Soc. 2018, 38, 1374–1380. [Google Scholar] [CrossRef]
  32. Koval, V.; Skorvanek, I.; Durisin, J.; Viola, G.; Kovalcikova, A.; Svec, P.; Saksl, K.; Yan, H. Terbium-induced phase transitions and weak ferromagnetism in multiferroic bismuth ferrite ceramics. J. Mater. Chem. C 2017, 5, 2669–2685. [Google Scholar] [CrossRef]
  33. Wang, T.; Song, S.H.; Wang, M.; Li, J.Q.; Ravi, M. Effect of annealing atmosphere on the structural and electrical properties of BiFeO3 multiferroic ceramics prepared by sol–gel and spark plasma sintering techniques. Ceram. Int. 2016, 42, 7328–7335. [Google Scholar] [CrossRef]
  34. Wang, T.; Xu, T.; Gao, S.; Song, S.H. Effect of Nd and Nb co-doping on the structural, magnetic and optical properties of multiferroic BiFeO3 nanoparticles prepared by sol-gel method. Ceram. Int. 2017, 43, 4489–4495. [Google Scholar] [CrossRef]
  35. Wang, T.; Wang, X.L.; Song, S.H.; Ma, Q. Effect of rare-earth Nd/Sm doping on the structural and multiferroic properties of BiFeO3 ceramics prepared by spark plasma sintering. Ceram. Int. 2020, 46, 15228–15235. [Google Scholar] [CrossRef]
  36. Zhang, S.X.; Wang, L.; Chen, Y.; Wang, D.L.; Yao, Y.B.; Ma, Y.W. Observation of room temperature saturated ferroelectric polarization in Dy substituted BiFeO3 ceramics. J. Appl. Phys. 2012, 111, 074105. [Google Scholar] [CrossRef]
  37. Khomchenko, V.A.; Paixao, J.A.; Shvartsman, V.V.; Borisov, P.; Kleemann, W.; Karpinsky, D.V.; Kholkin, A.L. Effect of Sm substitution on ferroelectric and magnetic properties of BiFeO3. Scr. Mater. 2010, 62, 238–241. [Google Scholar] [CrossRef]
  38. Islam, M.R.; Islam, M.S.; Zubair, M.A.; Usama, H.M.; Azam, M.S.; Sharif, A. Evidence of superparamagnetism and improved electrical properties in Ba and Ta co-doped BiFeO3 ceramics. J. Alloys Compd. 2018, 735, 2584–2596. [Google Scholar] [CrossRef]
  39. Durga Rao, T.; Raja Kandula, K.; Kumar, A.; Asthana, S. Improved magnetization and reduced leakage current in Sm and Sc co-substituted BiFeO3. J. Appl. Phys. 2018, 123, 244104. [Google Scholar] [CrossRef]
  40. Habib, M.; Tang, L.; Xue, G.; Rahman, A.; Kim, M.-H.; Lee, S.; Zhou, X.; Zhang, Y.; Zhang, D. Design and development of a new lead-free BiFeO3-BaTiO3 quenched ceramics for high piezoelectric strain performance. Chem. Eng. J. 2023, 473, 145387. [Google Scholar] [CrossRef]
  41. Park, T.-J.; Papaefthymiou, G.C.; Viescas, A.J.; Moodenbaugh, A.R.; Wong, S. Size-dependent magnetic properties of single-crystalline multiferroic BiFeO3 nanoparticles. Nano Lett. 2007, 7, 766–772. [Google Scholar] [CrossRef] [PubMed]
  42. Wu, X.; Luo, H.; Guo, H.; Liu, J.; Bai, Y.; Zhao, S. Tuning of multiferroic traits in BiFeO3 ceramics by electronic structure. Ceram. Int. 2024, 50, 18853–18867. [Google Scholar] [CrossRef]
  43. Wang, B.; Liu, W.; Zhao, T.L.; Peng, W.; Ci, P.H.; Dong, S.X. Promising lead-free BiFeO3-BaTiO3 ferroelectric ceramics: Optimization strategies and diverse device applications. Progress Mater. Sci. 2024, 146, 101333. [Google Scholar] [CrossRef]
  44. Radojković, A.; Luković-Golić, D.; Orsini, N.J.; Nikolić, N.; Ćirković, J.; Lazarević, S.; Despotović, Ž. Evolution of ferroelectric and piezoelectric properties of BiFeO3 ceramics doped with lanthanum and zirconium. J. Alloys Compd. 2024, 1009, 176901. [Google Scholar] [CrossRef]
  45. Zhang, L.; Ke, H.; Zhang, H.; Li, F.; Zhao, J.; Luo, H.; Cao, L.; Zeng, G.; Li, X.; Wang, W.; et al. Effects of morphotropic phase boundary on the electric behavior of Er/Ti co-doped BiFeO3 ceramics. Scr. Mater. 2019, 158, 71–76. [Google Scholar] [CrossRef]
  46. Zhang, W.; Zhu, X.; Wang, L.; Xu, X.; Yao, Q.; Mao, W.; Li, X. Study on the magnetic and ferroelectric properties of Bi0.95Dy0.05Fe0.95M0.05O3 (M=Mn, Co) ceramics. J. Supercond. Nov. Magn. 2017, 30, 3001–3005. [Google Scholar] [CrossRef]
  47. Hua, H.; Bao, G.; Li, C.; Zhu, Y.; Yang, J.; Li, X. Effect of Ho, Mn co-doping on the structural, optical and ferroelectric properties of BiFeO3 nanoparticles. J. Mater. Sci. Mater. Electron. 2017, 28, 17283–17287. [Google Scholar] [CrossRef]
  48. Chen, X.; Wang, J.; Yuan, G.; Wu, D.; Liu, J.; Yin, J.; Liu, Z. Structure, ferroelectric and piezoelectric properties of multiferroic Bi0.875Sm0.125FeO3 ceramics. J. Alloys Compd. 2012, 541, 173–176. [Google Scholar] [CrossRef]
  49. Jeong, J.; Le, M.D.; Bourges, P.; Petit, S.; Furukawa, S.; Kim, S.A.; Lee, S.; Cheong, S.W.; Park, J.G. Temperature-dependent interplay of Dzyaloshinskii-Moriya interaction and single-ion anisotropy in multiferroic BiFeO3. Phys. Rev. Lett. 2014, 113, 107202. [Google Scholar] [CrossRef]
  50. Yuan, G.L.; Or, S.W.; Liu, J.M.; Liu, Z.G. Structural transformation and ferroelectromagnetic behavior in single-phase Bi1−xNdxFeO3 multiferroic ceramics. Appl. Phys. Lett. 2006, 89, 052905. [Google Scholar] [CrossRef]
  51. Karpinsky, D.V.; Silibin, M.V.; Latushka, S.I.; Zhaludkevich, D.V.; Sikolenko, V.V.; Svetogorov, R.; Sayyed, M.I.; Almousa, N.; Trukhanov, A.; Trukhanov, S.; et al. Temperature-driven transformation of the crystal and magnetic structures of BiFe0.7Mn0.3O3 ceramics. Nanomaterials 2022, 12, 2813. [Google Scholar] [CrossRef] [PubMed]
  52. Abdel Monem, A.M.; Abd-Elmohssen, N.; El-Bahnasawy, H.H.; Makram, N.; Sedeek, K. Comparative studies by X-ray diffraction, Raman, vibrating sample magnetometer and Mössbauer spectroscopy of pure, Sr doped and Sr, Co co-doped BiFeO3 ceramic synthesized via tartaric acid-assisted technique. Ceram. Int. 2023, 49, 15213–15220. [Google Scholar] [CrossRef]
  53. Yadav, M.; Agarwal, A.; Sanghi, S.; Kotnala, R.K.; Shah, J.; Bhasin, T.; Tuteja, M.; Singh, J. Crystal structure refinement, dielectric and magnetic properties of A-site and B-site co-substituted Bi0.90Nd0.10Fe1-xTixO3 (x=0.00, 0.02, 0.05 & 0.07) ceramics. J. Alloys Compd. 2018, 750, 848–856. [Google Scholar]
Figure 1. (a) XRD patterns of undoped and Nd–Nb co-doped BiFeO3 ceramics; (b) magnified patterns around 2θ ~32°; (c) Rietveld refinement XRD patterns of 10NdNb samples.
Figure 1. (a) XRD patterns of undoped and Nd–Nb co-doped BiFeO3 ceramics; (b) magnified patterns around 2θ ~32°; (c) Rietveld refinement XRD patterns of 10NdNb samples.
Nanomaterials 15 00215 g001
Figure 2. SEM micrographs of the fresh fracture surfaces for (a) BFO, (b) 5NdNb, (c) 10NdNb, (d) 15NdNb, (e) 20NdNb, and (f) EDS; (g) Nd element mapping; and (h) Nb element mapping for 15NdNb samples (inset: the average grain size).
Figure 2. SEM micrographs of the fresh fracture surfaces for (a) BFO, (b) 5NdNb, (c) 10NdNb, (d) 15NdNb, (e) 20NdNb, and (f) EDS; (g) Nd element mapping; and (h) Nb element mapping for 15NdNb samples (inset: the average grain size).
Nanomaterials 15 00215 g002
Figure 3. Temperature dependences of (a) dielectric constant and (b) loss tangent at 1 MHz for undoped and Nd–Nb co-doped BiFeO3 ceramics.
Figure 3. Temperature dependences of (a) dielectric constant and (b) loss tangent at 1 MHz for undoped and Nd–Nb co-doped BiFeO3 ceramics.
Nanomaterials 15 00215 g003
Figure 4. (a) Room−temperature ferroelectric hysteresis loops (PE) and (b) ferroelectric parameters determined at 10 Hz for undoped and Nd–Nb co-doped BiFeO3 ceramics (Pr: remnant polarization and Ec: coercive field).
Figure 4. (a) Room−temperature ferroelectric hysteresis loops (PE) and (b) ferroelectric parameters determined at 10 Hz for undoped and Nd–Nb co-doped BiFeO3 ceramics (Pr: remnant polarization and Ec: coercive field).
Nanomaterials 15 00215 g004
Figure 5. (a) Leakage current density as a function of electric field (JE) and (b) leakage mechanism for undoped and Nd–Nb co-doped BiFeO3 ceramics, determined at room temperature.
Figure 5. (a) Leakage current density as a function of electric field (JE) and (b) leakage mechanism for undoped and Nd–Nb co-doped BiFeO3 ceramics, determined at room temperature.
Nanomaterials 15 00215 g005
Figure 6. (a) Magnetic hysteresis loops (MH) and (b) ferromagnetic parameters (remnant magnetization (Mr) and coercive field (Hc)) for undoped and Nd–Nb co-doped BiFeO3 ceramics, determined at room temperature.
Figure 6. (a) Magnetic hysteresis loops (MH) and (b) ferromagnetic parameters (remnant magnetization (Mr) and coercive field (Hc)) for undoped and Nd–Nb co-doped BiFeO3 ceramics, determined at room temperature.
Nanomaterials 15 00215 g006
Table 1. Phase constitution, lattice parameters, and R-factors estimated by Rietveld refinement for undoped and Nd–Nb co-doped samples.
Table 1. Phase constitution, lattice parameters, and R-factors estimated by Rietveld refinement for undoped and Nd–Nb co-doped samples.
SamplesSpace GroupPhase Constitution (%)Lattice Parameters (Å)R-Factors
BFOR3c100a = 5.571
c = 13.872
Rexp = 7.36 Rwp = 11.34
GOF = 1.54
5NdNbR3c97.45a = 5.578 c = 13.821Rexp = 6.57 Rwp = 8.67
GOF = 1.32
Pnma2.55a = 5.623 b = 7.946
c = 5.858
10NdNbR3c90.29a = 5.587 c = 13.796Rexp = 5.82 Rwp = 7.68
GOF = 1.32
Pnma9.71a = 5.572 b = 7.874
c = 5.869
15NdNbR3c85.46a = 5.592 c = 13.774Rexp = 6.13 Rwp = 8.65
GOF = 1.41
Pnma14.54a = 5.569 b = 7.866
c = 5.883
20NdNbR3c75.12a = 5.597 c = 13.721Rexp = 5.59 Rwp = 7.95
GOF = 1.42
Pnma24.88a = 5.546 b = 7.810
c = 5.912
Table 2. Ferroelectric and magnetic parameters for undoped and Nd–Nb co-doped BiFeO3 ceramics.
Table 2. Ferroelectric and magnetic parameters for undoped and Nd–Nb co-doped BiFeO3 ceramics.
ParameterBFO5NdNb10NdNb15NdNb20NdNb
Pr (μC/cm2)0.411.863.120.510.36
Ec (kV/cm)4.4511.4011.106.136.36
Mr (emu/g)5.67 × 10−40.0700.120.150.14
Hc (kOe)0.0506.178.849.809.05
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, T.; Ye, H.; Wang, X.; Cui, Y.; Mei, H.; Song, S.; Zhao, Z.; Wang, M.; Rosaiah, P.; Ma, Q. Improved Ferroelectric and Magnetic Properties of Bismuth Ferrite-Based Ceramics by Introduction of Non-Isovalent Ions and Grain Engineering. Nanomaterials 2025, 15, 215. https://doi.org/10.3390/nano15030215

AMA Style

Wang T, Ye H, Wang X, Cui Y, Mei H, Song S, Zhao Z, Wang M, Rosaiah P, Ma Q. Improved Ferroelectric and Magnetic Properties of Bismuth Ferrite-Based Ceramics by Introduction of Non-Isovalent Ions and Grain Engineering. Nanomaterials. 2025; 15(3):215. https://doi.org/10.3390/nano15030215

Chicago/Turabian Style

Wang, Ting, Huojuan Ye, Xiaoling Wang, Yuhan Cui, Haijuan Mei, Shenhua Song, Zhenting Zhao, Meng Wang, Pitcheri Rosaiah, and Qing Ma. 2025. "Improved Ferroelectric and Magnetic Properties of Bismuth Ferrite-Based Ceramics by Introduction of Non-Isovalent Ions and Grain Engineering" Nanomaterials 15, no. 3: 215. https://doi.org/10.3390/nano15030215

APA Style

Wang, T., Ye, H., Wang, X., Cui, Y., Mei, H., Song, S., Zhao, Z., Wang, M., Rosaiah, P., & Ma, Q. (2025). Improved Ferroelectric and Magnetic Properties of Bismuth Ferrite-Based Ceramics by Introduction of Non-Isovalent Ions and Grain Engineering. Nanomaterials, 15(3), 215. https://doi.org/10.3390/nano15030215

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