Next Article in Journal
Insight into Unsteady Separated Stagnation Point Flow of Hybrid Nanofluids Subjected to an Electro-Magnetohydrodynamics Riga Plate
Previous Article in Journal
Magnetization Dynamics in FexCo1-x in Presence of Chemical Disorder
Previous Article in Special Issue
Effect of Metal-Oxide Phase on the Magnetic and Magnetocaloric Properties of La0.7Ca0.3MnO3-MO (MO=CuO, CoO, and NiO) Composite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Bismuth on the Structure, Magnetic and Photocatalytic Characteristics of GdFeO3

Institute of Crystal Growth, School of Materials Science and Engineering, Shanghai Institute of Technology, Shanghai 201418, China
*
Author to whom correspondence should be addressed.
Magnetochemistry 2023, 9(2), 45; https://doi.org/10.3390/magnetochemistry9020045
Submission received: 25 November 2022 / Revised: 20 January 2023 / Accepted: 27 January 2023 / Published: 29 January 2023

Abstract

:
In this paper, a series of Gd1-xBixFeO3 (x = 0, 0.1, 0.3, 0.5, 0.7, 0.9, 1) nanoparticles have been readily synthesized by a green and facile sol–gel method. It gradually changed from the orthorhombic structure (space group Pbnm) to the rhombohedral perovskite structure (space group R3c). Weak ferromagnetic behavior was effectively induced by Bi3+, with reduced magnetization. It was closely related with the lattice distortion of the perovskite structure and modified interactions between Fe-O-Fe. Boosted photocatalytic activities of Gd1-xBixFeO3 were observed for the removal of methylene blue (MB) under the visible light irradiation. In particular, Gd0.5Bi0.5FeO3 showed the optimum photocatalytic efficiency, in which the degradation efficiency reached 82.1% after 180 min of visible light illumination, with good stability and repeatability. The improved performance was mainly ascribed to enhanced visible light absorption, decreased optical band gap from 2.21 to 1.8eV and stronger charge transfer efficiency. A possible photocatalytic mechanism is also proposed according to the band structure. The results indicate that this system will be a promising candidate for the degradation of organic pollutant as a novel magnetically recoverable photocatalyst.

1. Introduction

The rare earth orthoferrite (RFeO3) with perovskite structure has received intensive attention, thanks to its exotic magnetism, multiferroics and unique features for fast magneto–optical isolator, ultra-fast spin switching, etc. [1]. Moreover, due to the suitable band gap, RFeO3 has also been considered as a potential visible light photocatalyst for the degradation of organic pollutants [2,3,4]. Interestingly, these magnetic photocatalysts display advantages of magnetic separation and recyclability after photocatalytic reactions, which make them easy to recover from waste water to avoid secondary pollution. As a typical orthorferrite compound, GdFeO3 contains Gd3+ half-filled 4f shell and a stable structure with a Pbnm space group. It displays a narrow band gap (around 2.4 eV), capable of absorbing both visible and UV light. Some studies have been conducted to investigate the photocatalytic behaviors of GdFeO3, and effective strategies have also been proposed to boost the photocatalytic performance. For instance, the substitution of Mn at the Fe site of GdFeO3 introduces an extra electronic state between the conduction band and the valence band, which reduces the optical band gap and enhances the rhodamine-B (Rh-B) degradation efficiency [5]. The degradation rate of Rh-B is greatly improved by a factor of 2.5 for GdFe0.7Mn0.3O3 as compared to pure GdFeO3. Similarly, the introduction of Mn in ErFeO3 greatly increases the photocatalytic degradation efficiency of methylene blue (MB), due to the narrower band gap and smaller particle size [6]. Moreover, the degradation efficiency of Rh-B up to 92% at 120 min is achieved in GdFeO3/g-C3N4-3:7 composites, while minor weigh loss is observed during the consecutive runs [7].
Bismuth ferrites (BiFeO3), with distorted rhombohedral perovskite structure, shows fascinating multiferroic properties with simultaneous antiferromagnetic ordering (TN = 370 °C) and ferroelectric configurations (Tc = 830 °C) at room temperature [8]. BiFeO3 is also regarded as a novel photocatalyst under visible light irradiation, owing to its narrow band gap (2.0 eV–2.2 eV) and excellent chemical stability [9,10]. Meanwhile, the internal electric field from ferroelectric BiFeO3 possibly enhances surface reactions and photogenerated charge separation efficiency [11]. However, the photocatalytic performance is still restricted by the presence of impurity phases, structural instability, and slow mobility rate [12]. Various strategies have also been applied to boost the photocatalytic efficiency, such as surface engineering, morphology tuning, heterostructure formation, doping and crystal facet engineering. For example, La-doped BiFeO3-based multiferroics were synthesized to achieve a strong magnetic response and high piezo-photocatalytic activity [12]. The optical bandgap of BiFeO3 nanoparticles has been tuned from 2.06 eV to 1.75 eV by Cd-Ni co-doping. The best degradation activity of MB and Rh-B has been achieved in a Bi0.94Cd0.06Fe0.94Ni0.06O3 sample under UV–visible irradiation, which may be originated from the high surface area, smaller average particle size and bandgap [8]. MB dye was almost completely degraded in 120 min for a AgFe2O4/BiFeO3 heterostructure under visible light irradiation [13].
Now, enhanced photocatalytic performance of ferrites is still highly expected by effective regulation of the structure, morphology, band energy, etc. Dopants are proven to be feasible strategies to modify the magnetic and photocatalytic properties of orthoferrites. Substituents may act as surface traps for holes and electrons, reducing recombination rate and increasing photo-produced charge carrier lifetime, leading to enhanced photocatalysis of materials [11]. The lattice structure, electronic band gap and the dipole–dipole interaction of perovskites can also be effectively regulated by the substitution at the A site or B site, making it possible to effectively modulate the optical, magnetic, electrical and catalytic properties. Orthorhombic GdFeO3 and rhombohedral BiFeO3 show distinctive crystal structures. Cation substitution of Gd3+ for larger Bi3+ cations results in structural modifications mainly in FeO6 rotations and tilting, and the electric polar moment on GdFeO3 is also modified by the s2 lone pairs of electrons in the Bi cations [14]. Therefore, the influence of Bi3+ on the lattice structure, band energy, magnetic properties and catalytic behaviors of GdFeO3 need to be explored. Recently, it has been reported that the introduction of Bi3+ in YFeO3 does not generate significant changes in band gap values and photocatalytic efficiency, but modifies the crystal structure and the magnetic behavior [14]. The investigation towards the effect of Bi3+ on the structure, magnetic and photocatalytic characteristics of GdFeO3 is still scarce. Herein, a series of Gd1-xBixFeO3 (x = 0, 0.1, 0.3, 0.5, 0.7, 0.9, 1) powders were prepared by the sol–gel method. Sponge-like particles are usually formed during the sol–gel method, which is helpful for enhancing the catalytic reactions. The crystal structure, crystalline morphology, magnetic and optical properties of the as-prepared samples were fully characterized. The photocatalytic activity of the samples was investigated via the degradation of MB under visible light irradiation. The photocatalytic efficiency was effectively increased by the introduction of Bi3+. The optimum catalytic performance was observed for Gd0.5Bi0.5FeO3, which was mainly ascribed to the narrower band gap, increased optical absorption, and enhanced charge transfer efficiency.

2. Materials and Methods

A series of Gd1-xBixFeO3 (x = 0, 0.1, 0.3, 0.5, 0.7, 0.9, 1 marked as GFO, GBFO1, GBFO3, GBFO5, GBFO7, GBFO9, BFO) powders were been readily synthesized by the sol–gel method. The raw materials were Gd(NO3)·6H2O(99%), Fe(NO3)3·9H2O(99%) and Bi(NO3)3·5H2O(99%), and citric acid was chosen as the complexing agent. First, at the appropriate stoichiometric ratio, Gd(NO3)3·6H2O, Fe(NO3)3·9H2O and Bi(NO3)3·5H2O were dissolved in deionized water to form a transparent solution. Then, the appropriate amount of citric acid was added into the solution. Second, the pH of the solution was adjusted to 1–2 by adding an ammonia solution. The solution was magnetically stirred at about 80 °C until a sticky gel was formed. Subsequently, the sticky gel was ignited at a setting temperature of 300 °C. Thereafter, the samples were calcined at different temperatures (550 °C–700 °C) in air for 3 h.
The phase identification and crystal structures were checked by X-ray diffraction (XRD, D/Max-2550V, Rigaku, Tokyo, Japan) in the 2θ range of 10° to 80° with a step scanning at 0.02° per step to identify the phase and to calculate the lattice parameters. The Rietveld refinements of the XRD patterns were conducted by the GSAS program. The morphology and microstructure of the prepared samples were investigated by scanning electron microscopy (SEM, S4800, Hitachi, Chiyoda-ku, Japan). The surface composition and valence states of Bi, Fe and O were determined by X-ray photoelectron spectroscopy (XPS, Thermo Scientific K-Alpha, Thermo Fisher Scientific, Waltham, MA, USA) with AlKα as the radiation source ( = 1486.6 eV). The magnetic properties of the samples were studied by a vibrating sample magnetometer (VSM, Lakeshore7407, Lakeshore, Columbus, OH, USA). UV-vis absorption spectra of the samples were recorded with an UV-vis spectrophotometer (Cary 5000, Agilent Technologies, Santa Clara, CA, USA).
The photocatalytic activity of all the samples was evaluated by photodegradation of MB using visible light irradiation with a xenon lamp (500 W) with a 420 nm cutoff filter as a light source at room temperature. A total of 50 mg of photocatalyst was added to 50 mL of methylene blue (MB) solution with a concentration of 5 mg/L. Prior to the visible light irradiation, the reaction mixture was homogeneously stirred in the dark for 60 min to attain the adsorption–desorption equilibrium. The solution of the sample was positioned 20 cm away from the xenon lamp. Then, the solution was exposed to light with continuous stirring; at regular 20–30 min intervals, about 3 mL of solution was collected and centrifuged to remove the catalyst particles. The photocatalytic activities of the synthesized Gd1−xBixFeO3 were monitored by the change in the concentration of MB, which was evaluated by the variation of absorption band maximum (664 nm) using a UV-vis spectrophotometer. Electrochemical impedance spectroscopy (EIS) and Mott–Schottky (MS) curves were carried out on an electrochemical workstation (CHI 760E, CH Instrument, Shanghai Chenhua Instrument Company, Shanghai, China) using a three-electrode system. The saturated Ag/AgCl and platinum foil served as the reference electrode and the counter electrode, respectively. The working electrode was prepared by dip-coating the slurry (5 mg of photocatalyst powders dispersed in 50 µL of ethanol and 10 µL of Nafion solution (5 wt%) under sonication) on the FTO glass. A 0.1 M Na2SO4 aqueous solution (pH 6.8) was used as the electrolyte. EIS tests were recorded under visible light (λ ≥ 400 nm) irradiation by applying an AC voltage of 5 mV in the frequency range from 0.01 Hz to 100 kHz. The Mott–Schottky plots were measured with a fixed frequency of 2000 Hz.

3. Results

GdxBi1−xFeO3 (x = 0, 0.1, 0.3, 0.5, 0.7) samples were prepared and calcined at 700 °C, while GdxBi1−xFeO3 (x = 0.9, 1) powders were successfully obtained at 550 °C. The results indicate that Bi3+ is effective in lowering the formation temperature of orthoferrite, without any observable impurities. The Rietveld refined X-ray powder diffraction patterns of GdxBi1-xFeO3 are shown in Figure 1. In Figure 1a, when the content of Bi3+ was ≤0.5, the pure orthorhombic perovskite phase (GdFeO3, PDF No. 74-1476) was observed, without any impurities. Obviously, the mixed phase of the Pnma (wt ≈ 64.68%) and R3c (wt ≈ 35.32%) phase was observed at x = 0.7, indicating that the orthorhombic perovskite structure can remain when the content of bismuth was lower than x = 0.7. Figure 1b reveals an enlarged view of the (112) diffraction peak around 32° and 34°. As the Bi concentration increased, this diffraction peak shifted towards a lower 2θ value, suggesting enlarged lattice volumes. In Figure 1c, when the concentration of Bi3+ was ≥0.9, the pure rhombohedral perovskite (R3c) (BiFeO3, PDF No. 72-2112) was obtained. The phase transition was attributed to the structural distortion resulting from the incorporation of Bi3+ ions into the lattice of GFO, since the ionic radius of Bi3+ (1.17 Å) is larger than of Gd3+ (1.05 Å) [15]. Lattice distortion will cause changes in the lattice parameters, bond length, bond angle of Fe-O-Fe and other related parameters.
Table 1 shows the refined unit cell parameters, bond angle, bond length and goodness of fit. For the samples of GdxBi1−xFeO3 (x = 0, 0.1, 0.3, 0.5), the lattice parameters of the Pnma phase gradually increased with the increased concentration of Bi3+, suggesting the component-driven structural distortions. For the samples of GdxBi1−xFeO3 (x = 0.7, 0.9, 1) in the R3c phase, the lattice constants a and b decreased with the increase in Bi content while the lattice constant c increased with the increase in Bi. Overall, the unit volumes were enlarged with the introduction of Bi for these samples. Additionally, the change of the bond length and the bond angles were varied for the Pnma phase and the R3c phase. In particular, for GBFO5, the Fe-O1 bond length and the Fe-O2 bond length were expanded to 2.043 Å and 2.033 Å, respectively, while the bond angles of Fe-O1-Fe and Fe-O2-Fe were increased to be 142.943° and 146.985°, respectively. Moreover, it has been reported that the hybridization between Bi and O could possibly induce enhanced octahedral tilting distortion of the perovskite structure [16].
Figure 2a–d shows the SEM images of the prepared Gd1−xBixFeO3 (x = 0, 0.5, 0.7, 0.9) powders. The Gd1−xBixFeO3 (x = 0, 0.5, 0.7) samples were calcined at 700 °C and Gd0.1Bi0.9FeO3 was prepared at 550 °C. The statistical calculation of the grain size distribution of the corresponding samples is shown in Figure 2e–h. In Figure 2a, pure GFO particles were highly agglomerated and the crystalline grains were well defined by the grain boundaries, which showed irregular shapes with an average grain size around 190 nm. The continuous porous network was attributed from the releasing of a large amount of gas during the combustion reaction process. Plate-like particles were observed for Gd0.5Bi0.5FeO3, as seen in Figure 2b, with an average particle size around 123 nm. Figure 2c shows the SEM images of the prepared Gd0.3Bi0.7FeO3 powder, with a grain size of about 180 nm. Apparently, with increasing doping amounts, the grain size of the samples decreased gradually, indicating that the crystal growth was inhibited by the introduction of Bi3+. This behavior could be attributed to the formation of crystal defects due to the occupation of the lattice site of the host structure by dopant ions, which acts as a barrier to the crystal growth and thus reduces the particle size [17]. Figure 2d shows the SEM images of the Gd0.1Bi0.9FeO3 powder calcined at 550 °C. Due to the lower calcination temperature, the elliptical or spherical particles were partially agglomerated, with a particle size of around 120 nm. Smaller particles had a higher surface energy, which may stimulate the agglomeration process to occur [18].
Figure 3 demonstrates the result of the XPS analysis of the spectra of the elements Bi, Fe and O in the Gd0.5Bi0.5FeO3 sample. Figure 3a shows the two peaks located at 158.68 eV for Bi 4f7/2 and 163.98 eV for Bi 4f5/2. The energy difference between the Bi 4f7/2 and Bi 4f5/2 peaks was around 5.3 eV, which confirms the +3 oxidation state of bismuth [19]. In Figure 3b, the asymmetric peaks observed in the Fe 2p spectrum were resolved into Fe 2p3/2 and Fe 2p1/2. The typical peaks of Fe3+ were located at 711.03 eV and 724.83 eV, with a satellite peak at 718.22 eV. Meanwhile, the peaks at 709.72 eV and 723.25 eV were assigned to Fe2+. The appearance of defects broke the charge balance of the system in the process of Bi ion replacing Gd. In order to make the charge balance of the system, Fe3+ ions were converted to Fe2+ ions. The presence of Fe ions of varied valence states will be conducive to the formation of more oxygen vacancies, which is beneficial to the surface adsorption on catalysts of organic and oxygen species [20]. Figure 3c depicts the O 1s spectra of the GBFO5 powder; two peaks were found at 529.28 eV and 530.78 eV. The O-I peak is generally believed to be caused by the O2− ion in the crystal structure of the material, and O-II peaks may be formed by oxygen vacancies or defects [21].
Figure 4a,b shows the magnetic hysteresis loops of Gd1−xBixFeO3 (x = 0.0, 0.1, 0.3, 0.5, 0.7 and 0.9) at room temperature. In Figure 4a, the linear hysteresis curve was observed for GFO and GBFO1, showing characteristic features of antiferromagnetic (AFM) behavior, due to their canted antiferromagnetic spin structure and high magnetic anisotropic energy. Similar AFM characteristics of GdFeO3 was also reported by other groups [22,23]. Magnetic hysteresis loops were observed for Gd1−xBixFeO3 (x = 0.3, 0.5, 0.7, 0.9), indicating that the week ferromagnetic behavior was partly induced by Bi3+. Meanwhile, the magnetization was weakened for Gd1−xBixFeO3 (x = 0.0, 0.1, 0.3, 0.5, 0.7), and it was moderately strengthened for Gd1−xBixFeO3 (x = 0. 9) which belongs to the R3c phase of BiFeO3. Overall, compared with pure GdFeO3, weak ferromagnetic behaviors were observed for solid solution compounds, and the magnetization of Gd1−xBixFeO3 was effectively decreased. In view of the unsaturation of the M-H curves for Gd1−xBixFeO3 (x = 0–0.7), canted antiferromagnetic ordering is still considered to exist in these samples even though the ferromagnetism was enhanced.
Generally, the magnetic variations are influenced by a variety of factors, including the chemical composition, magnetic anisotropic field, structure, grain size, morphology, defects, domain wall motion, magnetic moment reversal, etc. [24]. With the Pbnm structure, the family of RFeO3 has two magnetic sublattices from the 4f-electrons of R3+ and 3d-electrons of Fe3+, and there exists three types of magnetic exchange interaction: Fe3+-Fe3+, R3+-R3+ and Fe3+-R3+ [1]. The spin-canting of RFeO3 is mainly caused by the Dzyaloshinskii–Moriya interaction between antiferromagnetically ordered Fe 3d spins, and the spin-canting angle is closely related to the lattice distortion including the bond angle and length of Fe-O-Fe. For GdFeO3, at room temperature, the Fe3+-Fe3+ interaction was dominant, which caused the Fe3+ spins usually order at 660 K to have a slightly canted G-type AFM structure. The substitution of Gd3+ ions by Bi3+ ions with larger radii can result in the modification of the lattice distortion. Considering the XRD refinement results, the Fe-O bond length and the Fe-O-Fe bond angle were highly dependent on the doping content of Bi. It can be seen that magnetization of Gd1−xBixFeO3 (x = 0, 0.1, 0.3, 0.5, 0.7) was directly proportional to the Fe-O2-Fe bond angle and inversely proportional to the Fe-O2 bond length, which is consistent with the conclusion of Raj et al. [25]. Generally, the introduction of bismuth usually results in decreased magnetization at room temperature, which also originates from decreased interactions between Fe-O-Fe and distortion of the FeO6 octahedral of the perovskite structure. Meanwhile, according to the SEM results, compared with GFO, smaller and more dispersed particles were observed for Gd1-xBixFeO3, which possibly led to the enhancement of the surface spin effect, thus reducing the saturation magnetization.
Light absorption is crucial for photocatalytic activities. The UV-vis absorbance spectra of the Gd1−xBixFeO3 samples are presented in Figure 5. The optical band gap of Gd1−xBixFeO3 was calculated by the Tauc relationship:
( a h v ) = A ( h v E g ) 1 2
where α, h, v, Eg and A are the absorption coefficient, Plank’s coefficient, light frequency, band gap energy and a constant, respectively. The Eg value was estimated by extrapolating the linear region of the plot in the energy axis (insets in Figure 5). According to linear extrapolation, the bandgap of GdFeO3 was calculated to be 2.21 eV, which is quite comparable with previous reports [26,27]. For the Bi3+-doped GdFeO3, the optical band gap was effectively decreased, with enhanced absorptions of both UV and visible light. In addition, the narrower band gap allowed more carrier excitation under visible light irradiation, implying improved photocatalytic activities. In particular, the smallest band gap and highest visible light absorption were observed in Gd0.5Bi0.5FeO3, and the band gap was around 1.8 eV.
In addition, the reduction in band gap for all the doped sample may be related to the distortion in the Fe-O-Fe octahedral rearrangement of molecular orbitals and decreased particles size [28]. The decrease of the band gap was also observed in Sm-doped BiFeO3 and Mn-doped GdFeO3 [5,29]. The hybridization between the Fe-3d and O-2p states depends on Fe-O-Fe exchange angles, which directly modify the band gap [14]. In addition, the band gap is also the result of some other factors, such as the defect levels, the crystallinity and the oxygen vacancies on the surface. Thus, the higher absorption and narrower band gap of GFO5 may originate from its smaller particle size, lattice distortion caused by appropriate Bi ion doping, higher crystallinity, etc. Proper band gap could promote the light absorption and increase photogenerated charge carriers, thus improving the catalytic activity [30]. The introduction of Bi dopants favors the light absorption in the GFO nanoparticles, implying that improved visible light-driven photocatalytic performance is expected.
The photocatalytic activity of Gd1−xBixFeO3 samples were evaluated by degradation of MB under visible light irradiation. The (Ct/C0) vs. visible light irradiation time (t) curve for Gd1−xBixFeO3 sample is shown in Figure 6a. The concentration of MB hardly changed when the photocatalyst was not added or the reaction was carried out in the dark. It indicates that MB is a stable pollutant and the solution reached the adsorption–desorption equilibrium. After 180 min of visible light illumination, the degradation efficiency reached about 74%, 79.7%, 76.7%, 82.1%, 69% and 63.2% for GFO, GBFO1, GBFO3, GBFO5, GBFO7 and GBFO9, respectively. Obviously, the concentration of Bi3+ ions doping affected the photocatalytic performance of the sample. With the increase of the Bi3+ doping amount, the degradation efficiency of MB first increased and then decreased, and GBFO5 showed exceptionally enhanced photocatalytic efficiency. The degradation performance of Gd1-xBixFeO3 was similar to the result of the Y1−xBixFeO3 system, in which the catalytic activities ranged from 50% to 80% MB degradation efficiency for 180 min for x = 0.1 to x = 0.4, respectively [14]. A BiFeO3-GdFeO3 heterojunction photocatalyst has been reported to degrade 56% MB and 98% MB in 180 min and 9 h, respectively, under sunlight irradiation [31]. The nanogranular BiFeO3 exhibited 53% MB dye degradation after irradiation for 180 min [32]. The nanotubes showed much improved catalytic activity with a degradation efficiency of 75% after irradiation for 180 min. The results also imply that the photocatalytic performance is partly influenced by the crystalline morphology.
Figure 6b represents the ln(C0/Ct) vs. irradiation time (t) curves, where the linearity nature indicated the first-order kinetics of the visible light-driven photocatalytic MB degradation. Therefore, the photocatalytic activities can be evaluated by the apparent first-order rate constant k, and it can be expressed by Equation (2) [33]:
ln ( C 0 C ) = kt  
where C0 represents the initial concentration of MB, C is the concentration of MB after the degradation with respect to time (t) and k is the rate constant. The obtained values of k were 7.3 × 10−3 min−1, 8.0 × 10−3 min−1, 8.8 × 10−3 min−1, 9.8 × 10−3 min−1, 6.4 × 10−3 min−1 and 5.3 × 10−3 min−1 for GFO, GBFO1, GBFO3, GBFO5, GBFO7 and GBFO9, respectively. This further confirmed that proper Bi3+ substitution of GFO can significantly improve the photocatalytic activity.
The charge separation of e-h+ pairs on the interface of the photocatalyst can be revealed by the electrochemical impedance spectra (EIS). The radii of the semicircles are proportional to the charge transfer resistance. For the photocatalytic reaction, a reduction in resistance always means a faster charge transfer and slower recombination of holes and photoelectrons occurred, which is helpful for the relevant photocatalytic efficiency. Figure 7a shows the EIS spectra obtained for GFO, GBFO1, GBFO3, GBFO5, GBFO7 and GBFO9. Compared with GFO, the radius of the semicircle for Gd1-xBixFeO3 (x = 0.1, 0.3, 0.5) decreased with increasing Bi3+, implying reduced charge transfer resistance. Thereafter, the radius of the semicircle increased for Gd1−xBixFeO3 (x = 0.7, 0.9). GBFO5 exhibited the smallest semicircle, which is indicative of the best charge transfer efficiency and the slowest recombination of holes and photoelectrons out of these five samples, which contribute to the photo-degradation organic pollutants.
Figure 7b shows the Mott–Schottky plots of GFO, GBFO1, GBFO3, GBFO5, GBFO7 and GBFO9 with a fixed frequency (2000 Hz). The positive slope featured typical n-type semiconductors for the samples. According to Equation (3), the flat band potential VFB of Gd1−xBixFeO3 was estimated by the Mott–Schottky (MS) measurement [34].
1 C s c 2 = 2 e N d ε ε 0 ( E E F B K T E )  
where, C s c 2 is the space charge capacitance (in F cm−2), e is electronic charge in C, ε0 is the permittivity of free space, ε is the dielectric constant of the semiconductors, Nd is the charge carrier density in cm−3, EFB is the flat band potential in V, K is the Boltzmann constant and T is the temperature in K. The flat-band potentials VFB were obtained from the extrapolation of the Mott–Schottky plots, which were −0.45, −0.53, −0.55, −0.68, −0.50 and −0.48 eV (vs. Ag/AgCl) for GFO, GBFO1, GBFO3, GBFO5, GBFO7 and GBFO9, respectively. Then, the values corresponded to be −0.25, −0.33, −0.35, −0.48, −0.30 and −0.28 eV versus the normal hydrogen electrode (NHE), respectively. Generally, the conduction band minimum (ECB) of n-type semiconductors is approximately more negative than EFB by 0.2 eV. Correspondingly, the conduction band minimum (ECB) of Gd1−xBixFeO3 (x = 0.0, 0.1, 0.3, 0.5, 0.7 and 0.9) were calculated to be −0.45, −0.53, −0.55, −0.68, −0.50 and −0.48 eV, respectively. The valence band maximum (EVB) can be determined with the relation Eg = EVB − ECB, and the optical bandgaps were obtained from the UV-vis absorption spectra. Based on these results, the band energy diagram was displayed in Figure 8. Compared with GFO, the conduction band of Bi-doped samples was more negative, making it easier to generate superoxide anions to participate in the reactions.
The scavenger studies were carried out for GBFO5 which was taken as the typical example. Tertbutyl-alcohol (TBA), p-benzoquinone(p-BQ), K2Cr2O7 and EDTA-2Na were used as scavengers for OH·, O2·−, e and h+, respectively. These scavengers were separately added to the solution. The photocatalytic experiments were conducted in the presence of the scavengers along with GBFO5. When irradiated for 180 min, with the addition of TBA and EDTA-2Na, the degradation efficiency reached 75.3% and 73.7%, respectively. Compared to the solution without any scavengers, the addition of TBA and EDTA-2Na had a negligible influence on the photocatalytic activity, implying that OH· and h+ are not the main active species. On the contrary, only 40.2% and 33.7% of MB was degraded with the addition of p-BQ and K2Cr2O7, respectively. Severe depression of the performance indicates that O2 and e were the dominant reactive species responsible for the degradation of MB.
According to the above results and discussions, a possible mechanism of the charge transfer and photocatalytic degradation was also proposed. As illustrated in the scheme in Figure 8, under visible light irradiation, photogenerated electrons in the valance band (VB) of the photocatalyst obtain sufficient energy to jump to the conduction band (CB), leaving holes in the VB (Equation (4)). Since the conduction band edge of Gd1-xBixFeO3 is more negative than the standard redox potential E(O2·−/O2)(−0.33 eV vs. NHE) [35], the excited electrons migrate to the catalyst surface, followed by reaction with dissolved O2 to generate superoxide radicals (O2·−) (Equation (5)). In addition, the standard redox potential E(O2/H2O2) (0.685 eV vs. NHE) [36] is located between the CB and VB of Gd1-xBixFeO3, indicating that electrons could also react with oxygen to form H2O2, which is subsequently reduced to form hydroxyl radicals (OH·) (Equations (6) and (7)) [7]. The holes formed at the VB may directly oxidize the organic pollutant [11]. Finally, the superoxide radicals, hydroxyl radical and holes decompose the MB into CO2, H2O and by-products, which could be benzenesulfonicacid 2-amino-5-(methyl amino)-hydroxybenzene sulfonic acid and 2-amino-5-(NN-methyl formamide) benzene sulfonic acid [37].
Gd 1 x Bi x FeO 3 + h ν e + h +
e + O 2 O 2 ·
e + O 2 + H + H 2 O 2
H 2 O 2 + e OH · + OH
OH · /   O 2 · / h + + MB CO 2 + H 2 O + byproducts   of   MB
In addition to the photocatalytic efficiency, the reusability and the stability of photocatalyst are two important factors for practical application. To evaluate the stability, we conducted photocatalytic experiments using Gd0.5Bi0.5FeO3 for five successive runs under the same conditions as mentioned above. The sample was irradiated for 180 min for every cycle. As shown in Figure 9a, almost no change in the photocatalytic activity was observed after five circulations, suggesting good stability. To further verify its stability, we performed XRD measurements of the GBFO5 nanoparticles before and after the photocatalysis reaction, which is displayed in Figure 9b. The crystal structure of the GBFO5 photocatalyst did not change after the photocatalytic reaction and no secondary phase was observed after the reaction, verifying good stability of this photocatalyst.
The photocatalytic activities are generally determined by numerous factors, including the crystalline phase, morphology, defects, band energy, surface activity, etc. Enhanced photocatalytic activities of Gd1−xBixFeO3 have been observed, with the optimum performance observed with Gd0.5Bi0.5FeO3. The SEM results indicated that GBFO5 samples show smaller particle sizes, which endows it with higher photocatalytic activity due to the increased number of reactive sites for surface reactions. Based on the absorption spectra, it can be seen that Bi doping leads to a lower band gap and higher absorbance in a wider visible light region, which contributed to the generation of more photogenerated carriers. Combined with the EIS and Mott-Schottky measurements, GBFO5 also showed high charge transfer efficiency and more negative redox potential over the O2·−/O2 redox potential. The photogenerated electrons were easier to react with oxygen to generate superoxide anions which were the dominant active species for the degradation of MB. Although GBFO9 was also characterized with smaller crystalline particles, the band structure of GBFO5 is more favorable for the redox reactions. All these factors possibly contribute to the optimum performance of Gd0.5Bi0.5FeO3. Experiments exploring deeper insights into the photocatalytic kinetics are still going on.

4. Conclusions

Polycrystalline nanoparticles of Gd1−xBixFeO3 (x = 0, 0.1, 0.3, 0.5, 0.7, 0.9, 1) were successfully synthesized by the sol–gel method. Gd1−xBixFeO3 (x = 0, 0.1, 0.3, 0.5) samples maintained the orthorhombic phase (GdFeO3, space group Pbnm). With further increases of Bi3+, the crystal structure changed to the rhombohedral perovskite (space group R3c). Pure GdFeO3 showed typical antiferromagnetic behavior, and weak ferromagnetism were observed for bismuth-substituted samples with reduced magnetization. The results were explained by the decreased interactions between Fe-O-Fe and distortion of the FeO6 octahedral of the perovskite structure. Moreover, enhanced photocatalytic activities of Gd1-xBixFeO3 were observed, with the optimum performance observed with Gd0.5Bi0.5FeO3. This behavior could be attributed to reduced particle size, increased UV/visible light absorption, decreased band gap and higher charge transfer efficiency. Combined with the characteristics of ferromagnetic nature, good stability and reusability, this system is regarded as an excellent candidate for a magnetic photocatalyst, favoring recyclable and reusable applications for the degradation of organic pollutants.

Author Contributions

Conceptualization, H.S.; methodology, Y.M. (Yudie Ma); software, Y.F.; validation, H.G. and Y.L.; formal analysis, Y.Z.; investigation, Y.M. (Yudie Ma); resources, J.X.; data curation, H.S.; writing—original draft preparation, Y.M. (Yudie Ma); writing—review and editing, H.S.; visualization, Y.Z.; supervision, H.S.; project administration, J.X.; funding acquisition, H.S., J.X. and Y.M. (Yunfeng Ma). All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the NSAF of China (Grant No. U2130124) and National Natural Science Foundation of China (Grant No. 51972213, No. 51802196).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shen, H.; Xian, Q.; Xie, T.; Wu, A.; Wang, M.; Xu, J.; Jia, R.; Kalashnikova, A.M. Modulation of magnetic transitions in SmFeO3 single crystal by Pr3+ substitution. J. Magn. Magn. Mater. 2018, 466, 81–86. [Google Scholar] [CrossRef]
  2. Kumar, A.; Sharma, K.; Thakur, M.; Pathania, D.; Sharma, A. Fabrication of high visible light active LaFeO3/Cl-g-C3N4/RGO heterojunction for solar assisted photo-degradation of aceclofenac. J. Environ. Chem. Eng. 2022, 10, 108098. [Google Scholar] [CrossRef]
  3. Ismael, M.; Elhaddad, E.; Taffa, D.H.; Wark, M. Synthesis of phase pure hexagonal YFeO3 perovskite as efficient visible light active photocatalyst. Catalysts 2017, 7, 326. [Google Scholar] [CrossRef] [Green Version]
  4. Mengying, X.; Peisong, T.; Junwei, C.; Li, F.; Xinyu, X. Preparation of ErFeO3 by Sol-gel Method and Its Visible-light Photocatalytic Activity. Rare Matal. Mat. Eng. 2018, 47, 228–231. [Google Scholar]
  5. Maity, R.; Dutta, A.; Halder, S.; Shannigrahi, S.; Mandal, K.; Sinha, T.P. Enhanced photocatalytic activity, transport properties and electronic structure of Mn doped GdFeO3 synthesized using the sol–gel process. PCCP 2021, 23, 16060–16076. [Google Scholar] [CrossRef] [PubMed]
  6. Shen, H.; Feng, P.; Jiang, G.; Xian, Q. Synthesis of Mn-doped ErFeO3 with enhanced photo and vibration catalytic activities. J. Sol-Gel Sci. Technol. 2020, 95, 230–238. [Google Scholar] [CrossRef]
  7. Li, L.; Wang, F.; Feng, J.; Guo, S.; Xu, M.; Wang, L.; Quan, G. Step-scheme GdFeO3/g-C3N4 heterostructures with outstanding photocatalytic activity. J. Mater. Sci.-Mater. 2021, 32, 16400–16410. [Google Scholar] [CrossRef]
  8. Kebede, M.T.; Devi, S.; Dillu, V.; Chauhan, S. Influence of novel Cd–Ni co-substitution on structural, magnetic, optical and photocatalytic properties of BiFeO3 nanoparticles. J. Alloys Compd. 2022, 894, 162552. [Google Scholar] [CrossRef]
  9. Dumitru, R.; Ianculescu, A.; Păcurariu, C.; Lupa, L.; Pop, A.; Vasile, B.; Surdu, A.; Manea, F. BiFeO3-synthesis, characterization and its photocatalytic activity towards doxorubicin degradation from water. Ceram. Int. 2019, 45, 2789–2802. [Google Scholar] [CrossRef]
  10. Irfan, S.; Zhuanghao, Z.; Li, F.; Chen, Y.X.; Liang, G.X.; Luo, J.T.; Ping, F. Critical review: Bismuth ferrite as an emerging visible light active nanostructured photocatalyst. J. Mater. Res. Technol. 2019, 8, 6375–6389. [Google Scholar] [CrossRef]
  11. Kebede, M.T.; Devi, S.; Tripathi, B.; Chauhan, S.; Dillu, V. Structural transition and enhanced magnetic, optical and photocatalytic properties of novel Ce–Ni co-doped BiFeO3 nanoparticles. Mater. Sci. Semicond. Process. 2022, 152, 107086. [Google Scholar] [CrossRef]
  12. Lan, S.; Yu, C.; Sun, F.; Chen, Y.; Chen, D.; Mai, W.; Zhu, M. Tuning piezoelectric driven photocatalysis by La-doped magnetic BiFeO3-based multiferroics for water purification. Nano Energy 2022, 93, 106792. [Google Scholar] [CrossRef]
  13. Mao, Y.; Qiu, J.; Zhang, P.; Fei, Z.; Bian, C.; Janani, B.J.; Fakhri, A. A strategy of silver Ferrite/Bismuth ferrite nano-hybrids synthesis for synergetic white-light photocatalysis, antibacterial systems and peroxidase-like activity. J. Photoch. Photobiol. A 2022, 426, 113756. [Google Scholar] [CrossRef]
  14. Rosales-González, O.; Sánchez-De, J.F.; Camacho-González, M.A.; Cortés-Escobedo, C.A.; Bolarín-Miró, A.M. Synthesis of magnetically removable photocatalyst based on bismuth doped YFeO3. Mater. Sci. Eng. 2020, 261, 114773. [Google Scholar] [CrossRef]
  15. Li, J.B.; Rao, G.H.; Xiao, Y.; Liang, J.K.; Luo, J.; Liu, G.Y.; Chen, J.R. Structural evolution and physical properties of Bi1−xGdxFeO3 ceramics. Acta Mater. 2010, 58, 3701–3708. [Google Scholar] [CrossRef]
  16. Hu, W.; Chen, Y.; Yuan, H.; Li, G.; Qiao, Y.; Qin, Y.; Feng, S. Structure, magnetic, and ferroelectric properties of Bi1-xGdxFeO3 nanoparticles. J. Phys. Chem. C 2011, 115, 8869–8875. [Google Scholar] [CrossRef]
  17. Mahdikhah, V.; Sheibani, S.; Abdi, M. Visible light photocatalytic performance of La-Fe co-doped SrTiO3 perovskite powder. Opt. Mater. 2020, 102, 109803. [Google Scholar]
  18. Aamir, M.; Bibi, I.; Ata, S.; Jilani, K.; Majid, F.; Kamal, S.; Alwadai, N.; Raza, M.A.S.; Bashir, M.; Iqbal, S.; et al. Ferroelectric, dielectric, magnetic, structural and photocatalytic properties of Co and Fe doped LaCrO3 perovskite synthesized via micro-emulsion route. Ceram. Int. 2021, 47, 16696–16707. [Google Scholar] [CrossRef]
  19. Patel, S.K.S.; Lee, J.H.; Kim, M.K.; Bhoi, B.; Kim, S.K. Single-crystalline Gd-doped BiFeO3 nanowires: R3c-to-Pn21 a phase transition and enhancement in high-coercivity ferromagnetism. J. Mater. Chem. C 2018, 6, 526–534. [Google Scholar] [CrossRef]
  20. Meng, W.; Hu, R.; Yang, J.; Du, Y.; Li, J.; Wang, H. Influence of lanthanum-doping on photocatalytic properties of BiFeO3 for phenol degradation. Chin. J. Catal. 2016, 37, 1283–1292. [Google Scholar] [CrossRef]
  21. Tong, Z.; Yao, Q.; He, Q.; Deng, J.; Wang, J.; Zhou, H.; Rao, G. Effect of microcosmic regulation of unit cells bonding on microwave absorption properties of perovskite structure GdFeO3. Mater. 2021, 20, 101263. [Google Scholar] [CrossRef]
  22. Karoblis, D.; Zarkov, A.; Mazeika, K.; Baltrunas, D.; Niaura, G.; Beganskiene, A.; Kareiva, A. YFeO3-GdFeO3 solid solutions: Sol-gel synthesis, structural and magnetic properties. Solid State Sci. 2021, 118, 106632. [Google Scholar] [CrossRef]
  23. Mangalmurti, K.; Shirbhate, S.; Acharya, S. Modulation of electric and magnetic ordering in GdFeO3 orthoferrites system by Ce-doping. Ferroelectrics 2022, 587, 158–173. [Google Scholar] [CrossRef]
  24. Akhtar, M.N.; Ali, K.; Umer, A.; Ahmad, T.; Khan, M.A. Structural elucidation, and morphological and magnetic behavior evaluations, of low-temperature sintered, Ce-doped, nanostructured garnet ferrites. Mater. Res. Bull. 2018, 101, 48–55. [Google Scholar] [CrossRef]
  25. Raj, R.A.; Meenu, S.; Joseph, A.; Jose, R.; Sajan, D.; Guha, A.; Joy, L.K. Study of the modified magnetic, dielectric, ferroelectric and optical properties in Ni substituted GdFe1−xNixO3 orthoferrites. Nanotechnology 2021, 33, 035705. [Google Scholar]
  26. Panchwanee, A.; Reddy, V.R.; Ajay, G.; Bharathi, A.; Phase, D.M. Study of local distortion and spin reorientation in polycrystalline Mn doped GdFeO3. J. Alloys Compd. 2018, 745, 810–816. [Google Scholar] [CrossRef]
  27. Ujwal, M.P.; Yashas, S.R.; Shivaraju, H.P.; Swamy, N.K. Gadolinium ortho-ferrite interfaced polyaniline: Bi-functional catalyst for electrochemical detection and photocatalytic degradation of acetaminophen. Surf. Interfaces 2022, 30, 101878. [Google Scholar] [CrossRef]
  28. Irfan, S.; Shen, Y.; Rizwan, S.; Wang, H.C.; Khan, S.B.; Nan, C.W. Band-Gap Engineering and Enhanced Photocatalytic Activity of Sm and Mn Doped BiFeO3 Nanoparticles. Am. Ceram. Soc. 2017, 100, 31–40. [Google Scholar] [CrossRef]
  29. Orudzhev, F.F.; Alikhanov, N.M.R.; Ramazanov, S.M.; Sobola, D.S.; Murtazali, R.K.; Ismailov, E.H.; Gasimov, R.D.; Aliev, A.S.; Ţălu, S. Morphotropic Phase Boundary Enhanced Photocatalysis in Sm Doped BiFeO3. Molecules 2022, 27, 7029. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, X.; Mao, H.; Shan, Y. Enhanced photocatalytic behavior and excellent electrochemical performance of hierarchically structured NiO microspheres. RSC Adv. 2014, 4, 35614–35619. [Google Scholar] [CrossRef]
  31. Subramanian, Y.; Ramasamy, V.; Karthikeyan, R.J.; Srinivasan, G.R.; Arulmozhi, D.; Gubendiran, R.K.; Sriramalu, M. Investigations on the enhanced dye degradation activity of heterogeneous BiFeO3-GdFeO3 nanocomposite photocatalyst. Heliyon 2019, 5, e01831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Li, Y.; Zhang, X.; Chen, L.; Sun, H.; Zhang, H.; Si, W.; Wang, W.; Wand, L.; Li, J. Enhanced magnetic and photocatalytic properties of BiFeO3 nanotubes with ultrathin wall thickness. Vacuum 2021, 184, 109867. [Google Scholar] [CrossRef]
  33. Patil, S.B.; Bhojya Naik, H.S.; Nagaraju, G.; Viswanath, R.; Rashmi, S.K. Synthesis of visible light active Gd3+-substituted ZnFe2O4 nanoparticles for photocatalytic and antibacterial activities. Eur. Phys. Plus 2017, 132, 328. [Google Scholar] [CrossRef]
  34. Bera, S.; Ghosh, S.; Shyamal, S.; Bhattacharya, C.; Basu, R.N. Photocatalytic hydrogen generation using gold decorated BiFeO3 heterostructures as an efficient catalyst under visible light irradiation. Sol. Energy Mater. Sol. Cells 2019, 194, 195–206. [Google Scholar] [CrossRef]
  35. Wei, K.; Faraj, Y.; Yao, G.; Xie, R.; Lai, B. Strategies for improving perovskite photocatalysts reactivity for organic pollutants degradation: A review on recent progress. Chem. Eng. J. 2021, 414, 128783. [Google Scholar] [CrossRef]
  36. Nkwachukwu, O.V.; Muzenda, C.; Ojo, B.O.; Zwane, B.N.; Koiki, B.A.; Orimolade, B.O.; Nkosi, D.; Mabuba, N.; Arotiba, O.A. Photoelectrochemical degradation of organic pollutants on a La3+ doped BiFeO3 perovskite. Catalysts 2021, 11, 1069. [Google Scholar] [CrossRef]
  37. Nunocha, P.; Kaewpanha, M.; Bongkarn, T.; Phuruangrat, A.; Suriwong, T. A new route to synthesizing La-doped SrTiO3 nanoparticles using the sol-gel auto combustion method and their characterization and photocatalytic application. Mater. Sci. Semicond. Process. 2021, 134, 106001. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) Gd1−xBixFeO3 (x = 0, 0.1, 0.3, 0.5, 0.7) annealed at 700 °C; (c) Gd1−xBixFeO3 (x = 0.9, 1.0) annealed at 550 °C; (b,d) enlarged view of 2θ from 31° to 35°.
Figure 1. XRD patterns of (a) Gd1−xBixFeO3 (x = 0, 0.1, 0.3, 0.5, 0.7) annealed at 700 °C; (c) Gd1−xBixFeO3 (x = 0.9, 1.0) annealed at 550 °C; (b,d) enlarged view of 2θ from 31° to 35°.
Magnetochemistry 09 00045 g001aMagnetochemistry 09 00045 g001b
Figure 2. SEM image of (a) GFO calcined at 700 °C; (b) GBFO5 calcined at 700 °C; (c) GBFO7 calcined at 700 °C; (d) GBFO9 calcined at 550 °C; (eh) statistical estimation of the particle size distribution of GFO, GBFO5, GBFO7 and GBFO9.
Figure 2. SEM image of (a) GFO calcined at 700 °C; (b) GBFO5 calcined at 700 °C; (c) GBFO7 calcined at 700 °C; (d) GBFO9 calcined at 550 °C; (eh) statistical estimation of the particle size distribution of GFO, GBFO5, GBFO7 and GBFO9.
Magnetochemistry 09 00045 g002
Figure 3. X-ray photoelectron spectroscopy (XPS) spectra of (a) Bi 4f, (b) Fe 2p and (c) O 1s for the Gd0.5Bi0.5FeO3 sample.
Figure 3. X-ray photoelectron spectroscopy (XPS) spectra of (a) Bi 4f, (b) Fe 2p and (c) O 1s for the Gd0.5Bi0.5FeO3 sample.
Magnetochemistry 09 00045 g003
Figure 4. Room temperature magnetic hysteresis loops of (a) GFO, GBFO1 and GBFO3 samples; (b) GBFO5, GBFO7 and GBFO9 samples.
Figure 4. Room temperature magnetic hysteresis loops of (a) GFO, GBFO1 and GBFO3 samples; (b) GBFO5, GBFO7 and GBFO9 samples.
Magnetochemistry 09 00045 g004
Figure 5. UV-vis absorption spectra of Gd1-xBixFeO3 nanoparticles; the inset shows the calculation of the band gaps.
Figure 5. UV-vis absorption spectra of Gd1-xBixFeO3 nanoparticles; the inset shows the calculation of the band gaps.
Magnetochemistry 09 00045 g005
Figure 6. (a) Photocatalytic degradation efficiencies of MB using Gd1−xBixFeO3 as a catalyst; (b) pseudo-first-order kinetics fitting data for the photodegradation of MB over Gd1−xBixFeO3 samples.
Figure 6. (a) Photocatalytic degradation efficiencies of MB using Gd1−xBixFeO3 as a catalyst; (b) pseudo-first-order kinetics fitting data for the photodegradation of MB over Gd1−xBixFeO3 samples.
Magnetochemistry 09 00045 g006
Figure 7. (a) Complex impedance spectra (Z’ vs. Z’’) and (b) Mott-Schottky plots of Gd1−xBixFeO3 samples at room temperature.
Figure 7. (a) Complex impedance spectra (Z’ vs. Z’’) and (b) Mott-Schottky plots of Gd1−xBixFeO3 samples at room temperature.
Magnetochemistry 09 00045 g007
Figure 8. Schematic diagram of the band structure and proposed mechanism of the photocatalytic process.
Figure 8. Schematic diagram of the band structure and proposed mechanism of the photocatalytic process.
Magnetochemistry 09 00045 g008
Figure 9. (a) Reusability of Gd0.5Bi0.5FeO3 nanoparticles in the degradation of MB under UV light irradiation over five cycles; (b) XRD patterns of the Gd0.5Bi0.5FeO3 sample before and after the photocatalytic experiment.
Figure 9. (a) Reusability of Gd0.5Bi0.5FeO3 nanoparticles in the degradation of MB under UV light irradiation over five cycles; (b) XRD patterns of the Gd0.5Bi0.5FeO3 sample before and after the photocatalytic experiment.
Magnetochemistry 09 00045 g009
Table 1. Refined structural parameters for Gd1−xBixFeO3 (x = 0, 0.1, 0.3, 0.5, 0.7, 0.9, 1.0) samples.
Table 1. Refined structural parameters for Gd1−xBixFeO3 (x = 0, 0.1, 0.3, 0.5, 0.7, 0.9, 1.0) samples.
GFOGBFO1GBFO3GBFO5GBFO7GBFO9BFO
Space Group PbnmPbnmPbnmPbnmPbnm
(Wt: 64.68%)
R3c
(Wt: 35.32%)
R3cR3c
Lattice parameters a (Å)5.321 (3)5.349 (3)5.373 (3)5.394 (3)5.398 (3)5.635 (3)5.598 (3)5.578 (3)
b (Å)5.578 (3)5.605 (3)5.620 (3)5.629 (3)5.609 (3)5.635 (3)5.598 (3)5.578 (3)
c (Å)7.632 (3)7.673 (3)7.711 (3)7.747 (3)7.753 (3)13.503 (3)13.723 (3)13.865 (3)
V (Å)226.568 (0)230.062 (0)232.846 (0)235.243 (0)234.774 (0)371.327 (0)372.387 (0)373.544 (0)
α = β90°90°90°90°90°90°90°90°
γ90°90°90°90°90°120°120°120°
Bond lengthGd/Bi-O1 (Å)2.264 (20)2.275 (19)2.283 (11)2.290 (7)2.287 (19)2.292 (29)2.304 (18)2.312 (37)
Gd/Bi-O2 (Å)2.505 (16)2.518 (15)2.528 (9)2.538 (5)2.537 (14)2.539 (5)2.526 (14)2.519 (6)
Fe-O1 (Å)2.013 (14)2.024 (13)2.033 (8)2.0425 (6)2.044 (18)2.023 (26)2.023 (13)2.025 (35)
Fe-O2 (Å)2.010 (14)2.020 (14)2.027 (8)2.033 (5)2.030 (13)
Bond angleFe-O1-Fe (°)142.881 (3)142.888 (3)142.910 (2)142.943 (12)142.968 (4)156.087 (1)156.156 (3)156.200 (1)
Fe-O2-Fe (°)147.042 (1)147.037 (1)147.015 (1)146.985 (5)146.947 (2)
Other parameterdensity (g·cm−3)7.6547.5387.4487.3727.3878.3948.369
Statistical fitχ21.1521.0391.3943.1791.8621.9811.78
wRp (%)4.804.745.869.447.648.458.48
Rp (%)3.823.744.516.215.946.786.52
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

Ma, Y.; Shen, H.; Fang, Y.; Geng, H.; Zhao, Y.; Li, Y.; Xu, J.; Ma, Y. Effect of Bismuth on the Structure, Magnetic and Photocatalytic Characteristics of GdFeO3. Magnetochemistry 2023, 9, 45. https://doi.org/10.3390/magnetochemistry9020045

AMA Style

Ma Y, Shen H, Fang Y, Geng H, Zhao Y, Li Y, Xu J, Ma Y. Effect of Bismuth on the Structure, Magnetic and Photocatalytic Characteristics of GdFeO3. Magnetochemistry. 2023; 9(2):45. https://doi.org/10.3390/magnetochemistry9020045

Chicago/Turabian Style

Ma, Yudie, Hui Shen, Yating Fang, Heyan Geng, Yu Zhao, Yasheng Li, Jiayue Xu, and Yunfeng Ma. 2023. "Effect of Bismuth on the Structure, Magnetic and Photocatalytic Characteristics of GdFeO3" Magnetochemistry 9, no. 2: 45. https://doi.org/10.3390/magnetochemistry9020045

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