Five samples of xNiFe2O4/(1 − x)Ba0.8Sr0.2TiO3 (where x = 0, 0.1, 0.2, 0.3, and 0.4) multiferroic nanocomposites have been successfully fabricated using ball milling combined with heat treatment in a short time. X-ray diffraction patterns indicated the coexistence of two phases, namely, NiFe2O4 (NFO) and Ba0.8Sr0.2TiO3 (BSTO). The average grain size obtained is about 50–100 nm, and NFO and BSTO phases are evenly distributed in the samples. With an increase in the content of NFO, the values characterizing the ferroelectric and ferromagnetic properties improve significantly. Furthermore, the bandgap energy (Eg) value was also strongly reduced. The results on the degradability of methylene orange show that the apparent first-order rate of composites containing NFO with x = 0.4 was found to be k = 0.0228 min−1, which is significantly higher than that of pure BSTO (k = 0.0166 min−1), suggesting that NFO/BSTO multiferroic nanocomposites could be considered as candidate photocatalysts for the degradation of pollutants.

Today, the research and application of multiferroics are increasingly developing, attracting many scientists worldwide. Although multiferroic materials have simultaneous existence of ferroelectric (or anti-ferroelectric) and ferromagnetic (or anti-ferromagnetic) properties, especially electromagnetic correlation in materials, the electrical properties can be controlled by magnetic fields, and vice versa.1–3 Multiferroic composites possess much stronger electromagnetic effects than structural single-phase multiferroics. Thus, they attract many application studies in logical memories, electromagnetic memories, magnetic field sensors, energy harvesters, magnetoelectric resonators, and the read head.3–5 Besides, multiferroic composites are also considered as potential materials for water pollution treatment, dye discoloration, and decomposition of organic compounds under light ultraviolet, sunlight, or ultrasonic vibration.6,7

Currently, dyes have properties that resist the effects of detergents, water, and sunlight [methyl orange (MO), Rhodamine B (RhB), Methylene Blue (MB), etc.], so they are increasingly difficult to decompose. Azo dyes such as MO account for over 50% of global dye production, equivalent to 700 000 tons per year. This dye is used in the pharmaceutical, textile, and paper industries. However, the remaining dye discharged into the environment accounts for about 15% of the total.8–10 MO dye is highly toxic, so it can pollute water sources, pollute the environment, destroy ecosystems, and even cause cancer among humans.10,11 Therefore, the treatment of water contaminated with MO dyes is considered urgent and necessary. Among the current treatment methods, photocatalysis is the dominant method to treat water contaminated by dyes and organic chemical products.

Photocatalysis is a process where a semiconductor absorbs photons from ultraviolet rays or sunlight, generating electrons and holes that can induce chemical reactions or generate an electric current. To enhance the efficiency of photocatalytic reactions, it is crucial to select materials with a bandgap energy that can absorb ultraviolet light or sunlight and exhibit low electron–hole recombination rates. Recent studies have shown that ABO3 (BaTiO3 and SrTiO3) compounds possess ferroelectric properties and have potential applications in photocatalytic degradation of synthetic dyes as noble metal-free catalysts for oxygen production at low cost.12 ABO3 nanoparticles have also been prepared and tested as photocathode catalytic components in microbial fuel cells.13 Among these, BaTiO3 (BTO) is a ferroelectric material with a simple and flexible structure, making it suitable for various applications. One of its notable properties is its ability to decompose organic pollutants, making it environmentally friendly.14–17 BTO is known to be an n-type semiconductor, with valence band (VB) positions at ∼2.31 eV and conduction band (CB) positions approximately at −0.83 eV. The bandgap energy (Eg) of BTO is typically in the range of 3.0–3.3 eV.18 

The photocatalytic reaction efficiency depends on the electrical polarization of the material as improved electrical polarization results in better separation of charge carriers inside the bulk or on the surface of the semiconductor photocatalyst. Therefore, improving the electrical polarization of BTO-based ferroelectric materials is an important issue. Recent studies have doped Sr into the sublattice of BTO to make Ba1−ySryTiO3 compounds. As previously reported, Ba1−ySryTiO3 is a ferroelectric perovskite ceramic with a high dielectric constant, exhibiting excellent dielectric properties and good electrical polarization, making it suitable for various applications, such as supercapacitors, ferroelectric memory, piezoelectric sensors, and gas sensors.19–22 

Ba1−ySryTiO3 is essentially a perovskite-based solid solution of BTO and SrTiO3 (STO), and factors such as the Ba/Sr ratio, phase purity, and particle size directly affect the characteristic properties of Ba1−ySryTiO3 materials.21–26 As the concentration of Sr substituting for Ba gradually increases, the Curie temperature of the material is gradually reduced.27–29 Among these, Ba0.8Sr0.2TiO3 (BSTO) is considered typical. The crystal structure of BSTO exhibits high flexibility, allowing for sensitive response to external stress under conditions near room temperature or below 120 °C.25,30 It has also shown that the energy band structure and CB and VB positions of BSTO are similar to those of BaTiO3.22,30,31 In 2018, Yuan et al.30 demonstrated that under the influence of ultrasonic vibration for 120 min, Ba1−ySryTiO3 nanowires could completely decompose the MO dyes. As the Ba1−ySryTiO3 crystal is excited by ultraviolet light or mechanical stress, electrons on the CB can be transferred to the solution and reduce O2 dissolved in water to generate •O2− radicals. At the same time, the holes on the VB will oxidize OH to generate •OH radicals.30 Furthermore, Kanamadi et al.25 demonstrated that the composites formed by the combination of ferrite spinel and BSTO exhibit excellent magnetic and dielectric properties. As excited with reasonable energy, at the same time, two types of charge carriers can appear in the material, specifically electrons and holes. In addition, on the octahedral sites, the electron exchange between Fe2+ and Fe3+ ions contributes to the local displacement in the direction of the external electric field and the dielectric polarization in ferromagnetic materials.25 Besides, MFe2O4 (M = Fe, Ni, Co) ferrite spinels are semiconductors with various advanced properties, such as good chemical stability, good magnetism, narrow bandgap, and visible light operation with high potential electrical performance.31–33 Dang et al.15 successfully decomposed MB synthetic dyes using a CoFe2O4–BaTiO3 composite. With a photocatalyst concentration of 0.4 g/l and under the effect of UV light, the decomposition efficiency reached 72% after irradiation for 180 min.15 Among magnetic semiconductors, NiFe2O4 (NFO) is considered a typical compound. NFO has an inverse spinel structure and exhibits properties such as high permeability, low anisotropy, large resistivity, and high piezomagnetism.34–37 It is a chemically stable and photocatalytic semiconductor with a narrow bandgap, corresponding to the CB and VB positions at approximately −0.6 and 1.28 eV, respectively.38,39 However, due to the rapid recombination of photoelectron–hole pairs, the photocatalytic potential of NFO is limited. The successful combination of NFO with BSTO to create NFO/BSTO multiferroic nanocomposites with strong photocatalytic ability is considered an innovative idea. Changing the NFO content in the composites can result in positive changes in the physical properties of NFO/BSTO multiferroics. These results have significance for the development of future research-oriented applications of multi-materials in the field of water pollution treatment contaminated with dyes and organic compounds by photocatalysis and for the recovery of photocatalyst materials using the magnetic field.

Multiferroic composites with component formula xNFO/(1 − x)BSTO (NFO/BSTO) were synthesized using raw chemicals, including high purity (≥ 99.9%) BaCO3, SrCO3, TiO2, NiO, and Fe2O3 powders. Initially, the particulate composites with BSTO ferroelectric and NFO ferromagnetic phases were fabricated separately using a Spex-8000D machine. The respective raw chemical powder mixtures of BSTO and NFO phases were ground and mixed separately in air for 10 h at 875 rpm. The powder mixtures obtained after grinding and mixing were then pressed into pellets under a pressure of 7000 kg/cm2. The pellets corresponding to each phase of the materials were sintered separately in the air at 1250 and 1000 °C to successfully fabricate BSTO and NFO samples, respectively. Subsequently, the NFO and BSTO samples were subjected to high-energy ball milling for 10 h in the air, resulting in the formation of fine powders as the final products.

Next, the NFO and BSTO powders were weighed according to the molar ratio of x = 0, 0.1, 0.2, 0.3, and 0.4 to create the xNFO/(1 − x)BSTO (NFO/BSTO) composites denoted as BSTO, BSNF1, BSNF2, BSNF3, and BSNF4, respectively. After weighing, the powder mixture was thoroughly mixed and then pressed into pellets under a pressure of 7000 kg/cm2. These pellets were then incubated for 6 h at 800 °C to form solid pellets of the same size. The structure of all samples was analyzed using x-ray diffraction (XRD) patterns obtained from an Equinox 5000 device from Thermo Fisher Scientific, using Cu-Kα radiation (λ = 1.54056 Å) with a voltage of 40 kV and a current of 40 mA. An acquisition time of 45 min per sample and 2θ step size of 0.015° were used. Using field emission scanning electron microscopy (FE-SEM) S4800 equipment (Hitachi), the morphology and particle size of the samples were determined. The composition of chemical elements and the distribution of substances in samples were analyzed by Aztec Energy dispersive spectroscopy (EDS) equipment from Oxford Instruments (UK). The characteristic ferromagnetic and ferroelectric properties of the samples were evaluated through the M(H) and P(E) hysteresis curves, which were measured on a vibrating sample magnetometer and a Precision LC II model 609 system at room temperature, respectively. The typical optical properties of the samples were specifically determined based on the analysis of the UV–Vis spectrum carried out on the Cary 5000 UV–Vis–near-infrared spectrophotometer instrument. MO solution was used to evaluate the photocatalytic performance of NFO/BSTO nanocomposites. Accordingly, 0.2 g of powder from each NFO/BSTO sample was stirred in 1000 ml of MO (0.2 mg/l, PH = 7) solution in the dark for 60 min to ensure equilibrium. After that, an ultraviolet lamp with a luminous flux intensity of 22 mJ/cm2 and a wavelength of 254 nm (model 30E, OAI, USA) was used to stimulate the photocatalytic reaction at different irradiation times.

Figures 1(a)1(e) present the XRD patterns of all the NFO/BSTO composites. It shows that the BSTO (x = 0) sample [Fig. 1(a)] exhibits XRD peaks at 2θ ≈ 22.94°, 32.19°, 32.21°, 39.56°, 45.72°, 45.88°, 51.61°, 56.76°, and 66.47°, which belong to the Ba0.8Sr0.2TiO3 phase according to ICDD card No. 96-151-2122. These peaks were denoted by solid triangles (▼) and indexed by the Miller indices (100), (101), (110), (111), (002), (200), (201), (211), and (202), respectively. On the XRD patterns of the remaining samples (x = 0.1–0.4), as shown in Figs. 1(b)1(e), some additional peaks appear at positions of 2θ ≈ 31.03°, 36.32°, 37.94°, 43.82°, 47.21°, 54.11°, 57.64°, and 63.09°, which are denoted by empty triangles (). These peaks are completely consistent with ICDD card No. 00-066-0245 for NiFe2O4 phase and indexed by the Miller indices (220), (311), (222), (400), (331), (422), (333), and (440), respectively. As the NFO concentration increases (x increases), the intensity of the XRD peaks related to the NFO phase tends to increase while the peaks related to the BSTO phase show a decreasing trend. These behaviors can be seen more clearly in Fig. 1(f), which shows the changes in the two strongest XRD peaks obtained at 2θ ≈ 32.19° and 36.32° corresponding to the BSTO and NFO phases with x (NFO content). Based on XRD data, the lattice parameters for both phases have been calculated. It shows that the BSTO phase belongs to a tetragonal structure (space group of P4mm) with a = 3.973 Å, c = 3.986 Å, and V = 62.918 Å3 and that the NFO phase belongs to a cubic structure (space group of Fd-3m) with a = 8.346 Å and V = 581.347 Å3. These values are almost unchanged with different NFO concentrations. Besides, the appearance of the strange peaks was not observed at all. These suggest that the BSTO and NFO phases could co-exist in the samples, and the heat treatment did not introduce a new phase. In addition, the average crystallite sizes (⟨D⟩) of the BSTO and NFO phases in the samples were determined according to the Scherrer formula,40,
D=kλβcosθ,
(1)
where β, θ, and k are the full width at half maximum, the diffraction angle of the XRD peak, and the particle shape factor (using k = 0.9), respectively. In this work, the full width half maximum caused by the nano-size effect (β) was corrected using the relation41,
β2=(βObs)2(βIns)2;
(2)
herein, βObs and βIns are the full width at half maximum of the XRD peak estimated for the sample and the instrumental broadening, respectively. To determine the value of βIns, a standard sample of Si powder was used, and the full width at half maximum of (111) peak of Si powder (β(Si111)) was assigned to βIns.41 Accordingly, the average crystallite sizes calculated for the BSTO and NFO phases are ∼35 and 29 nm, respectively, and they are virtually unaffected by NFO concentration.
FIG. 1.

XRD patterns of the powdered NFO/BSTO composites (a) x = 0; (b) x = 0.1; (c) x = 0.2; (d) x = 0.3; (e) x = 0.4. (f) The XRD intensity of the two strongest peaks corresponding to the BSTO and NFO phases vs with x.

FIG. 1.

XRD patterns of the powdered NFO/BSTO composites (a) x = 0; (b) x = 0.1; (c) x = 0.2; (d) x = 0.3; (e) x = 0.4. (f) The XRD intensity of the two strongest peaks corresponding to the BSTO and NFO phases vs with x.

Close modal

SEM images present the morphology and size of the particles in all samples. The SEM images of two representative samples of BSTO and BSNF4 are shown in Figs. 2(a) and 2(b), respectively. Based on the results of determining the particle size in various regions of the samples, the average value and particle size distribution were constructed and are presented representatively in the insets of Figs. 2(a) and 2(b). We can see that the particle sizes of the particles in the samples are distributed over a quite wide range with an average value of ∼69 nm. Because the sample conducts electricity poorly, the grain boundaries are unclear, and clumps between the particles occur. Besides, the particle size in the sample did not change with the sample’s composition, as well as the content of NFO. However, the particle size determined from SEM images is significantly larger than the crystallite size, suggesting that each particle contains more than one crystal.

FIG. 2.

SEM images (a) and (b) and EDX spectra (c) and (d) of BSTO and BSNF4 samples, respectively.

FIG. 2.

SEM images (a) and (b) and EDX spectra (c) and (d) of BSTO and BSNF4 samples, respectively.

Close modal

The energy dispersive x-ray (EDX) spectra demonstrate the presence of Ba, Sr, Ti, Ni, O, and Fe elements. These elements are all derived from the original raw chemicals, and no other foreign elements appear. This is demonstrated by the EDX spectra of two representative samples of BSTO and BSNF4, as presented in Figs. 2(c) and 2(d). The atomic percentages of chemical elements in the samples are quite consistent with the expected sample composition. Besides, chemical elements such as Ba, Sr, Ti, O, Fe, and Ni are also distributed quite uniformly as desired in the samples. This is shown in the EDX mappings for two representative samples of BSTO and BSNF4 in Figs. 3 and 4, respectively.

FIG. 3.

EDX element mapping images of the BSTO sample.

FIG. 3.

EDX element mapping images of the BSTO sample.

Close modal
FIG. 4.

EDX element mapping images of the BSNF4 sample.

FIG. 4.

EDX element mapping images of the BSNF4 sample.

Close modal

Figure 5 depicts the hysteresis M(H) curves for all samples at room temperature. It shows that all the samples exhibit ferromagnetic characteristics except the BSTO sample, which is a paramagnetic compound with very low value of magnetization. However, the magnetization of the NFO/BSTO composites increases rapidly with NFO concentration. For NFO content increased in the range corresponding to x = 0.1–0.4, the value of the saturation magnetization (Ms) increases in the range of 4.7–17.5 emu/g, and the value of the residual magnetization (Mr) increases in the range of 0.36–1.78 emu/g, see the inset in Fig. 5 and Table I. These can be explained by the vital contribution of the magnetic moments of the ferromagnetic phase of NFO. The results of this study are considered to be relatively consistent with the results of previous studies on this family of multiferroics.6,7,42–44 Improving the magnetization of NFO/BSTO composites makes them more useful for recovery and reuse by using external magnetic fields.15 

FIG. 5.

Hysteresis curves of NFO/BSTO samples at room temperature. The inset presents the dependence on the NFO content of Ms and Mr.

FIG. 5.

Hysteresis curves of NFO/BSTO samples at room temperature. The inset presents the dependence on the NFO content of Ms and Mr.

Close modal
TABLE I.

The values of Ms, Mr, Hc, Ps, Pr, and Ec obtained at room temperature for NFO/BSTO composites.

SamplesMs (emu/g)Mr (emu/g)Hc (Oe)Ps (μC/cm2)Pr (μC/cm2)Ec (kV/cm)
BSTO 0.58 ⋯ ⋯ 0.044 0.012 2.4 
BSNF1 4.8 0.36 88 0.064 0.028 3.6 
BSNF2 8.3 0.95 96 0.076 0.034 3.6 
BSNF3 13.1 1.38 67 0.089 0.038 3.8 
BSNF4 16.9 1.78 72 0.12 0.068 4.7 
SamplesMs (emu/g)Mr (emu/g)Hc (Oe)Ps (μC/cm2)Pr (μC/cm2)Ec (kV/cm)
BSTO 0.58 ⋯ ⋯ 0.044 0.012 2.4 
BSNF1 4.8 0.36 88 0.064 0.028 3.6 
BSNF2 8.3 0.95 96 0.076 0.034 3.6 
BSNF3 13.1 1.38 67 0.089 0.038 3.8 
BSNF4 16.9 1.78 72 0.12 0.068 4.7 

Figures 6(a)6(e) show the ferroelectric hysteresis P(E) loops of NFO/BSTO composites measured at room temperature in a maximum external voltage of 1 kV at a frequency of 50 Hz. It is clear that these P(E) hysteresis loops exhibit ferroelectric characteristics. As the NFO concentration increases, these P(E) curves are widened and the value of the electrical polarization increases significantly. For composites with high NFO content (x ≥ 0.3), there is an attenuation of the polarization in high electric field, which may be due to the much lower resistance of the NFO phase than the BSTO phase and could be related to the energy loss due to leakage currents in these composites.45–48 The ferroelectricity of NFO/BSTO composites is improved by increasing the value of the residual polarization (Pr), the saturation polarization (Ps), and the coercive field (Ec), as shown in Fig. 6(f). Here, the Pr value increased more than five times from 0.012 to 0.064 μC/cm2, the Pr value increased more than three times from 0.044 to 0.16 μC/cm2, and the Ec value increased nearly two times from 2.3 to 4.5 kV/cm for the NFO increasing from x = 0 to x = 0.4, respectively. Interestingly, the Pr and Ps values obtained for the BSNF4 sample (x = 0.4) are five and three times higher than those of the pure BSTO compound, respectively. Table I also lists the values of the characteristic parameters of NFO/BSTO multiferroic nanocomposites, including Ms, Mr, Hc, Ps, Pr, and Ec. It shows that composites with x = 0.1–0.4 are multiferroics with rather small Hc (below 100 Oe). The addition of the NFO phase not only improves the ferromagnetic properties but also improves the ferroelectric properties in the composites. Figure 7(a) shows the UV–Vis absorption spectra of NFO/BSTO composites. It implies that the absorption coefficient tends to increase gradually and the absorption region of the samples expands toward the longer wavelengths. Accordingly, their bandgap energy (Eg) values can be estimated as shown in Fig. 7(b). The Eg value is strongly reduced from 3.05 to 2.76 eV, corresponding to an increase in x from x = 0 to x = 0.4.

FIG. 6.

P(E) curves measured at room temperature in a maximum external voltage of 1 kV at a frequency of 50 Hz for BSTO (a), BSNF1 (b), BSNF2 (c), BSNF3 (d), and BSNF4 (e) samples and the dependence of Pr, Ps, and Ec on the NFO content (f).

FIG. 6.

P(E) curves measured at room temperature in a maximum external voltage of 1 kV at a frequency of 50 Hz for BSTO (a), BSNF1 (b), BSNF2 (c), BSNF3 (d), and BSNF4 (e) samples and the dependence of Pr, Ps, and Ec on the NFO content (f).

Close modal
FIG. 7.

UV–Vis diffuse reflectance spectra (a) and bandgap energies (b) of the BST0–BSNF4 samples.

FIG. 7.

UV–Vis diffuse reflectance spectra (a) and bandgap energies (b) of the BST0–BSNF4 samples.

Close modal

The value of Eg for NFO/BSTO is reduced corresponding to the gradual increase in the NFO content in the component from x = 0 to x = 0.4, which can be explained as follows: Ni(II) and Fe(III) ions co-exist in NFO/BSTO composites. The electrons in the d subshell of Ni2+ and Fe3+ are distributed across the t2g and eg orbitals. The energy level of the VB in NFO is higher than that of BSTO, while the value of the CB energy in NFO is lower than that of BSTO. As a result, a new energy level of Eg is formed. Internal electric fields have been created in ferroelectric materials; in the presence of ferromagnetism, transition metal 3d states in the Fermi energy band appear, reducing the band-gap and allowing for easier capture of energy phonons.6,33 Under irradiation by ultraviolet light, the energy required for electron transmission was reduced, and the maximum absorption of NFO/BSTO composites shifted toward a longer wavelength than that of BSTO. This effect is possible because the electrons can either move from the VB to the characteristic energy band of Ni(II) and Fe(III) or they can be transferred from the typical energy band of Ni(II) and Fe(III) to the CB.6 The enhancement in the catalytic ability of NFO/BSTO composites may be closely related to their electrical polarization, which is an important factor in promoting their photocatalytic ability. Reducing the electron–hole recombination and narrowing the bandgap are essential to increase the photocatalytic efficiency. These results are considered reasonable when compared with previous research findings, indicating that NFO/BSTO composites are more promising materials for photocatalytic applications.33 

To investigate the photocatalytic behavior of NFO/BSTO composites as a function of NFO content, we conducted a study where synthetic methyl orange (MO) dye was exposed to ultraviolet (UV) light. By monitoring the degradation of MO in water in the presence of the NFO/BSTO composite under UV irradiation, the photocatalytic properties of NFO/BSTO were systematically investigated. Figure 8 presents the absorption spectra of MO dye over different UV irradiation times for two typical samples, BSTO [Fig. 8(a)] and BSTO2 [Fig. 8(b)], at room temperature. Similar to other photocatalysts, the intensity of the MO absorption peak decreases with increasing irradiation time. We found that the photocatalytic degradation of MO followed the order BSTO < BSNF1 < BSNF2 < BSNF3 < BSNF4.

FIG. 8.

UV-visible absorption spectrum of MO dyes after irradiation with the participation of samples BST0 (a) and BSNF2 (b) at room temperature.

FIG. 8.

UV-visible absorption spectrum of MO dyes after irradiation with the participation of samples BST0 (a) and BSNF2 (b) at room temperature.

Close modal
Figure 9 shows the rate of MO degradation over time under UV irradiation with the participation of the NFO/BSTO composites at room temperature. The disappearance rate (DR) of MO was determined using the following formula:6,
DR(%)=C0CC0×100%,
(3)
where DR is the percent of MO removal by photocatalytic reaction and C0 and C are the MO’s initial and remaining concentrations at the measurement time, respectively. The results indicated that with the pure BSTO photocatalyst, the DR of MB degradation was 69.7% after 70 min of irradiation. However, when BSTO was replaced by NFO/BTO composites, we observed a significant improvement in the effective treatment of MO. After 70 min of irradiation, the DR of MB was ∼80% for x values ranging from 0.2 to 0.4. The photocatalytic efficiency of the NFO/BSTO composite fighting increases with the gradual increase in the NFO content in the component. This improvement in efficiency may be attributed to the enhancement of the ferroelectric properties of the material. Specifically, the increased polarity of the material can prevent recombination between electrons and holes, promote the formation of oxidizing agents, and ultimately increase the photocatalytic efficiency of NFO/BSTO composites.33 
FIG. 9.

Dependence of the ratio C/C0 after irradiation times in the presence of NFO/BSTO composites. The inset shows the value of the disappearance rate (DR) of samples after processing times of 30, 50, and 70 min.

FIG. 9.

Dependence of the ratio C/C0 after irradiation times in the presence of NFO/BSTO composites. The inset shows the value of the disappearance rate (DR) of samples after processing times of 30, 50, and 70 min.

Close modal
By investigating the behavior of photocatalytic degradation, the photocatalytic performance of NFO/BSTO composites will be evaluated and compared in detail. Photocatalysis was evaluated based on the photocatalytic rate by a pseudo-first-order equation as follows:6 
lnCC0=kpt,
(4)
where kp is the reaction rate constant and t is the reaction time. By plotting −ln(C0/C) vs reaction time, kp is defined as the value of the slope of the straight line, as shown in Fig. 10. The extracted slope value of kp = 0.0166 min−1 (R2 = 0.98), 0.0189 min−1 (R2 = 0.95), 0.0224 min−1 (R2 = 0.96), 0.0226 min−1 (R2 = 0.96), 0.0228 min−1 (R2 = 0.96) corresponds to samples BSTO, BSNF1, BSNF2, BSNF3, and BSNF4, respectively, which indicated that the presence of NFO has enhanced the photocatalytic activity of composites.
FIG. 10.

Dependence of −ln(C0/C) on UV irradiation time; the inset presents the dependence NFO (x) of the apparent first-order rate constant (kp).

FIG. 10.

Dependence of −ln(C0/C) on UV irradiation time; the inset presents the dependence NFO (x) of the apparent first-order rate constant (kp).

Close modal

The photocatalytic mechanisms that can account for the obtained results of the photocatalysis of NFO/BSTO have been described schematically in Fig. 11. The presence of NFO in the NFO/BSTO composite plays a crucial role in improving the photocatalytic activity. The process of MO decomposition by the photocatalyst under UV radiation can be summarized as follows: Under the effect of ultraviolet light, electrons are excited with enough energy and move to the conduction band from the valence band. Electron and hole pairs will be formed after this process. The generated electrons and holes are transferred to the surface by excluding the recombined electron–hole pairs. On the surface of the semiconductor, superoxide anion (•O2−) and hydroxyl (•OH) radicals will be formed by the reaction between electrons and holes with oxygen and water in the air.6,7,15 In addition, the electrical polarization of the NFO/BSTO composite may also contribute to the improved photocatalytic activity by preventing electron–hole recombination and promoting the formation of oxidizing agents. Overall, the proposed mechanism suggests that the NFO/BSTO composite can effectively utilize UV light to generate active species that decompose organic pollutants into safe materials, which can have important applications in environmental remediation and pollution control.

FIG. 11.

Possible photocatalytic mechanism for degradation of MO over the NFO/BSTO composite under ultraviolet light irradiation.

FIG. 11.

Possible photocatalytic mechanism for degradation of MO over the NFO/BSTO composite under ultraviolet light irradiation.

Close modal

In order for us to see the efficiency when using NFO/BSTO composites as photocatalysts to decompose MO synthetic dyes, statistical results of the photocatalytic processing ability of similar materials6,7,15,30,32,33,47,49,50 are shown in Table II. It shows that the efficiency of MO dye treatment by using NFO/BSTO composites under ultraviolet light is compared to that reported for various materials. In 2018, Lassoued et al.32 succeeded in decomposing MO synthetic dyes by using NFO.32 The MO solution was photocatalytic and decomposed at 10 mg/l. The solution has a concentration of the photocatalyst of 1.0 g/l. Using visible light with a wavelength in the region from 350 to 800 nm emitted by a tungsten filament lamp, the decomposition rate of MO was achieved at about 83.2% after irradiation for 140 min.32 Besides, with the same concentration of the photocatalyst and MO synthetic dye, Yu et al.50 used the photocatalyst 0.4BaTiO3/0.6CuO to degrade ∼90% of MO synthetic dyes in 90 min under light from a UV source with a power of 200 W and immersed in an ultrasonic cleaner.50 Especially when performing MO dye degradation under UV light, corresponding to a lower concentration of the photocatalyst, Mishra et al. succeeded in decomposing MO synthetic dyes by using CoFe2O4/Fe3O4 nanoparticles.33 The concentration of the catalyst used was 0.2 g/l, and the decomposition rate of MO reached about 93% under ultraviolet light for 300 min.8,38 Eskandari et al.6 used the CoFe2O4/SrTiO3 photocatalyst with a concentration of 0.4 g/l to successfully decompose the azo dye solution under the light generated by the UV light source and visible light source. The decomposition rate of the azo dyes solution reached ∼70% after irradiation of 180 min.6 The results confirm that the NFO/BSTO composite is a new photocatalyst that effectively decomposes the organic pollutant MO under UV light, using lower concentration than or equal photocatalyst concentration to those in previous publications.

TABLE II.

Statistical results of photocatalytic efficiency when treating synthetic dyes with NFO/BSTO and other materials. Boldface denotes for the results obtained in this work.

Initial
Type ofconcentrationPhotocatalystExperimental
pollutant(mg/l)Photocatalystconcentration (g/l)conditiont (min)DR (%)kp (min−1)R2References
MO 10 BSTO 0.2 UV lamp 70 69.7 0.0166 0.98 This work 
BSNF1 69.7 0.0189 0.95 
BSNF2 79.8 0.0224 0.96 
BSNF3 79.9 0.0226 0.96 
BSNF4 81.5 0.0228 0.96 
MO BSTO Ultrasonic power 120 100 ⋯ ⋯ 30  
MO 10 BTO/Fe3O4 0.4 UV–visible 1200 71.2 ⋯ ⋯ 7  
AZO 500 SrTiO3/CoFe2O4 UV + LED lamps 180 70 ⋯ ⋯ 6  
MB 3.2 CoFe2O4/BaTiO3 0.4 UV lamp 180 72 0.0071 0.99 15  
MO 10 NiFe2O4 Tungsten filament lamp 140 83 ⋯ ⋯ 32  
RhB SrTiO3 0.25 Xenon lamp 200 98 0.0038 0.98 47  
RhB Nd doped SrTiO3 0.25 Xenon lamp 200 87 0.0072 0.94 47  
MO BaTiO3/g-C3N4 0.5 Xenon lamp 360 76 ⋯ ⋯ 49  
MO 10 BaTiO3/0.6CuO Xenon lamp 90 90 0.05 ⋯ 50  
MO 10 CoFe2O4/Fe3O4 0.2 UV lamp 300 93 ⋯ ⋯ 33  
Initial
Type ofconcentrationPhotocatalystExperimental
pollutant(mg/l)Photocatalystconcentration (g/l)conditiont (min)DR (%)kp (min−1)R2References
MO 10 BSTO 0.2 UV lamp 70 69.7 0.0166 0.98 This work 
BSNF1 69.7 0.0189 0.95 
BSNF2 79.8 0.0224 0.96 
BSNF3 79.9 0.0226 0.96 
BSNF4 81.5 0.0228 0.96 
MO BSTO Ultrasonic power 120 100 ⋯ ⋯ 30  
MO 10 BTO/Fe3O4 0.4 UV–visible 1200 71.2 ⋯ ⋯ 7  
AZO 500 SrTiO3/CoFe2O4 UV + LED lamps 180 70 ⋯ ⋯ 6  
MB 3.2 CoFe2O4/BaTiO3 0.4 UV lamp 180 72 0.0071 0.99 15  
MO 10 NiFe2O4 Tungsten filament lamp 140 83 ⋯ ⋯ 32  
RhB SrTiO3 0.25 Xenon lamp 200 98 0.0038 0.98 47  
RhB Nd doped SrTiO3 0.25 Xenon lamp 200 87 0.0072 0.94 47  
MO BaTiO3/g-C3N4 0.5 Xenon lamp 360 76 ⋯ ⋯ 49  
MO 10 BaTiO3/0.6CuO Xenon lamp 90 90 0.05 ⋯ 50  
MO 10 CoFe2O4/Fe3O4 0.2 UV lamp 300 93 ⋯ ⋯ 33  

NiFe2O4/Ba0.8Sr0.2TiO3 multiferroic composites were successfully synthesized by combining high-energy ball milling with heat treatment. XRD measurements and analysis confirmed the presence of both NFO and BSTO phases in the NFO/BSTO composites. The electrical polarization and magnetization of the composites were found to be significantly improved with increasing NFO content. In addition, the ability to catalyze activity by the UV light of NFO/BSTO composites was shown to be considerably higher than with pure BSTO.

The MO degradation of NFO/BSTO composites reached ∼80% after being illuminated with UV radiation for 70 min. These composites have the advantage of being easily recovered by using an external magnetic field, representing a simple recovery method. In addition, the significant reduction in the bandgap energy and the enhancement in the apparent first-order rate constant (kp) have demonstrated the role of the NFO ferromagnetic phase in these composites. These suggest that NFO/BSTO composites have promising potential in degrading synthetic organic dyes and treating contaminated water by multiferroic composites under ultraviolet light.

This research was funded by the Institute of Materials Science, Vietnam Academy of Science and Technology, under Grant No. CS.10/21-22.

The authors have no conflicts to disclose.

Dao Son Lam: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Project administration (equal); Writing – original draft (equal); Writing – review & editing (equal). Nguyen Minh Hieu: Formal analysis (equal); Investigation (equal). Dang Duc Dung: Formal analysis (equal); Investigation (equal). Dinh Chi Linh: Formal analysis (equal); Investigation (equal). Nguyen Ngoc Tung: Investigation (equal). Nguyen Thi Viet Chinh: Formal analysis (equal); Investigation (equal). Ngo Thu Huong: Investigation (equal); Methodology (equal); Writing – review & editing (equal). Tran Dang Thanh: Conceptualization (equal); Investigation (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available within the article.

1.
A.
Plyushch
,
D.
Lewin
,
A.
Sokal
,
R.
Grigalaitis
,
V. V.
Shvartsman
,
J.
Macutkevič
,
S.
Salamon
,
H.
Wende
,
K. N.
Lapko
,
P. P.
Kuzhir
,
D. C.
Lupascu
, and
J.
Banys
, “
Magnetoelectric coupling in nonsintered bulk BaTiO3xCoFe2O4 multiferroic composites
,”
J. Alloys Compd.
917
,
165519
(
2022
).
2.
J.
Martínez-Pérez
,
A.
Bolarín-Miró
,
C.
Cortés-Escobedo
, and
F.
Sánchez-De Jesús
, “
Magnetodielectric coupling in barium titanate-cobalt ferrite composites obtained via thermally-assisted high-energy ball milling
,”
Ceram. Int.
48
(
7
),
9527
9533
(
2022
).
3.
D.
Liu
,
C.
Jin
,
F.
Shan
,
J.
He
, and
F.
Wang
, “
Synthesizing BaTiO3 nanostructures to explore morphological influence, kinetics, and mechanism of piezocatalytic dye degradation
,”
ACS Appl. Mater. Interfaces
12
,
17443
17451
(
2020
).
4.
V. S.
Puli
,
D. K.
Pradhan
,
G.
Sreenivasulu
,
S. N.
Babu
,
N. V.
Prasad
,
K.
Madgula
,
D. B.
Chrisey
, and
R. S.
Katiyar
, “
Magnetoelectric and multiferroic properties of BaTiO3/NiFe2O4/BaTiO3 heterostructured thin films grown by pulsed laser deposition technique
,”
Crystals
11
(
10
),
1192
(
2021
).
5.
G. H.
Rather
and
M.
Ikram
, “
Enhancement of magnetoelectric effect in multiferroic composites of dysprosium and zinc doped BaTiO3–CoFe2O4
,”
J. Mater. Sci.: Mater. Electron.
32
,
551
566
(
2021
).
6.
N.
Eskandari
,
G.
Nabiyouni
, and
S. D.
MasoumiGhanbari
, “
Preparation of a new magnetic and photo-catalyst CoFe2O4–SrTiO3 perovskite nanocomposite for photo-degradation of toxic dyes under short time visible irradiation
,”
Composites, Part B
176
,
107343
(
2019
).
7.
X.
Li
,
G.
Wang
, and
Y.
Cheng
, “
Preparation and characteristics of Fe3O4–BaTiO3 heterostructural nanocomposite as photocatalyst
,”
Res. Chem. Intermed.
41
,
3031
3039
(
2015
).
8.
R.
Kumar
,
G.
Kumar
,
M. S.
Akhtar
, and
A.
Umar
, “
Sonophotocatalytic degradation of methyl orange using ZnO nano-aggregates
,”
J. Alloys Compd.
629
,
167
172
(
2015
).
9.
N. B.
Bokhale
,
S. D.
Bomble
,
R. R.
Dalbhanjan
,
D. D.
Mahale
,
S. P.
Hinge
,
B. S.
Banerjee
,
A. V.
Mohod
, and
P. R.
Gogate
, “
Sonocatalytic and sonophotocatalytic degradation of rhodamine 6G containing wastewaters
,”
Ultrason. Sonochem.
21
,
1797
1804
(
2014
).
10.
D. T.
Sponza
and
M.
Işik
, “
Decolorization and azo dye degradation by anaerobic/aerobic sequential process
,”
Enzyme Microb. Technol.
31
,
102
110
(
2002
).
11.
K.
Vinodgopal
,
D. E.
Wynkoop
, and
P. V.
Kamat
, “
Environmental photochemistry on semiconductor surfaces: Photosensitized degradation of a textile azo dye, acid orange, on TiO2 particles using visible light
,”
Environ. Sci. Technol.
30
,
1660
1666
(
1996
).
12.
N.
Tanwar
,
S.
Upadhyay
,
R.
Priya
,
S.
Pundir
,
P.
Sharma
, and
O. P.
Pandey
, “
Eu-doped BaTiO3 perovskite as an efficient electrocatalyst for oxygen evolution reaction
,”
J. Solid State Chem.
317
,
123674
(
2023
).
13.
N.
Touach
,
A.
Benzaouak
,
J.
Toyir
,
Y.
El Hamdouni
,
M.
El Mahi
,
E. M.
Lotfi
,
N.
Labjar
,
M.
Kacimi
, and
L. F.
Liotta
, “
BaTiO3 functional perovskite as photocathode in microbial fuel cells for energy production and wastewater treatment
,”
Molecules
28
(
4
),
1894
(
2023
).
14.
Y.
Cui
,
J.
Briscoe
, and
S.
Dunn
, “
Effect of ferroelectricity on solar-light-driven photocatalytic activity of BaTiO3-influence on the carrier separation and stern layer formation
,”
Chem. Mater.
25
,
4215
4223
(
2013
).
15.
H. T. M.
Dang
,
D. V.
Trinh
,
N. K.
Nguyen
,
T. T. T.
Le
,
D. D.
Nguyen
,
H. V.
Tran
,
M. H.
Phan
, and
C. D.
Huynh
, “
Enhanced photocatalytic activity for degradation of organic dyes using magnetite CoFe2O4/BaTiO3 composite
,”
J. Nanosci. Nanotechnol.
18
,
7850
7857
(
2018
).
16.
C.
Yu
,
S.
Wang
,
K.
Zhang
,
M.
Li
,
H.
Gao
,
J. H.
ZhangYang
,
L.
Hu
,
A. V.
Jagadeesha
, and
D.
Li
, “
Visible-light-enhanced photocatalytic activity of BaTiO3/γ-Al2O3 composite photocatalysts for photodegradation of tetracycline hydrochloride
,”
Opt. Mater.
135
,
113364
(
2023
).
17.
D.
Masekela
,
N. C.
Hintsho-Mbita
,
S.
Sam
,
T. L.
Yusuf
, and
N.
Mabuba
, “
Application of BaTiO3-based catalysts for piezocatalytic, photocatalytic and piezo-photocatalytic degradation of organic pollutants and bacterial disinfection in wastewater: A comprehensive review
,”
Arabian J. Chem.
16
,
104473
(
2023
).
18.
Y.
Cui
,
J.
Briscoe
,
Y.
Wang
,
N. V.
Tarakina
, and
S.
Dunn
, “
Enhanced photocatalytic activity of heterostructured ferroelectric BaTiO3/α-Fe2O3 and the significance of interface morphology control
,”
ACS Appl. Mater. Interfaces
9
,
24518
24526
(
2017
).
19.
A. O.
Turky
,
M. M.
Rashad
, and
M.
Bechelany
, “
Tailoring optical and dielectric properties of Ba0.5Sr0.5TiO3 powders synthesized using citrate precursor route
,”
Mater. Des.
90
,
54
59
(
2016
).
20.
C.
Liu
and
P.
Liu
, “
Microstructure and dielectric properties of BST ceramics derived from high-energy ball-milling
,”
J. Alloys Compd.
584
,
114
118
(
2014
).
21.
J.
Ćirković
,
K.
Vojisavljević
,
N.
Nikolić
,
P.
Vulić
,
Z.
Branković
,
T.
Srećković
, and
G.
Branković
, “
Dielectric and ferroelectric properties of BST ceramics obtained by a hydrothermally assisted complex polymerization method
,”
Ceram. Int.
41
,
11306
11313
(
2015
).
22.
S.
Dupuis
,
S.
Sulekar
,
J. H.
Kim
,
H.
Han
,
P.
Dufour
,
C.
Tenailleau
,
J. C.
Nino
, and
S.
Guillemet-Fritsch
, “
Colossal permittivity and low losses in Ba1−xSrxTiO3−δ reduced nanoceramics
,”
J. Eur. Ceram. Soc.
36
,
567
575
(
2016
).
23.
T.
Kytae
,
K.
Sutjarittangtham
,
T.
Thurakitseree
, and
W.
Leenakul
, “
Phase transition and electrical properties of Ba0.8Sr0.2TiO3: Comparison of different preparation techniques
,”
Appl. Mech. Mater.
866
,
287
290
(
2017
).
24.
F. M.
Pontes
,
E. R.
Leite
,
D. S. L.
Pontes
,
E.
Longo
,
E. M. S.
Santos
,
S.
Mergulhão
,
P. S.
Pizani
,
F.
Lanciotti
, Jr.
,
T. M.
Boschi
, and
J. A.
Varela
, “
Ferroelectric and optical properties of Ba0.8Sr0.2TiO3 thin film
,”
J. Appl. Phys.
91
,
5972
5978
(
2002
).
25.
C. M.
Kanamadi
,
B. K.
Das
,
C. W.
Kim
,
D. I.
Kang
,
H. G.
Cha
,
E. S.
Ji
,
A. P.
Jadhav
,
B. E.
Jun
,
J. H.
Jeong
,
B. C.
Choi
,
B. K.
Chougule
, and
Y. S.
Kang
, “
Dielectric and magnetic properties of (x)CoFe2O4 + (1 − x)Ba0.8Sr0.2TiO3 magnetoelectric composites
,”
Mater. Chem. Phys.
116
,
6
10
(
2009
).
26.
S.
K Barik
,
R. N. P.
Choudhary
, and
A. K.
Singh
, “
AC impedance spectroscopy and conductivity studies of Ba0.8Sr0.2TiO3 ceramics
,”
Adv. Mater. Lett.
2
(
6
),
419
424
(
2011
).
27.
B.
Vigneshwaran
,
P.
Kuppusami
,
S.
Ajithkumar
, and
H.
Sreemoolanadhan
, “
Study of low temperature-dependent structural, dielectric, and ferroelectric properties of BaxSr1−xTiO3 (x = 0.5, 0.6, 0.7) ceramics
,”
J. Mater. Sci.: Mater. Electron.
31
,
10446
10459
(
2020
).
28.
Y.
Wang
,
Z. Y.
Shen
,
Y. M.
Li
,
Z. M.
Wang
,
W. Q.
Luo
, and
Y.
Hong
, “
Optimization of energy storage density and efficiency in BaxSr1−xTiO3 (x ≤ 0.4) paraelectric ceramics
,”
Ceram. Int.
41
,
8252
8256
(
2015
).
29.
A.
Kumari
and
B.
Dasgupta Ghosh
, “
Effect of strontium doping on structural and dielectric behaviour of barium titanate nanoceramics
,”
Adv. Appl. Ceram.
117
,
427
435
(
2018
).
30.
B.
Yuan
,
J.
Wu
,
N.
Qin
,
E.
Lin
, and
D.
Bao
, “
Enhanced piezocatalytic performance of (Ba, Sr)TiO3 nanowires to degrade organic pollutants
,”
ACS Appl. Nano Mater.
1
,
5119
5127
(
2018
).
31.
C. B.
Samantaray
,
H.
Sim
, and
H.
Hwang
, “
First-principles study of electronic structure and optical properties of barium strontium titanates (BaxSr1−xTiO3)
,”
Appl. Surf. Sci.
250
,
146
151
(
2005
).
32.
A.
Lassoued
,
M. S.
Lassoue
,
B.
Dkhil
,
S.
Ammar
, and
A.
Gadri
, “
Photocatalytic degradation of methyl orange dye by NiFe2O4 nanoparticles under visible irradiation: Effect of varying the synthesis temperature
,”
J. Mater. Sci.: Mater. Electron.
29
,
7057
7067
(
2018
).
33.
D.
Mishra
,
K. K.
Senapati
,
C.
Borgohain
, and
A.
Perumal
, “
CoFe2O4–Fe3O4 magnetic nanocomposites as photocatalyst for the degradation of methyl orange dye
,”
J. Nanotechnol.
2012
,
323145
(
2012
).
34.
Y.
Liu
,
W.
Jiang
,
L.
Xu
,
X.
Yang
, and
F.
Li
, “
Decoration of carbon nanotubes with nearly monodisperse MIIFe2O4 (MFe2O4, M = Fe, Co, Ni) nanoparticles
,”
Mater. Lett.
63
,
2526
2528
(
2009
).
35.
H.
Ji
,
X.
Jing
,
Y.
Xu
,
J.
Yan
,
H.
Li
,
Y.
Li
,
L.
Huang
,
Q.
Zhang
,
H.
Xu
, and
H.
Li
, “
Magnetic g-C3N4/NiFe2O4 hybrids with enhanced photocatalytic activity
,”
RSC Adv.
5
,
57960
57967
(
2015
).
36.
S. E.
Shirsath
,
B. G.
Toksha
, and
K. M.
Jadhav
, “
Structural and magnetic properties of In3+ substituted NiFe2O4
,”
Mater. Chem. Phys.
117
,
163
168
(
2009
).
37.
S.
Balaji
,
R.
Kalai Selvan
,
L.
John Berchmans
,
S.
Angappan
,
K.
Subramanian
, and
C. O.
Augustin
, “
Combustion synthesis and characterization of Sn4+ substituted nanocrystalline NiFe2O4
,”
Mater. Sci. Eng.: B
119
,
119
124
(
2005
).
38.
S.
Gautam
,
P.
Shandilya
,
V. P.
Singh
,
P.
Raizada
, and
P.
Singh
, “
Solar photocatalytic mineralization of antibiotics using magnetically separable NiFe2O4 supported onto graphene sand composite and bentonite
,”
J. Water Process Eng.
14
,
86
100
(
2016
).
39.
Y.
Xia
,
Z.
He
,
J.
Su
,
B.
Tang
,
K.
Hu
,
Y.
Lu
,
S.
Sun
, and
X.
Li
, “
Fabrication of magnetically separable NiFe2O4/BiOI nanocomposites with enhanced photocatalytic performance under visible-light irradiation
,”
RSC Adv.
8
,
4284
(
2018
).
40.
D.
Neupane
,
J.
Casey
,
J.
Sultana
,
A. K.
Pathak
,
S.
Karna
,
S.
Pollard
, and
S. R.
Mishra
, “
Magnetocaloric properties of shape-dependent nanostructured Gd2O3 oxide particles
,”
Adv. Nat. Sci.: Nanosci. Nanotechnol.
14
,
035002
(
2023
).
41.
T. D.
Thanh
,
C. T.
Anh Xuan
,
T. N.
Bach
,
B. X.
Khuyen
,
D. S.
Lam
,
D. C.
Linh
,
L. T.
Giang
, and
V. D.
Lam
, “
Magnetic and microwave absorbing properties of La0.7Sr0.3MnO3 nanoparticles
,”
AIP Adv.
12
,
035134
(
2022
).
42.
M.
Atif
and
M.
Nadeem
, “
Interplay between the ferromagnetic and ferroelectric phases on the magnetic and impendance analysis of (x)Pb(Zr0.52Ti0.48)O3–(1 − x)CoFe2O4 composites
,”
J. Alloys Compd.
623
,
447
453
(
2015
).
43.
L.
Zhang
,
J.
Zhai
,
W.
Mo
,
X.
Yao
,
X.
Cheng
,
X.
Yu
, and
Z.
Xing
, “
Electrical, dielectric and magnetic properties of CoFe2O4–BaTiO3 composite films with core-shell
,”
Ferroelectrics
406
,
213
220
(
2010
).
44.
R.
Grigalaitis
,
M.
Vijatović Petrović
,
J.
Bobić
,
A.
Dzunuzovic
,
R.
Sobiestianskas
,
A.
Brilingas
,
B.
Stojanović
, and
J.
Banys
, “
Dielectric and magnetic properties of BaTiO3–NiFe2O4 multiferroic composites
,”
Ceram. Int.
40
,
6165
6170
(
2014
).
45.
S. Q.
Liu
, “
Magnetic semiconductor nano-photocatalysts for the degradation of organic pollutants
,”
Environ. Chem. Lett.
10
,
209
216
(
2012
).
46.
R. D.
Ambashta
and
M.
Sillanpää
, “
Water purification using magnetic assistance: A review
,”
J. Hazard. Mater.
180
,
38
49
(
2010
).
47.
Y. X.
Song
,
W. Q.
Ma
,
J. J.
Chen
,
J.
Xu
,
Z. Y.
Mao
, and
D. J.
Wang
, “
Photocatalytic activity of perovskite SrTiO3 catalysts doped with variable rare earth ions
,”
Rare Met.
40
(
5
),
1077
1085
(
2021
).
48.
N.
Adhlakha
,
K.
Yadav
, and
R.
Singh
, “
Effect of BaTiO3 addition on structural, multiferroic and magneto-dielectric properties of 0.3CoFe2O4–0.7BiFeO3 ceramics
,”
Smart Mater. Struct.
23
,
105024
(
2014
).
49.
S.
Kappadan
,
S.
Thomas
, and
N.
Kalarikkal
, “
Enhanced photocatalytic performance of BaTiO3/g-C3N4 heterojunction for the degradation of organic pollutants
,”
Chem. Phys. Lett.
771
,
138513
(
2021
).
50.
C.
Yu
,
M.
Tan
,
C.
Tao
,
Y.
Hou
,
C.
Liu
,
H.
Meng
,
Y.
Su
,
L.
Qiao
, and
Y.
Bai
, “
Remarkably enhanced piezo-photocatalytic performance in BaTiO3/CuO heterostructures for organic pollutant degradation
,”
J. Adv. Ceram.
11
,
414
426
(
2022
).