banner
Home / Blog / Photocatalytic properties of BiFeO3 powders synthesized by the mixture of CTAB and Glycine
Blog

Photocatalytic properties of BiFeO3 powders synthesized by the mixture of CTAB and Glycine

Aug 27, 2023Aug 27, 2023

Scientific Reports volume 13, Article number: 12338 (2023) Cite this article

247 Accesses

Metrics details

Highly pure BiFeO3 (BFO) powders were prepared by the solution combustion synthesis method using cetyltrimethylammonium bromide (CTAB) and glycine as fuels at various fuel-to-oxidant (φ) ratios. Microstructural characteristics, morphology, optical properties, and thermal analysis were studied by X-ray diffraction (XRD), scanning electron microscopy (SEM), diffuse reflectance spectroscopy (DRS), and differential thermal/thermogravimetry (DTA/TGA), respectively. The combusted powders prepared at different fuel content contained a small amount of impurity phases such as Bi24Fe2O39 and Bi2Fe4O9. During the calcination of BFO powders at 600 °C for 1 h, a nearly pure BFO phase was produced. Combusted powders photodegraded about 80% of methylene blue dye at φ = 2 through 90 min of visible light irradiation.

Single-phase BiFeO3 (BFO) is a multiferroic material with distorted rhombohedral and perovskite structures exhibiting the R3c space group. Due to its ferroelectric performance at high Curie temperatures up to 830 °C and antiferromagnetic behavior under its Neel temperature of 370 °C, this material is considered for non-volatile memory devices, photovoltaics, sensors, and spintronics1,2,3,4. It is also known that these abundant and interesting compounds with the perovskite structure exhibit improved compositional and structural tunability5,6. Because of its narrow bandgap in the range of 2.2–2.8 eV and high chemical stability, BFO has been considered a visible light photocatalyst to degrade organic pollutants7. Many photocatalysts, such as TiO2, ZnO, CdS, ZnS, etc., have been used to photodegrade dyes under ultraviolet (UV) light irradiation8,9,10,11,12,13. However, UV only spans a small portion (~ 4%) of the sunlight spectrum; thus, many efforts have been made to develop visible-light catalysts covering a broader range14,15,16,17,18,19.

Impurity phases such as Bi2O3, Bi2Fe4O9, and Bi24Fe2O39 appear during the synthesis of BFO due to their phase formation kinetics. Therefore, many researchers developed various synthesis routes to remove these secondary phases. Hydrothermal methods20,21, polymer assisted hydrothermal22, sol–gel23, co-precipitation24,25,26, aerosol-spraying, electrospinning27, solvothermal route28, and solution combustion29 were used to synthesize pure BFO.

Developing simple, environmentally safe, and energy-efficient methods to synthesize a pure BFO powder is of great interest. Solution combustion synthesis (SCS) is a simple, relatively cheap, and fast chemical process to produce various nanomaterials30. A self-propagating exothermic reaction occurs between the mixture of metal nitrates and different organic fuels (e.g., glycine, citric acid, urea, etc.), releasing an enormous amount of gaseous products29.

Among different organic fuels, glycine is an amino acid that facilitates the formation of a metal ion complex in the solution owing to its carboxylic acid and amino groups at opposite ends of the molecule31. Likewise, Cetyltrimethylammonium bromide (CTAB) is a cationic surfactant with a high decomposition temperature that is extensively used to control particle shapes, size, and microstructure by minimizing the precursor’s surface tension32. BFO has been synthesized by glycine fuel through microwave-assisted solution combustion with some impurity phases such as Bi2Fe4O9 and Bi24Fe2O3933. In our previous works, BFO was synthesized using various single and mixed fuels at a constant fuel to oxidant ratio of 1, but in this work different fuel-to-oxidant ratios (φ) were varied from 0.5 to 232,33,34.

Nevertheless, combining different fuels might be more effective than individual fuels via improved control over the reaction temperature, the type, and the amount of gaseous products released. Therefore, in this work, glycine and CTAB were mixed at various fuel-to-oxidant amounts in the uni-molar ratio to synthesize nearly pure and single-phase BFO.

Analytical grades Fe(NO3)3.9H2O, Bi(NO3)2.5H2O, CTAB [(C16H33)N(CH3)3]Br (> 99%), glycine (C2H5NO2), were purchased from Merk Co. without any further purification. Whereby HNO3 (68 wt %) was added to dissolve bismuth nitrate. The required amount of Bi(NO3)0.5H2O and Fe(NO3)0.9H2O, cetyltrimethylammonium bromide ([(C16H33)N(CH3)3]Br), and glycine (C2H5NO2) were prepared by dissolving 15 mL of 3 mol L−1 of HNO3 in various fuel-to-oxidant ratios of (φ = 0.5, 0.75, 1 and 2). H2O, CO2, Br2, and N2 are assumed to be the gaseous products of the combustion reaction, where the type of gaseous products and adiabatic temperature are controlled by the fuel-to-oxidant ratio (φ).

8500S SHIMADZU spectrophotometer recorded IR spectra in the range of (400–4000 cm−1). Differential thermal (DTA) and thermogravimetry analysis (TGA) were used to study the combustion behavior in the air with a heating rate of 5 ˚C/min on an STA Ba¨HR 503 instrument. Microstructure and phase evolution were examined by X-ray diffraction (PANalytical X’pert, CuKa = 1.54060 A°). Crystallite sizes were also calculated using the raw data from XRD using the Williamson-Hall method. Field emission scanning electron microscopy (FESEM) was acquired to characterize the morphology of the as-combusted powders by TESCAN Vega II. UV–Vis diffuse reflectance spectrum (DRS) was acquired to measure the bandgap and visible light absorption of the powders by a Shimadzu UV–Vis 52550 spectrophotometer in the wavelength range of (300–800 nm). A Tauc’s plot is used to determine the optical bandgap of semiconductors. Typically, a Tauc’s plot shows the quantity hν (the photon energy) on the (x-coordinate) and the quantity (αhν)1/2 on the (y-coordinate), where α is the absorption coefficient of the material. Thus, extrapolating this linear region to the abscissa yields the energy of the optical bandgap of the semiconductor material.

In the presence of combusted BFO powders, the methylene blue (MB) dye was broken down by visible light irradiation from two 100 W Xenon lamps with a cutoff ultraviolet filter.100 mg of BFO catalyst was dispersed in 100 mL of methylene blue (15 mg/L) in the company of 0.1 mL of H2O2 (30%) and stirred for 60 min in the dark to obtain the adsorption/desorption equilibrium. Furthermore, the pH of the solution was adjusted by HCl (37 wt%). BFO powders were separated by centrifugation at 6000 rpm for 10 min, followed by MB concentration monitoring on a PG Instruments Ltd. T80-UV/Vis spectrophotometer.

Figure 1 illustrates the thermal analysis of dried gel produced by a mixture of glycine and CTAB fuels at φ = 1. A slight drop (~ 9%) in the gel’s mass is possibly due to the evaporation of absorbed water. A sharp decline at about 178 °C, possibly triggered by the exothermic reaction between metal nitrates and glycine CTAB fuels. This enormous drop (~ 70%) in the gel mass is because of a combustion reaction that released a large amount of gaseous products such as CO2, H2O, N2, Br2, etc.

TGA/DTA curves of the dried gel prepared by a mixture of CTAB and glycine fuels at φ = 1.

Based on previous findings34,35,36, glycine has a lower decomposition temperature compared to CTAB, with a fast combustion reaction rate showing sharp weight loss in the gel. Similarly, when the mixture of glycine and CTAB is used as a fuel, colossal weight loss seems to be dominated by the presence of glycine rather than CTAB. Furthermore, a smaller exothermic peak at 285 °C with a gradual weight loss might be attributed to the slow oxidation reaction of residual organics that remained in the gel37.

The FTIR spectra of dried gel and as-combusted powders prepared by the glycine and CTAB fuel mixture at φ = 1 are illustrated in Fig. 2. The broad vibrational stretching modes in the range of 3200–3700 cm−1 correspond to the absorption of hydroxyl groups of water molecules that are omitted in the combusted BFO powders38. The stretching vibration of C-H bonds in CTAB molecules can lead to the formation of bands at 2920 and 2850 cm−139. The vibrational band at 1350 cm−1 is due to the attachment of CO32− groups to the cations38. The adsorption bands at 1650 cm−1, 1360 cm−1, 902 cm−1, 802 cm−1, and 730 cm−1 confirm the formation of NO3− connected to the CTAB and glycine molecules in the dried gel40. Stretching bands at 1757 cm−1 and 1556 cm−1 resemble the existence of COO− groups formed through the oxidation of CTAB molecules41. Peaks at 1105 cm−1 and 1020 cm−1 confirm the presence of NH2 groups. Carboxylate groups can chelate cations, leading to the absorption band at 586 cm−1, which corresponds to metal–oxygen bonds42. Strong peaks at 557 cm−1 and 465 cm−1 of combusted powder can be assigned to the vibrational bending and stretching of Fe–O in the octahedral FeO6 groups in the perovskite structure43.

FTIR spectra of (a) dried gel and (b) the as-combusted BFO powders at φ = 1.

XRD patterns of conventionally combusted BFO powders at the various φ values are depicted in Fig. 3. The combusted powders at φ = 0.5 and φ = 0.75 are semi-amorphous due to their incomplete combustion reaction and low adiabatic combustion temperature. Bi2Fe4O9 (JCDPS Card No. 00-020-0836) impurity phase is presented at φ values of 0.5 and 0.75. However, maximum adiabatic temperature occurs at φ = 1, leading to a well crystalline pattern. Bi24Fe2O39 (JCDPS Card No. 00-042-0201) was the only impurity phase formed at the φ values of 1 and 2.

XRD patterns of the as-combusted BFO powders using glycine and CTAB fuel content. (filled down pointing triangle: BiFeO3, open down pointing triangle: Bi2Fe4O9, filled diamond: Bi24Fe2O39).

Bi2O3 and Fe2O3, as the transitional phases, can take part in the solid reaction of (Bi2O3 + Fe2O3 → 2 BiFeO3) to produce the BiFeO3 phase. Nevertheless, the formation of impurity phases such as Bi2Fe4O9 and Bi24Fe2O39 can be ascribed to the insufficiency of Bi2O3 and Fe2O3 phases initiated by the phase segregation44:

The inferior crystallinity of as-combusted powders due to their lower combustion temperatures can be improved by further calcination at higher temperatures (Fig. 4). Impurity phases were mainly eliminated by one-hour calcination at 600 °C due to the reaction of the residual Bi2O3 and Fe2O3 phases.

XRD patterns of the as-calcined BFO powders using glycine and CTAB fuel content (filled down pointing triangle: BiFeO3, open down pointing triangle: Bi2Fe4O9).

SEM micrographs of BFO powders synthesized at different φ values are illustrated in Fig. 5. As combusted powders display a bulky microstructure, the particle sizes are reduced from 37 to 18 nm with the increase in fuel content presented in Table 1, as calculated from XRD data using the Williamson-Hall technique. Particle size mainly depends on the combustion temperature and reaction rate, where the combustion rate would influence the number of nucleation sites and higher combustion temperature enhances the particle growth45. When the fuel content is increased, a higher amount of generated heat is consumed by the combustion gases and hence reduces the adiabatic temperature. This decline in the adiabatic temperature resulted in particle size refinement. However, as seen in Fig. 5d–f, calcined powders exhibit particle size growth as a result of the rise in the temperature.

SEM micrographs of the as-combusted BFO powders at (a) φ = 0.75, (b) φ = 1 and (c) φ = 2 and the as-calcined BFO powders at (d) φ =0.75, (e) φ =1 and (f) φ =2.

Another proposed reason behind this particle size refinement at higher CTAB/glycine content, as illustrated schematically in Fig. 6, could be due to the interaction of CTAB micelles with the cationic head inside the solution precursor, separated a large amount of the cationic ends apart and produced smaller BFO nanoparticles when the CTAB amount was maximized. Figure 7a displays diffuse reflectance spectra of the as combusted BFO powders. The amount of visible light absorption mainly depends on the crystallinity, strain, particle size, impurity phases, etc. The crystal field and metal–metal transitions affect the absorption spectra46. The increase in light absorption at higher fuel content (φ = 2) is possibly due to the decrease in the impurity phase Bi24Fe2O39. However, powders synthesized at fuel content of (φ = 0.75) significantly absorbed a higher amount of visible light, probably due to the formation of distinct impurity phase of Bi2Fe4O9, as previously discussed in XRD data. The Tauc’s plot measured bandgap energy of the combusted powders ((αhυ)2 vs. hυ), as shown in Fig. 7b and summarized in Table 1. The bandgap energies of the combusted BFO powders are in the range of 1.85–1.96 eV, in good agreement with the bandgap of powders and thin films reported in the literature46. At higher fuel contents, the decrease in bandgap energy is mainly due to the particle size refinement, while the increase in the bandgap energy at φ = 2 could be owed to the presence of the impurity phase Bi24Fe2O39.

Schematic of BFO nanoparticle synthesis via micells formation at high CTAB content.

(a) UV–Vis diffuse reflectance spectra and (b) Tauc’s plot of the as-combusted BFO powders.

The relative concentrations (C/C0) of MB, as an organic pollutant, versus visible light irradiation is illustrated in Fig. 8. The MB is photodegraded by about 80% during 90 min of visible light illumination for combusted powders at φ values of 2 and 0.75. However, the degradation rate is slightly higher for the combusted powders synthesized at φ = 2. Based on our previous findings32,33, the combusted powders synthesized by pure glycine or pure CTAB only showed the photodegradation of MB at about 50 and 30 percent, respectively. Therefore, by mixing the glycine and CTAB fuels, the photodegradation of MB is profoundly enhanced possibly due to the particle size refinement and higher crystallinity with less amount of impurity phases being present. The oxidation of MB dye molecule to CO2 and H2O species mainly depends on the presence of active species such as O−2, OH radicals47. The photogenerated electrons and holes during the light absorption of powders would react with the oxygen and water molecules to produce the active species. Thus, the optical properties of combusted powders such as bandgap energy, absorption coefficient, and band edge position play an important role in photocatalytic performance48.

C/C0 versus irradiation time in the presence of the BFO powders at different fuel content and type.

High purity BFO powders were synthesized via solution combustion synthesis by the mixture of glycine and CTAB as fuels at different fuel contents. The amount of the impurity phase was reduced by increasing fuel content from 0.5 to 2. Bi2Fe4O9 impurity phase was present at φ values of 0.5 and 0.75. However, at higher fuel ratios (φ values of 1 and 2), the impurity phase was transformed into the Bi24Fe2O39 phase.

The combusted powders at φ values of 2 and 0.75 showed the highest MB photodegradation of about 80% under 90 min of visible light illumination mainly due to the particle size refinement, higher visible light absorption, and less amount of impurities. Furthermore, the photodegradation rate of combusted powders synthesized at φ = 2 was somewhat enhanced.

The datasets used and/or analysed during the current study available from the corresponding author on reasonable request.

Shokrollahi, H. Magnetic, electrical and structural characterization of BiFeO3 nanoparticles synthesized by co-precipitation. Powder Technol. 235, 953–958 (2013).

Article CAS Google Scholar

Vanga, P. R., Mangalaraja, R. V. & Ashok, M. Structural, magnetic and photocatalytic properties of La and alkaline co-doped BiFeO3 nanoparticles. Mater. Sci. Semicond. Process. 40, 796–802 (2015).

Article Google Scholar

Chen, S. et al. Hand-fabricated CNt/AgNps electrodes using Wax-on-plastic platforms for electro-immunosensing application. Sci. Rep. 9(1), 6131 (2019).

Article ADS PubMed PubMed Central Google Scholar

Xu, X., Wang, W., Zhou, W. & Shao, Z. Recent advances in novel nanostructuring fthrumethods of perovskite electrocatalysts for energy-related applications. Small Methods 2, 1800071 (2018).

Article Google Scholar

He, J., Xiaomin, Xu., Li, M., Zhou, S. & Zhou, W. Recent advances in perovskite oxides for non-enzymatic electrochemical sensors: A review. Anal. Chim. Acta 1251, 341007 (2023).

Article CAS PubMed Google Scholar

Yang, L. et al. ACS Sustain. Chem. Eng. 10(5), 1899–1909 (2022).

Article CAS Google Scholar

Guo, R., Fang, L., Dong, W., Zheng, F. & Shen, M. Enhanced photocatalytic activity and ferromagnetism in Gd doped BiFeO3 nanoparticles. J. Phys. Chem. C 114(49), 21390–21396 (2010).

Article CAS Google Scholar

Chen, Y., Tang, Y., Luo, S., Liu, C. & Li, Y. TiO2 nanotube arrays co-loaded with Au nanoparticles and reduced graphene oxide: facile synthesis and promising photocatalytic application. J. Alloys Compd. 578, 242–248 (2013).

Article CAS Google Scholar

Zhang, Y. et al. Combination of photoelectrocatalysis and adsorption for removal of bisphenol A over TiO2-graphene hydrogel with 3D network structure. Appl. Catal. B Environ. 221, 36–46 (2018).

Article CAS Google Scholar

Nag, A., Sapra, S., Chakraborty, S., Basu, S. & Sarma, D. D. Synthesis of CdSe nanocrystals in a noncoordinating solvent: effect of reaction temperature on size and optical properties. J. Nanosci. Nanotechnol. 7(6), 1965–1968 (2007).

Article CAS PubMed Google Scholar

Rajeshwar, K. et al. Heterogeneous photocatalytic treatment of organic dyes in air and aqueous media. J. Photochem. Photobiol. C Photochem. Rev. 9(4), 171–192 (2008).

Article CAS Google Scholar

Herrmann, J.-M. Heterogeneous photocatalysis: fundamentals and applications to the removal of various types of aqueous pollutants. Catal. Today 53(1), 115–129 (1999).

Article CAS Google Scholar

Fu, Y. & Wang, X. Magnetically separable ZnFe2O4–graphene catalyst and its high photocatalytic performance under visible light irradiation. Ind. Eng. Chem. Res. 50(12), 7210–7218 (2011).

Article CAS Google Scholar

Vahdat Vasei, H., Masoudpanah, S. M., Adeli, M. & Aboutalebi, M. R. Photocatalytic properties of solution combustion synthesized ZnO powders using mixture of CTAB and glycine and citric acid fuels. Adv. Powder Technol. 30(2), 284–291 (2019).

Article CAS Google Scholar

Singh, H., Garg, N., Arora, P. & Rajput, J. K. Sucrose chelated auto combustion synthesis of BiFeO3 nanoparticles: Magnetically recoverable catalyst for the one-pot synthesis of polyhydroquinoline. Appl. Organometal Chem. 32, e4357 (2018).

Article Google Scholar

Kumar, S. S., Rao, K., Krishnaveni, T., Goud, A. & Reddy, P. Solution Combustion Synthesis and Characterization of Nanosized Bismuth Ferrite. AIP Conf. Proc. 1447, 339–340 (2011).

Article Google Scholar

Ilić, N. et al. Auto-combustion synthesis as a method for preparing BiFeO3 powders and flexible BiFeO3/PVDF films with improved magnetic properties Influence of doping ion position, size and valence on electric properties. Mater. Sci. Eng.: B 280, 115686 (2022).

Article Google Scholar

Paborji, F., Shafiee Afarani, M., Arabi, A. M. & Ghahari, M. Solution combustion synthesis of FeCr2O4 powders for pigment applications: Effect of fuel type. Int. J. Appl. Ceram. Technol. 19, 2406–2418 (2022).

CAS Google Scholar

Paborji, F., Shafiee Afarani, M., Arabi, A. M. & Ghahari, M. Solution combustion synthesis of FeCr2O4:Zn, Al pigment powders. Int. J. Appl. Ceram. Technol. 20, 2203–2216 (2023).

Article CAS Google Scholar

Gao, T. et al. Shape-controlled preparation of bismuth ferrite by hydrothermal method and their visible-light degradation properties. J. Alloys Compd. 648, 564–570 (2015).

Article CAS Google Scholar

Wang, X. et al. PVP assisted hydrothermal fabrication and morphology-controllable fabrication of BiFeO3 uniform nanostructures with enhanced photocatalytic activities. J. Alloys Compd. 677, 288–293 (2016).

Article CAS Google Scholar

Wang, Y. et al. Low temperature polymer assisted hydrothermal synthesis of bismuth ferrite nanoparticles. Ceram. Int. 34(6), 1569–1571 (2008).

Article CAS Google Scholar

Gao, T. et al. Synthesis of BiFeO3 nanoparticles for the visible-light induced photocatalytic property. Mater. Res. Bull. 59, 6–12 (2014).

Article CAS Google Scholar

Ke, H. et al. Factors controlling pure-phase multiferroic BiFeO3 powders synthesized by chemical co-precipitation. J. Alloys Compd. 509(5), 2192–2197 (2011).

Article CAS Google Scholar

Bo, H. Y., Tan, G. Q., Miao, H. Y. & Xia, A. Co-precipitation synthesis of BiFeO3 powders. Adv. Mater. Res. 105, 286–288 (2010).

Article Google Scholar

Liu, Z., Qi, Y. & Lu, C. High efficient ultraviolet photocatalytic activity of BiFeO3 nanoparticles synthesized by a chemical coprecipitation process. J. Mater. Sci. Mater. Electron. 21(4), 380–384 (2010).

Article ADS CAS Google Scholar

Wang, W. et al. Electrospinning of magnetical bismuth ferrite nanofibers with photocatalytic activity. Ceram. Int. 39(4), 3511–3518 (2013).

Article CAS Google Scholar

Huo, Y., Jin, Y. & Zhang, Y. Citric acid assisted solvothermal synthesis of BiFeO3 microspheres with high visible-light photocatalytic activity. J. Mol. Catal. A Chem. 331(1–2), 15–20 (2010).

Article CAS Google Scholar

Yang, J. et al. Factors controlling pure-phase magnetic BiFeO3 powders synthesized by solution combustion synthesis. J. Alloys Compd. 509(37), 9271–9277 (2011).

Article CAS Google Scholar

Aruna, S. T. & Mukasyan, A. S. Combustion synthesis and nanomaterials. Curr. Opin. Solid State Mater. Sci. 12(3–4), 44–50 (2008).

Article ADS CAS Google Scholar

Motevalian, A. & Salem, S. Effect of glycine–starch mixing ratio on the structural characteristics of MgAl2O4 nano-particles synthesized by sol–gel combustion. Particuology 24, 108–112 (2016).

Article CAS Google Scholar

Asefi, N., Hasheminiasari, M. & Masoudpanah, S. M. Solution combustion synthesis of BiFeO3 powders using CTAB as fuel. J. Electron. Mater. 48, 1–7 (2018).

Google Scholar

Asefi, N., Masoudpanah, S. M. & Hasheminiasari, M. Microwave-assisted solution combustion synthesis of BiFeO3 powders. J. Sol-Gel Sci. Technol. 86, 1–9 (2018).

Article Google Scholar

Asefi, N., Masoudpanah, S. M. & Hasheminiasari, M. Photocatalytic performances of BiFeO3 powders synthesized by solution combustion method: The role of mixed fuels. Mater. Chem. Phys. 228, 168–174 (2019).

Article CAS Google Scholar

Huang, J. et al. Microwave hydrothermal synthesis of BiFeO3: Impact of different surfactants on the morphology and photocatalytic properties. Mater. Sci. Semicond. Process. 25, 84–88 (2014).

Article CAS Google Scholar

Joshi, U. A., Jang, J. S., Borse, P. H. & Lee, J. S. Microwave synthesis of single-crystalline perovskite BiFeO3 nanocubes for photoelectrode and photocatalytic applications. Appl. Phys. Lett. 92(24), 242106 (2008).

Article ADS Google Scholar

Radpour, M., Masoudpanah, S. M. & Alamolhoda, S. Microwave-assisted solution combustion synthesis of Fe3O4 powders. Ceram. Int. 43(17), 14756–14762 (2017).

Article CAS Google Scholar

Socrates, G. Infrared and Raman Characteristic Group Frequencies, 3rd edn., (Wiley, 2001).

Google Scholar

Hamedani, S. F. N., Masoudpanah, S. M., Bafghi, M. S. & Baloochi, N. A. Solution combustion synthesis of CoFe2O4 powders using mixture of CTAB and glycine fuels. J. Sol-Gel Sci. Technol. 86(3), 743–750 (2018).

Article Google Scholar

Naderi, P., Masoudpanah, S. M. & Alamolhoda, S. Magnetic properties of Li0.5Fe2.5O4 nanoparticles synthesized by solution combustion method. Appl. Phys. A 123(11), 702 (2017).

Article ADS Google Scholar

Shabani, S., Mirkazemi, S. M., Masoudpanah, S. M. & Abadi, P. T. D. Synthesis and characterization of pure single phase BiFeO3 nanoparticles by the glyoxylate precursor method. J. Supercond. Nov. Magn. 27(12), 2795–2801 (2014).

Article CAS Google Scholar

Ortiz-Quiñonez, J. L. et al. Easy synthesis of high-purity BiFeO3 nanoparticles: New insights derived from the structural, optical, and magnetic characterization. Inorg. Chem. 52(18), 10306–10317 (2013).

Article PubMed Google Scholar

Masoudpanah, S. M., Mirkazemi, S. M., Bagheriyeh, R., Jabbari, F. & Bayat, F. Structural, magnetic and photocatalytic characterization of Bi1–xLaxFeO3 nanoparticles synthesized by thermal decomposition method. Bull. Mater. Sci. 40(1), 93–100 (2017).

Article CAS Google Scholar

Feng, K., Wang, L.-C., Lu, J., Wu, Y. & Shen, B.-G. Experimentally determining the intrinsic center point of Bi2O3–Fe2O3 phase diagram for growing pure BiFeO3 crystals. CrystEngComm 15(24), 4900–4904 (2013).

Article CAS Google Scholar

Deganello, F. & Tyagi, A. K. Solution combustion synthesis, energy and environment: Best parameters for better materials. Prog. Cryst. Growth Charact. Mater. 64(2), 23–61 (2018).

Article CAS Google Scholar

Lam, S.-M., Sin, J.-C. & Mohamed, A. R. A newly emerging visible light-responsive BiFeO3 perovskite for photocatalytic applications: A mini review. Mater. Res. Bull. 90, 15–30 (2017).

Article CAS Google Scholar

Prashanthi, K., Thakur, G. & Thundat, T. Surface enhanced strong visible photoluminescence from one-dimensional multiferroic BiFeO3 nanostructures. Surf. Sci. 606(19–20), L83–L86 (2012).

Article ADS CAS Google Scholar

Soltani, T. & Entezari, M. H. Photolysis and photocatalysis of methylene blue by ferrite bismuth nanoparticles under sunlight irradiation. J. Mol. Catal. A Chem. 377, 197–203 (2013).

Article CAS Google Scholar

Download references

School of Metallurgy and Materials Engineering, Iran University of Science and Technology (IUST), Narmak, Tehran, Iran

N. Asefi, M. Hasheminiasari & S. M. Masoudpanah

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

N.A. performed the experimental part and wrote the draft of the paper. M.H. supervised the research and wrote and edited the paper. S.M.M. supervised the research and wrote and edited the paper.

Correspondence to M. Hasheminiasari.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Asefi, N., Hasheminiasari, M. & Masoudpanah, S.M. Photocatalytic properties of BiFeO3 powders synthesized by the mixture of CTAB and Glycine. Sci Rep 13, 12338 (2023). https://doi.org/10.1038/s41598-023-39622-4

Download citation

Received: 15 June 2023

Accepted: 27 July 2023

Published: 31 July 2023

DOI: https://doi.org/10.1038/s41598-023-39622-4

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.