Next Article in Journal
Conformational-Dependent Photodissociation of Glycolic Acid in an Argon Matrix
Previous Article in Journal
Water-Soluble Dicationic Deuteroporphyrin Derivative for Antimicrobial PDT: Singlet Oxygen Generation, Passive Carrier Interaction and Nosocomial Bacterial Strains Photoinactivation
Previous Article in Special Issue
Application of Zirconia/Alumina Composite Oxide Ceramics as Photocatalysts for Removal of 2,4,6-Trichlorophenol from Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Synthesis, Characterization, and Photocatalytic Performance of BiOF/BiFeO3 Hybrid Heterojunction for Benzylamine Coupling under Simulated Light Irradiation

1
Department of Chemistry, United Arab Emirates University, Al-Ain P.O. Box 15551, United Arab Emirates
2
Department of Chemical Engineering, United Arab Emirates University, Al-Ain P.O. Box 15551, United Arab Emirates
*
Author to whom correspondence should be addressed.
Photochem 2023, 3(1), 187-196; https://doi.org/10.3390/photochem3010012
Submission received: 5 February 2023 / Revised: 10 March 2023 / Accepted: 14 March 2023 / Published: 21 March 2023

Abstract

:
Under simulated light irradiation, the aerobic oxidation of benzylamine to N,N-benzylidenebenzylamine was carried out as a model reaction to investigate the photocatalytic activity of a hydrothermally prepared composite based on BiOF and BiFeO3 materials. The prepared photocatalysts were characterized using several spectroscopic techniques, such as powder X-ray diffraction (PXRD), diffuse reflectance spectroscopy (DRS), scanning electron microscopy (SEM), energy-dispersive X-ray spectroscopy (EDX), and Fourier transform infrared spectroscopy (FTIR). Band gap analysis showed that the composite exhibits a band gap that lies in the UV region (3.5 eV). Nonetheless, pristine BiOF and BiFeO3 exhibited band gaps of 3.8 eV and 2.15 eV, respectively. N,N-benzylidenebenzylamine was selectively achieved with a high conversion yield of ~80% under atmospheric conditions in which the product was confirmed using 1H-NMR, 13C-NMR, and FTIR spectroscopic techniques. Various control experiments were conducted to further confirm the enhanced photocatalytic performance of the reported composite.

1. Introduction

Since its discovery in 1960, bismuth ferrite has garnered considerable interest due to its rarity as one of the few multiferroics with the simultaneous presence of ferroelectric and antiferromagnetic order characteristics in a perovskite structure [1,2]. An inorganic semiconductor, bismuth ferrite (BiFeO3, BFO) is a promising multifunctional material with a wide range of intriguing applications, including spintronics [3,4], sensors [5,6], photocatalysis [7,8,9], optical devices [10,11], and data storage [12]. Due to the narrow band gap energy of BFO and the potential for internal polarization suppression of the electron–hole recombination rate, it has unique advantages as a heterogenous photocatalyst [13,14]. Compared to the commonly used TiO2 photocatalyst [15], which absorbs in the UV region due to its broad band gap, BFO narrow band gap enables the greatest and most efficient exploitation of visible light from solar radiation energy [16,17]. In photocatalytic applications, the behavior of BFO in the presence of visible light is of particular interest due to the number of distinctive properties it holds, including a low cost [18], nontoxicity [19,20], chemical stability [18], discrete crystalline structure [18], and special electro-optical properties [21].
BFO perovskites can only stabilize over a limited range of temperatures, and, until now, the problem of obtaining single-phase pure BFO crystallite remains challenging. Nonetheless [8], a range of wet chemical techniques have been employed to create pure single-phase BiFeO3. A hydrothermal procedure can create pure single-phase BiFeO3 at low temperatures, but pressure is necessary [22]. These procedures include the microemulsion technique [23], the ferrioxalate precursor method [24], the solution combustion method [25], and the citrate method [26], and using sol–gel techniques based on ethylene glycol [27].
It has been proven in many previous studies that metal-oxides-based photocatalysts perform better when heterojunctions are formed; thus, attention to the assembly of various heterojunction structures has grown in the past few years. For example, heterojunctions such as BiOF/BiOI [28], MoS2/BiVO4 [29], BiOI/TiO2 [30], NiS/CdS [31], and SnO2/TiO2 [32] were used as photocatalysts for the removal of diclofenac potassium, the photodegradation of methylene blue, the photodegradation of methyl orange, photocatalytic H2-production, and the decolorization of rhodamine-B, respectively.
In this study, the coprecipitation method was used to generate pure single-phase BFO crystallites at a low temperature of about 70 °C. This method had the ability to create BFO powders that were well crystallized, had regulated morphology, and had a limited dispersion of particle size. Notably, all previous studies on BFO materials have focused on magnetic and dielectric properties. Nevertheless, our present study focuses on the photocatalytic aspect of this material and a newly prepared heterojunction based on BiFeO3 and BiOF. The photocatalytic activity has been investigated using the photocatalytic benzyl amine coupling reaction as a model reaction.

2. Experimental

2.1. Materials and General Procedures

Bismuth nitrate pentahydrate (BiNO3∙5H2O), iron nitrate nonahydrate (Fe(NO3)3∙9H2O), ammonium fluoride (NH4F), deuterated chloroform (CDCl3), acetonitrile (CH3CN, ACN), glacial acetic acid (CH3COOH), and benzylamine (C7H9N) were purchased from Sigma-Aldrich (St. Louis Missouri, MO, USA) and used as received. The FTIR spectrum was obtained using ATR-IR spectroscopy (Agilent Technologies Cary 600 Series FTIR Spectrometer, Santa Clara, CA, USA) across a range of 4000 to 500 cm−1 (512 scan average), in which the background was first collected using potassium bromide where the oily product was sandwiched between two KBr disks. Using a benchtop Rigaku MiniFlex X-ray diffractometer (Neu-Isenburg, Germany) with a CuKα radiation tube (λ = 1.542 Å) operating at 40 kV across a range of 3–50° (2θ) and a rate of 2 °C min−1, powder X-ray diffraction (PXRD) (Neu-Isenburg, Germany) measurements were made. The sample surface morphology and elemental analysis were analyzed using Quattro ESEM instrument scanning electron microscopy (SEM) (Waltham, MA, USA), equipped with an energy-dispersive X-ray (EDX) detector, that was operated at high vacuum with 30 kV accelerating voltage. Diffuse reflectance (DRS) of BiFeO3, BiOF, and the composite was measured using a Shimadzu UV-3600 spectrophotometer (Kyoto, Japan) over wavelength ranges from 200 nm to 800 nm after a baseline measurement using barium sulfate. Proton nuclear magnetic resonance, 1H-NMR, and 13C-NMR spectra were acquired using a Varian 400 MHz (Palo Alto, CA, USA) in d-chloroform as the solvent to validate the identity of the product. Nitrogen sorption experiments at 77 K were used to assess the surface area and porosity of the photocatalysts.

2.2. Preparation of the Photocatalysts

2.2.1. Preparation of BiFeO3

Bismuth ferrite was synthesized using the coprecipitation method. Bismuth nitrate and iron nitrate were dissolved separately in 2-methoxyethanol and left under stirring until completely dissolved. In detail, 6.18 × 10−3 moles of bismuth nitrate and iron nitrate equivalent to 3 g and 2.5 g were used, respectively. A total of 30 mL of 2-methoxyethanol was used to dissolve bismuth nitrate and an additional 5 mL of ethylene glycol was added to enhance the solubility of bismuth nitrate. For iron nitrate, 8 mL of 2-methoxyethanol was needed. The two solutions were mixed, and the pH was set to 5 using acetic acid and ammonium hydroxide. The solution was then left under continuous stirring and heating at 70 °C for around 7 h. The product was then collected and washed with DI water and ethanol several times. It was subsequently dried for 2 h at 80 °C, then calcined at 600 °C for 4 h (Scheme 1).

2.2.2. Preparation of BiOF

According to previously reported procedure, BiOF was synthesized using simple coprecipitation method [28]. Bismuth nitrate and ammonium fluoride were utilized in quantities of 0.002 mols each, or 970 mg and 74.08 mg, respectively. Bismuth nitrate was dissolved in 20 mL of 4 M acetic acid with constant stirring and heating at 80 °C. Then, ammonium fluoride was dissolved in 2 mL of deionized water (DI) and added to bismuth nitrate solution. The mixture was left under continuous stirring and heating at 80 °C for 7 h. The product was then collected and washed with DI water and ethanol several times. It was subsequently dried for 2 h at 80 °C and calcined for 4 h at 400 °C (Scheme 1).

2.2.3. Preparation of BiFeO3/BiOF Composite

BiFeO3/BiOF composite was synthesized using the simple hydrothermal method. A molar ratio of 1:1 was used from each of BiFeO3 and BiOF, which was added to a 25 mL Teflon line autoclave with 15 mL of DI water. The mixture was sonicated for 1 h before heating at 140 °C for 72 h (Scheme 1).

2.2.4. Photocatalytic Activity

In a 25 mL round-bottom flask connected to a condenser, the photocatalytic reaction was conducted under aerobic conditions. An optimum of 5.0 mg (0.16, 0.02, 0.009 mmol for BiOF, BiFeO3, and BiOF/BiFeO3 composite, respectively) of each sample was added to 51 μL (0.50 mmol) of benzyl amine in 2 mL of acetonitrile (ACN). The reaction was then irradiated by a 400 W halogen lamp for 24 h. The photocatalysts were washed with ACN and separated from the product by syringe filtration. Product purification was completed by simple column chromatography using silica gel and a mixture of ethyl acetate and hexane (1:2) to remove any possible impurities or byproducts. To isolate the product (N-benzylidenebenzylamine) that was dissolved in ACN, the solvent was evaporated using a rotary evaporator. Finally, 1H-NMR, 13C-NMR, and FTIR techniques were used to confirm the identity of the product. All control experiments followed the same procedure with required changes in some variables, as outlined in Table 1.

3. Results and Discussion

3.1. Characterization of the Photocatalysts

The purity and phase structure of prepared BiFeO3 and BiOF were confirmed by the PXRD diffraction pattern and compared to the standard PXRD database (JCPDS file No. 01-073-0548 and 73-1595 for BiFeO3 and BiOF, respectively), as shown in Figures S1 and S2 in the Supporting Information. All diffraction peaks were matched using Match!3 software (version 3.0) to the pure trigonal and tetragonal phases of BiFeO3 and BiOF, respectively, with no extra peaks determining any impurities the samples may contain. The BiOF diffraction peaks indexed to (001), (101), (002), (110), (102), (112), (200), (103), and (211), which showed at 2θ of 14.16°, 27.64°, 28.58°, 33.66°, 37.42°, 44.82°, 48.46°, 50.15°, and 56.77°, respectively. On the other hand, the BiFeO3 diffraction peaks indexed to (002), (012), (110), (013), (004), (020), (114), and (122), which appeared at 2θ of 19.34°, 29.2°, 31.93°, 37.21°, 39.62°, 45.93°, 51.5°, and 55.56°, respectively. Moreover, upon comparing the PXRD pattern of the prepared photocatalyst BiOF/BiFeO3 composite with both mixed metal oxides (Figure 1), the obtained patterns showed a combination of similar peaks from both mixed materials, which confirmed the preparation of the composite.
The band gaps of BiFeO3, BiOF, and BiOF/BiFeO3 composite were measured using the Tauc method [33], with which conduction band positions and band potential measurements could be estimated as previously reported [34,35]. When plotting (α) vs. photon energy (), as shown in Figure 2, the BiFeO3 exhibited a band of 2.15 eV lying in the visible region, and BiOF showed a larger band gap of 3.8 eV in the UV region. The BiOF/BiFeO3 composite exhibited a band gap of 3.5 eV lying in the UV light region compared to pure BiOF, indicating that the synthesized composite forms a heterojunction interface.
SEM images of pure BiFeO3, BiOF, and composite BiOF/BiFeO3 photocatalysts are shown in Figure 3. BiOF showed agglomerated small particles on an irregular surface, whereas BiFeO3 possessed a spherical particle shape. As for the BiOF/BiFeO3 composite, there is no specific adopted morphology from either material, but the composite has an irregular surface with agglomerated particles. The EDX data (Figures S3–S5, Supporting Information) were used to confirm the elemental composition of the synthesized photocatalyst. Atomic percent confirmed the presence of Bi, Fe, F, and O in the photocatalyst with a higher presence in Bi and O compared to both BiFeO3 and BiOF (Tables S1–S3, Supporting Information).
The FT-IR spectra of the three materials are shown in Figure 4. FT-IR data were used to determine the bond present in the materials. BiFeO3 has bands around 530, 550, and 750 cm−1 that represent the Fe-O stretching and bending vibrations. Bi−O/F vibrations in BiOF occurred at 415−558 cm−1. For the BiOF/BiFeO3 composite, the 1405 cm−1 and 553 cm−1 bands are for the Bi-O and Fe-O bending and stretching bands, respectively.

3.2. Photocatalytic Activities

Based on their band gap values, the photocatalytic performance of the reported photocatalysts was tested using the photocatalytic oxidative coupling reaction of benzylamine as a model photocatalytic reaction. The photocatalytic experiments were carried out using 51 µL of benzyl amine placed in a 25 mL round-bottom flask irradiated from the side (10 cm from the radiation source) using light from a 400 W halogen lamp with a wavelength range of 340–850 nm. The round-bottom flask was attached to a condenser open at the top to the atmosphere. A pale-yellow oily product of N-benzylidenebenzylamine was formed and confirmed using 1H-NMR and 13C-NMR (Figure S6, Supporting Information). Peak shifts were as follows: 1H-NMR (400 MHz, CDCl3): δ 8.40 (s, 1H), 7.80–7.77 (m, 2H), 7.43–7.41 (m, 3H), 7.35–7.34 (m, 4H), 7.29–7.26 (m, 1H), 4.83 (s, 2H); 13C-NMR (400 MHz, CDCl3): δ 162, 139.23, 136.09, 130.78, 128.61, 128.49, 128.27, 127.98, 126.99, 65.06. Because of the newly appeared double bond formed after the coupling of another benzyl amine, the peak at 8.4 ppm was the most deshielded signal. The 1H-NMR spectrum of the starting material benzyl amine was compared to the products in which the signal at 3.9 ppm (-CH2 protons) was downfielded to 4.9 in N-benzylidenebenzylamine, which further confirms the coupling step. In agreement with this finding, 13C-NMR data showed the highest shift at δC = 162 ppm, which belongs to the characteristic peak of the N=C bond of the products. Having the expected shifts along with the integration of each peak, the conversion percent yields using the three photocatalysts were calculated to be 18, 39, and 80% for BiFeO3, BiOF, and BiOF/BiFeO3 composite, respectively, with no starting material or any other byproduct detected (Table 1). Due to rapid charge/hole recombination at lower band gaps, BiFeO3 shows low conversion, while the composite had the highest conversion because the recombination was suppressed.
The enhanced yield of formation under UV light irradiation for the prepared BiOF/BiFeO3 composite heterojunction in comparison to pure BiOF or BFO is mainly explained by the fact that a heterojunction formed by adding BiOF could absorb more photons and act as an efficient transfer of excited electrons from the BiOF conduction band to the conduction band of BFO particles; moreover, a transfer from the valence band of BiOF to the valence band of BFO suppresses recombination and allows greater electron transfer to benzyl amine.
To provide more evidence regarding the completion of the coupling reaction, FTIR spectra were recorded for the starting material and compared to the product as well. Upon coupling, the characteristic band of -NH stretch in benzyl amine (3200 and 3500 cm−1) disappeared, and a new band at 1643 cm−1 appeared, which corresponded to the C=N of the product. Moreover, the bands of the benzene ring were obvious and appeared between 1650 and 1800 cm−1 (Figure S7, Supporting Information).
Different control experiments have been carried out to investigate other reaction conditions. As illustrated in Table 2, using the BiOF/BiFeO3 composite photocatalyst showed the highest conversion. The percent yields were calculated as previously reported [33]. The highest yields of N-benzylidenebenzylamine products were obtained in an open-air and acetonitrile environment, while 0, 7, and 13% were obtained using BiFeO3, BiOF, and BiOF/BiFeO3 composite, respectively, in a solvent-free condition (Entries II–IV), which confirms that our photocatalysts work best in the solvent environment. Many other control experiments have shown no conversion yields in the absence of either light (Entries V–VII) or photocatalysts (entries VIII), which strongly supports the high photocatalytic activity of the reported photocatalysts under visible or UV irradiation.
A plausible mechanism of the oxidative coupling of benzylamine is explained in Scheme 2. The process happens through an O2 mediated pathway when the photoexcited electron of the photocatalyst first rises to the conduction band (CB) upon radiation. It then reacts with atmospheric molecular oxygen to produce an O2 radical, which was previously confirmed using electron spin resonance (EPR) measurements [36,37,38]. An aminium radical cation intermediate is then formed by oxidized benzylamine by the photoinduced hole in the valence band (VB). Hydrogen-atom abstraction from the O2 radical then occurs at this radical cation, producing phenylmethanamine (Ph–CH2–NH2), and, thus, forming Ph–CH=NH and H2O2 [36]. H2O2 is then dissociated to ·OH radicals, which act as a supporting material for benzylamine oxidation to the produced imine that finally reacts with another free benzylamine to produce the N-benzylidenebenzylamine product, liberating ammonia.

4. Conclusions

In this study, benzylamine was photocatalytically coupled to N-benzylidenebenzylamine using the newly synthesized BiOF/BiFeO3 composite photocatalyst, which showed a high conversion yield of 80% compared to other materials. The composite’s construction was proven by multiple characterization methods, including XRD, DRS, SEM, EDX, and FT-IR. According to the measured band gaps, the BiOF/BiFeO3 composite exhibited a lower band gap compared to BiOF (3.5 eV vs. 3.8 eV), which can be successfully employed as a photocatalyst in the UV region. The reported synthesis procedure provides an easy, eco-friendly, and low-cost route for developing various active materials for photocatalytic applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/photochem3010012/s1, Figure S1: Matched X-ray spectroscopy of the BiOF using Match!3 software (version 3.0). Figure S2: Matched X-ray spectroscopy of the BiFeO3 using Match!3 software (version 3.0); Figure S3: EDX elemental analysis of BiOF; Figure S4: EDX elemental analysis of BiFeO3; Figure S5: EDX elemental analysis of BiOF/ BiFeO3 composite; Table S1: Weight and atomic percentages of BiOF; Table S2: Weight and atomic percentages of BiFeO3; Table S3: Weight and atomic percentages of BiOF/ BiFeO3 composite; Figure S6: Spectra of N-benzylidenebenzylamine product using 1H-NMR (A) and 13C-NMR (B); Figure S7: FTIR spectrum of N-benzylidenebenzylamine.

Author Contributions

Conceptualization, A.A. and A.S.A.; methodology, A.S.A., R.H.A. and L.A.S.; software, A.S.A., R.H.A. and L.A.S.; validation, A.A.; formal analysis, A.S.A., R.H.A. and L.A.S.; investigation, A.S.A., R.H.A., L.A.S., A.E., D.A., Z.K., H.Z. and R.E.; resources, A.A.; data curation, A.S.A., R.H.A., L.A.S. and A.A.; writing—original draft preparation, A.S.A., R.H.A., L.A.S. and A.A.; writing—review and editing, A.A.; visualization, A.S.A.; supervision, A.A.; project administration, A.A.; funding acquisition, A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by United Arab Emirates University, grant no. 31R238 and SURE Plus 2022 project, grant code G00003232.

Data Availability Statement

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

Acknowledgments

This work was financially supported by the National Water and Energy Center and by the United Arab Emirates University.

Conflicts of Interest

All authors certify that the submission is original work and is not under review at any other publication.

References

  1. Smith, R.T.; Achenbach, G.D.; Gerson, R.; James, W.J. Dielectric Properties of Solid Solutions of BiFeO3 with Pb(Ti, Zr)O3 at High Temperature and High Frequency. J. Appl. Phys. 1968, 39, 70–74. [Google Scholar] [CrossRef]
  2. Bhide, V.G.; Multani, M.S. Mössbauer Effect in Ferroelectric-Antiferromagnetic BiFeO3. Solid State Commun. 1965, 3, 271–274. [Google Scholar] [CrossRef]
  3. Béa, H.; Bibes, M.; Cherifi, S.; Nolting, F.; Warot-Fonrose, B.; Fusil, S.; Herranz, G.; Deranlot, C.; Jacquet, E.; Bouzehouane, K.; et al. Tunnel Magnetoresistance and Robust Room Temperature Exchange Bias with Multiferroic BiFeO3 Epitaxial Thin Films. Appl. Phys. Lett. 2006, 89, 3–6. [Google Scholar] [CrossRef] [Green Version]
  4. Béa, H.; Bibes, M.; Sirena, M.; Herranz, G.; Bouzehouane, K.; Jacquet, E.; Fusil, S.; Paruch, P.; Dawber, M.; Contour, J.P.; et al. Combining Half-Metals and Multiferroics into Epitaxial Heterostructures for Spintronics. Appl. Phys. Lett. 2006, 88, 4–7. [Google Scholar]
  5. Tong, T.; Chen, J.; Jin, D.; Cheng, J. Preparation and Gas Sensing Characteristics of BiFeO3 Crystallites. Mater. Lett. 2017, 197, 160–162. [Google Scholar] [CrossRef]
  6. Das, S.; Rana, S.; Mursalin, S.M.; Rana, P.; Sen, A. Sonochemically Prepared Nanosized BiFeO3 as Novel SO2 Sensor. Sensors Actuators B Chem. 2015, 218, 122–127. [Google Scholar] [CrossRef]
  7. Gao, F.; Chen, X.Y.; Yin, K.B.; Dong, S.; Ren, Z.F.; Yuan, F.; Yu, T.; Zou, Z.G.; Liu, J.-M. Visible-Light Photocatalytic Properties of Weak Magnetic BiFeO3 Nanoparticles. Adv. Mater. 2007, 19, 2889–2892. [Google Scholar] [CrossRef]
  8. Xian, T.; Yang, H.; Dai, J.F.; Wei, Z.Q.; Ma, J.Y.; Feng, W.J. Photocatalytic Properties of BiFeO3 Nanoparticles with Different Sizes. Mater. Lett. 2011, 65, 1573–1575. [Google Scholar] [CrossRef]
  9. Mushtaq, F.; Chen, X.; Hoop, M.; Torlakcik, H.; Pellicer, E.; Sort, J.; Gattinoni, C.; Nelson, B.J.; Pané, S. Piezoelectrically Enhanced Photocatalysis with BiFeO3 Nanostructures for Efficient Water Remediation. iScience 2018, 4, 236–246. [Google Scholar] [CrossRef]
  10. Sando, D.; Yang, Y.; Bousquet, E.; Carrétéro, C.; Garcia, V.; Fusil, S.; Dolfi, D.; Barthélémy, A.; Ghosez, P.; Bellaiche, L.; et al. Large Elasto-Optic Effect and Reversible Electrochromism in Multiferroic BiFeO3. Nat. Commun. 2016, 7, 10718. [Google Scholar] [CrossRef] [Green Version]
  11. Chu, S.H.; Singh, D.J.; Wang, J.; Li, E.-P.; Ong, K.P. High Optical Performance and Practicality of Active Plasmonic Devices Based on Rhombohedral BiFeO3. Laser Photonics Rev. 2012, 6, 684–689. [Google Scholar] [CrossRef]
  12. Zeches, R.J.; Rossell, M.D.; Zhang, J.X.; Hatt, A.J.; He, Q.; Yang, C.-H.; Kumar, A.; Wang, C.H.; Melville, A.; Adamo, C.; et al. A Strain-Driven Morphotropic Phase Boundary in BiFeO3. Science 2009, 326, 977–980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Kumar, A.; Rai, R.C.; Podraza, N.J.; Denev, S.; Ramirez, M.; Chu, Y.H.; Martin, L.W.; Ihlefeld, J.; Heeg, T.; Schubert, J.; et al. Linear and Nonlinear Optical Properties of BiFeO3. Appl. Phys. Lett. 2008, 92, 67–70. [Google Scholar] [CrossRef] [Green Version]
  14. Li, S.; Lin, Y.-H.; Zhang, B.-P.; Wang, Y.; Nan, C.-W. Controlled Fabrication of BiFeO3 Uniform Microcrystals and Their Magnetic and Photocatalytic Behaviors. J. Phys. Chem. C 2010, 114, 2903–2908. [Google Scholar] [CrossRef]
  15. Alzard, R.H.; Siddig, L.A.; Alhatti, N.; Abdallah, I.; Aljabri, L.; Alblooshi, A.; Alzamly, A. Titania Derived from NH2-MIL-125(Ti) Metal—Organic Framework for Selective Photocatalytic Conversion of CO2 to Propylene Carbonate. Comments Inorg. Chem. 2022, 43, 1–15. [Google Scholar] [CrossRef]
  16. 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. 2017, 90, 15–30. [Google Scholar] [CrossRef]
  17. Rouhani, Z.; Karimi-Sabet, J.; Mehdipourghazi, M.; Hadi, A.; Dastbaz, A. Response Surface Optimization of Hydrothermal Synthesis of Bismuth Ferrite Nanoparticles under Supercritical Water Conditions: Application for Photocatalytic Degradation of Tetracycline. Environ. Nanotechnol. Monit. Manag. 2019, 11, 100198. [Google Scholar] [CrossRef]
  18. Haruna, A.; Abdulkadir, I.; Idris, S.O. Photocatalytic Activity and Doping Effects of BiFeO3 Nanoparticles in Model Organic Dyes. Heliyon 2020, 6, e03237. [Google Scholar] [CrossRef] [Green Version]
  19. Kumar, V.; Soam, A.; Sahoo, P.K.; Panda, H.S. Enhancement of Electrochemical Properties of Carbon Solution Doped Bismuth Ferrite for Supercapacitor Application. Mater. Today Proc. 2021, 41, 165–171. [Google Scholar] [CrossRef]
  20. Sankar Ganesh, R.; Sharma, S.K.; Sankar, S.; Divyapriya, B.; Durgadevi, E.; Raji, P.; Ponnusamy, S.; Muthamizhchelvan, C.; Hayakawa, Y.; Kim, D.Y. Microstructure, Structural, Optical and Piezoelectric Properties of BiFeO3 Nanopowder Synthesized from Sol-Gel. Curr. Appl. Phys. 2017, 17, 409–416. [Google Scholar] [CrossRef]
  21. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides. Science. 2001, 293, 269–271. [Google Scholar] [CrossRef] [PubMed]
  22. Cho, C.M.; Noh, J.H.; Cho, I.S.; An, J.S.; Hong, K.S.; Kim, J.Y. Low-Temperature Hydrothermal Synthesis of Pure BiFeO3 Nanopowders Using Triethanolamine and Their Applications as Visible-Light Photocatalysts. J. Am. Ceram. Soc. 2008, 91, 3753–3755. [Google Scholar] [CrossRef]
  23. Sharif, M.K.; Khan, M.A.; Hussain, A.; Iqbal, F.; Shakir, I.; Murtaza, G.; Akhtar, M.N.; Ahmad, M.; Warsi, M.F. Synthesis and Characterization of Zr and Mg Doped BiFeO3 Nanocrystalline Multiferroics via Micro Emulsion Route. J. Alloys Compd. 2016, 667, 329–340. [Google Scholar] [CrossRef]
  24. Shami, M.Y.; Awan, M.S.; Anis-ur-Rehman, M. Phase Pure Synthesis of BiFeO3 Nanopowders Using Diverse Precursor via Co-Precipitation Method. J. Alloys Compd. 2011, 509, 10139–10144. [Google Scholar] [CrossRef]
  25. Azam, A.; Jawad, A.; Ahmed, A.S.; Chaman, M.; Naqvi, A.H. Structural, Optical and Transport Properties of Al3+ Doped BiFeO3 Nanopowder Synthesized by Solution Combustion Method. J. Alloys Compd. 2011, 509, 2909–2913. [Google Scholar] [CrossRef]
  26. Wani, W.A.; Kundu, S.; Ramaswamy, K.; Venkataraman, B.H. Optimizing Phase Formation of BiFeO3 and Mn-Doped BiFeO3 Nanoceramics via Thermal Treatment Using Citrate Precursor Method. SN Appl. Sci. 2020, 2, 1969. [Google Scholar] [CrossRef]
  27. Haruna, A.; Abdulkadir, I.; Idris, S.O. Synthesis, Characterization and Photocatalytic Properties of Bi0.85-XMXBa0.15FeO3 (M = Na and K, X = 0, 0.1) Perovskite-like Nanoparticles Using the Sol-Gel Method. J. King Saud Univ.-Sci. 2020, 32, 896–903. [Google Scholar] [CrossRef]
  28. Alzamly, A.; Bakiro, M.; Ahmed, S.H.; Sallabi, S.M.; Al Ajeil, R.A.; Alawadhi, S.A.; Selem, H.A.; Al Meshayei, S.S.M.; Khaleel, A.; Al-Shamsi, N.; et al. Construction of BiOF/BiOI Nanocomposites with Tunable Band Gaps as Efficient Visible-Light Photocatalysts. J. Photochem. Photobiol. A Chem. 2019, 375, 30–39. [Google Scholar] [CrossRef]
  29. Li, H.; Yu, K.; Lei, X.; Guo, B.; Fu, H.; Zhu, Z. Hydrothermal Synthesis of Novel MoS2/BiVO4 Hetero-Nanoflowers with Enhanced Photocatalytic Activity and a Mechanism Investigation. J. Phys. Chem. C 2015, 119, 22681–22689. [Google Scholar] [CrossRef]
  30. Dai, G.; Liu, S.; Liang, Y.; Luo, T. Synthesis and Enhanced Photoelectrocatalytic Activity of P-n Junction Co3O4/TiO2 Nanotube Arrays. Appl. Surf. Sci. 2013, 264, 157–161. [Google Scholar] [CrossRef]
  31. Zhang, J.; Qiao, S.Z.; Qi, L.; Yu, J. Fabrication of NiS Modified CdS Nanorod P-n Junction Photocatalysts with Enhanced Visible-Light Photocatalytic H2-Production Activity. Phys. Chem. Chem. Phys. 2013, 15, 12088–12094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Zhou, M.; Yu, J.; Liu, S.; Zhai, P.; Jiang, L. Effects of Calcination Temperatures on Photocatalytic Activity of SnO2/TiO2 Composite Films Prepared by an EPD Method. J. Hazard. Mater. 2008, 154, 1141–1148. [Google Scholar] [CrossRef] [PubMed]
  33. Alzard, R.H.; Siddig, L.A.; Abdelhamid, S.A.; Alzamly, A. Visible-Light-Driven Photocatalytic Coupling of Neat Benzylamine over a Bi-Ellagate Metal–Organic Framework. ACS Omega 2022, 7, 36689–36696. [Google Scholar] [CrossRef] [PubMed]
  34. Guo, J.; Ma, D.; Sun, F.; Zhuang, G.; Wang, Q.; Al-Enizi, A.M.; Nafady, A.; Ma, S. Substituent Engineering in G-C3N4/COF Heterojunctions for Rapid Charge Separation and High Photo-Redox Activity. Sci. China Chem. 2022, 65, 1704–1709. [Google Scholar] [CrossRef]
  35. Fu, Y.; Tan, M.; Guo, Z.; Hao, D.; Xu, Y.; Du, H.; Zhang, C.; Guo, J.; Li, Q.; Wang, Q. Fabrication of Wide-Spectra-Responsive NA/NH2-MIL-125(Ti) with Boosted Activity for Cr(VI) Reduction and Antibacterial Effects. Chem. Eng. J. 2023, 452, 139417. [Google Scholar] [CrossRef]
  36. Li, J.; Chang, B.; Zhao, H.; Meng, Q.; Li, M.; Han, Q. Visible-Light-Responsive Polyoxometalate-Based Metal–Organic Framework for Highly Efficient Photocatalytic Oxidative Coupling of Amines. J. Mater. Sci. 2021, 56, 6676–6688. [Google Scholar] [CrossRef]
  37. Su, F.; Mathew, S.C.; Möhlmann, L.; Antonietti, M.; Wang, X.; Blechert, S. Aerobic Oxidative Coupling of Amines by Carbon Nitride Photocatalysis with Visible Light. Angew. Chem. Int. Ed. 2011, 50, 657–660. [Google Scholar] [CrossRef]
  38. Zheng, S.; Du, H.; Yang, L.; Tan, M.; Li, N.; Fu, Y.; Hao, D.; Wang, Q. PDINH Bridged NH2-UiO-66(Zr) Z-Scheme Heterojunction for Promoted Photocatalytic Cr(VI) Reduction and Antibacterial Activity. J. Hazard. Mater. 2023, 447, 130849. [Google Scholar] [CrossRef]
Scheme 1. Schematic syntheses of the hydrothermal method for preparing BiFeO3, BiOF, and BiOF/BiFeO3 composite using hydrothermal methods.
Scheme 1. Schematic syntheses of the hydrothermal method for preparing BiFeO3, BiOF, and BiOF/BiFeO3 composite using hydrothermal methods.
Photochem 03 00012 sch001
Figure 1. XRD patterns of the synthesized photocatalysts. BiFeO3 is presented by (⬥) and BiOF is presented by (*).
Figure 1. XRD patterns of the synthesized photocatalysts. BiFeO3 is presented by (⬥) and BiOF is presented by (*).
Photochem 03 00012 g001
Figure 2. UV–Vis DRS spectra (A) and Tauc plots of prepared photocatalysts (B). The intercept at the x axis of the extended linear dashed lines in the graph determines the band gap values.
Figure 2. UV–Vis DRS spectra (A) and Tauc plots of prepared photocatalysts (B). The intercept at the x axis of the extended linear dashed lines in the graph determines the band gap values.
Photochem 03 00012 g002
Figure 3. SEM images of the photocatalysts material at 5 µm scale bar.
Figure 3. SEM images of the photocatalysts material at 5 µm scale bar.
Photochem 03 00012 g003
Figure 4. FTIR spectra of the prepared photocatalysts with corresponding band assignments.
Figure 4. FTIR spectra of the prepared photocatalysts with corresponding band assignments.
Photochem 03 00012 g004
Scheme 2. Proposed mechanism of benzylamine coupling over reported photocatalyst.
Scheme 2. Proposed mechanism of benzylamine coupling over reported photocatalyst.
Photochem 03 00012 sch002
Table 1. Photocatalytic coupling of benzylamine to N-benzylidenebenzylamine over BiFeO3, BiOF, and BiFeO3/BiOF composite.
Table 1. Photocatalytic coupling of benzylamine to N-benzylidenebenzylamine over BiFeO3, BiOF, and BiFeO3/BiOF composite.
Photochem 03 00012 i001
PhotocatalystConversion Yield (%)
BiFeO318
BiOF39
BiOF/BiFeO3 composite80
Reaction optimum conditions: 0.5 mmol (51 µL) benzylamine, and 0.160, 0.020, and 0.009 mmol (5 mg) of BiFeO3, BIOF, and BiFeO3/BiOF composite, respectively; open air; irradiated for 24 h.
Table 2. Photocatalytic activities under different control conditions.
Table 2. Photocatalytic activities under different control conditions.
EntryControl ConditionsConversion Yield (%)
I5 mg BiOF/BiFeO3 composite, light, solvent80
IIBiOF/BiFeO3 composite, no solvent13
IIIBiFeO3, no solventNil
IVBiOF, no solvent7
VBiOF/BiFeO3 composite, no light, no heatNil
VIBiOF/BiFeO3 composite, heat, no light2
VIIHeat, no light, no BiOF/BiFeO3 compositeNil
VIIILight, no BiOF/BiFeO3 compositeNil
Reaction optimum conditions: 0.5 mmol (51 µL) benzylamine, 0.16, 0.02, and 0.009 mmol (5 mg) of BiFeO3, BIOF, and BiOF/BiFeO3 composite, respectively; open air; irradiated for 24 h.
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

Abdelhamid, A.S.; Alzard, R.H.; Siddig, L.A.; Elbahnasy, A.; Aljazmati, D.; Kadoura, Z.; Zeidane, H.; Elshikh, R.; Alzamly, A. Facile Synthesis, Characterization, and Photocatalytic Performance of BiOF/BiFeO3 Hybrid Heterojunction for Benzylamine Coupling under Simulated Light Irradiation. Photochem 2023, 3, 187-196. https://doi.org/10.3390/photochem3010012

AMA Style

Abdelhamid AS, Alzard RH, Siddig LA, Elbahnasy A, Aljazmati D, Kadoura Z, Zeidane H, Elshikh R, Alzamly A. Facile Synthesis, Characterization, and Photocatalytic Performance of BiOF/BiFeO3 Hybrid Heterojunction for Benzylamine Coupling under Simulated Light Irradiation. Photochem. 2023; 3(1):187-196. https://doi.org/10.3390/photochem3010012

Chicago/Turabian Style

Abdelhamid, Abdalla S., Reem H. Alzard, Lamia A. Siddig, Aya Elbahnasy, Duha Aljazmati, Zaina Kadoura, Hind Zeidane, Rufaida Elshikh, and Ahmed Alzamly. 2023. "Facile Synthesis, Characterization, and Photocatalytic Performance of BiOF/BiFeO3 Hybrid Heterojunction for Benzylamine Coupling under Simulated Light Irradiation" Photochem 3, no. 1: 187-196. https://doi.org/10.3390/photochem3010012

Article Metrics

Back to TopTop