Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Construction of a 1D/0D/2D BiFeO3/Ag/g-C3N4 Z-scheme heterojunction for enhanced visible light photocatalysis of methylene blue

Donghai Li, Yunrui Xu, Shilin Zhang and Linping Wang*
College of Chemical Engineering, Qinghai University, Xining-810016, China. E-mail: wanglp@qhu.edu.cn

Received 6th July 2025 , Accepted 19th August 2025

First published on 2nd September 2025


Abstract

To improve the utilization of solar energy and the efficiency of photocatalytic organic pollutant degradation, novel Z-scheme heterojunctions with high visible light catalytic performances have been widely developed. Herein, a novel Z-scheme BiFeO3/Ag/g-C3N4 heterojunction with a hierarchical 1D/0D/2D structure and visible light absorption was constructed by matching the suitable band structure between 1D BiFeO3 and 2D g-C3N4 and employing the localized surface plasmon resonance (LSPR) effect of 0D Ag nanoparticles. BiFeO3 nanofibers were synthesized via the electrospinning technique, providing short electron transport paths and visible light absorption range for g-C3N4. Plasmonic Ag nanoparticles were photodeposited on the surface of BiFeO3 to enhance the separation efficiency of the photogenerated electron–hole pairs between the bulk interfaces. UV-vis DRS, PL, photocurrent and EIS spectra confirmed the important roles of Ag and BiFeO3 in improving the photocatalytic activity of the Z-scheme heterojunction. The working principle of the BiFeO3/Ag/g-C3N4 Z-scheme heterojunction was proposed, indicating that ˙O2 and ˙OH are the main active species for the photocatalytic degradation of methylene blue (MB). The prepared BiFeO3/Ag/g-C3N4 heterojunction exhibited the highest photocatalytic activity, where its rate constant was 5.84 times higher than that of the pristine BiFeO3 and 3.26 times higher than that of the pristine g-C3N4. This study offers a new means for the design of novel high-performance photocatalysts, which can be a promising candidate in industrial applications.


1. Introduction

Nowadays, solar-based photocatalytic technology is a promising solution to the global energy crisis and environmental remediation, including photocatalytic hydrogen production, photocatalytic reduction of CO2 and photocatalytic degradation of organic pollutants.1 Methylene blue (MB), a sulphur-containing cationic dye, is regarded as a common pollutant in wastewater discharged from textile industries.2–5 MB poses a carcinogenic risk to humans and many aquatic organisms and affects the balance of the entire aquatic ecosystem.6 Photocatalytic degradation of MB is considered a high-efficiency, low-cost and environmentally friendly water treatment technology, which needs high-performance photocatalysts for the decomposition of MB into CO2, H2O and some harmless organic molecules.7 However, single photocatalytic materials, such as TiO2, suffer from the drawbacks of low utilization of the visible light, easy recombination of the photogenerated electron–hole pairs, and low photocatalytic activity.8,9 As a typical 2D conjugated polymer photocatalyst, g-C3N4 possesses the advantages of low cost, non-toxicity, and high specific surface area.10,11 However, g-C3N4 can absorb sunlight only at wavelengths below 460 nm and has a low valence band level (EVB = 1.50 eV), which is lower than the oxidation potential for the generation of hydroxyl radicals (˙OH), and thus, its efficiency for the photocatalytic degradation of pollutants is significantly weakened.12

To enhance the photocatalytic activity of photocatalysts, numerous strategies have been explored, such as defect engineering, ion doping, and heterojunction construction.13–15 Among the numerous approaches, heterojunction construction is regarded as a promising strategy due to its merits in promoting photogenerated carrier separation and enhancing the light absorption and stability.16 Generally, the interface structures of heterojunctions are composed of two or more different semiconductor materials, which can be divided into type I, type II, and Z-schemes. Compared with type I and type II heterojunctions, Z-scheme heterojunctions enable efficient electron–hole separation with great retention of redox capability. Zheng et al. prepared a WO3/g-C3N4 Z-scheme photocatalyst by calcination and hydrothermal treatment to achieve high redox capacity, efficient charge separation and large reaction surface area for the photocatalytic degradation of dodecylmorpholine (DMP), and its degradation efficiency was 73% in 60 min.17 However, the electron transfer rate between heterojunction interfaces greatly affects the efficiency for the photocatalytic degradation of pollutants. In this case, the incorporation of metal nanoparticles exhibiting localized surface plasmon resonance (LSPR) within Z-scheme heterojunctions can accelerate the electron transfer rate across the interface and broaden the absorption spectrum to visible light with enhanced the solar energy utilisation. The Schottky junction at the interface between the metal and the semiconductor inhibits the compounding of photogenerated carriers.

As a typical multiferroic perovskite material, BiFeO3 exhibits excellent photovoltaic, thermal, specific ferroelectric and magnetic properties. Especially in the field of photocatalysis, its polarization effect leads to band bending, resulting in the effective separation of excitons.18,19 In this study, BiFeO3 nanofibres with a suitable band gap (ECB = 0.51 eV, EVB = 2.61 eV, and Eg = 2.10 eV) are matched with g-C3N4 (ECB = −1.22 eV, EVB = 1.50 eV, and Eg = 2.72 eV) to form an ideal Z scheme heterojunction.20,21 The BiFeO3 nanofibres with a large aspect ratio can largely shorten the photogenerated electron transfer pathways.22 The Ag nanoparticles acted as a direct electron bridge between BiFeO3 and g-C3N4 to construct an efficient Z-scheme heterojunction, which reduced the electron transfer resistance between the interfaces.23 The LSPR effect of the Ag nanoparticles further broadened the light absorption range and lowered the Schottky barrier to enhance the solar energy utilization. We achieved the degradation of MB using the 1D/0D/2D BiFeO3/Ag/g-C3N4 Z-scheme heterojunction and proposed its rational mechanism via free radical trapping experiments. This study provides new insights into the development of high-performance Z-scheme photocatalysts for the efficient photocatalytic degradation of pollutants in wastewater.

2. Experimental

2.1. Materials

Bismuth nitrate pentahydrate (Bi(NO3)3·5H2O, 99.9%), melamine (C3H6N6, 99%), iron(III) nitrate nonahydrate (Fe(NO3)3·9H2O, 99.9%) and polyvinylpyrrolidone (PVP) with an average molecular weight of 1[thin space (1/6-em)]200[thin space (1/6-em)]000 were purchased from Shanghai Aladdin Reagent Co., Ltd. Silver nitrate (AgNO3, 99.8%) was purchased from Oubokai Chemical Co., Ltd. N,N-Dimethylformamide (DMF, 99.5%) and acetone (AC, C3H6O, 99.5%) were purchased from Shanghai Guoyao Group Chemical Reagent Co., Ltd. Anhydrous ethanol (C2H5OH, 99.7%) and methanol (CH3OH, 99.5%) were purchased from Chengdu Chron Chemical Co., Ltd. Methylene blue (C16H18N3ClS, 98.5%) was purchased from Shanghai Zhanyun Chemical Co., Ltd.

2.2. Synthesis of g-C3N4

g-C3N4 was prepared by pyrolyzing melamine.24 In detail, 10 g of melamine powder was initially placed in a lidded ceramic crucible and subjected to heating in a muffle furnace. The furnace was heated at a rate of 5 °C min−1 until the temperature reached 550 °C and was maintained at this temperature for 2 h. At the end of the calcination process, the material was cooled naturally to room temperature and ground with a mortar and pestle to obtain the final g-C3N4 powder.

2.3. Synthesis of BiFeO3

BiFeO3 nanofibres were prepared by the electrospinning method. 1.00 g Bi(NO3)3·5H2O and 0.75 g Fe(NO3)3·9H2O were dissolved in 2.50 mL ultrapure water, and then 1.50 mL glacial acetic acid was added and stirred for 2 h. In parallel, 0.93 g PVP was dissolved in 5.63 mL mixed solution of AC and DMF (AC/DMF = 1[thin space (1/6-em)]:[thin space (1/6-em)]2). These two solutions were mixed and stirred for 4 h to obtain a golden-yellow precursor solution for electrostatic spinning. An applied voltage of 16.50 kV and an advancement speed of 0.05 mm min−1 were used to obtain the BiFeO3 nanofibre precursor. The BiFeO3 nanofibre precursor was placed in a muffle furnace and heated to 550 °C for 2 h at a ramp rate of 1 °C min−1. The calcined samples were cooled to room temperature in an air environment and placed in an onyx mortar for grinding to obtain pure BiFeO3 nanofibres.

2.4. Synthesis of BiFeO3/Agx

The BiFeO3/Agx compound was synthesised via the photodeposition of Ag nanoparticles on BiFeO3 nanofibres by using CH3OH as a reducing agent. 0.1 g BiFeO3 nanofibres and a certain amount of AgNO3 were added to a mixed solution of 10 mL CH3OH and 40 mL H2O, and sonicated for 10 min. The mixed solution was irradiated by a xenon lamp for 2 h. After irradiation, the mixed solution was centrifuged at 12[thin space (1/6-em)]000 rpm for 10 min, washed twice with ethanol, and dried under vacuum at 60 °C for 12 h to obtain the BiFeO3/Agx samples, where x denotes the mass fraction of Ag.

2.5. Synthesis of (BiFeO3/Agx)/(g-C3N4)y

0.1 g of synthesised g-C3N4 was added to 50 mL of ultrapure water and sonicated for 0.5 h to form a homogeneous solution. A certain amount of BiFeO3/Agx was dissolved in the solution of g-C3N4 and ultrasonicated for 2 h. The mixed solution was subsequently centrifuged at 12[thin space (1/6-em)]000 rpm, and the resulting precipitate was dried under vacuum at 60 °C for 24 h to obtain the (BiFeO3/Agx)/(g-C3N4)y composites, where the ratio of y is defined as the mass fraction of g-C3N4, as shown in Fig. 1.
image file: d5ra04825g-f1.tif
Fig. 1 Schematic of the preparation of BiFeO3/Ag/g-C3N4.

2.6. Characterization

The morphologies of the samples were characterized using a scanning electron microscope (SEM, JSM-7900F, JEOL). The elemental composition of the samples was characterized using an energy-dispersive X-ray spectrometer (EDX, JSM-7900F, JEOL). The specific surface area of the photocatalyst was determined using a fully automatic specific surface area and porosity analyzer (BET, ASAP 2460, Micromeritics). The phase identification of the samples was performed on a powder X-ray diffractometer (XRD, Ultima IV, Rigaku). Elemental states and surface chemistry were examined by X-ray photoelectron spectroscopy (XPS, Nexsa G2, Thermo Fisher). The absorbance of the pollutants was determined by ultraviolet-visible spectroscopy (UV-vis, UV-2550, Shimadzu). The functional groups of the materials were tested using a Fourier transform infrared spectrometer (FTIR, VERTEX70V, Bruker). The photogenerated electron–hole pair recombination rates were analysed using a photoluminescence spectrometer (PL, Cary Eclipse, Agilent), operating at room temperature with an excitation wavelength of 300 nm. The light absorption performances of the samples were tested using a solid-state ultraviolet-visible diffuse reflectance spectrometer (DRS, Cary 5000, Agilent). The intermediate products of the photocatalytic degradation of MB were determined by liquid chromatography-mass spectrometry (LC-MS, Integrion + TSQ Fortis Plus, Thermo Scientific).

2.7. Electrochemical measurements

Photoelectrochemical properties were measured by means of a conventional three-electrode system. Sample-modified indium tin oxide-coated glass was used as the working electrode, a carbon rod as the counter electrode, and a saturated calomel electrode as the reference electrode. In the photocurrent tests, a 0.1 mol L−1 ascorbic acid (AA) solution was used as the electrolyte and a 300 W xenon lamp was used as the light source, which was switched on and off at 20 s intervals during the test period of 0–160 s. In the Nyquist plot measurements, a 2.0 mM 1[thin space (1/6-em)]:[thin space (1/6-em)]1 K3[Fe(CN)6]/K4[Fe(CN)6] and 0.1 mol L−1 Na2SO4 solution was used as the electrolyte with an applied bias voltage of 0.2 V. In the Mott–Schottky tests, a 0.5 mol L−1 Na2SO4 solution was used as the electrolyte.

2.8. Photocatalytic activity evaluation

The catalytic performance of the prepared samples was evaluated using MB solution under simulated sunlight (xenon lamp = 300 W). 10 mg catalyst was added to 50 mL of MB solution (c = 10 mg L−1) at room temperature and magnetically stirred in the dark for 30 min to ensure adsorption and desorption equilibrium. Under light illumination, 2 mL of the suspension was collected and filtered through a 0.45 μm mixed cellulose ester (MCE) membrane. The photocatalytic activity of the catalyst was analyzed by measuring its absorbance at 664 nm on a UV-vis spectrophotometer.

3. Results

3.1. Morphology and structure analysis

3.1.1. XRD analysis. The XRD analysis confirmed the successful preparation of g-C3N4, BiFeO3, and (BiFeO3/Ag0.05)/(g-C3N4)0.3, as shown in Fig. 2a. The main diffraction peak located at the 2θ value of 27.5° is attributed to the (002) crystal plane of g-C3N4 (JCPDS no. 50-1512).25 The main diffraction peaks for the prepared BiFeO3 at 2θ = 22.5°, 32.1°, 32.3°, 39.8°, 45.8°, and 52.3° correspond to the (012), (104), (110), (202), (024), and (116) crystal planes of BiFeO3, respectively (JCPDS no. 86-1518).26 Comparatively, (BiFeO3/Ag0.05)/(g-C3N4)0.3 exhibits the main characteristic peaks of BiFeO3, g-C3N4, and Ag for the (111), (200), and (220) crystal planes at 2θ = 37.6°, 44.0° and 63.8°, respectively (JCPDS no. 65-2871).27 We performed Rietveld refinement on (BiFeO3/Ag0.05)/(g-C3N4)0.3. As shown in Fig. 2b, peaks corresponding to the g-C3N4 phase, BiFeO3 phase, and Ag0 phase can clearly be observed. By simulating g-C3N4 (a = 6.4505 Å, b = 6.4505 Å, c = 2.4231 Å and V = 87.3171 Å3),28 BiFeO3 (a = 5.5881 Å, b = 5.5881 Å, c = 13.9072 Å and V = 376.0994 Å3),29 and Ag (a = 4.0842 Å, b = 4.0842 Å, c = 4.0842 Å and V = 68.1923 Å3),30 we obtained the lattice parameters and phase composition of the material. As shown in Table 1, the lattice parameters of g-C3N4 (a = 6.4456 Å, b = 6.4456 Å, c = 2.4206 Å and V = 87.1032 Å3), BiFeO3 (a = 5.5786 Å, b = 5.5786 Å, c = 13.8601 Å and V = 373.5562 Å3), and Ag (a = 4.0683 Å, b = 4.0683 Å, c = 4.0683 Å and V = 68.1923 Å3) are almost identical to their standard card values, indicating that the synthesized material has high crystallinity and purity.
image file: d5ra04825g-f2.tif
Fig. 2 (a) XRD patterns of g-C3N4, BiFeO3, and (BiFeO3/Ag0.05)/(g-C3N4)0.3. (b) Rietveld refinement XRD patterns of (BiFeO3/Ag0.05)/(g-C3N4)0.3.
Table 1 Microstructural parameters obtained from the Rietveld analysis of (BiFeO3/Ag0.05)/(g-C3N4)0.3
Sample Phase Space group a (Å) b (Å) c (Å) V3) RWP (%) RP (%) Volume (%)
  JCPDS no. 50-1512 P63/m 6.4505 6.4505 2.4231 87.3171
JCPDS no. 86-1518 R3c 5.5881 5.5881 13.9072 376.0994
JCPDS no. 65-2871 Fm[3 with combining macron]m 4.0842 4.0842 4.0842 68.1923
(BiFeO3/Ag0.05)/(g-C3N4)0.3 g-C3N4 P63cm 6.4456 6.4456 2.4206 87.1032 1.20 0.82 33.1
BiFeO3 R3c 5.5786 5.5786 13.8601 373.5562 65.1
Ag Fm[3 with combining macron]m 4.0683 4.0683 4.0683 68.1923 1.8


3.1.2. SEM analysis. The micromorphology of the catalysts was investigated by SEM. As shown in Fig. 3a, BiFeO3 exhibits a fibrous structure that is continuous and uniformly dispersed with a diameter of around 100 nm. The Ag nanoparticles are decorated on the BiFeO3 nanofibers with a particle size in the range of 10 to 20 nm (Fig. 3b). Non-uniform broken BiFeO3 nano-fibers are formed due to the photodeposition of Ag nanoparticles on BiFeO3. After the combination of BiFeO3/Ag with g-C3N4, flake-like g-C3N4 appeared in the nanofibers (Fig. 3c). The 1D/0D/2D BiFeO3/Ag/g-C3N4 multidimensional structure increases the specific surface area of the single materials, which was further confirmed by the N2 adsorption–desorption isotherm tests. The N2 adsorption–desorption isotherm curves are displayed in Fig. S1, and the specific surface area and pore volume values are displayed in Table S1. The surface area of (BiFeO3/Ag0.05)/(g-C3N4)0.3 is obviously higher than that of g-C3N4 and BiFeO3/Ag, facilitating the greater exposure of reactive sites to further improve the efficient ion transport and photocatalytic degradation ability. The EDX analysis confirms the presence of Bi, Fe, O, C, N and Ag elements in the (BiFeO3/Ag0.05)/(g-C3N4)0.3 composite, indicating the successful preparation of the BiFeO3/Ag/g-C3N4 heterojunction (Fig. 3d–k). The presence of Bi, Fe, O, C, N and Ag elements evenly distributed with mass ratios of 28.21%, 15.76%, 16.99%, 24.43%, 14.31%, and 0.29%, respectively, indicates the successful incorporation of a certain amount of Ag in the composite.
image file: d5ra04825g-f3.tif
Fig. 3 SEM images of (a) BiFeO3, (b) BiFeO3/Ag0.05, and (c) (BiFeO3/Ag0.05)/(g-C3N4)0.3 and (d–k) EDX element mapping of (BiFeO3/Ag0.05)/(g-C3N4)0.3.
3.1.3. FTIR analysis. FTIR spectroscopy was employed to determine the surface functional groups in g-C3N4, BiFeO3, BiFeO3/Ag0.05, and (BiFeO3/Ag0.05)/(g-C3N4)0.3, as shown in Fig. 4. In the spectrum of g-C3N4, the absorption peaks at 805 cm−1, 887 cm−1, 1230 cm−1, and 3140 cm−1 are attributed to the triazine breathing mode, the deformation mode of N–H, and the stretching vibration of the C–N and N–H bond, respectively.31–34 The peaks at 440 cm−1 and 544 cm−1 are attributed to the bending vibrations of Fe–O in the octahedral FeO6 groups, and the stretching vibrations of Fe–O and Bi–O.35 After Ag nanoparticles were deposited on the surface of BiFeO3, a blue-shift in the absorption peak was observed from 544 cm−1 to 538 cm−1.36 The spectrum of (BiFeO3/Ag0.05)/(g-C3N4)0.3 contains peaks related to g-C3N4, and two characteristic BiFeO3/Ag0.05 peaks at 440 and 538 cm−1, confirming the successful fabrication of the heterojunction structure.
image file: d5ra04825g-f4.tif
Fig. 4 FT-IR spectra of g-C3N4, BiFeO3, BiFeO3/Ag0.05, and (BiFeO3/Ag0.05)/(g-C3N4)0.3.
3.1.4. XPS analysis. To further pursue the chemical state and composition of the elements in the as-prepared (BiFeO3/Ag0.05)/(g-C3N4)0.3, its XPS spectrum was recorded, as shown in Fig. 5. Fig. 5a shows the main elements of Fe, Bi, O, N, C and Ag in (BiFeO3/Ag0.05)/(g-C3N4)0.3. Fig. 5b–g represent the high-resolution XPS spectra of Bi 4f, Fe 2p, O 1s, C 1s, N 1s and Ag 3d, respectively. The Bi 4f spectrum exhibits doublet peaks at 163.4 eV and 159.0 eV, corresponding to Bi 4f5/2 and Bi 4f7/2 of Bi3+, respectively.37 The Fe 2p spectrum of (BiFeO3/Ag0.05)/(g-C3N4)0.3 exhibits three peaks at 710.7 eV, 712.8 eV, and 724.2 eV, which correspond to Fe2+ 2p3/2, Fe3+ 2p3/2, and Fe3+ 2p1/2, respectively. The other peaks centered at 732.0 eV and 718.7 eV are designated as Fe3+ satellite peaks.38,39 The XPS O 1s spectrum displays three peaks at 529.8 eV, 532.0 eV and 533.4 eV, corresponding to the surface lattice oxygen, surface hydroxyl groups and surface adsorbed oxygen.40 In the C 1s XPS spectrum, the peak at 289.0 eV is typically associated with the N–C[double bond, length as m-dash]N in the triazine units, the peak at 284.8 eV is attributed to C–O species on the surface of g-C3N4, and the peak at 292.8 eV can be assigned to π electronic excitation.41,42 In the XPS N 1s spectrum, the peak at 397.9 eV is typically attributed to C–N[double bond, length as m-dash]C, and the peak at 399.9 eV corresponds to N–H.42 Another peak centered at 403.4 eV is related to pi-excitations on heterocycles.43 In the Ag 3d XPS spectrum, the peaks at 374.2 eV and 368.2 eV are attributed to Ag 3d5/2 and Ag 3d3/2, respectively. The difference between these two peaks is 6 eV, indicating that silver atoms are in a metallic state.44 To further confirm the electron transfer in the heterojunction, the peak positions of Bi 4f in BiFeO3, and N 1s of g-C3N4 were examined by XPS. As shown in Fig. 5h and i, compared with BiFeO3 and g-C3N4, the position of Bi 4f in (BiFeO3/Ag0.05)/(g-C3N4)0.3 shifted positively by approximately 0.2 eV, while the position of N 1s shifted negatively by approximately 1.1 eV. Similarly, as shown in Fig. 5g, the Ag 3d peak in (BiFeO3/Ag0.05)/(g-C3N4)0.3 exhibits a positive shift of approximately 0.2 eV relative to the reported values for Ag0 (368.0 and 374.0 eV).45 The peak shifting in (BiFeO3/Ag0.05)/(g-C3N4)0.3 demonstrates that electrons could migrate from BiFeO3 to Ag, and then to g-C3N4, resulting a fixed Z-scheme heterojunction photocatalytic system at the double interfaces of BiFeO3/Ag and g-C3N4/Ag.46
image file: d5ra04825g-f5.tif
Fig. 5 (a) XPS spectrum of (BiFeO3/Ag0.05)/(g-C3N4)0.3. (b–g) Narrow-scan XPS of (BiFeO3/Ag0.05)/(g-C3N4)0.3 in the regions of Bi 4f, Fe 2p, O 1s, C 1s, N 1s, and Ag 3d core levels, respectively. (h) Comparison of the XPS spectra of BiFeO3 and (BiFeO3/Ag0.05)/(g-C3N4)0.3 for Bi 4f. (i) Comparison of the XPS spectra of g-C3N4 and (BiFeO3/Ag0.05)/(g-C3N4)0.3 for N 1s.
3.1.5. UV-vis DRS and PL analysis. The UV-vis absorbance of g-C3N4, BiFeO3, BiFeO3/Ag0.05 and (BiFeO3/Ag0.05)/(g-C3N4)0.3 was tested using UV-visible diffuse reflectance spectroscopy. As shown in Fig. 6a, g-C3N4 has narrow visible light absorption ability, and the absorption edge of g-C3N4 is located at 458 nm. The pure-phase BiFeO3 exhibits strong visible light absorption properties in the range of 400–550 nm, and its absorption boundary is located at 619 nm. After loading Ag nanoparticles, the absorption edge of BiFeO3/Ag0.05 is markedly red-shifted to 660 nm. Due to the LSPR effect of the Ag nanoparticles, ‘hot’ electrons were injected into the BiFeO3 interface, further broadening the light absorption range and lowering the Schottky barrier.47 The LSPR effect is largely dependent on the size and shape of the Ag nanoparticles, and the LSPR peaks could be tuned to ∼450 nm based on the size of Ag nanoparticles in the range of 10–20 nm.48,49 The (BiFeO3/Ag0.05)/(g-C3N4)0.3 heterojunction has an excellent visible light absorption performance with an absorption edge at 594 nm.
image file: d5ra04825g-f6.tif
Fig. 6 (a) UV-vis DRS spectra of the as-prepared g-C3N4, BiFeO3, BiFeO3/Ag0.05 and (BiFeO3/Ag0.05)/(g-C3N4)0.3. (b) PL spectra of g-C3N4, BiFeO3, BiFeO3/Ag(0.01–0.07), and (BiFeO3/Ag0.05)/(g-C3N4)0.3 (300 nm excitation).

PL spectra were used to evaluate the separation rate of photogenerated electron–hole pairs. As shown in Fig. 6b, the PL emission intensity of BiFeO3/Ag is markedly reduced than that of pristine BiFeO3. This indicates that the Ag nanoparticles deposited on the surface of BiFeO3 formed a Schottky junction, which can suppress the recombination of electron–hole pairs at the interface of BiFeO3/Ag. In addition, BiFeO3/Ag0.05 exhibits the lowest intensity, and the intensity of its PL emission peak decreases with an increase in the concentration of Ag nanoparticles. Given that excessive Ag nanoparticles may accumulate, resulting in the formation of composite centers, the separation rate of photogenerated electron–hole pairs decreases.47 The PL intensity of (BiFeO3/Ag0.05)/(g-C3N4)0.3 was significantly reduced compared with that of pure BiFeO3 and g-C3N4, indicating that the separation efficiency of the photogenerated electron–hole pairs was higher than that of pure BiFeO3 and g-C3N4.

3.1.6. Photocurrent and EIS analysis. Transient photocurrent measurements were performed to investigate the photogenerated carrier separation and interfacial charge transfer efficiency of the samples. As shown in Fig. 7a, the photocurrent intensity of g-C3N4 is 0.61 μA, while that of BiFeO3 is 0.65 μA. With the addition of Ag nanoparticles, the photocurrent intensity of BiFeO3/Ag0.05 significantly increased to 1.65 μA, which is attributed to the good conductive ability of the Ag nanoparticles. The photocurrent intensity of (BiFeO3/Ag0.05)/(g-C3N4)0.5 is 4.62 μA, which is about 7.57-fold higher than that of g-C3N4 and 7.10-fold higher than that of BiFeO3. The heterojunction structure of (BiFeO3/Ag0.05)/(g-C3N4)0.3 provides a more efficient transport path for photogenerated electrons by the synergistic effect of LSPR Ag, BiFeO3 and g-C3N4.
image file: d5ra04825g-f7.tif
Fig. 7 (a) Photocurrent spectra of g-C3N4, BiFeO3, BiFeO3/Ag0.05, and (BiFeO3/Ag0.05)/(g-C3N4)0.3. (b) EIS spectra of g-C3N4, BiFeO3, BiFeO3/Ag0.05 and (BiFeO3/Ag0.05)/(g-C3N4)0.3, inset: equivalent circuit (photocurrent responses measured at a bias voltage of 0.0 V (vs. SCE) in 0.1 M ascorbic acid and EIS responses measured in 0.1 M Na2SO4 containing 2.0 mM 1[thin space (1/6-em)]:[thin space (1/6-em)]1 K3[Fe(CN)6]/K4[Fe(CN)6] after biasing at 0.2 V (vs. SCE) for 200 s). Impedance fitting results of (c) Rct and (d) Cdl data versus the applied potential. (e) Electron lifetimes of the photocatalysts versus applied potential.

Electrochemical impedance spectroscopy (EIS) of the g-C3N4, BiFeO3, BiFeO3/Ag0.05, and (BiFeO3/Ag0.05)/(g-C3N4)0.3 modified ITO electrodes was also carried out to evaluate the kinetics of interfacial charge transmission and fitted to the circuit equivalent, the Randles and Ershler model (inset of Fig. 7b).50 As shown in Fig. 7b, g-C3N4 exhibits the largest semicircle with the charge transfer resistance (Rct) value of about 3694 Ω, indicating that g-C3N4 possesses poor electron conductive ability. Compared to BiFeO3 (3113 Ω), BiFeO3/Ag0.05 has a significantly lower electrochemical impedance with an Rct value of 1350 Ω for the rapid transfer of carriers by LSPR Ag and formed Schottky junctions. The small amount of Ag loading can significantly lower the charge transfer resistance and exhibit better photogenerated carrier mobility. The (BiFeO3/Ag0.05)/(g-C3N4)0.3 heterojunction exhibits a significantly reduced arc radius than that of the other materials with an Rct value of 825 Ω. The lower charge transfer resistance of the (BiFeO3/Ag0.05)/(g-C3N4)0.3 heterojunction implies a faster electron transfer rate and smaller electron–hole recombination effect. In addition, the EIS equivalent circuit can be used to evaluate the charge-transfer abilities of photoelectrochemical systems, where the double-layer capacitance (Cdl) is the double layer charge storage capacity at the semiconductor/electrolyte interface and Rct is the charge-transfer resistance at the same interface.51 To verify the superior photo-generated carrier separation performance of (BiFeO3/Ag0.05)/(g-C3N4)0.3, EIS measurements were carried out by varying the potential from 0.15 V to 0.25 V (versus saturated calomel electrode (SCE)) in the dark. Fitting to the circuit equivalent, the Rct and Cdl parameters of the different photocatalysts were obtained, as displayed in Fig. 7c and d, respectively. The electron lifetime (τ) at the depletion layer of the photocatalysts was calculated according to eqn (1), as follows:52

 
τ = RctCdl (1)

Fig. 7e presents the electron lifetime of g-C3N4, BiFeO3, BiFeO3/Ag0.05, and (BiFeO3/Ag0.05)/(g-C3N4)0.3, where a shorter lifetime means faster charge transfer ability given that the amount of electrons in the depletion layer of the photocatalysts is very low. The electron lifetime follows the order of g-C3N4 > BiFeO3 > BiFeO3/Ag0.05 > (BiFeO3/Ag0.05)/(g-C3N4)0.3. The Z-scheme (BiFeO3/Ag0.05)/(g-C3N4)0.3 photocatalytic system has superior photogenerated electron–hole pair separation ability, retaining a greater number of photo-generated carriers for producing radicals, which will enhance the photocatalytic efficiency for the degradation of pollutants.

3.2. Evaluation of photocatalytic activity

3.2.1. Efficiency of photocatalytic degradation of MB. The photocatalytic activity of the (BiFeO3/Ag0.05)/(g-C3N4)0.3 heterojunction was assessed by the photocatalytic degradation of MB and calculated using eqn (2), as follows:53
 
Degradation% = 1 − A/A0 = 1 − C/C0 (2)
where A0 is the initial absorbance of the MB solution, A is the final absorbance of the MB solution after illumination, C0 is the initial concentration of MB solution, and C is the final concentration of MB solution after illumination. Before the photocatalytic degradation of MB, adsorption equilibrium occurs for 30 min in the dark reaction, and the maximum adsorption of MB reached 13% degradation by g-C3N4, while (BiFeO3/Ag0.05)/(g-C3N4)0.3 caused 6% degradation, indicating that g-C3N4 had stronger dark adsorption ability (Fig. 8a). The corresponding photocatalytic kinetic studies proved that the dark reactions followed a pseudo-second-order kinetic reaction (Fig. S2 and S3). The adsorption kinetic studies were evaluated using eqn (3) and (4), as follows:54
 
ln(qeq) = ln[thin space (1/6-em)]qek1t (3)
 
t/q = 1/(k2qe2) + t/qe (4)
where qe and q are the adsorption capacities at equilibrium and at time t and k1 and k2 are rate constants for the pseudo-first-order and pseudo-second-order kinetics, respectively. In the case of g-C3N4, the pseudo-second-order kinetic model (R2 = 0.9948) provides a better fit than the pseudo-first-order kinetic model (R2 = 0.9002). Similarly, (BiFeO3/Ag0.05)/(g-C3N4)0.3 was also significantly better fitted by the pseudo-second-order kinetic model (R2 = 0.9802) than the pseudo-first-order kinetic model (R2 = 0.7908). Therefore, both g-C3N4 and (BiFeO3/Ag0.05)/(g-C3N4)0.3 follow the pseudo-second-order kinetic model during the dark reaction periods. After 40 min of dark adsorption and 150 min illumination, the maximum removal efficiency of MB reached 84.60% by the (BiFeO3/Ag0.05)/(g-C3N4)0.3 heterojunction, while that of g-C3N4, BiFeO3, BiFeO3/Ag0.05, and BiFeO3/(g-C3N4)0.3 was only 49.2%, 31.3%, 60.0% and 66.3%, respectively. The corresponding UV-vis spectra are shown in Fig. S4, and the MB degradation efficiency was calculated at the maximum absorption wavelength of 664 nm. In addition, the optimum mass ratio of BiFeO3/Ag/g-C3N4 in photocatalytic MB degradation was investigated, as shown in Fig. S5. According to Fig. S5a, the BiFeO3/Ag composite achieved the best catalytic effect when the amount of Ag is 5 wt%. Additionally, with a mass fraction of g-C3N4 at 30 wt%, the BiFeO3/Ag/g-C3N4 composite showed the best catalytic effect (Fig. S5b). The photodegradation rate constant was fitted by the first-order reaction kinetics using eqn (5), as follows:55
 
ln(C/C0) = −kt (5)
where k is the photodegradation rate constant. As shown in Fig. 8b, the rate constant of (BiFeO3/Ag0.05)/(g-C3N4)0.3 (k = 0.01163 min−1) is 5.84-times than that of BiFeO3 and 3.26-times than that of g-C3N4, demonstrating its superior photocatalytic activity towards MB. The MB removal efficiency of BiFeO3/Ag/g-C3N4 is 2.7-times than that of pure BiFeO3 and 1.7-times than that of pure g-C3N4, confirming the advantages of the 1D/0D/2D layered architecture and synergistic effect in the BiFeO3/Ag/g-C3N4 Z-scheme heterojunction. (i) The 1D/0D/2D multilayered architecture of BiFeO3/Ag/g-C3N4 significantly enhanced the specific surface area of single materials, which could provide more accessible surface sites for the adsorption and degradation of MB. (ii) The LSPR enhancement effect of Ag offers the advantages of low cost, improved visible-light absorption and photoelectron transfer.56 In addition, Ag acts as an electronic bridge, promoting electron transfer between BiFeO3 and g-C3N4. (iii) The formation of a Z-scheme heterojunction can retain strong redox potentials by generating holes in the VB of BiFeO3 and electrons in the CB of g-C3N4, and simultaneously form ˙O2 and ˙OH radicals to directly react with pollutants. Table S2 presents a comparison with other semiconductor photocatalysts for the degradation of pollutants. The present BiFeO3/Ag/g-C3N4 Z scheme heterojunction displayed an appropriate photo-degradation efficiency, further providing a green and simple method for the construction of visible-light Z-scheme heterojunctions.

image file: d5ra04825g-f8.tif
Fig. 8 (a) Catalytic efficiency (inset: the variation of MB solution color by time); (b) catalytic kinetic plots of g-C3N4, BiFeO3, BiFeO3/Ag0.05, (BiFeO3/Ag0.05)/(g-C3N4)0.3 and BiFeO3/(g-C3N4)0.3 for MB (catalyst = 10 mg, MB = 10 mg L−1, xenon lamp = 300 W).
3.2.2. Reusability analysis. The photochemical stability and reusability of photocatalysts are key indicators for assessing their catalytic efficiency and long-term performance in practical applications. In this study, the reusability of the (BiFeO3/Ag0.05)/(g-C3N4)0.3 heterojunction was evaluated, as shown in Fig. S6. The degradation efficiency of (BiFeO3/Ag0.05)/(g-C3N4)0.3 remained 70.3% after four photocatalytic degradation cycling experiments, indicating its satisfactory photocatalytic stability and reusability for the degradation of MB in water.
3.2.3. Photocatalytic mechanism. The bandgap of different photocatalysts can be calculated using the UV-vis DRS and Mott–Schottky methods. Fig. 9a presents the plots of (αhν)2 versus hν for g-C3N4, BiFeO3 and BiFeO3/Ag. The bandgap energies of g-C3N4, BiFeO3 and BiFeO3/Ag are determined to be 2.79 eV, 2.12 eV, and 1.99 eV, respectively, which are consistent with the literature values.57 As shown in Fig. 9b and c, the flat-band potentials (Vfb) of g-C3N4, BiFeO3 and BiFeO3/Ag are −1.51 V, 0.37 V and −0.30 V versus SCE, respectively. These flat band potential values are modulated to the normal hydrogen electrode (NHE) values using eqn (6), as follows:58
 
Vfb(NHE, pH = 7) = Vfb(SCE) + 0.2415 − 0.059 (7 – pH of the electrolyte) (6)

image file: d5ra04825g-f9.tif
Fig. 9 (a) Plots of (αhν)2 versus hν for g-C3N4, BiFeO3 and BiFeO3/Ag. (b and c) Mott–Schottky plots of pure g-C3N4 and BiFeO3 and BiFeO3/Ag, respectively. (The Mott–Schottky effect was measured in a 0.5 M Na2SO4 solution at 1.0 kHz.) (d) Schematic of the band structures of g-C3N4, BiFeO3 and BiFeO3/Ag.

It has been reported that the conduction band position (ECB) of n-type semiconductors is very close to Vfb; therefore, the ECB values of g-C3N4, BiFeO3 and BiFeO3/Ag are −1.27 eV, 0.61 eV and −0.06 eV, respectively.59 The valence-band edge (EVB) was subsequently calculated using eqn (7), as follows:

 
Eg = EVBECB (7)

The band structure diagram of the materials is shown in Fig. 9d. The CB and VB potentials of BiFeO3/Ag are more negative than that of BiFeO3, and thus the accumulated photoelectrons on the CB of BiFeO3 easily flow to the Ag nanoparticles. Owing to the lower Fermi level of the Ag nanoparticles, Ag acted as an electron reservoir, surmounting the Schottky barriers at the BiFeO3/Ag interface. The LSPR effect of the Ag nanoparticles induced by the oscillation of their surface electrons could further enhance the visible light harvesting of photocatalysts and promote electron–hole separation.60,61

As shown in Fig. 10a, in the photocatalytic systems, p-benzoquinone (p-BQ), isopropanol (IPA), and methanol (MeOH) were added to the MB solution to capture the superoxide (˙O2), ˙OH, and h+, respectively.62–64 After adding IPA and p-BQ, separately, the degradation efficiency of MB was severely suppressed, indicating that ˙OH and ˙O2 play a major role in the degradation process compared to h+, respectively. To ensure effective quenching of target free radicals, we investigated the effect of different scavenger concentrations on the photocatalytic degradation of MB (Fig. S7). As the scavenger concentration increased from 10 mM to 30 mM, the free radical scavenging efficiency remained nearly constant, and thus we selected a scavenger concentration of 30 mM for comparison.


image file: d5ra04825g-f10.tif
Fig. 10 (a) Effect of scavengers on the photocatalytic removal of MB by (BiFeO3/Ag0.05)/(g-C3N4)0.3 (scavengers' content: 30 mM p-BQ, 30 mM IPA, and 30 mM MeOH). EPR spectra of (BiFeO3/Ag0.05)/(g-C3N4)0.3 for (b) DMPO–˙OH and (c) DMPO–˙O2.

Electron paramagnetic resonance (EPR) experiments were conducted for (BiFeO3/Ag0.05)/(g-C3N4)0.3 to further verify the production of ˙OH and ˙O2. As shown in Fig. 10b and c, no EPR signals were detected in the dark. The EPR signals of ˙OH and ˙O2 can be clearly seen under light irradiation, which is consistent with the results of the active species scavenging experiments. This suggests that (BiFeO3/Ag0.05)/(g-C3N4)0.3 can produce ˙OH and ˙O2 radicals, confirming the hypothesis of constructing a Z-scheme photocatalytic system.

Furthermore, the main intermediates yielded during the photocatalytic degradation of MB and the possible degradation pathways were investigated using liquid chromatography-mass spectrometry (LC-MS). According to the m/z ratios obtained in positive mode, a total of 15 dominant structures was identified under 0, 30, 90, 150 min light irradiation, as shown in Fig. S8, and the possible three degradation pathways under ˙OH and ˙O2 induction are shown in Fig. 11. The mass spectrum of MB shows a prominent molecular ion peak at m/z = 284, which is consistent with a previous report.65 The ˙OH and ˙O2 radicals may further cause the MB degradation pathways to proceed through demethylation, oxidative, and ring opening to the final fragmentation, and eventually mineralization was achieved.66,67 In pathway I, the original MB is demethylated to form intermediate P2 (m/z = 270), followed by the loss of another methyl group to form P3 (m/z = 256), and finally converted to the product P4 (m/z = 198) through demethylation and deamination. In pathway II, the MB molecule is converted to two oxidative products, P5 (m/z = 317) and P8 (m/z = 273), by the formation of a sulfoxide group and demethylation process. According to the easily broken bond of N–CH3, and the C–N or C–S bond on the central heterocycle of MB, the P5 and P8 intermediates can be degraded into small forms by desulfurization, demethylation, hydroxylation reactions.68 In pathway III, the MB molecule is demethylated to form P12 (m/z = 228), subsequently converting to P13 (m/z = 280) by the oxidation of the C–S bond on the central heterocycle to sulfoxide, which is further converted to P14 (m/z = 173) and P15 (m/z = 141) through the cleavage of the central ring. Eventually, the short open ring structures were completely mineralized into CO2, H2O, SO42−, NO3 and some inorganic compounds.69


image file: d5ra04825g-f11.tif
Fig. 11 Possible photocatalytic degradation pathways of MB.

Considering the radical scavenging experiments and the bandgap configuration of the prepared g-C3N4 and BiFeO3, a possible photocatalytic mechanism is proposed, as shown in Fig. 12. g-C3N4 is unable to produce ˙OH (its EVB is more negative than 2.38 eV), while BiFeO3 is unable to produce ˙O2 (its ECB is more positive than −0.33 eV). The Z-scheme heterojunction of BiFeO3/Ag/g-C3N4 is formed to achieve the photocatalytic degradation of MB. Given that the ECB of g-C3N4 (−1.27 eV) is more negative than that of O2/˙O2 (−0.33 eV vs. NHE), it can capture adsorbed oxygen to form ˙O2 radicals. The EVB of BiFeO3 (−2.73 eV) is more positive than the reduction potential of H2O/˙OH (2.38 eV vs. NHE), and ˙OH radicals are created.70 The photogenerated electrons from the CB of BiFeO3 can rapidly transfer to g-C3N4 through plasmonic Ag, and combine with the holes from the VB of g-C3N4. Consequently, the electrons from the CB of g-C3N4 and the holes from the VB of BiFeO3 are effectively separated. The MB in solution is oxidized and degraded in the presence of ˙OH and ˙O2 to CO2, H2O, mineral acids.71 The possible reactions are as follows:

 
image file: d5ra04825g-t1.tif(8)
 
H2O + h+ → ˙OH + H+ (9)
 
O2 + e → ˙O2 (10)
 
˙OH + MB → degraded product (11)
 
˙O2 + MB → degraded product (12)


image file: d5ra04825g-f12.tif
Fig. 12 Possible photocatalytic mechanism of the Z-scheme (BiFeO3/Ag0.05)/(g-C3N4)0.3 system.

4. Conclusion

In summary, a Z-scheme heterojunction of 1D/0D/2D BiFeO3/Ag/g-C3N4 was constructed to enhance the photocatalytic activity in degradation of organic dyes under visible light irradiation. The structural parameters, optical properties, photocatalytic dye degradation, and dielectric properties of the materials were assessed by XRD, XPS, BET, SEM, EDX, UV-vis DRS, PL, electrochemical and photocatalytic degradation testing. The EPR and free radical scavenging experiments confirmed the active species (˙OH and ˙O2) for the decomposition of MB, demonstrating the possible photocatalytic mechanism. This study provides a feasible way for improving the visible light degradation of dyes in wastewater, and has significant reference for the design and synthesis of novel Z-scheme heterojunction photocatalysts with multidimensional contact interfaces.

Author contributions

Donghai Li: investigation, writing – original draft, data curation and validation; Yunrui Xu: visualization, investigation; Shilin Zhang: visualization, investigation; Linping Wang: writing – review & editing, writing – original draft, supervision, investigation, funding acquisition.

Conflicts of interest

We declare that there is no conflict of interest to this work.

Data availability

The data supporting this article have been included as part of the SI. No additional datasets were generated or analyzed during this study.

Supplementary information: The authors declare that the data supporting the findings of this study are available within the paper and its SI. Isotherm linear plots, dark reaction kinetic curves, effects of Ag and g-C3N4 mass fractions on efficiency, LC-MS spectra, photocatalytic stability and comparison table of photocatalytic degradation. See DOI: https://doi.org/10.1039/d5ra04825g.

Acknowledgements

This work was funded by the Natural Science Foundation of Qinghai Province (no. 2021-ZJ-964Q) and the Youth Research Fund Project of Qinghai University (no. 2020-QGY-8).

References

  1. F. Deng, J. Peng, X. Li, X. Luo, P. Ganguly, S. C. Pillai, B. Ren, L. Ding and D. D. Dionysiou, Metal sulfide-based Z-scheme heterojunctions in photocatalytic removal of contaminants, H2 evolution and CO2 reduction: Current status and future perspectives, J. Cleaner Prod., 2023, 416, 137957–137976 CrossRef.
  2. K. Parida, N. Baliarsingh, B. S. Patra and J. Das, Copperphthalocyanine immobilized Zn/Al LDH as photocatalyst under solar radiation for decolorization of methylene blue, J. Mol. Catal. A: Chem., 2007, 267, 202–208 CrossRef.
  3. A. C. Pradhan and K. Parida, Facile synthesis of mesoporous composite Fe/Al2O3–MCM-41: an efficient adsorbent/catalyst for swift removal of methylene blue and mixed dyes, J. Mater. Chem., 2012, 22, 7567–7579 RSC.
  4. P. Mohapatra and K. Parida, Photocatalytic activity of sulfate modified titania 3: Decolorization of methylene blue in aqueous solution, J. Mol. Catal. A: Chem., 2006, 258, 118–123 CrossRef.
  5. S. Nayak, K. K. Das and K. Parida, Indulgent of the physiochemical features of MgCr-LDH nanosheets towards photodegradation process of methylene blue, J. Colloid Interface Sci., 2023, 634, 121–137 CrossRef PubMed.
  6. S. Y. Kim, I. Y. Kim, S.-H. Park, M. Hwangbo and S. Hwangbo, Novel ultrasonic technology for advanced oxidation processes of water treatment, RSC Adv., 2024, 14, 11939–11948 RSC.
  7. S. M. H. Al-Jawad, K. H. Aboud, N. J. Imran and S. Y. Taher, Copper doping of CdS nanoflakes and nanoflowers for efficient photocatalytic degradation of MB and MV dyes, Plasmonics, 2025, 20, 983–1002 CrossRef.
  8. C. Zhao, X. Li, L. Yue, S. Yuan, X. Ren, Z. Zeng, X. Hu, Y. Wu and Y. He, One-step preparation of novel Bi-Bi2O3/CdWO4 Z-scheme heterojunctions with enhanced performance in photocatalytic NH3 synthesis, J. Alloys Compd., 2023, 968, 171956–171969 CrossRef.
  9. M. A. Gatou, A. Syrrakou, N. Lagopati and E. A. Pavlatou, Photocatalytic TiO2-based nanostructures as a promising material for diverse environmental applications: a review, Reactions, 2024, 5, 135–194 CrossRef.
  10. M. Alhaddad and M. S. Amin, Removal of ciprofloxacin applying Pt@BiVO4-g-C3N4 nanocomposite under visible light, Opt. Mater., 2022, 124, 111976–111986 CrossRef.
  11. B. Malathi, Y. Mori, S. Harish and A. Nakamura, Investigation of 2D-layered photocatalytic semiconductors with enhanced heterojunction region for photocatalytic degradation, J. Mater. Sci.: Mater. Electron., 2024, 35, 1236–1250 CrossRef.
  12. X. Li, Y. Huang, W. Ho, S. Han, P. Wang, S. Lee and Z. Zhang, Modulation of sulfur vacancies at ZnIn2S4-δ/g-C3N4 heterojunction interface for successive C-H secession in photocatalytic gaseous formaldehyde complete oxidation, Appl. Catal., B, 2023, 338, 123048–123060 CrossRef.
  13. M. Atighi, G. Falahati, M. Hasanzadeh, N. Gholami, H. Ahmadi and M. Najafi, Fabrication of a graphene oxide/zeolitic imidazolate framework-8 (GO/ZIF-8) composite for enhanced adsorptive and piezo-assisted photocatalytic removal of organic dyes, New J. Chem., 2025, 49, 10751–10765 RSC.
  14. Y. Xu, Y. Liang, Q. He, R. Xu, D. Chen, X. Xu and H. Hu, Review of doping SrTiO3 for photocatalytic applications, Bull. Mater. Sci., 2022, 46, 6 CrossRef.
  15. N. A. Neto, A. Lima, R. Wilson, T. Nicacio, M. Bomio and F. Motta, Heterostructures obtained by ultrasonic methods for photocatalytic application: A review, Mater. Sci. Semicond. Process., 2022, 139, 106311–106321 CrossRef.
  16. C. Nie, X. Wang, P. Lu, Y. Zhu, X. Li and H. Tang, Advancements in S-scheme heterojunction materials for photocatalytic environmental remediation, J. Mater. Sci. Technol., 2024, 169, 182–198 CrossRef.
  17. L. Zheng, Y. Wei, C. Wang, H. Liu, L. Li, M. Huang, Y. Huang, L. Fan and J. Wu, Construction of direct WO3/g-C3N4 Z-scheme heterojunction for degrading flotation agent effectively, Ceram. Int., 2024, 50, 38860–38870 CrossRef.
  18. S. D. Lakshmi and I. B. S. Banu, Multiferroism and magnetoelectric coupling in single-phase Yb and X (X = Nb, Mn, Mo) co-doped BiFeO3 ceramics, J. Sol-Gel Sci. Technol., 2019, 89, 713–721 CrossRef.
  19. S. Mansingh, S. Sultana, R. Acharya, M. K. Ghosh and K. M. Parida, Efficient photon conversion via double charge dynamics CeO2–BiFeO3 p–n heterojunction photocatalyst promising toward N2 fixation and phenol–Cr(VI) detoxification, Inorg. Chem., 2020, 59, 3856–3873 CrossRef.
  20. C. Li, Z. Li, S. Guo, X. Li, Q. Cheng and S. Meng, Sensitivity enhancement by employing BiFeO3 and graphene hybrid structure in surface plasmon resonance biosensors, Opt. Mater., 2021, 121, 111618–111627 CrossRef.
  21. Y. Nassereddine, M. Benyoussef, B. Asbani, M. El Marssi and M. Jouiad, Recent advances toward enhanced photocatalytic proprieties of BiFeO3-based materials, Nanomaterials, 2024, 14, 51–81 CrossRef PubMed.
  22. I. Shakir, Synthesis of Gd doped BiFeO3/g-C3N4 composite: Enhancement of solar mediated photocatalytic performance, Mater. Sci. Eng., B, 2024, 302, 117252–117263 CrossRef.
  23. M. P. McOyi, K. T. Mpofu, M. Sekhwama and P. Mthunzi-Kufa, Developments in localized surface plasmon resonance, Plasmonics, 2024, 20, 5481–5520 CrossRef.
  24. S. Ni, Z. Fu, L. Li, M. Ma and Y. Liu, Step-scheme heterojunction g-C3N4/TiO2 for efficient photocatalytic degradation of tetracycline hydrochloride under UV light, Colloids Surf., A, 2022, 649, 129475–129487 CrossRef.
  25. J. Tang, R. Wang, M. Liu, Z. Zhang, Y. Song, S. Xue, Z. Zhao and D. D. Dionysiou, Construction of novel Z-scheme Ag/FeTiO3/Ag/BiFeO3 photocatalyst with enhanced visible-light-driven photocatalytic performance for degradation of norfloxacin, Chem. Eng. J., 2018, 351, 1056–1066 CrossRef.
  26. Z. H. Jabbar, B. H. Graimed, M. A. Issa, S. H. Ammar, S. E. Ebrahim, H. J. Khadim and A. A. Okab, Photocatalytic degradation of Congo red dye using magnetic silica-coated Ag2WO4/Ag2S as Type I heterojunction photocatalyst: Stability and mechanisms studies, Mater. Sci. Semicond. Process., 2023, 153, 107151–107166 CrossRef.
  27. V. Kashyap, H. Pawar, I. Sihmar, C. Kumar, A. Kumar, S. Kumar, N. Chaudhary, N. Goyal and K. Saxena, X-ray analysis of Ag nanoparticles on Si wafer and influence of Ag nanoparticles on Si nanowire-based gas sensor, Appl. Phys. A, 2024, 130, 238–251 CrossRef.
  28. Y. Guo, L. Zhu, J. Wang and J. Li, Comprehensive computational investigation of structural and mechanical characteristics of crystalline C3N4 under high pressure based on experimental data, Phys. B, 2024, 674, 415604–415611 CrossRef CAS.
  29. D. A. Rusakov, A. M. Abakumov, K. Yamaura, A. A. Belik, G. Van Tendeloo and E. Takayama-Muromachi, Structural evolution of the BiFeO3–LaFeO3 system, Chem. Mater., 2011, 23, 285–292 CrossRef CAS.
  30. M. Ghosh, S. Mandal, A. Roy, S. Chakrabarty, G. Chakrabarti and S. K. Pradhan, Enhanced antifungal activity of fluconazole conjugated with Cu-Ag-ZnO nanocomposite, Mater. Sci. Eng., C, 2020, 106, 110160–110170 CrossRef CAS PubMed.
  31. Q. Xiang, J. Yu and M. Jaroniec, Preparation and enhanced visible-light photocatalytic H2-production activity of graphene/C3N4 composites, J. Phys. Chem. C, 2011, 115, 7355–7363 CrossRef CAS.
  32. F. Dong, Z. Wang, Y. Sun, W.-K. Ho and H. Zhang, Engineering the nanoarchitecture and texture of polymeric carbon nitride semiconductor for enhanced visible light photocatalytic activity, J. Colloid Interface Sci., 2013, 401, 70–79 CrossRef CAS.
  33. J. Liu, T. Zhang, Z. Wang, G. Dawson and W. Chen, Simple pyrolysis of urea into graphitic carbon nitride with recyclable adsorption and photocatalytic activity, J. Mater. Chem., 2011, 21, 14398–14401 RSC.
  34. H. Yan and H. Yang, TiO2–g-C3N4 composite materials for photocatalytic H2 evolution under visible light irradiation, J. Alloys Compd., 2011, 509, L26–L29 CrossRef CAS.
  35. S. J. J. Kay, N. Chidhambaram, A. Thirumurugan, S. Shanavas, P. Sakthivel and R. S. Rimal Isaac, Stoichiometric balancing of bismuth ferrite-perovskite nanoparticles: comparative investigations on biogenic versus conventional chemical synthesis, J. Mater. Sci.: Mater. Electron., 2023, 34, 2034–2048 CrossRef CAS.
  36. M. B. Islam, M. J. Haque, N. M. Shehab and M. S. Rahman, Synthesis and characterization (optical and antibacterial) of silver doped zinc oxide nanoparticles, Open Ceram., 2023, 14, 100370–100376 CrossRef CAS.
  37. Z. Du, Y. Li, D. Kuang, W. Wang, F. Yang, J. Yang and L. Hou, Insight into the synergistic effect of adsorption and photocatalysis for the removal of organic dye pollutants by novel BiFeO3@GO fibers, J. Mater. Sci.: Mater. Electron., 2023, 34, 589–607 CrossRef CAS.
  38. D. A. Upar, D. Gogoi, M. R. Das, B. Naik and N. N. Ghosh, Facile synthesis of g-C3N4-exfoliated BiFeO3 nanocomposite: a versatile and efficient S-scheme photocatalyst for the degradation of various textile dyes and antibiotics in water, ACS Omega, 2023, 8, 38524–38538 CrossRef CAS.
  39. S. Sharma and M. Kumar, Structural and magnetic properties and graphene-induced photocatalytic activity of BiFeO3 ceramics, J. Supercond. Novel Magn., 2023, 36, 1193–1202 CrossRef CAS.
  40. K. Saravanakumar and C. M. Park, Rational design of a novel LaFeO3/g-C3N4/BiFeO3 double Z-scheme structure: Photocatalytic performance for antibiotic degradation and mechanistic insight, Chem. Eng. J., 2021, 423, 130076–130088 CrossRef CAS.
  41. Q. Huang, J. Yu, S. Cao, C. Cui and B. Cheng, Efficient photocatalytic reduction of CO2 by amine-functionalized g-C3N4, Appl. Surf. Sci., 2015, 358, 350–355 CrossRef CAS.
  42. J. Zhang, Y. Zheng, H. Zheng, T. Jing, Y. Zhao and J. Tian, Porous oxygen-doped g-C3N4 with the different precursors for excellent photocatalytic activities under visible light, Materials, 2022, 15, 1391–1405 CrossRef CAS.
  43. D. Gao, Q. Xu, J. Zhang, Z. Yang, M. Si, Z. Yan and D. Xue, Defect-related ferromagnetism in ultrathin metal-free g-C3N4 nanosheets, Nanoscale, 2014, 6, 2577–2581 RSC.
  44. A. Joseph, R. Jayakrishnan, A. M. Anand and V. Thomas, Ag-doped BiOBr nanoparticles: a grander photocatalyst, J. Mater. Sci.: Mater. Electron., 2025, 36, 637–657 CrossRef CAS.
  45. P. Duan, Q. Xu, S. Shen, Y. Zhang, L. Zhang, F. Fu and X. Liu, One-pot modification on cotton fabric using an emulsion of Ag NPs protected by mercaptosuccinic acid to achieve durably antibacterial effect, Fibers Polym., 2019, 20, 1803–1811 CrossRef CAS.
  46. S. Nayak and K. Parida, Dynamics of charge-transfer behavior in a plasmon-induced quasi-type-II p–n/n–n dual heterojunction in Ag@Ag3PO4/g-C3N4/NiFe LDH nanocomposites for photocatalytic Cr(VI) reduction and phenol oxidation, ACS Omega, 2018, 3, 7324–7343 CrossRef CAS.
  47. J. Xu, T. Qin, W. Chen, J. Lv, X. Zeng, J. Sun, Y.-y. Li and J. Zhou, Synergizing piezoelectric and plasmonic modulation of Ag/BiFeO3 fibrous heterostructure toward boosted photoelectrochemical energy conversion, Nano Energy, 2021, 89, 106317–106326 CrossRef CAS.
  48. W. Su, S. S. Wei, S. Q. Hu and J. X. Tang, Preparation of TiO2/Ag colloids with ultraviolet resistance and antibacterial property using short chain polyethylene glycol, J. Hazard. Mater., 2009, 172, 716–720 CrossRef CAS.
  49. K.-C. Lee, S.-J. Lin, C.-H. Lin, C.-S. Tsai and Y.-J. Lu, Size effect of Ag nanoparticles on surface plasmon resonance, Surf. Coat. Technol., 2008, 202, 5339–5342 CrossRef CAS.
  50. L. Wang, Y. Meng, Y. Zhang, C. Zhang, Q. Xie and S. Yao, Photoelectrochemical aptasensing of thrombin based on multilayered gold nanoparticle/graphene-TiO2 and enzyme functionalized graphene oxide nanocomposites, Electrochim. Acta, 2017, 249, 243–252 CrossRef CAS.
  51. J. Ângelo, P. Magalhães, L. Andrade and A. Mendes, Characterization of TiO2-based semiconductors for photocatalysis by electrochemical impedance spectroscopy, Appl. Surf. Sci., 2016, 387, 183–189 CrossRef.
  52. L. Andrade, S. M. Zakeeruddin, M. K. Nazeeruddin, H. A. Ribeiro, A. Mendes and M. Grätzel, Influence of sodium cations of N3 dye on the photovoltaic performance and stability of dye-sensitized solar cells, ChemPhysChem, 2009, 10, 1117–1124 CrossRef CAS PubMed.
  53. Y. Zhang, Y. Wang, J. Li, J. Xie, W. Wang and Z. Fu, Enhanced photocatalytic activity of Z-scheme meso-BiVO4-Au-CdS for degradation of Rhodamine B, J. Wuhan Univ. Technol., Mater. Sci. Ed., 2024, 39, 869–876 CrossRef CAS.
  54. A. K. Sasmal, J. Pal, R. Sahoo, P. Kartikeya, S. Dutta and T. Pal, Superb dye adsorption and dye-sensitized change in Cu2O–Ag crystal faces in the dark, J. Phys. Chem. C, 2016, 120, 21580–21588 CrossRef CAS.
  55. Z. Long, X. Zheng and H. Shi, Construction of BiVO4/CoPc S-scheme heterojunctions with enhanced photothermal-assisted photocatalytic activity, Sci. China Mater., 2024, 67, 550–561 CrossRef CAS.
  56. S. Patnaik, G. Swain and K. M. Parida, Highly efficient charge transfer through a double Z-scheme mechanism by a Cu-promoted MoO3/g-C3N4 hybrid nanocomposite with superior electrochemical and photocatalytic performance, Nanoscale, 2018, 10, 5950–5964 RSC.
  57. X. Zhou, L. Peng, L. Xu, J. Luo, X. Ning, X. Zhou, F. Peng and X. Zhou, Pd(II), Pt(II) metallosupramolecular complexes as Single-Site Co-Catalyst for photocatalytic H2 evolution, Chem. Eng. J., 2023, 474, 145967–145978 CrossRef CAS.
  58. A. Balakrishnan, E. S. Kunnel, R. Sasidharan, M. Chinthala and A. Kumar, Tailored citric acid-functionalized carbon nitride homojunction-immobilized carboxymethyl cellulose 3D photocatalytic hydrogels: A multifaceted approach toward environmental remediation, ACS Sustainable Chem. Eng., 2024, 12, 5169–5185 CrossRef CAS.
  59. L. Su, N. Qin, W. Xie, J. Fu and D. Bao, The surface-plasmon-resonance and band bending effects on the photoluminescence enhancement of Ag-decorated ZnO nanorods, J. Appl. Phys., 2014, 116, 063108–063116 CrossRef.
  60. L. Panda, A. Pradhan, E. Subudhi, R. K. Sahoo and B. Nanda, Ag-loaded BiFeO3/CuS heterostructured based composite: an efficient photocatalyst for removal of antibiotics and antibacterial activities, Environ. Sci. Pollut. Res., 2024, 31, 5540–5554 CrossRef CAS PubMed.
  61. Y. K. Lee, C. H. Jung, J. Park, H. Seo, G. A. Somorjai and J. Y. Park, Surface plasmon-driven hot electron flow probed with metal-semiconductor nanodiodes, Nano Lett., 2011, 11, 4251–4255 CrossRef PubMed.
  62. H. Xing, J. Shi, Y. Li and J. Wu, Visible light driven generation of dual active oxygen species on Zr-MOF/g-C3N4 photocatalyst for highly selective photocatalytic oxidation of sulfides to sulfoxides, Inorg. Chem. Commun., 2024, 162, 112129–112136 CrossRef.
  63. K. Peng, Z. Fan, Y. Wang, Y. Xie and Y. Ling, Construction of TiO2/SnIn4S8 heterojunction to promotion photogenerated carrier separation and enhance photocatalytic hydrogen production via superoxide radical active species, Int. J. Hydrogen Energy, 2024, 60, 835–844 CrossRef.
  64. S. Asgari, G. Mohammadi Ziarani, A. Badiei and Y. Vasseghian, Zr-UiO-66, ionic liquid (HMIM+TFSI), and electrospun nanofibers (polyacrylonitrile): All in one as a piezo-photocatalyst for degradation of organic dye, Chem. Eng. J., 2024, 487, 150600–150611 CrossRef.
  65. Y. Zhong, X. Wan, X. Lian, W. Cheng, X. Ma and D. Wang, Hydroxylamine facilitated catalytic degradation of methylene blue in a Fenton-like system for heat-treatment modified drinking water treatment residues, Environ. Sci. Pollut. Res., 2023, 30, 79282–79296 CrossRef PubMed.
  66. M. V. Nikolic, Z. Z. Vasiljevic, M. Dimitrijevic, N. Radmilovic, J. Vujancevic, M. Tanovic and N. B. Tadic, Natural sunlight driven photocatalytic degradation of methylene blue and Rhodamine B over nanocrystalline Zn2SnO4/SnO2, Nanomaterials, 2025, 15, 1138–1163 CrossRef PubMed.
  67. W. Ahmad, N. Ahmad, S. Rasheed, M. I. Nabeel, A. Mohyuddin, M. T. Riaz and D. Hussain, Silica-based superhydrophobic and superoleophilic cotton fabric with enhanced self-cleaning properties for oil–water separation and methylene blue degradation, Langmuir, 2024, 40, 5639–5650 CrossRef.
  68. Y. Cai, C. Qiu, K. Yang, B. Tian and Y. Bi, Adsorption–degradation of methylene blue by natural manganese ore: kinetics, characterization, and mechanism, Int. J. Environ. Sci. Technol., 2024, 21, 1817–1830 CrossRef.
  69. A. Garg, A. Chauhan, C. Agnihotri, B. P. Singh, V. Mondem, S. Basu and S. Agnihotri, Sunlight active cellulose/g-C3N4/TiO2 nano-photocatalyst for simultaneous degradation of methylene blue dye and atenolol drug in real wastewater, Nanotechnology, 2023, 34, 505705–505719 CrossRef PubMed.
  70. Q. Xia, X. Liu, H. Li, Y. Guan, J. Chen, Y. Chen, Z. Hu and W. Gao, Construction of the Z-scheme Cu2O-Ag/AgBr heterostructures to enhance the visible-light-driven photocatalytic water disinfection and antibacterial performance, J. Alloys Compd., 2024, 980, 173665–173676 CrossRef.
  71. K. Parvathalu, K. Rajitha, B. Chandrashekar, K. Sathvik, K. Pranay Bhasker, B. Sreenivas, M. Pritam, P. Pushpalatha, K. Moses and P. Bala Bhaskar, Biomimetic synthesis of copper nanoparticles using Tinospora cordifolia plant Leaf extract for photocatalytic activity applications, Plasmonics, 2024, 19, 825–834 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.