Next Article in Journal
Enhanced Electrocatalytic Properties of Co3O4 Nanocrystals Derived from Hydrolyzed Polyethyleneimines in Water/Ethanol Solvents for Electrochemical Detection of Cholesterol
Next Article in Special Issue
Improved the Light Adsorption and Separation of Charge Carriers to Boost Photocatalytic Conversion of CO2 by Using Silver Doped ZnO Photocatalyst
Previous Article in Journal
A Mechanistic Study of Methanol Steam Reforming on Ni2P Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of a Well-Defined S-Scheme Heterojunction Based on Bi-ZnFe2O4/S-g-C3N4 Nanocomposite Photocatalyst to Support Photocatalytic Pollutant Degradation Driven by Sunlight

1
School of Chemistry and Materials Engineering, Huizhou University, Huizhou 516007, China
2
Guangdong Provincial Key Lab of Green Chemical Product Technology, Guangzhou 510640, China
3
Department of Chemistry, School of Science, University of Management and Technology, Lahore 54770, Pakistan
4
Guangdong Engineering & Technology Research Centre of Graphene-like Materials and Products, Department of Chemistry, College of Chemistry and Materials Science, Jinan University, Guangzhou 510632, China
5
Pharmaceutics and Pharmaceutical Technology Department, College of Pharmacy, Taibah University, Medina 42353, Saudi Arabia
6
Department of Chemistry, Faculty of Applied Science, Umm Al-Qura University, Mecca 24230, Saudi Arabia
7
Department of Chemistry, College of Science, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
8
Department of Biotechnology, College of Science, Taif University, P.O. Box 11099, Taif 21944, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Submission received: 16 August 2022 / Revised: 11 September 2022 / Accepted: 30 September 2022 / Published: 5 October 2022

Abstract

:
Currently, organic dyes and other environmental contaminants are focal areas of research, with considerable interest in the production of stable, high-efficiency, and eco-friendly photocatalysts to eliminate these contaminants. In the present work, bismuth-doped zinc ferrite (Bi-ZnFe2O4) nanoparticles (NPs) and bismuth-doped zinc ferrites supported on sulfur-doped graphitic carbon nitride (Bi-ZnFe2O4/S-g-C3N4) (BZFG) photocatalysts were synthesized via a hydrothermal process. SEM, XRD, and FTIR techniques were used to examine the morphological, structural, and bonding characteristics of the synthesized photocatalysts. The photocatalytic competence of the functional BZFG nanocomposites (NCs) was studied against MB under sunlight. The influence of Bi (0.5, 1, 3, 5, 7, 9, and 11 wt.%) doping on the photocatalytic performance of ZnFe2O4 was verified, and the 9%Bi-ZnFe2O4 nanoparticles exhibited the maximum MB degradation. Then, 9%Bi-ZnFe2O4 NPs were homogenized with varying amounts of S-g-C3N4 (10, 30, 50, 60, and 70 wt.%) to further enhance the photocatalytic performance of BZFG NCs. The fabricated Bi-ZnFe2O4/30%S-g-C3N4 (BZFG-30) composite outperformed ZnFe2O4, S-g-C3N4 and other BZFG NCs in terms of photocatalytic performance. The enriched photocatalytic performance of the BZFG NCs might be ascribed to a more efficient transfer and separation of photo-induced charges due to synergic effects at the Bi-ZnFe2O4/S-g-C3N4 interconnection. The proposed modification of ZnFe2O4 using Bi and S-g-C3N4 is effective, inexpensive, and environmentally safe.

1. Introduction

Water is an abiotic factor but is very important for all biotic factors, and living organisms cannot survive without it. Toxic dyes are produced and discharged directly into water bodies from homes and industries, such as the textile, dyeing, and printing industries. Therefore, water becomes contaminated and contains bacteria, viruses, and heavy metals, which cause various diseases in human beings, animals, and plants [1]. With increasing water demands and dwindling supplies, these harmful compounds must be isolated from wastewater and kept out of the aquatic environment. Nanotechnology and nanoparticles can be used to eliminate organic chemicals, inorganic pollutants, and microorganisms from wastewater via the photocatalytic process [2]. Photocatalytic removal of organic pollutants with semiconductor materials, such as ZnO, TiO2, CdS, and ZnS, for decontamination waste products has received considerable attention during the last two decades [3,4]. However, photocatalysts (such as ZnO, TiO2, CdS, and ZnS) exhibit limitations during photocatalysis under visible radiations due to their high band gap-energies and fast electron–hole recombination [5,6].
Consequently, the development of novel visible-light-induced photocatalysts with increased activity has been a hot research topic for a long time. Among photocatalysts, ferrite (Fe2O4) nanoparticles are of unique significance because of their excellent adsorption properties for cations and anions, low band-gap energy, and low electron–hole recombination [7,8,9]. These ferrite NPs have potential applications in wastewater treatment; in the biomedical field for diagnosis, drug delivery, drug release, etc.; in electronic devices, such as sensors and biosensors; energy storage; electron magnetic resonance shielding; and recording media. They are highly stable, with a band gap of approximately 1.9 eV; relatively non-toxic; cheap; highly electronically conductive; eco-friendly; and recyclable [10]. Unfortunately, their low quantum efficiency limits their photocatalytic activity for practical use. Therefore, modification has been attempted to enhance the spectral response of these UV-active oxides into the visible range by doping ZnFe2O4 with Bi and developing its heterojunction with S-g-C3N4. Many studies have shown that doping ZnFe2O4 with appropriate metal ions and its composite formation with another suitable semiconductor material improve optical and photocatalytic characteristics.
Savunthari et al. reported the synthesis of (Cu, Bi) codoped ZnFe2O4 nanoparticles via a solution combustion route. The synthesized codoped NPs exhibited excellent bisphenol A degradation compared to undoped NPs [11]. Patil et al. used the coprecipitation approach to manufacture Gd3+-doped ZnFe2O4 nanoparticles, demonstrating improved MB degradation of roughly 99% compared to pure ZnFe2O4 (95% degradation in 240 min) [12]. According to Ajithkumar et al., yttrium-doped zinc ferrite made by solution combustion showed 95% MB degradation in 180 min. Y-ZnFe2O4 has higher photocatalytic effectiveness than pure zinc ferrite. Under visible light, cobalt-doped zinc ferrite decomposed methylene blue more efficiently than ZnFe2O4. Many researchers have concluded that ZnFe2O4 has finite band-gap energy and might form an effective heterojunction when combined with another semiconductor or metal [13,14,15]. The auto-combustion approach was used to synthesize visible-light-driven photocatalyst BiFeO3 (BFO) nanoparticles. The 2%BFO demonstrated superior antibacterial activity (against S. aureus and E. coli.) and photocatalytic efficiency against MB (81%) compared to undoped and other BFO samples [16]. Similarly, Amini et al. used BFO to remove chlorobenzene (CB). The produced BFO outperformed TiO2 nanomaterial and achieved a maximum of 87.92% removal of CB under UV light [17].
The g-C3N4 semiconductor has shown significant photocatalytic proficiency under visible light due to its favorable characteristics, such as high stability and reduced band-gap energy, improving its capacity to absorb visible radiation [18,19]. However, the quick recombination of photoinduced e/h+ pairs in g-C3N4 makes it unsuitable for use as a photocatalyst [18]. S-doping alters the bandgap of g-C3N4 and enhances the mobility and separation of e-h pairs. Hong et al. reported that the photocatalytic H2 production efficiency of mesoporous S-g-C3N4 is 30 times that of pure g-C3N4. Similarly, S-g-C3N4 was reported to exhibit an approximately 1.38 times greater photocatalytic CO2 reduction rate than that of pure g-C3N4 [20]. Under visible light, porous S-g-C3N4 had exhibited better photocatalytic degradation and adsorption of Rhodamine B dye than pure g-C3N4 [21]. By using a template-free chemical coprecipitation technique, Sukanya et al. effectively synthesized ZnFe2O4@g-C3N4 nanocomposites. The nanocomposite exhibited nine times better photocatalytic efficiency than pure ZnFe2O4 and degraded 89% RhB in 30 min under visible light [22]. Pure and modified mesoporous TiO2 nanoparticles with varying loadings of NiO (3–20.0 wt.%) were created using surfactant-assisted sol–gel technique with cetyltrimethylammonium bromide as a template. The photocatalytic performance of the as-synthesized samples was evaluated for the photodegradation of brilliant green (BG) and phenol, as well as hydrogen generation under visible light. Ten percent NiO/TiO2 had resulted in the highest photocatalytic activity [23]. According to Sohier et al., a mesoporous silica/bismuth vanadate composite was created using the ultrasonication technique. The most effective photocatalytic material against MB and BG was mSiO2/BiVO4, with 10.0 weight percent mesoporous SiO2 [24].
Motivated by the enhanced photocatalytic performance of metal-doped ZnFe2O4, S-g-C3N4, and ZnFe2O4/g-C3N4, we fabricated ZnFe2O4, Bi-doped ZnFe2O4, and Bi-ZnFe2O4/S-g-C3N4 via the hydrothermal method and examined their photocatalytic properties. S-g-C3N4 was synthesized by polycondensation of thiourea. The photocatalytic characteristics of synthesized materials were investigated using MB, a model dye. The study was comprised of two phases. The photocatalytic properties of (0.5, 1, 3, 5, 7, and 9 wt.%) of Bi-doped zinc ferrite (Bi-ZnFe2O4) NPs were examined in phase one. The 9% Bi-ZnFe2O4 NPs manifested the highest dye degradation efficiency. In the second phase, the 9% Bi-ZnFe2O4 NPs were mixed with 10, 30, 50, and 70 wt.% concentrations of S-g-C3N4 to produce Bi- ZnFe2O4/S-g-C3N4 NCs to further enhance the photocatalytic activity. The 9% Bi-ZnFe2O4/30% S-g-C3N4 nanocomposite exhibited the best photocatalytic activity compared to pure ZnFe2O4, 9% Bi-ZnFe2O4, and S-g-C3N4. To the best of our knowledge, the synthesis of Bi-ZnFe2O4/S-g-C3N4 heterojunctions via the hydrothermal approach has not been reported to date. This synthetic substance be applied in the wastewater treatment industry.

2. Experimental Section

2.1. Materials

Iron (III) nitrate nonahydrate Fe(NO3)3.9H2O (99.99%), Zinc nitrate hexahydrate (Zn(NO3)2·6H2O) (98%), bismuth nitrate pentahydrate Bi(NO3)3·5H2O (99%), polyvinyl pyrrolidone (PVP), sodium hydroxide (NaOH) (98%), thiourea (CH4N2S) (99.99%), and methylene blue (C16H18ClN3S) were purchased from Merck and used as received.

2.2. Synthesis of Bismuth-Doped Zinc Ferrites

The hydrothermal method was used to synthesize ZnFe2O4 and Bi-ZnFe2O4 NPs [25]. Briefly, to synthesize ZnFe2O4 NPs, 1.61 g of Fe(NO3)3·9H2O (0.2 M) and 1.19 g of Zn(NO3)2·6H2O (0.2 M) were combined in a beaker, to which 50 mL water was added. The mixture was then allowed to stir for about 1 h, at which point 10 mL of PVP solution (1%) was added and stirred. Once the pH of the reaction mixture reached a constant of 11–12, NaOH solution (6 M) was gradually added to the reaction mixture on a magnetic stirrer. Then, the reaction mixture was placed into an autoclave with a Teflon coating and baked for about 16 h at 180 °C. The reaction mixture was filtered, rinsed several times in distilled water, dried in an oven at 100 °C for roughly four hours, and crushed after the autoclave had been running for 16 h. The orange–brown NPs were collected. For the synthesis of Bi-doped zinc ferrites varying weight percentages (0.5%, 1%, 3%, 5%, 7%, 9%, and 11%) of Bi(NO3)3·5H2O were used. For the synthesis of 0.5% Bi-doped zinc ferrite NPs, 0.24 g of Bi(NO3)3·5H2O was dissolved in a beaker containing 100 mL of 0.2 M Fe(NO3)3·9H2O and 0.2 M of Zn(NO3)2·6H2O mixture and stirred for 1 h. The next steps of the synthesis of Bi-ZnFe2O4 were the same as those for ZnFe2O4 NPs. The same procedure was followed during the synthesis of other (1%, 3%, 5%, 7%, 9%, and 11%) Bi-ZnFe2O4 NPs.

2.3. Synthesis of Sulfur-Doped Graphitic Carbon Nitride

In order to prepare sulfur-doped graphitic carbon nitride, 8 g of thiourea was added to an alumina crucible and covered with a lid. The thiourea was heated to 550 °C for about 5 h at a rate of 5 °C/min. a muffle furnace. The prepared sulfur-doped graphitic carbon nitride was then allowed to cool, crushed, and stored for further use [26].

2.4. Synthesis of Bi-ZnFe2O4/S-gC3N4 Nanocomposites

A series of Bi-ZnFe2O4/S-g-C3N4 nanocomposites was made by incorporating 9% Bi-ZnFe2O4 with varying concentrations of S-g-C3N4 (10, 30, 50, 60, and 70 wt.%) via a surfactant-assisted hydrothermal process (Figure 1). For the preparation of 9% Bi-ZnFe2O4/10% S-g-C3N4, 0.19 g of S-g-C3N4, 1.61 g of Fe(NO3)3·9H2O, 1.19 g of Zn(NO3)2·6H2O, and 0.24 g of Bi(NO3)3·5H2O were combined in a beaker and dissolved in 100 mL of water. This mixture was allowed to stir on a magnetic stirrer for 1 h with the addition of 10 mL of PVP solution. Finally, orange–brown precipitates were obtained by following the same steps as those described in Section 2.2. The detailed composition of Bi-ZnFe2O4/S-g-C3N4 with corresponding catalytic efficiency is presented in Table 1.

2.5. Photocatalytic Activity

The photocatalytic activities of zinc ferrites, Bi-doped zinc ferrites, sulfur-doped graphitic carbon nitride, and nanocomposites were studied under solar light irradiation. An aqueous solution of an organic dye, methylene blue, was used as the standard contaminant. An amount of 0.2 g of each photocatalyst was added to a beaker, to which 100 mL of MB solution (10 mg L−1) was added and allowed to stir in the dark for about 45 min to attain the adsorption–desorption equilibrium. Then, the suspension was positioned under solar light in an open atmosphere, and aliquots of 5 mL were collected every 30 min. A UV-Vis spectrophotometer was used to evaluate the photocatalytic activity of the collected samples after centrifugation.

2.6. Characterization

The XRD (Bruker AXS, D8-S4, Billerica, MA, USA) pattern was determined by applying Cu Kα radiation (k = 1.54056 Å) at 40 kV, and 30 mA at room temperature was used to identify the structure of catalysts, whereas SEM-EDS detected the morphology (Hitachi, S-4800, Tokyo, Japan) and the elements present. A UV-vis-NIR spectrophotometer measured the photocatalytic and UV-visible absorption spectra in the wavelength range of 800 nm to 200 nm (UV-770, Jasco, Tokyo, Japan). FT-IR spectrometers measured functional groups in the 4000–400 cm−1 range with a resolution of 1 cm−1 (Perkin 400 FT-IR, Peterborough, England). The surface morphologies of the photocatalysts were examined using a transmission electron microscope (TEM, JEOL-JEM-1230, Tokyo, Japan).

3. Results and Discussion

3.1. XRD Analysis

Figure 2 shows the X-ray diffractograms of ZnFe2O4, 9% Bi-ZnFe2O4, S-g-C3N4, and 9% Bi-ZnFe2O4/30% S-g-C3N4 samples. Seven peaks were observed in the case of pure ZnFe2O4 with crystal facets (220), (311), (400), (422), (511), (440), and (533) at 2θ = 30°, 34.8°, 42.6°, 53°, 56°, 61.8°, and 73.2°, respectively, that fit well with the pattern of standard zinc ferrites with JCPDS file 01-077-0011 [12,27]. The characteristic peaks of zinc ferrites were also present in bismuth-doped zinc ferrite, indicating that no significant change in the structure of zinc ferrite occurred when it was doped with bismuth metal. Two characteristic peaks at 2θ = 13.1° and 27.4° were observed in the XRD pattern of S-g-C3N4; the crystal plane (002) was attributed to the interlayer assembly of aromatic systems, and the (100) plane was ascribed to the interplanar arrangement of aromatic systems (JCPDS file # 00-087-1526) [28,29]. After coupling with S-g-C3N4, the crystal phase of Bi-ZnFe2O4 remained unchanged, and the (002) crystal plane of the S-g-C3N4 (weak) was indicated in the composite systems. In 9%Bi-ZnFe2O4/30% S-g-C3N4 composites, owing to the high crystallinity of Bi-ZnFe2O4, the characteristic peaks of Bi-ZnFe2O4 are prominent. The appearance of the characteristic peaks of both Bi-ZnFe2O4 and S-g-C3N4 in composites indicates the successful fabrication of Bi-ZnFe2O4/S-g-C3N4 composites [30,31].

3.2. TEM and EDX Analyses

The form, size, and distribution of the particles in the as-prepared nanomaterials, as well as their elemental makeup, were disclosed by TEM profiles with EDX spectra. As shown in Figure 3a, pure S-g-C3N4 has an uneven, wrinkled, sheet-like shape. Figure 3b depicts the nanoparticles of ZnFe2O4, which range in size from 24 to 38 nm. With particle sizes ranging from 17 to 26 nm, Bi-ZnFe2O4 is visible as a spherical nanoparticle agglomerated structure (Figure 3c). TEM was conducted as described in Figure 3d to further demonstrate the heterojunction of Bi-ZnFe2O4/S-g-C3N4. The S-g-C3N4 nanosheets were completely covered by spherical Bi-ZnFe2O4 NPs with equal distribution, as shown in the TEM image of the NCs. The sheet-like S-g-C3N4 2D materials aid in enhancing photo-absorption, improving electron transport capabilities, and facilitating longer lifetimes for photo-excited charge carriers.

3.3. FTIR and XPS Analyses

A comparison between FT-IR spectra of prepared photocatalysts is shown in the wavenumber range 4000–450 cm−1 (Figure 4). The FT-IR spectrum of ZnFe2O4 confirms the absence of all functional groups and the presence of an M-O bond. The peak at 3452 cm−1 is the result of O-H bond stretching, whereas the peak at 1488 cm−1 is the result of bending of the O-H bond. The peak at 534 cm−1 confirms M-O bond stretching [32]. The FT-IR spectrum of 9% Bi-ZnFe2O4 shows a small change in peaks compared to pure ZnFe2O4. This slight variation in peak positions confirms the successful formation of Bi-doped ZnFe2O4. In S-g-C3N4, the broadband at 3172 cm−1 is the result of O-H bond stretching, the peaks between 1500–2000 cm−1 are the result of the stretching vibrations of C=N, and the peaks between 1500 and 1000 cm−1 are due to C-N bond stretching [33]. The peak at 805 cm−1 indicates the triazine unit. The FT-IR spectrum of Bi-ZnFe2O4/S-g-C3N4 contains peaks corresponding to both Bi-ZnFe2O4 and S-gC3N4, confirming the successful formation of the composite (Bi-ZnFe2O4/S-g-C3N4) [34]. Bi-ZnFe2O4/S-g-C3N4 was analyzed using XPS (Figure S1) to determine its chemical makeup and the electronic states of each of its constituent components. XPS analysis confirmed that Bi-ZnFe2O4/S-g-C3N4 contains Bi, ZnFe2O4, and S-g-C3N4, as also indicated by TEM and EDX.

3.4. Photocatalytic Degradation Study

Two stages were employed to determine the photocatalytic activity of synthesized samples in sunlight. In the presence of light, an aqueous methylene blue solution was used to examine the photocatalytic activity of pure and Bi-doped zinc ferrite NPs (0, 0.5, 1, 3, 5, 7, 9, and 11 wt.%). A UV-Vis spectrophotometer with a wavelength of 200–800 nm was used to track the dye degradation rate. Among the prepared nanoparticles, 9% Bi-ZnFe2O4 showed the highest degradation of MB dye in sunlight and exhibited 65% dye degradation within 150 min (Figure S2 and Figure 5a). There was a gradual increase in photocatalytic activity of Bi-ZnFe2O4 NPs up to 9% Bi doping, which may have occurred because Bi-doping decreased the bandgap of ZnFe2O4 and facilitated e/h+ pair generation, inhibiting e/h+ pair recombination. 9% Bi doping was determined as the optimal concentration of Bi+3 ions. Increasing the Bi+3 ion concentration beyond this <9 wt.%. leads to a decrease in photocatalytic activity of Bi-ZnFe2O4 NPs (Figure 5a,b). The observed photocatalytic activity of all the Bi-ZnFe2O4 NPs was better than that of ZnFe2O4 NPs (Figure 5b) [11]. A linear relationship was observed between the radiation period and ln (C/Co) for Bi-ZnFe2O4 NPs in the MB degradation reaction, and the accompanying degradation curves are in accordance with pseudo-first-order kinetics (Figure S3).
In the next step, the 9% Bi-ZnFe2O4 NPs were homogenized with varying amounts of S-g-C3N4 (as shown in Table 2) to develop Bi-ZnFe2O4/S-g-C3N4 NCs, the photocatalytic activity was determined in 15 min intervals. The constructed NCs were placed in the dark before exposure to sunlight to induce an equilibrium between dye adsorption and desorption on the ZF, BZFG10, BZFG30, BZFG50, and BZFG70 catalysts. The corresponding amounts of MB adsorbed on the catalysts are shown in Figure 6a.
The graph shown in Figure 6a unequivocally demonstrates that the samples absorbed only modest amounts of dye. Samples were exposed to sunshine, and the BZFG30 NCs exhibited maximum dye degradation compared to other samples (Figure 6b). According to the % degradation plots (Figure 6c) and degradation contours (Figure 6d), upon enhancement of SG contents in the BZFG NCs, dye degradation was increased to BZFG30 (containing 30% S-g-C3N4) and decreased for BZFG50 and BZFG70 (<30% S-g-C3N4). The observed degradation efficiencies of SG, Bi-ZnFe2O4, BZFG10, BZFG30, BZFG50, and BZFG70 catalysts were 41%, 58%, 63%, 99%, 81%, and 72%, respectively, after 120 min of sunlight irradiation.
The reduced photocatalytic efficacy of S-g-C3N4 sheets, which resemble structured catalysts, is caused by the photocatalyst being lost during the recycling process. The catalytic activity is increased by integrating the magnetic NPs into the S-g-C3N4 layer structure. As a result of the robust contact between the NPs, whereby the electrons are diffused in the trap state, the photo-induced electron in the Bi-ZnFe2O4 was transported to S-g-C3N4. To encourage the transition of electron holes from one NP to another and to stop photo-excited electron–hole pairs from recombining, the composite creates a heterojunction. Contrary to pure materials, the heterojunction between the particles contributes to the improvement of the photocatalytic active sites and an increase in capacitive nature. The produced nanocomposite sample contains both S-g-C3N4 and Bi-ZnFe2O4 NPs, as shown by the TEM images. The EDX spectrum (Figure 3e) evidences that Zn and Fe, as well as Bi, S, O, and C, are present in the nanocomposite. Improved charge separation and transfer via Bi-ZnFe2O4 and S-g-C3N4 coupling, as well as increased visible light absorption due to Bi doping in ZnFe2O4, may account for the improved degradation by BZFG [35,36,37]. Figure 6b depicts the % photocatalytic degradation of MB by NCs.
The Langmuir–Hinshelwood model was applied to explain the kinetics [37]. It is evident that the dye degradation by the NCs under sunlight is fit to pseudo-first-order kinetics (Figure 6c). The calculated values of the rate constant (k) of each are shown in Figure S4. BZFG30 (0.0169 min−1) and SG (0.0063 min−1) had the highest and lowest “k” values, respectively. The BZFG30 NCs completely decolorized the MB within 120 min, with a “k” value 1.92 and 2.7 times that of Bi-ZnFe2O4 (0.0088 min−1) and SG (0.0063 min−1), respectively. As the S-g-C3N4 concentration increases from 10% to 30% in BZFG NCs, dye degradation also enhances before reducing in association with concentrations of more than 30%. Thus, 30% S-g-C3N4 is the ideal concentration for BZFG NCs to exhibiting maximum photocatalytic activity. Furthermore, an increase in S-g-C3N4 concentration (<30%) might produce e-h pair combination centers, successively decreasing the photocatalytic efficiency [35,38]. To further analyze this rationalization, a preliminary investigation is required. The photocatalytic efficiency of BZFG50 NC is significantly higher than that reported in previous research, as shown in Table 3.
The stability of the BZFG NCs was investigated by the gradual degradation of MB five times. The obtained findings demonstrate effective MB breakdown; even after the fifth recycling, the percentage was still above 92 percent (Figure 7a). Possible explanations for the small decrease in catalytic effectiveness include partial blockage of active sites. Therefore, the BZFG NC catalysts might function as reliable, effective, and reusable NCs. Figure 7b, which represents the first and fifth cycles of dye photodegradation assessments, shows the XRD patterns of BZFG NCs. Stability analysis revealed that the crystal phase structure of the BZFG NCs did not change noticeably before or after the organic pollutant recycling investigations, demonstrating chemical structural robustness.

4. Photocatalytic Degradation Mechanism

The photocatalytic degradation mechanism follows the same step scheme (S-scheme) as that shown in the schematic sketch in Figure 8. The enhanced degradation of methylene blue by photocatalysts may be ascribed by to the generation of e/h+ pairs. When solar light is irradiated on Bi-ZnFe2O4/S-g-C3N4, both Bi-ZnFe2O4 and S-g-C3N4 are energized, and e/h+ pairs are generated on their conduction band (CB) and valence band (VB), respectively [47]. According to the CB/VB edge potentials, the photo-induced electrons can be easily migrated from the conduction band (CB) of Bi-ZnFe2O4 to the VB of S-g-C3N4, improving the spatial separation of the charge carriers. Additionally, the h+ generated in the VB of S-g-C3N4 could migrate to Bi-ZnFe2O4 [48]. The photogenerated electrons in the CB of ZnFe2O4 and the photogenerated h+ in the VB of S-g-C3N4 likewise rejoin due to Coulombic attraction. Consequently, the inefficient electrons and holes are recombined, but the potent photogenerated electrons in the CB of S-g-C3N4 and the holes in the VB of Bi-ZnFe2O4 are conserved to carry out the photocatalysis reaction. The generated e and h+ react with the water and oxygen molecules absorbed on the surface of the photocatalyst and produce radicals (OH and O−2) [49]. These radicals are utilized to break down MB by transforming it into low-molecular-weight intermediates, which are then changed into H2O, CO2, and inorganic ions via an oxidative mechanism [50].

5. Conclusions

In this study, we used a simple hydrothermal approach to produce ZnFe2O4 and Bi-ZnFe2O4 nanoparticles, as well as a series of Bi-ZnFe2O4/S-g-C3N4 nanocomposites. The formation and purity of samples were examined using XRD, EDX, and FTIR characterizations. The photocatalytic efficiency of the fabricated samples was examined by degrading MB at ambient temperature under sunlight. In a comparative investigation of the photocatalytic effectiveness of the synthesized samples against MB, 9%Bi-ZnFe2O4/30%S-g-C3N4 exhibited maximum catalytic efficiency, and dye degradation was found to have a pseudo-first-order rate constant. The 9%Bi-ZnFe2O4/30%S-g-C3N4 NCs completely mineralized the MB within 120 min, with a “k” 3.5 times that of ZnFe2O4. Therefore, the heterojunction of 9 percent Bi-ZnFe2O4 and 30 percent S-g-C3N4 is a viable option for the photocatalytic removal of organic pollutants from contaminated waste.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal12101175/s1, Figure S1: High-resolution XPS spectra of BZFG NCs; (a) Zn 2p, (b) O 1s, (c) Fe 2p, (d) Bi 4f, (e) C 1s and (f) N 1s; Figure S2: Photodegradation of MB by Bi-ZnFe2O4 NPs at 0 minutes (a) and after 150 minutes (b) of sunlight irradiation (Degradation contours); Figure S3: First-order kinetics plot of S-g-C3N4, Bi-ZnFe2O4 Ns, Bi-ZnFe2O4/S-g-C3N4 NCs; Figure S4: Photodegradation rate constant k (min−1) of Methylene Blue by S-g-C3N4, Bi-ZnFe2O4 NPs, Bi-ZnFe2O4/S-g-C3N4 NCs.

Author Contributions

Conceptualization, writing-original draft Preparation, M.L. and M.J.; methodology, K.J.; software, resources, S.T.; validation, project administration, writing review and editing, S.I., G.L. and W.B.K.; writing review and editing, M.A.Q.; investigation, H.A.; resources, R.A.P.; data curation, E.A.; writing-original draft preparation, A.-E.F. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thank the Deanship of Scientific Research at Umm Al-Qura University for supporting this work by Grant Code: (22UQU4320141DSR33). The authors also gratefully acknowledge the Research Fund Program of Guangdong Provincial Key Lab of Green Chemical Product Technology (grant no. GC202122), the Natural Science Key Foundation of Guangdong Province (2019B1515120056), the Huizhou Science and Technology Project (grant no. 2021WC010308), the major project of Fundamental and Application Research of the Department of Education of Guangdong Province (2017KZDXM080, 2022ZDXM255) and Key Projects of Guangdong Education Department (grant no. 2021ZDJS081).

Data Availability Statement

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

Acknowledgments

The authors would like to thank the Deanship of Scientific Research at Umm Al-Qura University for supporting this work by Grant Code: (22UQU4320141DSR33). The authors also gratefully acknowledge the Research Fund Program of Guangdong Provincial Key Lab of Green Chemical Product Technology (grant no. GC202122), the Natural Science Key Foundation of Guangdong Province (2019B1515120056), the Huizhou Science and Technology Project (grant no. 2021WC010308), the major project of Fundamental and Application Research of the Department of Education of Guangdong Province (2017KZDXM080, 2022ZDXM255) and Key Projects of Guangdong Education Department (grant no. 2021ZDJS081).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, X.; Zhao, Y.; Tian, E.; Li, J.; Ren, Y. Graphene oxide-based polymeric membranes for water treatment. Adv. Mater. Interfaces 2018, 5, 1701427. [Google Scholar] [CrossRef]
  2. Saleh, T.A.; Mustaqeem, M.; Khaled, M. Water treatment technologies in removing heavy metal ions from wastewater: A review. Environ. Nanotechnol. Monit. Manag. 2022, 17, 100617. [Google Scholar] [CrossRef]
  3. Dhasmana, A.; Uniyal, S.; Kumar, V.; Gupta, S.; Kesari, K.K.; Haque, S.; Lohani, M.; Pandey, J. Scope of nanoparticles in environmental toxicant remediation. In Environmental Biotechnology: For Sustainable Future; Springer: New York, NY, USA, 2019; pp. 31–44. [Google Scholar]
  4. Alshorifi, F.T.; Alswat, A.A.; Mannaa, M.A.; Alotaibi, M.T.; El-Bahy, S.M.; Salama, R.S. Facile and Green Synthesis of Silver Quantum Dots Immobilized onto a Polymeric CTS–PEO Blend for the Photocatalytic Degradation of p-Nitrophenol. ACS Omega 2021, 6, 30432–30441. [Google Scholar] [CrossRef] [PubMed]
  5. Muhammad Irfan, R.; Hussain Tahir, M.; Maqsood, M.; Lin, Y.; Bashir, T.; Iqbal, S.; Zhao, J.; Gao, L.; Haroon, M. CoSe as Non-Noble-Metal Cocatalyst Integrated with Heterojunction Photosensitizer for Inexpensive H2 Production under Visible Light. J. Catal. 2020, 390, 196–205. [Google Scholar] [CrossRef]
  6. Iqbal, S. Spatial charge separation and transfer in L-cysteine capped NiCoP/CdS nano-heterojunction activated with intimate covalent bonding for high-quantum-yield photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2020, 274, 119097. [Google Scholar] [CrossRef]
  7. Sun, J.; Wu, T.; Liu, Z.; Shao, B.; Liang, Q.; He, Q.; Luo, S.; Pan, Y.; Zhao, C.; Huang, D. Peroxymonosulfate activation induced by spinel ferrite nanoparticles and their nanocomposites for organic pollutants removal: A review. J. Clean. Prod. 2022, 346, 131143. [Google Scholar] [CrossRef]
  8. Amin, A.M.; Rayan, D.A.; Ahmed, Y.M.; El-Shall, M.S.; Abdelbasir, S.M. Zinc ferrite nanoparticles from industrial waste for Se (IV) elimination from wastewater. J. Environ. Manag. 2022, 312, 114956. [Google Scholar] [CrossRef]
  9. Bahadur, A.; Saeed, A.; Shoaib, M.; Iqbal, S.; Anwer, S. Modulating the burst drug release effect of waterborne polyurethane matrix by modifying with polymethylmethacrylate. J. Appl. Polym. Sci. 2019, 136, 47253. [Google Scholar] [CrossRef]
  10. Naik, C.C.; Salker, A. Fractional substitution of Mn ions in cobalt-copper ferrite: Effect on its magnetic, dielectric and microstructural properties. Inorg. Chem. Commun. 2022, 142, 109684. [Google Scholar] [CrossRef]
  11. Savunthari, K.V.; Shanmugam, S. Effect of co-doping of bismuth, copper and cerium in zinc ferrite on the photocatalytic degradation of bisphenol A. J. Taiwan Inst. Chem. Eng. 2019, 101, 105–118. [Google Scholar] [CrossRef]
  12. Patil, S.; Bhojya Naik, H.; Nagaraju, G.; Viswanath, R.; Rashmi, S. Synthesis of visible light active Gd3+-substituted ZnFe2O4 nanoparticles for photocatalytic and antibacterial activities. Eur. Phys. J. Plus 2017, 132, 328. [Google Scholar] [CrossRef]
  13. Wu, K.; Li, J.; Zhang, C. Zinc ferrite based gas sensors: A review. Ceram. Int. 2019, 45, 11143–11157. [Google Scholar] [CrossRef]
  14. Fan, G.; Tong, J.; Li, F. Visible-light-induced photocatalyst based on cobalt-doped zinc ferrite nanocrystals. Ind. Eng. Chem. Res. 2012, 51, 13639–13647. [Google Scholar] [CrossRef]
  15. Shoaib, M.; Bahadur, A.; ur Rahman, M.S.; Iqbal, S.; Arshad, M.I.; Tahir, M.A.; Mahmood, T. Technology, Sustained drug delivery of doxorubicin as a function of pH, releasing media, and NCO contents in polyurethane urea elastomers. J. Drug Deliv. Sci. Technol. 2017, 39, 277–282. [Google Scholar] [CrossRef]
  16. Sharmila, M.; Mani, R.J.; Parvathiraja, C.; Kader, S.A.; Siddiqui, M.R.; Wabaidur, S.M.; Islam, M.A.; Lai, W.-C. Visible Light Photocatalyst and Antibacterial Activity of BFO (Bismuth Ferrite) Nanoparticles from Honey. Water 2022, 14, 1545. [Google Scholar] [CrossRef]
  17. Amini Herab, A.; Salari, D.; Tseng, H.-H.; Niaei, A.; Mehrizadeh, H.; Rahimi Aghdam, T. Synthesis of BiFeO3 nanoparticles for the photocatalytic removal of chlorobenzene and a study of the effective parameters. React. Kinet. Mech. Catal. 2020, 131, 437–452. [Google Scholar] [CrossRef]
  18. Chen, Z.; Zhang, S.; Liu, Y.; Alharbi, N.S.; Rabah, S.O.; Wang, S.; Wang, X. Synthesis and fabrication of g-C3N4-based materials and their application in elimination of pollutants. Sci. Total Environ. 2020, 731, 139054. [Google Scholar] [CrossRef]
  19. Bai, X.; Wang, L.; Zong, R.; Zhu, Y. Photocatalytic activity enhanced via g-C3N4 nanoplates to nanorods. J. Phys. Chem. C 2013, 117, 9952–9961. [Google Scholar] [CrossRef]
  20. Hong, J.; Xia, X.; Wang, Y.; Xu, R. Mesoporous carbon nitride with in situ sulfur doping for enhanced photocatalytic hydrogen evolution from water under visible light. J. Mater. Chem. 2012, 22, 15006–15012. [Google Scholar] [CrossRef]
  21. Fan, Q.; Liu, J.; Yu, Y.; Zuo, S.; Li, B. A simple fabrication for sulfur doped graphitic carbon nitride porous rods with excellent photocatalytic activity degrading RhB dye. Appl. Surf. Sci. 2017, 391, 360–368. [Google Scholar] [CrossRef]
  22. Borthakur, S.; Saikia, L. ZnFe2O4@g-C3N4 nanocomposites: An efficient catalyst for Fenton-like photodegradation of environmentally pollutant Rhodamine B. J. Environ. Chem. Eng. 2019, 7, 103035. [Google Scholar] [CrossRef]
  23. Mannaa, M.A.; Qasim, K.F.; Alshorifi, F.T.; El-Bahy, S.M.; Salama, R.S. Role of NiO nanoparticles in enhancing structure properties of TiO2 and its applications in photodegradation and hydrogen evolution. ACS Omega 2021, 6, 30386–30400. [Google Scholar] [CrossRef]
  24. El-Hakam, S.A.; ALShorifi, F.T.; Salama, R.S.; Gamal, S.; El-Yazeed, W.A.; Ibrahim, A.A.; Ahmed, A.I. Application of nanostructured mesoporous silica/bismuth vanadate composite catalysts for the degradation of methylene blue and brilliant green. J. Mater. Res. Technol. 2022, 18, 1963–1976. [Google Scholar] [CrossRef]
  25. Shanavas, S.; Ahamad, T.; Alshehri, S.M.; Acevedo, R.; Munusamy Anbarasan, P. Hydrothermal assisted synthesis of ZnFe2O4 embedded g-C3N4 nanocomposite with enhanced charge transfer ability for effective removal of nitrobenzene and Cr (VI). ChemistrySelect 2020, 5, 5117–5127. [Google Scholar] [CrossRef]
  26. Wang, K.; Li, Q.; Liu, B.; Cheng, B.; Ho, W.; Yu, J. Sulfur-doped g-C3N4 with enhanced photocatalytic CO2-reduction performance. Appl. Catal. B Environ. 2015, 176–177, 44–52. [Google Scholar] [CrossRef]
  27. Askari, M.B.; Salarizadeh, P.; Seifi, M.; Di Bartolomeo, A. ZnFe2O4 nanorods on reduced graphene oxide as advanced supercapacitor electrodes. J. Alloys Compd. 2021, 860, 158497. [Google Scholar] [CrossRef]
  28. Jourshabani, M.; Shariatinia, Z.; Badiei, A. Facile one-pot synthesis of cerium oxide/sulfur-doped graphitic carbon nitride (g-C3N4) as efficient nanophotocatalysts under visible light irradiation. J. Colloid Interface Sci. 2017, 507, 59–73. [Google Scholar] [CrossRef] [PubMed]
  29. Kalisamy, P.; Lallimathi, M.; Suryamathi, M.; Palanivel, B.; Venkatachalam, M. ZnO-embedded S-doped gC3N4 heterojunction: Mediator-free Z-scheme mechanism for enhanced charge separation and photocatalytic degradation. RSC Adv. 2020, 10, 28365–28375. [Google Scholar] [CrossRef]
  30. Dai, Z.; Zhen, Y.; Sun, Y.; Li, L.; Ding, D. ZnFe2O4/g-C3N4 S-scheme photocatalyst with enhanced adsorption and photocatalytic activity for uranium (VI) removal. Chem. Eng. J. 2021, 415, 129002. [Google Scholar] [CrossRef]
  31. Kuai, S.; Zhang, Z.; Nan, Z. Synthesis of Ce3+ doped ZnFe2O4 self-assembled clusters and adsorption of chromium (VI). J. Hazard. Mater. 2013, 250, 229–237. [Google Scholar] [CrossRef]
  32. Kuang, M.; Zhang, J.; Wang, W.; Chen, J.; Liu, R.; Xie, S.; Wang, J.; Ji, Z. Synthesis of octahedral-like ZnO/ZnFe2O4 heterojunction photocatalysts with superior photocatalytic activity. Solid State Sci. 2019, 96, 105901. [Google Scholar] [CrossRef]
  33. Qamar, M.A.; Shahid, S.; Javed, M. Synthesis of dynamic g-C3N4/Fe@ZnO nanocomposites for environmental remediation applications. Ceram. Int. 2020, 46, 22171–22180. [Google Scholar] [CrossRef]
  34. Palanivel, B.; Maiyalagan, T.; Jayarman, V.; Ayyappan, C.; Alagiri, M. Rational design of ZnFe2O4/g-C3N4 nanocomposite for enhanced photo-Fenton reaction and supercapacitor performance. Appl. Surf. Sci. 2019, 498, 143807. [Google Scholar] [CrossRef]
  35. Qamar, M.A.; Shahid, S.; Javed, M.; Iqbal, S.; Sher, M.; Bahadur, A.; AL-Anazy, M.M.; Laref, A.; Li, D. Designing of highly active g-C3N4/Ni-ZnO photocatalyst nanocomposite for the disinfection and degradation of the organic dye under sunlight radiations. Colloids Surf. A Physicochem. Eng. Asp. 2021, 614, 126176. [Google Scholar] [CrossRef]
  36. Renukadevi, S.; Jeyakumari, A.P. Rational design of ZnFe2O4/g-C3N4 heterostructures composites for high efficient visible-light photocatalysis for degradation of aqueous organic pollutants. Inorg. Chem. Commun. 2020, 118, 108047. [Google Scholar] [CrossRef]
  37. Farouq, R. Coupling Adsorption-Photocatalytic Degradation of Methylene Blue and Maxilon Red. J. Fluoresc. 2022, 32, 1381–1388. [Google Scholar] [CrossRef]
  38. Sher, M.; Shahid, S.; Javed, M. Synthesis of a novel ternary (g-C3N4 nanosheets loaded with Mo doped ZnOnanoparticles) nanocomposite for superior photocatalytic and antibacterial applications. J. Photochem. Photobiol. B Biol. 2021, 219, 112202. [Google Scholar] [CrossRef]
  39. Dariani, R.; Esmaeili, A.; Mortezaali, A.; Dehghanpour, S. Photocatalytic reaction and degradation of methylene blue on TiO2 nano-sized particles. Optik 2016, 127, 7143–7154. [Google Scholar] [CrossRef]
  40. Wang, C.; Zhu, L.; Chang, C.; Fu, Y.; Chu, X. Preparation of magnetic composite photocatalyst Bi2WO6/CoFe2O4by two-step hydrothermal method and itsphotocatalytic degradation of bisphenol A. Catal. Commun. 2013, 37, 92–95. [Google Scholar] [CrossRef]
  41. Messih, M.A.; Ahmed, M.; Soltan, A.; Anis, S.S. Facile approach for homogeneous dispersion of metallic silver nanoparticles on the surface of mesoporous titania for photocatalytic degradation of methylene blue and indigo carmine dyes. J. Photochem. Photobiol. A Chem. 2017, 335, 40–51. [Google Scholar] [CrossRef]
  42. Ye, Z.; Kong, L.; Chen, F.; Chen, Z.; Lin, Y.; Liu, C. A comparative study of photocatalytic activity of ZnS photocatalyst for degradation of various dyes. Optik 2018, 164, 345–354. [Google Scholar] [CrossRef]
  43. Lin, X.P.; Huang, F.Q.; Wang, W.D.; Zhang, K.L. A novel photocatalyst BiSbO4 for degradation of methylene blue. Appl. Catal. A Gen. 2006, 307, 257–262. [Google Scholar] [CrossRef]
  44. Wang, R.; Shi, K.; Huang, D.; Zhang, J.; An, S. Synthesis and degradation kinetics of TiO2/GO composites with highly efficient activity for adsorption and photocatalytic degradation of MB. Sci. Rep. 2019, 9, 18744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Kokilavani, S.; Syed, A.; Kumar, B.H.; Elgorban, A.M.; Bahkali, A.H.; Ahmed, B.; Das, A.; Khan, S.S. Facile synthesis of MgS/Ag2MoO4 nanohybrid heterojunction: Outstanding visible light harvesting for boosted photocatalytic degradation of MB and its anti-microbial applications. Colloids Surf. A Physicochem. Eng. Asp. 2021, 627, 127097. [Google Scholar] [CrossRef]
  46. Min, S.; Wang, F.; Jin, Z.; Xu, J. Cu2O nanoparticles decorated BiVO4 as an effective visible-light-driven pn heterojunction photocatalyst for methylene blue degradation. Superlattices Microstruct. 2014, 74, 294–307. [Google Scholar] [CrossRef]
  47. Zhang, H.; Zhu, C.; Zhang, G.; Li, M.; Tang, Q.; Cao, J. Palladium modified ZnFe2O4/g-C3N4 nanocomposite as an efficiently magnetic recycling photocatalyst. J. Solid State Chem. 2020, 288, 121389. [Google Scholar] [CrossRef]
  48. Shi, Y.; Li, L.; Xu, Z.; Sun, H.; Amin, S.; Guo, F.; Shi, W.; Li, Y. Engineering of 2D/3D architectures type II heterojunction with high-crystalline g-C3N4 nanosheets on yolk-shell ZnFe2O4 for enhanced photocatalytic tetracycline degradation. Mater. Res. Bull. 2022, 150, 111789. [Google Scholar] [CrossRef]
  49. Wu, Y.; Wang, Y.; Di, A.; Yang, X.; Chen, G. Enhanced photocatalytic performance of hierarchical ZnFe2O4/g-C3N4 heterojunction composite microspheres. Catal. Lett. 2018, 148, 2179–2189. [Google Scholar] [CrossRef]
  50. Javed, M.; Qamar, M.A.; Shahid, S.; Alsaab, H.O.; Asif, S. Highly efficient visible light active Cu–ZnO/Sg-C3N4 nanocomposites for efficient photocatalytic degradation of organic pollutants. RSC Adv. 2021, 11, 37254–37267. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the synthesis of Bi-ZnFe2O4/S-g-C3N4.
Figure 1. Schematic representation of the synthesis of Bi-ZnFe2O4/S-g-C3N4.
Catalysts 12 01175 g001
Figure 2. XRD spectra of composites of ZnFe2O4, S-g-C3N4, 9% Bi-ZnFe2O4, and 9% Bi-ZnFe2O4/30% S-g-C3N4.
Figure 2. XRD spectra of composites of ZnFe2O4, S-g-C3N4, 9% Bi-ZnFe2O4, and 9% Bi-ZnFe2O4/30% S-g-C3N4.
Catalysts 12 01175 g002
Figure 3. TEM profiles of (a) S-g-C3N4, (b) ZnFe2O4, (c) 9% Bi-ZnFe2O4, and (d) 9% Bi-ZnFe2O4/30% S-g-C3N4 NCs. (e) EDX of 9% Bi-ZnFe2O4/30% S-g-C3N4 NCs.
Figure 3. TEM profiles of (a) S-g-C3N4, (b) ZnFe2O4, (c) 9% Bi-ZnFe2O4, and (d) 9% Bi-ZnFe2O4/30% S-g-C3N4 NCs. (e) EDX of 9% Bi-ZnFe2O4/30% S-g-C3N4 NCs.
Catalysts 12 01175 g003
Figure 4. FTIR spectra of composites of ZnFe2O4, S-gC3N4, 9% Bi-ZnFe2O4, and 9% Bi-ZnFe2O4/30S-gC3N4.
Figure 4. FTIR spectra of composites of ZnFe2O4, S-gC3N4, 9% Bi-ZnFe2O4, and 9% Bi-ZnFe2O4/30S-gC3N4.
Catalysts 12 01175 g004
Figure 5. Photocatalytic activity of Bi-ZnFe2O4 NPs against MB (a) and % degradation of MB by Bi-ZnFe2O4 NPs (b).
Figure 5. Photocatalytic activity of Bi-ZnFe2O4 NPs against MB (a) and % degradation of MB by Bi-ZnFe2O4 NPs (b).
Catalysts 12 01175 g005
Figure 6. Photocatalytic degradation rate (a), % degradation (b), kinetic characteristics (c), MB degradation contours by Bi-ZnFe2O4/S-g-C3N4, and NCs after 120 min (d).
Figure 6. Photocatalytic degradation rate (a), % degradation (b), kinetic characteristics (c), MB degradation contours by Bi-ZnFe2O4/S-g-C3N4, and NCs after 120 min (d).
Catalysts 12 01175 g006
Figure 7. (a) Cyclic stability of BZFG NC photocatalysts through the fifth cycle. (b) The structural stability of BZFG NCs was determined by comparing XRD patterns obtained before the first cycle and those obtained after the fifth recycling experiment.
Figure 7. (a) Cyclic stability of BZFG NC photocatalysts through the fifth cycle. (b) The structural stability of BZFG NCs was determined by comparing XRD patterns obtained before the first cycle and those obtained after the fifth recycling experiment.
Catalysts 12 01175 g007
Figure 8. Schematic representation of the MB sunlight catalytic degradation mechanism over the BZFG NCs.
Figure 8. Schematic representation of the MB sunlight catalytic degradation mechanism over the BZFG NCs.
Catalysts 12 01175 g008
Table 1. Composition of Bi-ZnFe2O4/S-g-C3N4 nanocomposites and their corresponding photocatalytic activities.
Table 1. Composition of Bi-ZnFe2O4/S-g-C3N4 nanocomposites and their corresponding photocatalytic activities.
Sr. NoNanocompositesWt. in gm.Photocatalytic Activity
S-g-C3N4Fe(NO3)3Zn(NO3)2Bi(NO3)3%
19% Bi-ZnFe2O4/10% S-g-C3N4 0.191.611.190.1763
29% Bi-ZnFe2O4/30% S-g-C3N4 0.581.611.190.1799
39% Bi-ZnFe2O4/50% S-g-C3N4 0.961.611.190.1781
49% Bi-ZnFe2O4/70% S-g-C3N4 1.351.611.190.1772
5S-g-C3N4 10000000041
79%Bi-ZnFe2O4001.611.190.1758
Table 2. Rate constant (k) values of the 9% Bi-ZnFe2O4/S-g-C3N4 nanocomposites.
Table 2. Rate constant (k) values of the 9% Bi-ZnFe2O4/S-g-C3N4 nanocomposites.
Sr. No.NanocompositeS-g-C3N4
(wt.%)
k (min−1)Nanocomposite CodeR2
1S-g-C3N41000.00625SG0.921
29% Bi-ZnFe2O4-0.0088BZF0.941
39% Bi-ZnFe2O4/10S-g-C3N4100.0095BZFG100.974
49% Bi-ZnFe2O4/30S-g-C3N4300.0169BZFG300.996
59% Bi-ZnFe2O4/50S-g-C3N4500.0156BZFG500.989
69% Bi-ZnFe2O4/70S-g-C3N4700.0129BZFG700.978
Table 3. Comparison of photocatalytic efficiency of the Bi-ZnFe2O4/S-g-C3N4 NCs with those reported in previous works.
Table 3. Comparison of photocatalytic efficiency of the Bi-ZnFe2O4/S-g-C3N4 NCs with those reported in previous works.
Sr. NoPhotocatalystContaminantLight SourceRadiation Time (min.)Degradation %Ref.
1TiO2MBXe lamp12099[39]
2FeOOH-LDOMBXe lamp18092[40]
3AgTi5MBUV20097.1[41]
4ZnSMBXe lamp12077.2[42]
5BiSbO4MBXe lamp10 h96.7[43]
5TiO2/GOMBSolar14097.5[44]
7MgS/Ag2MoO4MBVisible20090[45]
8Cu2O/BiVO4MBVisible18073[46]
9Bi-ZnFe2O4/S-g-C3N4MBSolar12097Present Work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lu, M.; Javed, M.; Javed, K.; Tan, S.; Iqbal, S.; Liu, G.; Khalid, W.B.; Qamar, M.A.; Alrbyawi, H.; Pashameah, R.A.; et al. Construction of a Well-Defined S-Scheme Heterojunction Based on Bi-ZnFe2O4/S-g-C3N4 Nanocomposite Photocatalyst to Support Photocatalytic Pollutant Degradation Driven by Sunlight. Catalysts 2022, 12, 1175. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12101175

AMA Style

Lu M, Javed M, Javed K, Tan S, Iqbal S, Liu G, Khalid WB, Qamar MA, Alrbyawi H, Pashameah RA, et al. Construction of a Well-Defined S-Scheme Heterojunction Based on Bi-ZnFe2O4/S-g-C3N4 Nanocomposite Photocatalyst to Support Photocatalytic Pollutant Degradation Driven by Sunlight. Catalysts. 2022; 12(10):1175. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12101175

Chicago/Turabian Style

Lu, Ming, Mohsin Javed, Kainat Javed, Shaozao Tan, Shahid Iqbal, Guocong Liu, Waleed Bin Khalid, Muhammad Azam Qamar, Hamad Alrbyawi, Rami Adel Pashameah, and et al. 2022. "Construction of a Well-Defined S-Scheme Heterojunction Based on Bi-ZnFe2O4/S-g-C3N4 Nanocomposite Photocatalyst to Support Photocatalytic Pollutant Degradation Driven by Sunlight" Catalysts 12, no. 10: 1175. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12101175

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop