Next Article in Journal
A Review of Noble Metal Catalysts for Catalytic Removal of VOCs
Next Article in Special Issue
Hydrogen Evolution Reaction Activities of Room-Temperature Self-Grown Glycerol-Assisted Nickel Chloride Nanostructures
Previous Article in Journal
Highly Dispersed Nickel Nanoparticles on Hierarchically Ordered Macroporous Al2O3 and Its Catalytic Performance for Steam Reforming of 1-Methyl Naphthalene
Previous Article in Special Issue
Effects of Preparation Methods of Pd Supported on (001) Crystal Facets Exposed TiO2 Nanosheets for Toluene Catalytic Combustion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Augmenting the Photocatalytic Performance of Direct Z-Scheme Bi2O3/g-C3N4 Nanocomposite

by
Krishnasamy Mahalakshmi
1,
Rajendran Ranjith
2,
Pazhanivel Thangavelu
3,
Matheshwaran Priyadharshini
3,
Baskaran Palanivel
3,4,
Mohamed Aslam Manthrammel
5,
Mohd Shkir
5,6 and
Barathi Diravidamani
1,*
1
Department of Physics, N.K.R Government Arts College for Women, Namakkal 637001, India
2
Department of Physics, K.S.R. College of Engineering, Tiruchengode, Namakkal 637215, India
3
Smart Materials Lab, Department of Physics, Periyar University, Salem 636011, India
4
Department of Physics, Kings Engineering College, Kancheepuram 602117, India
5
Department of Physics, Faculty of Science, King Khalid University, P.O Box-9004, Abha 61413, Saudi Arabia
6
Department of Chemistry and University Centre for Research & Development, Chandigarh University, Mohali 140413, India
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1544; https://doi.org/10.3390/catal12121544
Submission received: 8 October 2022 / Revised: 28 November 2022 / Accepted: 28 November 2022 / Published: 1 December 2022
(This article belongs to the Special Issue Advanced Nanostructured Materials for Modern Catalysis Applications)

Abstract

:
Huge demands for photocatalytically efficient visible-light-induced catalysts have spurred widespread interest in building adaptable heterojunctions. Here, we used in situ thermal polymerization to synthesise the Z-scheme Bi2O3/g-C3N4 heterojunction. The optical, structural, chemical, compositional and photocatalytic behaviours of the samples were analysed through various analytical techniques and photocatalytic methylene blue (MB) dye degradation reaction. Among the various ratios of Bi2O3/g-C3N4 heterojunction composites, the 1:1 ratio showed improved visible-light-induced catalytic activity, which attained 91.2% degradation efficiency after 120 min of visible-light exposure. The dye degradation efficiency was calculated under various environmental conditions by varying the dye concentration, solution pH and catalyst dosage. A improved Z-scheme photocatalytic mechanism was proposed in light of the results. A potential mechanism was suggested to explain the photocatalytic activity, and trapping experiments supported it. Last but not least, this strategy might be helpful to prepare the heterojunction photocatalyst for the degradation of organic pigments.

Graphical Abstract

1. Introduction

Overpopulation and the depletion of natural resources have had a serious impact on ecosystems and human life. Conventional physicochemical approaches have not adequately addressed the issue of ecosystem quality and health [1]. Rapid urbanization and industrialization have created significant global environmental issues with the release of large amounts of untreated wastewater into the ecosystem. Generally, wastewater contains pollutants, including pesticides, organic dyes, heavy metals and pharmaceutical waste. In particular, several synthetic dyes, such as methylene blue (MB), methyl orange (MO), rhodamine B (RhB), congo-red (CR), malachite green (MG) and methyl red (MR), are widely used in the paint, printing and textile industries. These organic dyes are highly stable due to their complex structure; they are not bio-degradable and have a carcinogenic effect on the water environment and human health [2,3,4]. Researchers have developed several treatment methods to reduce the impact of wastewater. However, most waste treatment methods have failed to remove these toxic pollutants from wastewater. Finding a potential alternative has become crucial for this purpose. Various eco-friendly advanced oxidation processes have been widely used in multiple areas of toxic agent removal. Specifically, semiconductor-based photocatalysts have gained much attention due to their high degradation efficiency cost-effectiveness and wide utilisation range of the solar spectrum [5]. In particular, metal oxides and their composites have been used as catalysts for the photocatalytic pollutant removal process, using solar energy utilization for the conversion of organic dyes into harmless products [2]. Several nanostructured metal oxides, such as TiO2, ZnO, CdO, CeO2, CuO, Co3O4, Fe2O3, WO3, V2O5, and NiO, have been developed for photocatalytic degradation applications. Among these materials, TiO2 and ZnO have received much attention due to their high redox ability. However, traditional TiO2 and ZnO photocatalysts are excited by only 2 to 4% of the solar spectrum. In addition, faster recombination of photo-generated excitons in single-phase components forbids the extension of activity [6]. So, great interest has been paid to the development of photocatalysts comprising larger exciton separation, wider absorption of visible spectra and better efficiency retention over multiple cycles.
The adjustable electronic structure of Bi2O3, from around 2 to 3.96 eV, has stimulated researchers’ attention. It offers easy synthesis procedures, is earth-abundant, has a favourable electronic arrangement, and demonstrates a high harvesting efficiency of the solar spectrum. Its efficiency in the removal of organic dyes has been analysed [7]. Unfortunately, the material fails to meet requirements due to the rapid recombination of charge carriers, which greatly affects its widespread application at the industrial level. To overcome the stated drawback, several strategies have been used. Some of them are (i) structural modification, (ii) doping, (iii) surface modification and (iv) developing a heterostructure [8]. Developing a heterostructure was found to be the most successful way of reducing the rate of recombination. In addition, it also enhanced photocatalytic degradation efficiency. As of now, a number of semiconductors, such as CeO2, MoS2 and BiOCl, have successfully been linked with Bi2O3 to create a stable structure that is conducive to the degradation of organic dyes such as methylene blue, methyl orange and malachite green [9].
Following the publication of a study by Wang et al. showing that graphitic carbon nitride demonstrates photocatalytic activity for the generation of H2 and O2 through water splitting and the decolorization of organic pollutants using visible spectrum, g-C3N4 gained widespread use as a novel metal-free visible-light photocatalyst [10,11]. It potentially withstands chemicals and temperature because of the s-triazine ring and high condensation. The presence of abundant nitrogen provides a greater number of active sites compared to other carbon, which is conductive in the photocatalytic mechanism [12,13].
It is anticipated that the efficient combination of g-C3N4 and Bi2O3 will answer the demand for an effective photocatalyst. The combination also offers a Z-scheme photocatalytic system, which is more effective in photocatalytic activity than single-phase materials. Compared with the traditional composite systems, the Z-scheme heterojunction systems offer stronger redox capability and higher photocatalytic efficiency [14]. A Z-scheme heterojunction system is formed between two semiconductors due to the short distance between the CB of one semiconductor and VB of the another semiconductor. In this circumstance, the electron at the CB of one semiconductor would be quickly combined with the VB holes of the another semiconductor. Therefore, the excited CB electrons in one semiconductor exhibit strong reduction reaction, while the holes in VB of the another semiconductor produce an excellent oxidation reaction [15]. In 2014, Jinfeng Zhang and co-authors synthesised a Z-scheme Bi2O3/g-C3N4 nanocomposite by the ball milling assisted calcination method. They reported that the prepared nanocomposites showed better photocatalytic activity compared to those of pure nanoparticles. Moreover, they reported that the optimal composite exhibited 78.1% and 45.6 % of TOC removal efficiency against MB and RhB dye, respectively. Further, they proved that the nanocomposite degraded the dye molecules through the Z-scheme mechanism. He et.al reported the Z-scheme type degradation of phenol pollutants using in situ fabricated Bi2O3/g-C3N4 nanocomposite at room temperature conditions. They showed that the room temperature in situ process helps to produce quantum-sized Bi2O3 on g-C3N4 nanosheets. This heterojunction helps to enhance the light absorption ability and phenol degradation efficiency [14].
From the literature survey, Bi2O3 has a valence band (VB) and conduction band (CB) at 3.13 and 0.33 eV, respectively. In addition g-C3N4 has a valence band and conduction band at 1.57 and −1.13 eV, respectively. When g-C3N4 and Bi2O3 are combined in this way, the electrons from Bi2O3’s conduction band quickly interact with the holes on g-C3N4’s valence band. As a result, holes in the valence band of Bi2O3 demonstrate stronger oxidation ability than photogenerated electrons on the conduction band of g-C3N4 [16,17,18,19].
In this study, we used a straightforward approach to successfully couple Bi2O3 with g-C3N4, an oxidation type semiconductor with a reduction type semiconductor. Th single-step solid state thermal polymerization method is used to prepare the nanocomposites. This process helps to intercalate the Bi2O3 nanoparticles into the g-C3N4 interlayers, so that the composite helps to enhance the optical and catalytic behaviour compared with pure nanoparticles. The impact of Bi2O3 content on the physical and chemical characteristics as well as on the photocatalytic activity of the composites is thoroughly examined. In light of this, a direct Z-scheme mechanism is discovered for produced composites, which mostly accounts for the augmentation of photoactivity in Bi2O3/g-C3N4 composites [20]. This research offers a practical method for designing and building g-C3N4-based Z-scheme heterojunction photocatalysts for use in solar energy conversion and water purification.

2. Results and Discussion

The synthesised samples’ structure and phase purity was analysed by XRD diffraction pattern and is shown in Figure 1. According to the XRD pattern shown in the figure, pure g-C3N4 has two distinctive peaks at 27.73° (002) and 13.1° (100), which correspond to long-range interplanar stacking of the aromatic system and in-plane structural packing, or the sequential distance between the tris triazine units. The existence of new peaks that correspond to Bi2O3 nanoparticles, as shown in the Bi2O3/g-C3N4 XRD pattern, confirms the incorporation of Bi2O3. Obviously, the prepared samples are crystalline in nature, and the different peaks are in good arrangement with the Joint Committee on Powder Diffraction Standards (JDPDS) Card No (87-1526) and (78-1793) for g-C₃N₄ and Bi₂O₃, respectively. The absence of peaks related to potential contaminants suggests that the processed samples were pure. We observed a slight decrement in the intensity of the spectra of Bi2O3 with the increase in the ratio of g-C3N4. The average crystallite size of the prepared nanoparticles was calculated by Scherrers’ formula [2]:
Crystallite   Size   ( D ) =   k λ β cos θ
The calculated crystallite size of the pure Bi2O3 nanoparticle is ~34 nm, which is little larger than the 1:1 (23 nm) and 3:1 (30 nm) nanocomposite. This result indicates that the addition of g-C3N4 reduces the crystallite size of the Bi2O3 nanoparticle in the composite system. This reduction of crystallite size is reflected in the diffraction peak intensity reduction and peak broadening.
Morphology of the as-synthesized samples was examined by SEM analysis and is shown in Figure 2. The SEM image of g-C3N4 describes that it has a stacked nanosheet structure and it is well matched with the literature [15]. The pure Bi2O3 nanoparticle exhibits sub-micron level particle–like morphology and is shown in Figure 2b. The nanocomposite shows the combination of nanosheets and nanoparticles, which reflects the presence of both g-C3N4 nanosheets and Bi2O3 nanoparticles. It can be clearly seen from the SEM analysis that the nanoparticles are deposited on the g-C3N4 nanosheets. Further, the EDS analysis was performed on 1:1 nanocomposite to determine the elemental distribution and is shown in Figure 2. The EDS spectrum and elemental distribution micrographs show the presence of C, N, Bi and O in the nanocomposite. The EDS spectrum does not show any other peaks other than the above-mentioned elements, which indicates the purity of the prepared sample.
Further, TEM analysis (Figure 3) was performed to illustrate the heterojunction between the nanoparticles (g-C3N4 and Bi2O3) in the nanocomposite and particle size of the nanoparticle. The TEM image clearly shows that the Bi2O3 nanoparticles are deposited on the g-C3N4 nanosheet and form a strong heterojunction. The measured particle size of the Bi2O3 nanoparticle is about ~20–35 nm in size, which is in good agreement with the crystallitie size of the Bi2O3 nanoparticle in the 1:1 g-C3N4/Bi2O3 nanocomposite. The TEM image clearly shows that the Bi2O3 nanoparticles are highly agglomerated, exhibit particle-like morphology, and are anchored on the g-C3N4 nanosheet. The HRTEM image of the nanocomposite confirms that a strong interfacial interaction exists between the nanoparticles. The measured d-space value of 0.29 nm is related to the plane (220) of Bi2O3 nanoparticle. The polycrystalline nature of the as-prepared nanocomposite is clearly shown by the SAED pattern. The heterojunction between the nanoparticles can be helpful to enhance the optical, electrical and photocatalytic activity.
Figure 4 shows the FT-IR spectra of pure and composite materials to identify the functional groups presented in the nanoparticles. The breathing mode of the s-triazine units and the stretching vibration of the C-N and C=N in heterocycles are attributed to the main absorption band in the range of 805, and 1240–1600 cm−1 in the spectrum of g-C3N4 and its composites. The broad band seen in the composite spectrum, which is around 3184 cm−1 and 3400 cm−1, is attributed to the N-H and O-H stretching vibration, respectively. Compared with g-C3N4, a small hump was observed at ~1408 cm⁻¹ due to the bending vibrations of C-O and C=O [18]. The peaks at 530 cm⁻¹ correspond to the bending vibration of Bi₂O₃. The broad transmittance peak positioned at 3400 cm−1 is attributed to the surface adsorbed water molecule.
The UV-Vis-DRS spectrum is used to analyse the optical absorption behaviour of the prepared nanoparticles. The absorption spectrum for pure and 1:1 nanocomposites is shown in Figure 5a. The pure g-C3N4 shows an absorption cutoff wavelength around 450 nm, whereas the pure Bi2O3 nanoparticle is absorbed at 480 nm. From the UV-Vis absorption spectrum, it is clear that the pure and composite photocatalyst can be activated by visible light irradiation. The visible-light absorption behaviour of g-C3N4 reflects the n-π* transition of the nitrogen atom lone pair of electrons in the heptazine unit [21]. It is noticed that the absorption wavelength of the 1:1 nanocomposite exhibits red shift compared to pure g-C3N4 and shows blue shift compared with Bi2O3. The blue shift of the nanocomposite compared to Bi2O3 may reflect the existence of the quantum confinement effect or interaction of low bandgap materials (g-C3N4). However, the nanocomposite shows an absorption cutoff wavelength of around 470 nm, which is comparatively higher than bare g-C3N4. This optical absorption enhancement reveals that the interfacial interaction between the nanoparticles helps to improve the light utilisation behaviour, and it can be helpful to increase the photocatalytic efficiency. The bandgap of the pure nanoparticles is calculated by Tauc’s plot and is shown in Figure 5. The calculated bandgap of g-C3N4 and Bi2O3 is 2.91 eV and 2.78 eV, respectively.
Further, the PL analysis was conducted to determine the charge separation efficiency of the nanocomposite for effective photocatalytic activity. The emission spectrum for g-C3N4, Bi2O3 and 1:1 ratio nanocomposites is shown in Figure 5b. When compared to pure g-C3N4 and Bi2O3, the nanocomposite’s PL emission intensity was significantly lower; this indicates that the photo-excited electron–hole pairs are separated for the photocatalytic degradation reaction. Thus, it is confirmed that the nanocomposite is an effective visible-light active material for a photocatalytic reaction by optical analyses.

Evaluation of Photocatalytic Activity

The degradation of MB dye under ambient LED light irradiation allowed us to examine the photocatalytic activity of the as-prepared samples. Figure 6 shows the relationship between time and dye concentration at various times (t). Figure 6a makes it clear that the breakdown of MB dyes by the photolysis process exhibits minimal efficiency in the absence of catalysts. The MB dye degradation is increased in nanocomposites compared to pure photocatalysts. Pure g-C3N4 and Bi2O3 nanoparticles show relatively lower degradation efficiency, which reflects the lower charge separation behaviour and poor visible-light utilization behaviour. The nanocomposites show higher degradation efficiency compared to pure samples; in particular, the 1:1 nanocomposite exhibits maximum degradation efficiency compared to any other photocatalysts. The 1:1 nanocomposite achieved 91.2% efficiency after 120 min of visible-light irradiation. This efficiency is about 4.8- and 2.85-fold higher than g-C3N4 and Bi2O3 photocatalysts, respectively. The recorded MB degradation efficiency of 1:3 and 3:1 nanocomposite is 46.8% and 64.2%, respectively. The improved MB dye degradation efficiency demonstrates that the interfacial interaction between the nanoparticles helps to improve the light utilization efficiency and helps to separate the photo-induced electron–hole pairs.
The following relation was used to compute the dye degradation process’ rate constant [22,23].
ln (C/C0) = −kt
where C0 represents the dye solution concentration at time t = 0, C represents the dye solution concentration at time ‘t’ and k signifies the rate constant.
The calculated rate constant (Figure 6b) of the 1:1 nanocomposite is 0.016 min−1 for MB dye degradation after 120 min of LED light irradiation. The “k” value for MB dye degradation using g-C3N4 and Bi2O3 is 0.001 min−1 and 0.003 min−1, respectively.
Therefore, it can be inferred from the degradation process that the binary composite performs best in visible light when compared to pure catalysts. The heterojunction between g-C3N4 and Bi2O3, which functions as an electron acceptor to enhance the passage of electrons and effectively inhibits the photo exciton recombination, is what gives the binary composite its improved photocatalytic performance. As a result, the charge carrier recombination effect was significantly diminished, and the photo-induced electron hole pairs were separated. The heterojunctional g-C3N4/Bi2O3 photocatalyst would therefore be a promising candidate for the oxidation of organic dyes when exposed to visible light.
In addition, an elemental trapping experiment using scavengers helped to identify the active radical responsible for the MB dye degradation process. We employed benzoquinone (BQ), triethanolamine (TEOA) and isopropanol (IPA) as scavengers for superoxide, holes and hydroxyl radicals. Figure 6c, showing the trapping experimental plot, demonstrates that the addition of IPA considerably reduced the dye degradation process and decreased the degradation efficiency of binary composites. Efficiency can be reduced by up to 29% when using a hydroxyl radical scavenger. The degradation reaction’s efficiency was slightly altered by the addition of BQ. Therefore, it was determined from the scavenger trapping experiment that the *OH radicals are the main active ingredients in the breakdown of methylene blue. The obtained scavenging data indicate that reactive species played an order of OH* > h+ > O2* function in the photocatalytic degradation of MB.
Further, the recycling test was conducted using the 1:1 Bi2O3/g-C3N4 nanocomposite for MB degradation under LED illumination. The sample was recovered by the centrifugal process and washed with ethanol/water several times after each cycle. Then, the sample was dried in a hot air oven for further recycling. To estimate the stability of the as-prepared nanocomposite, four successive recycling tests were conducted, and the results are shown in Figure 6d. From the graph, it is clear that the MB dye degradation efficiency is slightly decreased with increasing the recycle count. However the as-prepared nanocomposite exhibited 83.4% of MB dye degradation efficiency under visible-light irradiation. The decreased degradation efficiency is attributed to the loss of photocatalysts during the recycling process or active site reduction in every recycling process. However, the 1:1 nanocomposite exhibits enough stability for the recycling process of MB dye degradation. To highlight the presented work, a comparison table was plotted, as shown in Table 1.
Furthermore, the degradation experiment was conducted under different environmental conditions to elucidate the behaviour of the prepared nanocomposite, and the results are shown in Figure 7. Initially, MB dye degradation experiments were conducted under different solution pH and are shown in Figure 7a. The solution pH was adjusted with the help of diluted HCl solution and diluted NH3 solution. The prepared nanocomposite degraded the MB dye up to 93.8% after 120 min of LED irradiation under alkaline conditions. This is slightly higher than the efficiency obtained under neutral pH conditions. This enhancement is attributed to formation of more superoxide radicals under alkaline conditions compared to neutral or acidic conditions.
Industrial wastewater contains different concentrations of organic pollutants; therefore, it is necessary to study the influence of dye concentration in photocatalytic degradation reaction. The photocatalytic experiment was conducted using different concentrations of dye solution with neutral pH conditions and 50 mg of photocatalyst. The result clearly shows that the MB degradation decreases with increasing dye concentration. This result reveals that a higher concentration of dye molecules may be adsorbed on the photocatalyst surface, so that it can reduce the active sites of the photocatalyst or limit the visible-light utilization.
Furthermore, the dye degradation study was extended for different concentrations of photocatalyst, and the results are shown in Figure 7c. The MB dye degradation efficiency is increased with increasing the catalyst dosage, and it attains 98.9% efficiency at 100 mg of photocatalyst concentration. The increasing MB dye degradation efficiency is attributed to the availability of more active sites for the degradation reaction.
Based on the experimental results, it is necessary to elucidate the possible MB dye degradation mechanism using the Bi2O3/g-C3N4 nanocomposite. The possible mechanism is illustrated in Figure 8. The band potential of the g-C3N4 and Bi2O3 nanoparticle was calculated using the following relation [3]:
ECB = χ − Ee − 0.5Eg
EVB = ECB + Eg
where χ signifies the electronegativity, Ee represents the energy of the free electron in the hydrogen scale, and Eg denotes the bandgap energy. ECB and EVB reflect the energy level of the conduction band and valance band, respectively. The calculated valance band potential of g-C3N4 and Bi2O3 is 1.75 eV and 3.11 eV, respectively. The conduction band of the nanoparticle is −1.16 eV for g-C3N4 and 0.33 eV for Bi2O3.
According to the band potential of the nanoparticles, bandgap energy and scavengers’ test, a possible MB dye degradation mechanism was proposed, and the schematic of the heterojunction charge transfer is shown in Figure 8. From the band potential values, it can be clearly noticed that the VB potential of Bi2O3 and CB potential of g-C3N4 is adequate to execute the oxidation and reduction reaction, respectively. The VB of Bi2O3 has a higher value, oxidizing the water molecule into •OH radicals, compared to the standard oxidation (1.98 V vs. NHE) potential. The CB potential of g-C3N4 has higher reduction potential to reduce the oxygen molecule into superoxide radicals in comparison with standard reduction potential (−0.33 V vs. NHE) [24]. However, the CB of Bi2O3 and VB of g-C3N4 are not suitable for reduction and oxidation reaction due to lesser band potential compared to the standard redox potential values. Therefore, it not possible to generate the superoxide radicals and hydroxyl radicals at CB of Bi2O3 and VB of Bi2O3, respectively. The conventional type-Ⅱ heterojunction is not suitable for demonstrating the MB dye degradation mechanism of the Bi2O3/g-C3N4 nanocomposite due to inconsistency with the scavengers’ result. Therefore the reaction mechanism is described by the Z-scheme process with the help of the trapping experiment and literature reports. Both of the semiconductors are visible-light active materials and can produce electron–hole pairs during the light irradiation. According to the Z-scheme mechanism, the excited conduction band electrons from the Bi2O3 are transferred to and combined with the valance band holes of the g-C3N4 through the interfacial contact. Thereafter, the combined electrons are again excited to the CB of g-C3N4 during the light incident. In this way, the electrons and holes are separated for photocatalytic reaction. Therefore, the accumulated CB electrons at g-C3N4 and holes at VB of Bi2O3 generate the superoxide radicals and hydroxyl radicals for the degradation reaction. This element of the Z-scheme mechanism is well matched with the scavengers’ experiment and the literature reports [23,24,25,26]. The possible charge transport and degradation mechanism of the Bi2O3/g-C3N4 nanocomposite is given below [1,2,3,4]:
Bi2O3 + hυ → e (Bi2O3) + h+(Bi2O3)
g-C3N4 + hυ → e (g-C3N4) + h+(g-C3N4)
e (g-C3N4) + O2 → •O2
•O2 + e + 2H+ → H2O2
•O2 + H2O2 → •OH + OH + O2
2H2O2 → 2•OH
h+ (Bi2O3) + OH-/H2O → Bi2O3 + •OH
•O2/•OH + dye → Intermediate product → CO2 + H2O

3. Materials and Methods

Melamine was directly polymerized at 500 °C for 4 h and 520 °C for 2 h in order to prepare g-C3N4. A stoichiometric quantity of Bi(NO3) and melamine was first dissolved in 10 mL CH3OH to synthesis the Bi2O3/g-C3N4 nanocomposite. After stirring for two hours, the homogenous suspension was dried at 100 °C. The dry mixture was then ground and annealed for 4 h at a rate of 5 °C/min at 500 °C and 520 °C for 2 h. The sample procedure was followed to prepare pure Bi2O3. The sample was then designated as Bi2O3/g-C3N4 x:x, where x is the mass of Bi2O3 added to g-C3N4. By changing the mass of Bi2O3 under the same reaction conditions, samples of 1:3, 1:1 and 3:1 Bi2O3 to g-C3N4 were created [27,28].

Photocatalysis Experiments

In a quartz reactor, the samples’ photocatalytic activity was tested on the degradation of MB dye. First, the as-prepared photocatalyst (50 mg) was added to 50 mL of MB (20 mg/L) dye solution. A water circulation system was used to maintain the solution’s constant stirring in a magnetic stirrer at a reaction temperature of 27 to 30 °C in order to achieve homogenous dispersion. Prior to irradiation, the dye and photocatalyst suspension was constantly agitated for 30 min in the absence of radiation to achieve adsorption and desorption equilibrium. Due to its high-intensity emission in the visible region, the commercial 40 W LED bulb was employed as a source of visible light. The average LED light intensity that was observed was around 37,032 lux. Samples were removed and evaluated using UV-Spectroscopy at 664 nm for MB, which corresponds to the maximal absorption wavelength of dyes, at regular intervals [29]. Under LED illumination, a trapping experiment using various scavengers was conducted to assess the involvement of reactive spices in the degrading process of dye. As scavengers for superoxide anion, hole (h+), hydroxyl (•OH) radicals, 0.1 mmol of benzoquinone, 0.1 mmol of ethylenediaminetetraacetic acid, and 5 mL of isopropanol were utilised [30].

4. Conclusions

In conclusion, the binary heterojunction was synthesised using a straightforward technique. The XRD result clearly shows that the crystallite size of the Bi2O3 nanoparticles was reduced, while the heterojunction formation enables strong optical utilization behaviour compared to the pure samples. Further, the interaction between the nanoparticles helps to increase the electron–hole separation, which promotes higher MB dye degradation efficiency. Among all the as-synthesised samples, the 1:1 ratio of the Bi2O3/g-C3N4 nanocomposite exhibits the highest degradation efficiency after 120 min of LED light illumination. The synergistic effect between g-C3N4 and Bi2O3, which actively slows down the rate of charge carrier recombination and has a broad absorption spectrum for visible light, was responsible for the improved photocatalytic performance. As a result, the heterojunction photocatalyst degraded the dye molecule through the Z-scheme charge transfer process, which was supported by the scavengers’ experiment and band potential calculation. Hence, this simple method can be helpful to for the large-scare preparation of photocatalyst with higher efficiency.

Author Contributions

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

Funding

This research was funded by Deanship of Scientific Research at King Khalid University through a Research Groups Program under Grant No. R.G.P. 2/215/43.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time as the data also form part of an ongoing study.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through a Research Groups Program under Grant No. R.G.P. 2/215/43.

Conflicts of Interest

The authors declare that there is no conflict of interest.

References

  1. Bhuvaneswari, K.; Bharathi, R.D.; Pazhanivel, T. Silk fibroin linked Zn/Cd-doped SnO2 nanoparticles to purify the organically polluted water. Mater. Res. Express. 2018, 5, 024004. [Google Scholar] [CrossRef]
  2. Mukhtar, F.; Munawar, T.; Nadeem, M.S.; Rehman, M.N.U.; Riaz, M.; Iqbal, F. Dual S-scheme heterojunction ZnO–V2O5–WO3 nanocomposite with enhanced photocatalytic and antimicrobial activity. Mater. Chem. Phys. 2021, 263, 124372. [Google Scholar] [CrossRef]
  3. Mukhtar, F.; Munawar, T.; Nadeem, M.S.; Rehman, M.N.U.; Khan, S.A.; Koc, M.; Batool, S.; Hasan, M.; Iqbal, F. Dual Z-scheme core-shell PANI-CeO2-Fe2O3-NiO heterostructured nanocomposite for dyes remediation under sunlight and bacterial disinfection. Environ. Res. 2022, 215, 114140. [Google Scholar] [CrossRef]
  4. Mukhtar, F.; Munawar, T.; Nadeem, M.S.; Khan, S.A.; Koc, M.; Batool, S.; Hasan, M.; Iqbal, F. Enhanced sunlight-absorption of Fe2O3 covered by PANI for the photodegradation of organic pollutants and antimicrobial inactivation. Adv. Powder Technol. 2022, 33, 103708. [Google Scholar] [CrossRef]
  5. Bhuvaneswari, K.; Palanisamy, G.; Pazhanivel, T.; Bharathi, G.; Nataraj, D. Photocatalytic Performance on Visible Light Induced ZnS QDs-MgAl Layered Double Hydroxides Hybrids for Methylene Blue Dye Degradation. Chem. Sel. 2018, 3, 13419–13426. [Google Scholar] [CrossRef]
  6. Kalisamy, P.; Lallimathi, M.; Suryamathi, M.; Palanivel, B.; Venkatachalam, M. ZnO-embedded S-doped g-C3N4 heterojunction: Mediator-free Z-scheme mechanism for enhanced charge separation and photocatalytic degradation. RSC Adv. 2020, 10, 28365–28375. [Google Scholar] [CrossRef] [PubMed]
  7. Palanivel, B.; Mudisoodum Perumal, S.; Maiyalagan, T.; Jayaraman, 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]
  8. Palanivel, B.; Ayappan, C.; Jayaraman, V.; Chidambaram, S.; Maheshwaran, R.; Alagiri, M. Inverse spinel NiFe2O4 deposited g-C3N4 nanosheet for enhanced visible light photocatalytic activity. Mater. Sci. Semicond. 2019, 100, 87–97. [Google Scholar] [CrossRef]
  9. Li, B.; Nengzi, L.C.; Guo, R.; Cui, Y.; Zhang, Y.; Cheng, X. Novel synthesis of Z-scheme α-Bi2O3/g-C3N4 composite photocatalyst and its enhanced visible light photocatalytic performance: Influence of calcination temperature. Chin. Chem. Lett. 2020, 31, 2705–2711. [Google Scholar] [CrossRef]
  10. Fan, G.; Ma, Z.; Li, X.; Deng, L. Coupling of Bi2O3 nanoparticles with g-C3N4 for enhanced photocatalytic degradation of methylene blue. Ceram. Int. 2021, 47, 5758–5766. [Google Scholar] [CrossRef]
  11. Palanive, B.; Alagiri, M. Construction of rGO supported integrative NiFe2O4/g-C3N4 nanocomposite: Role of charge transfer for boosting the OH• radical production to enhance the photo-Fenton degradation. ChemistrySelect 2020, 5, 9765–9775. [Google Scholar] [CrossRef]
  12. Matheswaran, P.; Thangavelu, P.; Palanivel, B. Carbon dot sensitized integrative g-C3N4/AgCl Hybrids: An synergetic interaction for enhanced visible light driven photocatalytic process. Adv. Powder Technol. 2019, 30, 1715–1723. [Google Scholar] [CrossRef]
  13. Matheswaran, P.; Karuppiah, P.; Chen, S.M.; Thangavelu, P.; Ganapathi, B. Fabrication of g-C3N4 Nanomesh-Anchored Amorphous NiCoP2O7: Tuned Cycling Life and the Dynamic Behavior of a Hybrid Capacitor. ACS Omega 2018, 3, 18694–18704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. He, R.; Zhou, J.; Fu, H.; Zhang, S.; Jiang, C. Room-temperature in situ fabrication of Bi2O3/g-C3N4 direct Z-scheme photocatalyst with enhanced photocatalytic activity. Appl. Surf. Sci. 2018, 430, 273–282. [Google Scholar] [CrossRef]
  15. Zhang, J.; Hu, Y.; Jiang, X.; Chen, S.; Meng, S.; Fu, X. Design of a direct Z-scheme photocatalyst: Preparation and characterization of Bi2O3/g-C3N4 with high visible light activity. J. Hazard. Mater. 2014, 280, 713–722. [Google Scholar] [CrossRef]
  16. Zhang, J.; Qian, H.; Liu, W.; Chen, H.; Qu, Y.; Lin, Z. The construction of the heterostructural Bi2O3/g-C3N4 composites with an enhanced photocatalytic activity. Nano 2018, 13, 1–9. [Google Scholar] [CrossRef]
  17. Liu, S.; Chen, J.; Xu, D.; Zhang, X.; Shen, M. Enhanced photocatalytic activity of direct Z-scheme Bi2O3/g-C3N4 composites via facile one-step fabrication. J. Mater. Res. 2018, 33, 1391–1400. [Google Scholar] [CrossRef]
  18. Zhang, L.; Wang, G.; Xiong, Z.; Tang, H.; Jiang, C. Fabrication of flower-like direct Z-scheme β-Bi2O3/g-C3N4 photocatalyst with enhanced visible light photoactivity for Rhodamine B degradation. Appl. Surf. Sci. 2018, 436, 162–171. [Google Scholar] [CrossRef]
  19. Cui, Y.; Zhang, X.; Guo, R.; Zhang, H.; Li, B.; Xie, M.; Cheng, Q.; Cheng, X. Construction of Bi2O3/g-C3N4 composite photocatalyst and its enhanced visible light photocatalytic performance and mechanism. Sep. Purif. Technol. 2018, 203, 301–309. [Google Scholar] [CrossRef]
  20. Zhu, Z.; Huo, P.; Lu, Z.; Yan, Y.; Liu, Z.; Shi, W.; Li, C.; Dong, H. Fabrication of magnetically recoverable photocatalysts using g-C3N4 for effective separation of charge carriers through like-Z-scheme mechanism with Fe3O4 mediator. Chem. Eng. J. 2018, 331, 615–625. [Google Scholar] [CrossRef]
  21. Palanivel, B.; Shkir, M.; Alshahrani, T.; Mani, A. Novel NiFe2O4 deposited S-doped g-C3N4 nanorod: Visible-light-driven heterojunction for photo-Fenton like tetracycline degradation. Diam. Relat. Mater. 2021, 112, 108148. [Google Scholar] [CrossRef]
  22. De Almeida, M.F.; Bellato, C.R.; Mounteer, A.H.; Ferreira, S.O.; Milagres, J.L.; Miranda, L.D.L. Enhanced photocatalytic activity of TiO2—Impregnated with MgZnAl mixed oxides obtained from layered double hydroxides for phenol degradation. Appl. Surf. Sci. 2015, 357, 1765–1775. [Google Scholar] [CrossRef]
  23. Miao, X.; Ji, Z.; Wu, J.; Shen, X.; Wang, J.; Kong, L.; Liu, M.; Song, C. g-C3N4/AgBr nanocomposite decorated with carbon dots as a highly efficient visible-light-driven photocatalyst. J. Colloid Interface Sci. 2017, 502, 24–32. [Google Scholar] [CrossRef] [PubMed]
  24. Palanivel, B.; Mani, A. Conversion of a Type-II to a Z-Scheme heterojunction by intercalation of a 0D electron mediator between the integrative NiFe2O4/g-C3N4 composite nanoparticles: Boosting the radical production for photo-Fenton degradation. ACS Omega 2020, 5, 19747–19759. [Google Scholar] [CrossRef] [PubMed]
  25. Palanivel, B.; Jayaraman, V.; Ayyappan, C.; Alagiri, M. Magnetic binary metal oxide intercalated g-C3N4: Energy band tuned pn heterojunction towards Z-scheme photo-Fenton phenol reduction and mixed dye degradation. J. Water Process Eng. 2019, 32, 100968. [Google Scholar] [CrossRef]
  26. Palanivel, B.; Hossain, M.S.; Macadangdang, R.R., Jr.; Ayappan, C.; Krishnan, V.; Marnadu, R.; Kalaivani, T.; Alharthi, F.A.; Sreedevi, G. Activation of persulfate for improved naproxen degradation using FeCo2O4@g-C3N4 heterojunction photocatalysts. ACS Omega 2021, 6, 34563–34571. [Google Scholar] [CrossRef]
  27. Yan, X.; Xu, R.; Guo, J.; Cai, X.; Chen, D.; Huang, L.; Xiong, Y.; Tan, S. Enhanced photocatalytic activity of Cu2O/g-C3N4 heterojunction coupled with reduced graphene oxide three-dimensional aerogel photocatalysis. Mater. Res. Bull. 2017, 96, 18–27. [Google Scholar] [CrossRef]
  28. Wang, L.; Zhu, Z.; Wang, F.; Qi, Y.; Zhang, W.; Wang, C. State-of-the-art and prospects of Zn-containing layered double hydroxides (Zn-LDH)-based materials for photocatalytic water remediation. Chemosphere 2021, 278, 130367. [Google Scholar] [CrossRef]
  29. Liu, X.; Ji, X.; Liu, M.; Liu, N.; Tao, Z.; Dai, Q.; Wei, L.; Li, C.; Zhang, X.; Wang, B. High-performance ge quantum dot decorated graphene/zinc-oxide heterostructure infrared photodetector. ACS Appl. Mater. Interfaces 2015, 7, 2452–2458. [Google Scholar] [CrossRef]
  30. George, A.; Raj, D.M.A.; Raj, A.D.; Nguyen, B.-S.; Phan, T.-P.; Pazhanivel, T.; Sivashanmugan, K.; Josephine, R.; Irudayaraj, A.A.; Arumugam, J.; et al. Morphologically tailored CuO nanostructures toward visible-light-driven photocatalysis. Mater. Lett. 2020, 281, 128603. [Google Scholar] [CrossRef]
Figure 1. XRD pattern of (a,c) pristine g-C₃N₄ and Bi₂O₃; (b,d,e) different weight percentages of Bi₂O₃/g-C₃N₄ nanocomposite.
Figure 1. XRD pattern of (a,c) pristine g-C₃N₄ and Bi₂O₃; (b,d,e) different weight percentages of Bi₂O₃/g-C₃N₄ nanocomposite.
Catalysts 12 01544 g001
Figure 2. SEM image of (a) g-C3N4, (b) Bi2O3, (c) 1:1 Bi2O3/g-C3N4; (d) EDS spectrum/elemental mapping of 1:1 Bi2O3/g-C3N4 nanocomposite.
Figure 2. SEM image of (a) g-C3N4, (b) Bi2O3, (c) 1:1 Bi2O3/g-C3N4; (d) EDS spectrum/elemental mapping of 1:1 Bi2O3/g-C3N4 nanocomposite.
Catalysts 12 01544 g002
Figure 3. TEM (a), HRTEM (b) and SAED (c) images of 1:1 Bi2O3/g-C3N4 nanocomposite.
Figure 3. TEM (a), HRTEM (b) and SAED (c) images of 1:1 Bi2O3/g-C3N4 nanocomposite.
Catalysts 12 01544 g003
Figure 4. FTIR spectra of prepared samples.
Figure 4. FTIR spectra of prepared samples.
Catalysts 12 01544 g004
Figure 5. (a) UV-Vis-DRS spectra and (b) PL emission spectra.
Figure 5. (a) UV-Vis-DRS spectra and (b) PL emission spectra.
Catalysts 12 01544 g005
Figure 6. (a) Degradation plot, (b) pseudo 1st order kinetics, (c) scavenger’s test and (d) recycling test.
Figure 6. (a) Degradation plot, (b) pseudo 1st order kinetics, (c) scavenger’s test and (d) recycling test.
Catalysts 12 01544 g006
Figure 7. MB dye degradation under different conditions: (a) solution pH, (b) dye concentration and (c) catalyst dosage.
Figure 7. MB dye degradation under different conditions: (a) solution pH, (b) dye concentration and (c) catalyst dosage.
Catalysts 12 01544 g007
Figure 8. Possible reaction mechanism.
Figure 8. Possible reaction mechanism.
Catalysts 12 01544 g008
Table 1. Comparison table.
Table 1. Comparison table.
No.CatalystPollutant (ppm)Irradiation Time (min)Efficiency (%)Reference
1ZnO-V2O5-WO3MB (10 ppm)80 min99.8%[2]
2PANI-CeO2-Fe2O3-NiOMB (5 ppm)50 min99%[3]
3Bi2O3/g-C3N4Phenol (10 ppm)150 min>70%[14]
4Bi2O3/g-C3N4MB (10 ppm)120 min>90%[15]
5Bi2O3/g-C3N4Rh B (10 ppm)80 min98%[18]
6Bi2O3/g-C3N4Amido Black (10 ppm)120 min88.71%[19]
7NiFe2O4/CD/g-C3N4Rh B (10 ppm)120 min97%[21]
8Bi2O3/g-C3N4MB (10 ppm)120 min91.2%This 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

Mahalakshmi, K.; Ranjith, R.; Thangavelu, P.; Priyadharshini, M.; Palanivel, B.; Manthrammel, M.A.; Shkir, M.; Diravidamani, B. Augmenting the Photocatalytic Performance of Direct Z-Scheme Bi2O3/g-C3N4 Nanocomposite. Catalysts 2022, 12, 1544. https://doi.org/10.3390/catal12121544

AMA Style

Mahalakshmi K, Ranjith R, Thangavelu P, Priyadharshini M, Palanivel B, Manthrammel MA, Shkir M, Diravidamani B. Augmenting the Photocatalytic Performance of Direct Z-Scheme Bi2O3/g-C3N4 Nanocomposite. Catalysts. 2022; 12(12):1544. https://doi.org/10.3390/catal12121544

Chicago/Turabian Style

Mahalakshmi, Krishnasamy, Rajendran Ranjith, Pazhanivel Thangavelu, Matheshwaran Priyadharshini, Baskaran Palanivel, Mohamed Aslam Manthrammel, Mohd Shkir, and Barathi Diravidamani. 2022. "Augmenting the Photocatalytic Performance of Direct Z-Scheme Bi2O3/g-C3N4 Nanocomposite" Catalysts 12, no. 12: 1544. https://doi.org/10.3390/catal12121544

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