Next Article in Journal
Failure Analysis of a C-276 Alloy Pipe in a Controlled Decomposition Reactor
Previous Article in Journal
Atomic Force Microscopy Nanoindentation Method on Collagen Fibrils
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrication and Characterization of Highly Efficient As-Synthesized WO3/Graphitic-C3N4 Nanocomposite for Photocatalytic Degradation of Organic Compounds

by
Mai S. A. Hussien
1,2,
Abdelfatteh Bouzidi
3,
Hisham S. M. Abd-Rabboh
4,5,
Ibrahim S. Yahia
6,7,8,
Heba Y. Zahran
6,7,8,
Mohamed Sh. Abdel-wahab
9,*,
Walaa Alharbi
10,
Nasser S. Awwad
4 and
Medhat A. Ibrahim
11,12
1
Department of Chemistry, Faculty of Education, Ain Shams University, Roxy, Cairo 11757, Egypt
2
Nanoscience Laboratory for Environmental and Bio-Medical Applications (NLEBA), Department of Physics, Faculty of Education, Ain Shams University, Roxy, Cairo 11757, Egypt
3
Laboratory of Materials for Energy and Environment, and Modeling (LMEEM), Faculty of Sciences of Sfax, University of Sfax, B.P. 1171, Sfax 3038, Tunisia
4
Department of Chemistry, Faculty of Science, King Khalid University, P.O. Box 9004, Abha 61421, Saudi Arabia
5
Department of Chemistry, Faculty of Science, Ain Shams University, Abbassia, Cairo 11566, Egypt
6
Laboratory of Nano-Smart Materials for Science and Technology (LNSMST), Department of Physics, Faculty of Science, King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia
7
Research Center for Advanced Materials Science (RCAMS), King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia
8
Semicondcuotr Laboratory, Department of Physics, Faculty of Education, Ain Shams University, Roxy, Cairo 11757, Egypt
9
Materials Science and Nanotechnology Department, Faculty of Postgraduate Studies for Advanced Sciences, Beni-Suef University, Beni-Suef 62511, Egypt
10
Department of Chemistry, Science and Arts College, Rabigh Campus, King Abdulaziz University, P.O. Box 80200, Jeddah 21589 , Saudi Arabia
11
Nanotechnology Research Centre (NTRC), The British University in Egypt (BUE), Suez Desert Road, El-Sherouk City 11837, Cairo, Egypt
12
Molecular Spectroscopy and Modeling Unit, Spectroscopy Department, National Research Centre, 33 El-Bohouth St., Dokki, Giza 12622, Egypt
*
Author to whom correspondence should be addressed.
Materials 2022, 15(7), 2482; https://doi.org/10.3390/ma15072482
Submission received: 21 February 2022 / Revised: 9 March 2022 / Accepted: 23 March 2022 / Published: 28 March 2022
(This article belongs to the Topic Catalysis: Homogeneous and Heterogeneous)

Abstract

:
The incorporation of tungsten trioxide (WO3) by various concentrations of graphitic carbon nitride (g-C3N4) was successfully studied. X-ray diffraction (XRD), Scanning Electron Microscope (SEM), and Diffused Reflectance UV-Vis techniques were applied to investigate morphological and microstructure analysis, diffused reflectance optical properties, and photocatalysis measurements of WO3/g-C3N4 photocatalyst composite organic compounds. The photocatalytic activity of incorporating WO3 into g-C3N4 composite organic compounds was evaluated by the photodegradation of both Methylene Blue (MB) dye and phenol under visible-light irradiation. Due to the high purity of the studied heterojunction composite series, no observed diffraction peaks appeared when incorporating WO3 into g-C3N4 composite organic compounds. The particle size of the prepared composite organic compound photocatalysts revealed no evident influence through the increase in WO3 atoms from the SEM characteristic. The direct and indirect bandgap were recorded for different mole ratios of WO3/g-C3N4, and indicated no apparent impact on bandgap energy with increasing WO3 content in the composite photocatalyst. The composite photocatalysts’ properties better understand their photocatalytic activity degradations. The pseudo-first-order reaction constants (K) can be calculated by examining the kinetic photocatalytic activity.

1. Introduction

The photocatalytic performance of semiconductor materials, based on oxide semiconductors in their crystalline phases, has recently become desirable for producing hydrogen and oxygen via visible-light water splitting. Designing semiconductor composites produced between two semiconductor materials necessitates crystalline phase engineering [1,2,3]. The use of heterogeneous photocatalysis to convert water into hydrogen gas is regarded as one of the most promising options for addressing global energy and pollution problems [1,4,5]. In the present century, the rapid decline in energy sources, and increased power, waste, and renewable H2 energy sources, are attributable to worldwide demand for fossil fuels. A. Fujishima et al. studied for a long time how to establish a clean and renewable photocatalytic hydrogen-production method [6]. The absorption wavelength range of semiconductor materials such as oxides, sulfides, nitrides, and solid solutions was examined under visible light, to improve the absorption wavelength range [7,8].
Heterojunction photocatalysts were used extensively to improve the separation efficiency of photoexcited electron-hole pairs. A single-phase photocatalyst has significant limitations in a photocatalytic response, because photogenerated electrons and holes are quickly mixed [9] to develop optical properties. Jiang et al. doped phosphorus nanosheets with g-C3N4 and added carbon defects, which significantly increased the rate of hydrogen evolution through the photocatalytic method [9]. Furthermore, studies report an improved hydrogen production rate through using a mineral acid or phosphoric acid etching of g-C3N4 nanosheets to increase the number of active sites [10,11].
Graphite carbon nitride (g-C3N4) has recently gained new interest as a promising material in different applications such as photocatalysts, fuel cell electrodes, light-emitting devices, and chemical sensors. g-C3N4 stands out amongst many types of photocatalysts. A polymer photocatalyst bandgap is about 2.7 eV, allowing visible light to absorb up to 460 nm. Nevertheless, g-C3N4 can absorb visible light effectively since graphitic C3N4 has an adequate conduction band under illumination conditions that is able to be more harmful than protons formed by hydrogen [12,13]. Tungsten oxide (WO3) is considered a promising material [13,14]. It has been reported that synthesis of WO3/g-C3N4 composite organic compounds shapes a heterostructural composite photocatalyst [15,16],provided that the electron donor is an aqueous solution for triethanolamine. In the color-sensitization method, we looked at the composite catalyst’s hydrogen production behavior under visible-light irradiation. g-C3N4 composite, blended with WO3 using a planetary mill, was prepared using hydrothermal treatment to improve its photocatalytic activity [15]. The photocatalytic activity incorporated WO3 in g-C3N4 composite organic compounds [17]. J. Liang et al. [18] investigated the use of crystalline phase engineering in WO3/g-C3N4 composite organic compounds to improve photocatalytic activity under visible light. They demonstrated that WO3/g-C3N4 composites with h-WO3 show better dispersion of WO3, higher charge separation, and higher photocatalytic activity through visible-light photocatalytic degradation of Rhodamine B (RhB).
In this research, incorporating WO3 into g-C3N4, composite organic compound photocatalysts were prepared, and exhibited improved photocatalytic activity under visible light. Their morphological, diffused reflectance UV-Vis optical, and (Methylene blue (MB) and phenol) photocatalytic activity properties were analyzed and discussed in detail, to understand the effects of WO3 on g-C3N4 composite organic compounds.

2. Experimental Parts

2.1. Regents

In the experiment, an analytical grade of melamine, sodium tungstate dihydrate (Na2WO4·2H2O) ACS reagent, ≥99%, and other chemicals purchased from Sigma-Aldrich were used. These chemicals did not need further purification as they were of analytical grade. De-ionized water was utilized to avoid contamination.

2.2. Synthesis of Pure g-C3N4 and WO3/g-C3N4 Composite Organic Compounds

Simple g-C3N4 synthesis involves melamine pyrolysis prepared in an air atmosphere. Initially, 8 g of melamine was ground inside a crucible, then heated at a rate of 5 °C per min until it attained 550 °C, and then adjusted at this temperature for 2 h. Afterwards, the samples were left to cool down to ambient temperature. In the mortar, the resultant powder was then ground with a pestle. Following the first stage, Na2WO4·2H2O was added to melamine to synthesize the doped WO3/g-C3N4 composite organic compounds, after which the procedure for pure g-C3N4 was repeated.

2.3. Characterization Techniques

The effect of WO3 on the physical structure of the g-C3N4 composite organic compounds for the 13 structures was characterized by X-ray diffraction. Shimadzu X-ray Diffractometer (XRD-6000 Series) was utilized with the standard copper X-ray tube at 30 kV and 30 mA with a wavelength equal to 1.5406 Å. The surface structure of all samples was evaluated by scanning electron microscopy (SEM-JSM6360 Series, with an acceleration voltage = 20 kV).
The effect of WO3 on g-C3N4 nanocomposite organic compounds was tested employing a UV-vis-NIR spectrophotometer model (Shimadzu UV-3600) with diffused reflectance in a wide range between 200 and 800 nm. The device was designed with the BaSO4 built-in sphere attachment as reference material. To counteract the WO3/g-C3N4 composites within the holder, as a guide for the BaSO4 and the other holder, a specifically constructed holder attached to the integrating sphere device was employed. The thickness of the holder corresponds to the thickness of the WO3/g-C3N4 nanocomposites. To track the photo-removal phase, a UV-visible spectrophotometer was used.
The diffused reflectance was measured in ambient conditions using a JASCO UV-Vis-NIR-V-570 double beam spectrophotometer.

2.4. Photocatalytic Measurements

A simple wooden photoreactor tested the photocatalytic activity of WO3/g-C3N4 nanocomposites under visible-light spectrum radiation of samples, using both methylene blue (MB) and aqueous phenol as a pollutant example in wastewater. I.S. Yahia and his group designed the photoreactor in NLEBA, Ain Shams University (ASU), Egypt, consisting of two parts: the outer part was in the form of a box made from wood (height: 100 cm, width: 65 cm); the inner part had seven visible lamps (18 W, 60 cm length, 425 to 600 nm), and could be separately controlled. Each set of nanopowders was placed in a beaker containing 50 mL of MB (20 mg/L) (i.e., for the phenol photodegradation). The sample was magnetically stirred in the dark system to reach equilibrium between dye and photocatalyst. The sample was removed from the solution after a specific time (every 15 min), and the amount of sample remaining was then exposed to visible light. A UV-Vis spectrophotometer was used to analyze the samples.

3. Results and Discussions

3.1. Structural XRD Measurements

The crystalline structure phase was analyzed with the XRD technique to reveal the effect of doping WO3 in the final products. XRD patterns are shown in Figure 1 for pure g-C3N4 and its doping with WO3 on the g-C3N4 composite organic compounds, with various WO3 effects. As visualized in Figure 1, the g-C3N4 powder sample highlights two peaks at 2θ = 13.1°, and 2θ = 27.4°, which can be indexed to a small peak, (100) plane, pertinent to the in-plane structural packing of graphitic material, and a sharp peak, (002) plane, assigned to the interlayer spacing of the conjugated aromatic system. The (002) peak indicates a higher-density packing of the g-C3N4 molecules. This corresponds to the characteristic interplanar staking peaks of aromatic systems and the inter-layer structural packing, respectively [19,20]. These results are compared with Mo et al. [21]. The peak at = 27.4° (JCPDS 87-1526) confirms the formation of hexagonal phase g-C3N4 powders [18,20,21]. The introduction of doping WO3 on g-C3N4 composite organic compounds reduces the maximum intensity of peaks. In Figure 1, we notice that only two peaks appear in the entirety of the prepared samples related to pure g-C3N4 except for the last sample, noted as 0.5 g WO3-doped g-C3N4, showing the other two peaks related to WO3. The two peaks related to WO3 appear at = 16.9° and 32.51°, related to (101) and (022), according to JCPDS 01-083-0950.
The average crystallite size for the prepared samples was measured using the Debye–Scherrer formula, as follows [22]:
D = 0.94 λ / β cos θ
The calculated crystallite size D depends on the broadening diffraction peak β, the diffraction angle θ , and the X-ray wavelength λ. Both dislocation density δ and lattice strain ε were calculated using the following equations [23,24]:
δ = n / D 2
ε = β cos θ / 4
The XRD structural parameters for the prepared samples summarized in Table 1 show that the average crystallite size for the pure g-C3N4 is 39.17 nm, and this increases from 35.91, 35.93, 40.84, 44.83, to 81.67 nm for the 0.5, 0.1, 0.05, 0.01, and 0.001 WO3-doped g-C3N4 composites, respectively.
The lower dislocation density values reflect the higher quality of the prepared samples.

3.2. Morphologies and Microstructure Analysis

To assess the effect of WO3 in g-C3N4 composite organic photocatalyst compounds, the morphology and microstructure of the pure g-C3N4 and its WO3-doped g-C3N4 composite organic compounds was studied by using SEM image analysis. Figure 2a–f presents the SEM images of the investigated composite photocatalysts. The morphology of pure g-C3N4 exhibited apparent granular aggregates and a typical sheet that consisted of small particles created from some irregular particles. Figure 2c–f shows that the WO3 particles were attached to the surface of the sheet g-C3N4. The particle sizes obtained from SEM images of the investigated composite organic compounds were gathered in Table 2. The particle sizes varied between 1.65 and 1.51 µm. As the WO3 doping particle contents increased, the g-C3N4 composite photocatalyst had no noticeable influence on the obtained particle size. A similar result is reported by X. Chu et al. [25].

3.3. Optical Properties of WO3 Doped g-C3N4 Nanocomposites

Figure 3a shows the optical UV-Vis diffused reflectance spectroscopy study to identify the bandgap energy of pure g-C3N4 and WO3-doped g-C3N4 composite organic photocatalyst compounds. The optical UV-Vis diffused reflectance spectra for other WO3/g-C3N4 photocatalyst composite samples should be the superimposed signals of g-C3N4. All diffused reflection spectra increased with increasing wavelengths up to 400 nm. The absorption edge spectra start to redshift towards the visible region, allocated to the intrinsic g-C3N4 bandgap [25].
The redshift suggested that the photocatalyst composite compounds could use sunlight and produce more electron-hole pairs, which would help the photocatalytic response. The above findings may be caused by the interactions between the WO3 and g-C3N4 incorporated into the heterojunction materials. Figure 3b,c displays the direct and indirect energy bandgap of pure g-C3N4 and its WO3 atom-doped g-C3N4 composite organic photocatalyst compounds, which can be estimated from the plots of (ahυ)2 and (ahυ)1/2 vs. the incident photon energy () using Tauc’s formula [26]. The Kubelka–Munk function and its related absorption coefficient (α) are as follows [27,28,29]:
A = (F(R) hυ/d) = A(hυEg)r
where A is a pre-factor, Eg is the energy bandgap, and r limits the transition band type. When r = 2 is related to the indirect allowed band, r = 1/2 is associated with the direct allowed band. The bandgap of the photocatalyst was determined by Tauc’s plot, where (αhυ)1/2 with axis (i.e., the linear portion of the plots (αhυ)2 with axis). By drawing a tangent line of each curve, each photocatalyst composite material’s bandgap was obtained from the intercept axis. Therefore, the estimated bandgaps of pure g-C3N4 and its WO3 atom-doped g-C3N4 composite organic photocatalyst compounds, with various WO3 content, confirmed that the direct and indirect bandgaps for g-C3N4 were 2.83 [30] and 2.56 eV, respectively. This result was also reported by J.Y. Tai et al. [31]. For the indirect bandgap, we noted a slight change from 2.56 eV for the pure g-C3N4 changing to 2.6, 2.65, 2.66, 2.86, to 2.71 eV for the 0.5, 0.1, 0.05, 0.01, and 0.001 WO3-doped g-C3N4 composite, respectively. For the direct bandgap, a small change was also noted, from 2.83 eV for the pure g-C3N4 changing to 2.86. 2.88, 2.89, 2.90, 2.92 eV for the 0.5, 0.1, 0.05, 0.01, and 0.001 WO3-doped g-C3N4 composite, respectively. This indicated that the increase in WO3 content in the composite photocatalyst had no evident influence on bandgap energy. However, it was found that the movement and lifetime of the photogenerated charge carriers in g-C3N4 could be significantly affected by such a combination.

3.4. Kinetics of the Photodegradation Process of WO3-Doped g-C3N4 Nanocomposites

The photocatalytic performances of pure g-C3N4 and its WO3/g-C3N4 composites, with various concentrations, were assessed using MB and phenol contaminants under visible-light irradiation. The extent of the concentration of MB and phenol contaminants on the photocatalyst was measured in the dark system until equilibrium. The equilibrium between the prepared photocatalyst and the investigated organic pollutants, MB, as a colored dye, and phenol, as a colorless organic compound, was recognized after 30 min of stirring. The photocatalytic performances of pure g-C3N4 and its WO3/g-C3N4 composites with different photocatalysts were investigated for the photodegradation of MB and phenol contaminants. The achieved results are shown in Figure 4a,b. Figure 4a shows a limited photocatalytic performance of the pure g-C3N4 toward MB. However, the improved photocatalytic performance of WO3/g-C3N4 composites was recorded, indicating a major effect on the photocatalytic activity of g-C3N4 when modified with WO3. In general, the composite’s photocatalytic activity increased after WO3 was applied, and enhanced photocatalytic activity of 0.05% WO3 was noted.
Nevertheless, the increase in WO3 (0.1 and 0.5) has been found to reduce photocatalytic efficiency, which may be due to the decrease in visible-light absorption catalytic sites. Additionally, the same behavior was observed for the degradation of phenol. The composite material containing 0.5% WO3 showed the maximum performance for phenol degradation, which can be attributed to the production of the WO3-semiconductor heterojunction effectively, which leads to transfer of charge from g-C3N4 under the visible-light irradiation. Moreover, the kinetics of the degradation of MB and phenol was noticed to follow the pseudo-first-order model, and the reaction rates were calculated by Equation (2) [32]:
ln(Co/C) = Kt,
Here, Co and C are the concentrations at the initial step and regular time intervals, respectively. Additionally, K is the pseudo-first-order reaction-rate constant (min−1) for investigated organic contaminants, and t is the time (min). Figure 5a,b shows the linear relationship between ln(Co/C) and t. According to these results, the photodegradation of MB and phenol pollutants follows pseudo-first-order reaction kinetics. Additionally, the degradation rate relies on the concentration of the organic substrate. Figure 6 shows the constant rate values for the degradation of MB and phenol contaminants. Percentage of degradation can be calculated to determine the efficiency of the prepared catalysts in the photocatalytic reaction, as illustrated in Figure 5. The efficiency of degradation increases with WO3 content to 0.5%, then the efficiency decreases. It can be noted that the photocatalytic degradation rates of MB had better rate-constant characteristics than the photocatalytic degradation rates of phenol. The photodegradation using MB and phenol contaminants in aqueous solution improvement onto pure g-C3N4 and its WO3/g-C3N4 photocatalyst composites with various concentrations was evaluated under visible light. Dependant on variations in the WO3 content, the plot of the irradiation time (t) against −ln(C/Co) was nearly a straight line. This result has also been reported by G. Lui et al. [33]. The corresponding estimated degradation rate constants (K) were added to Table 2.

3.5. Photodegradation Conduction Mechanism of WO3-Doped g-C3N4 Nanocomposites

The proposed mechanism for photocatalytic behavior of WO3/g-CN composite under the visible spectrum is interpreted in Figure 6. It is a successfully designed WO3/g-CN composite with various WO3. The Kubelka–Munk method was used to calculate the valence band (VB) and conduction band (CB) edge potentials of the prepared pure g-CN and WO3/g-CN with different concentrations of WO3. The direct and indirect bandgaps for g-CN are 2.83 and 2.56 eV, respectively. The increase in WO3 content in the WO3/g-CN composite significantly influences bandgap energy [34]. The proposed system suggests that both g-CN and WO3 will likely offer photogenerated charge carriers. The photogenerated electron in CB of WO3 then flows to VB of g-C3N4 due to electrostatic forces of attraction between the electron in CB of WO3 and the hole in VB g-CN, reducing the recombination of electrons and holes in the WO3/g-C3N4 composite and assisting in enhancing the charge separation spatially.
Furthermore, the electrons in the CB of g-C3N4 can also be retained by reducing the molecular oxygen O2 to form O−2•, due to the negative nature of the CB edge potential compared to the standard redox capability of O2/O−2• [35]. Thus, the hole created in the VB of WO3 leads interacts with H2O to generate OH radicals, giving the more positive potential of the VB than the standard redox capability of OH/OH. In this manner, the production of O−2• and OH radicals plays a significant role in the photodegradation of MB and phenol degradation under visible light. With both OH and O−2, the phenol and Methylene Blue dyes produced CO2 and H2O [36,37].
WO3-g-C3N4 + hυ → e + h+
O2 + e → O−2•
H2O + h+ → OH
OH + O−2• + organic dye (MB or Ph) → CO2+ H2O
In comparison to the previous work, our 0.5% WO3/g-C3N4 has the highest photocatalytic activity of both MB and phenol in the presence of visible light, as shown in Table 3. Using the hydrothermal impregnation and calcination method, Cong Zhao et al. [38] obtained simple and inexpensive C-doped g-C3N4/WO3 as highly active photocatalysts. Tao Pan et al. [39] prepared WO3/g-C3N4 that showed improved photocatalytic activity for TC degradation. Perhaps most importantly, the UV-C-induced–TC degradation activity outperformed all other degradations regardless of concentration or pH level. According to Zhao et al. [40], the Ag/WO2.9/g-C3N4 composite outperforms pure g-C3N4 and pure WO3 in terms of light-selective adsorption and photocatalysis. Hong Yan et al. [41] proved that WO3/g-C3N4-modified nanocomposites had more significant photocatalytic activity than pure WO3 and pure g-C3N4. Junling Zhao et al. [19] discovered that WO3/g-C3N4 photocatalysts have increased degradation activities towards RhB dye when exposed to simulated sunlight. Minji Yoon et al. [42] demonstrated that WO3/g-C3N4 does not exhibit any photocatalytic activity on the degradation of p-nitrophenol. On the other hand, the degradation rate of p-nitrophenol was remarkably increased by the addition of Fenton to WO3/g-C3N4, which means that photocatalytic activity was further enhanced in the presence of hydrogen peroxide (H2O2) because of the generation of hydroxyl radicals.

3.6. Recycling of WO3-Doped g-C3N4 Nanocomposites

A 0.05 WO3/g-CN was subjected re-use in photodegradation of MB and phenol for 4 cycles. Figure 7 represents the recycling and reusability of WO3/g-C3N4. The samples exhibited high performance without further decreased reaction rate, which means that WO3/g-C3N4 is a promising material in MB and phenolic compound photodegradation.

4. Conclusions

Novel, efficient WO3/g-C3N4 nanocomposite with various concentrations of graphitic carbon nitride (g-C3N4) was prepared. The prepared composites exhibited improved photocatalytic activity on the degradation of MB and phenol under visible-light irradiation. The composite material containing 0.5% WO3 showed the maximum MB and phenol degradation performance. The kinetic study showed that the pseudo-first-order kinetic best fitted the photocatalytic process. A mechanism was proposed to degrade organic pollutants using WO3/g-C3N4 composites under visible-light irradiation.

Author Contributions

Conceptualization, I.S.Y., M.S.A.H. and M.A.I.; methodology, M.S.A.H., A.B., H.S.M.A.-R. and H.Y.Z.; investigation, M.S.A.-w., W.A., N.S.A. and I.S.Y.; resources, I.S.Y. and M.S.A.H.; data curation, M.S.A.-w., W.A., N.S.A. and I.S.Y.; writing—original draft preparation, M.S.A.-w., M.S.A.H. and H.Y.Z.; supervision, I.S.Y., M.A.I., and N.S.A.; funding acquisition, M.S.A.-w. and I.S.Y. All authors have read and agreed to the published version of the manuscript.

Funding

The authors express their appreciation to the Deputyship for Research and Innovation, Ministry of Education, in Saudi Arabia, for funding this research work through the project number: (IFP-KKU-2020/10).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data supporting this study’s findings are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ngullie, R.C.; Alaswad, S.O.; Bhuvaneswari, K.; Shanmugam, P.; Pazhanivel, T.; Arunachalam, P. Synthesis and characterization of efficient ZnO/g-C3N4 nanocomposites photocatalyst for Photocatalytic degradation of methylene blue. Coatings 2020, 10, 500. [Google Scholar] [CrossRef]
  2. Ismael, M.; Elhadad, E.; Wark, M. Construction of SnO2/g-C3N4 composite photocatalyst with enhanced interfacial charge separation and high efficiency for hydrogen production and Rhodamine B degradation. Colloids Surf. A Physicochem. Eng. Asp. 2022, 638, 128288. [Google Scholar] [CrossRef]
  3. Wei, H.; Zhang, Y.; Zhang, Y.; Zhang, Y. Design and synthesis of a new high-efficiency CeO2/SnS2/polyaniline ternary composite visible-light photocatalyst. Colloid Interface Sci. Commun. 2021, 45, 100550. [Google Scholar] [CrossRef]
  4. Moniz, S.J.A.; Shevlin, S.A.; Martin, D.J.; Guo, Z.X.; Tang, J. Visible-light driven heterojunction photocatalysts for water splitting-a critical review. Adv. Funct. Mater. 2015, 8, 2421–2440. [Google Scholar] [CrossRef]
  5. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splittingw. Chem. Soc. Rev. 2008, 48, 1685–1687. [Google Scholar] [CrossRef]
  6. Hisatomi, T.; Kubota, J.; Domen, K. Recent advances in semiconductors for photocatalytic and photoelectrochemical water splitting. Chem. Soc. Rev. 2014, 43, 7520. [Google Scholar] [CrossRef]
  7. Chen, X.; Shangguan, W. Hydrogen production from water splitting on CdS-based photocatalysts using solar light. Front. Energy 2013, 7, 111–118. [Google Scholar] [CrossRef]
  8. Honda, K. One and Two-dimensional Structure of Alpha-Helix and Beta-Sheet Forms of Poly(L-Alanine) shown by Specific Heat Measurements at Low Temperatures (1.5–20 K), 1910.
  9. Kong, L.; Jiang, Z.; Lai, H.H.; Nicholls, R.J.; Xiao, T.; Jones, M.O.; Edwards, P.P. Unusual reactivity of visible-light-responsive AgBr–BiOBr heterojunction photocatalysts. J. Catal. 2012, 293, 116–125. [Google Scholar] [CrossRef]
  10. Gao, Y.; Hou, F.; Hu, S.; Wu, B.; Wang, Y.; Zhang, H.; Jiang, B.; Fu, H. Graphene quantum-dot-modified hexagonal tubular carbon nitride for visible-light photocatalytic hydrogen evolution. ChemCatChem 2018, 10, 1330–1335. [Google Scholar] [CrossRef]
  11. Ding, C.; Shi, J.; Wang, Z.; Li, C. Photoelectrocatalytic water splitting: Significance of cocatalysts, Electrolyte, and Interfaces. ACS Catal. 2017, 7, 675–688. [Google Scholar] [CrossRef]
  12. Xing, P.; Zhou, F.; Li, Z. Preparation of WO3/g-C3N4 composites with enhanced photocatalytic hydrogen production performance. Appl. Phys. A 2019, 125, 788. [Google Scholar] [CrossRef]
  13. Nakajima, T.; Kitamura, T.; Tsuchiya, T. Visible light photocatalytic activity enhancement for water purification in Cu(II)-grafted WO3 thin films grown by photoreaction of nanoparticles. Appl. Catal. B Environ. 2011, 108–109, 47–53. [Google Scholar] [CrossRef]
  14. Tong, T.; Zhu, B.; Jiang, C.; Cheng, B.; Yu, J. Mechanistic insight into the enhanced photocatalytic activity of single-atom Pt, Pd or Au-embedded g-C3N4. Appl. Surf. Sci. 2018, 433, 1175–1183. [Google Scholar] [CrossRef]
  15. Li, Z.; Ma, Y.; Hu, X.; Liu, E.; Fan, J. Enhanced photocatalytic H2 production over dual-cocatalyst-modified g-C3N4 heterojunctions. Chin. J. Catal. 2019, 40, 434–445. [Google Scholar] [CrossRef]
  16. Kim, J.; Lee, C.W.; Choi, W. Platinized WO3 as an environmental photocatalyst that generates OH radicals under visible light. Environ. Sci. Technol. 2010, 44, 6849–6854. [Google Scholar] [CrossRef]
  17. Jin, Z.; Murakami, N.; Tsubota, T.; Ohno, T. Complete oxidation of acetaldehyde over a composite photocatalyst of graphitic carbon nitride and tungsten(VI) oxide under visible-light irradiation. Appl. Catal. B Environ. 2014, 150–151, 479–485. [Google Scholar] [CrossRef] [Green Version]
  18. Liang, J.; Li, Y.; Zhang, J.; Li, C.; Yang, X.; Chen, X.; Wang, F.; Chen, C. Crystalline phase engineering in WO3/g-C3N4 composites for improved photocatalytic performance under visible light. Mater. Res. Express. 2020, 7, 065503. [Google Scholar] [CrossRef]
  19. Zhao, J.; Ji, Z.; Shen, X.; Zhou, H.; Ma, L. Facile synthesis of WO3 nanorods/g-C3N4 composites with enhanced photocatalytic activity. Ceram. Int. 2015, 41, 5600–5606. [Google Scholar] [CrossRef]
  20. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2008, 8, 76–80. [Google Scholar] [CrossRef]
  21. Wang, S.; Li, Y.; Wang, X.; Zi, G.; Zhou, C.; Liu, B.; Liu, G.; Wang, L.; Huang, W. One-step supramolecular preorganization constructed crinkly graphitic carbon nitride nanosheets with enhanced photocatalytic activity. J. Mater. Sci. Technol. 2022, 104, 155–162. [Google Scholar] [CrossRef]
  22. Ansari, A.R.; Hammad, A.H.; Abdel-wahab, M.S.; Shariq, M.; Imran, M. Structural, optical and photoluminescence investigations of nanocrystalline CuO thin films at different microwave powers. Opt. Quantum Electron. 2020, 52, 1–16. [Google Scholar] [CrossRef]
  23. Hammad, A.H.; Abdel-wahab, M.S.; Vattamkandathil, S.; Ansari, A.R. Influence the oxygen flow rate on the film thickness, structural, optical and photoluminescence behavior of DC sputtered NiOx thin films. Phys. B Condens. Matter 2019, 568, 6–12. [Google Scholar] [CrossRef]
  24. Alzahrani, A.O.M.; Abdel-wahab, M.S.; Alayash, M.; Aida, M.S. Effect of ZnO layer thickness upon optoelectrical properties of NiO/ZnO heterojunction prepared at room temperature. J. Mater. Sci. Mater. Electron. 2018, 29, 16317–16324. [Google Scholar] [CrossRef]
  25. Mo, Z.; She, X.; Li, Y.; Liu, L.; Huang, L.; Chen, Z.; Zhang, Q.; Xu, H.; Li, H. Synthesis of g-C3N4 at different temperatures for superior visible/UV photocatalytic performance and photoelectrochemical sensing of MB solution†. RSC Adv. 2015, 123, 101552–101562. [Google Scholar] [CrossRef]
  26. Kadi, M.W.; Mohamed, R.M. Increasing visible light water splitting efficiency through synthesis route and charge separation in measoporous g-C3N4 decorated with WO3 nanoparticles. Ceram. Int. 2019, 45, 3886–3893. [Google Scholar] [CrossRef]
  27. Tauc, J.; Grigorovici, R.; Vancu, A. Optical properties and electronic structure of amorphous germanium. Phys. Status Solidi 1996, 15, 627–637. [Google Scholar] [CrossRef]
  28. Chu, X.; Liu, J.; Liang, S.; Bai, L.; Dong, Y.; Epifani, M. Facile preparation of g-C3N4-WO3 composite gas sensing materials with enhanced gas sensing selectivity to acetone. J. Sens. 2019, 2019, 6074046. [Google Scholar] [CrossRef] [Green Version]
  29. Liu, Z.X.; Wang, X.L.; Hu, A.P.; Tang, Q.L.; Xu, Y.L.; Chen, X.H. 3D Se-doped NiCoP nanoarrays on carbon cloth for efficient alkaline hydrogen evolution. J. Cent. South Univ. 2021, 28, 2345–2359. [Google Scholar] [CrossRef]
  30. Huang, L.; Xu, H.; Li, Y.; Li, H.; Cheng, X.; Xia, J.; Xu, Y.; Cai, G. Visible-light-induced WO3/g-C3N4 composites with enhanced photocatalytic activity. Dalton Trans. 2013, 42, 8606. [Google Scholar] [CrossRef]
  31. Tai, J.Y.; Leong, K.H.; Saravanan, P.; Sim, L.C. Bioinspired Synthesis of Carbon Dots/g-C3N4 Nanocomposites for Photocatalytic Application. E3S Web Conf. 2018, 65, 05015. [Google Scholar] [CrossRef] [Green Version]
  32. Abdel-wahab, M.S. Substrate temperature impact on the structural, optical and photo-catalytic activity of sputtered Cu-Doped ZnO thin films. J. Electron. Mater. 2021, 50, 4364–4372. [Google Scholar] [CrossRef]
  33. Liu, G.; Han, J.; Zhou, X.; Huang, L.; Zhang, F.; Wang, X.; Ding, C.; Zheng, X.; Han, H.; Li, C. Enhancement of visible-light-driven O2 evolution from water oxidation on WO3 treated with hydrogen. J. Catal. 2013, 307, 148–152. [Google Scholar] [CrossRef]
  34. Sun, X.Y.; Zhang, F.J.; Kong, C. Porous g-C3N4/WO3 photocatalyst prepared by simple calcination for efficient hydrogen generation under visible light. Colloids Surf. A Physicochem. Eng. Asp. 2020, 594, 124653. [Google Scholar] [CrossRef]
  35. Hussien, M.S.A.; Yahia, I.S. Hybrid multifunctional core/shell g-C3N4 @TiO2 heterojunction nano-catalytic for photodegradation of organic dye and pharmaceutical compounds. Environ. Sci. Pollut. Res. 2021, 28, 29665–29680. [Google Scholar] [CrossRef] [PubMed]
  36. Deng, S.Z.; Yang, G.L.; Zhu, Y.H.; Li, F.W.; Zhang, X. WO3 nanosheets/g-C3N4 nanosheets’ nanocomposite as an effective photocatalyst for degradation of rhodamine B. Appl. Phys. A Mater. Sci. Process. 2019, 125, 1–11. [Google Scholar] [CrossRef]
  37. Al-Ghamdi, A.A.; Khedr, M.H.; Ansari, M.S.; Hasan, P.M.Z.; Abdel-Wahab, M.S.; Farghali, A.A. RF sputtered CuO thin films: Structural, optical and photocatalytic behavior. Phys. E Low-Dimens. Syst. Nanostructures 2016, 81, 83–90. [Google Scholar] [CrossRef]
  38. Zhao, C.; Ran, F.; Dai, L.; Li, C.; Zheng, C.; Si, C. Cellulose-assisted construction of high surface area Z-scheme C-doped g-C3N4/WO3 for improved tetracycline degradation. Carbohydr. Polym. 2021, 255, 117343. [Google Scholar] [CrossRef]
  39. Pan, T.; Chen, D.; Xu, W.; Fang, J.; Wu, S.; Liu, Z.; Wu, K.; Fang, Z. Anionic polyacrylamide-assisted construction of thin 2D-2D WO3/g-C3N4 Step-scheme heterojunction for enhanced tetracycline degradation under visible light irradiation. J. Hazard. Mater. 2020, 393, 122366. [Google Scholar] [CrossRef]
  40. Zhao, X.; Zhang, X.; Han, D.; Niu, L. Ag supported Z-scheme WO2.9/g-C3N4 composite photocatalyst for photocatalytic degradation under visible light. Appl. Surf. Sci. 2020, 501, 144258. [Google Scholar] [CrossRef]
  41. Yan, H.; Zhu, Z.Y.; Long, W.L. Single-source-precursor-assisted synthesis of porous WO3/g-C3N4 with enhanced photocatalytic property. Colloids Surf. A Physicochem. Eng. Asp. 2019, 582, 123857. [Google Scholar] [CrossRef]
  42. Yoon, M.Y.; Oh, S.; Hong, J.S.; Lee, R.; Boppella, S.H.K.; Mota, F.M.; Kim, S.O.; Kim, D.H. Synergistically enhanced photocatalytic activity of graphitic carbon nitride and WO3 nanohybrids mediated by photo-Fenton reaction and H2O2. Appl. Catal. B Environ. 2017, 206, 263–270. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of pure g-C3N4 powder and its WO3 doping on g-C3N4 composite organic compounds (0.001, 0.01, 0.05, 0.1, and 0.5 g of WO3).
Figure 1. XRD patterns of pure g-C3N4 powder and its WO3 doping on g-C3N4 composite organic compounds (0.001, 0.01, 0.05, 0.1, and 0.5 g of WO3).
Materials 15 02482 g001
Figure 2. (af) SEM images of pure g-C3N4 and its WO3/g-C3N4 nanocomposite organic compounds with various amounts of tungsten oxide (0.001, 0.01, 0.05, 0.1, and 0.5 g of WO3, respectively).
Figure 2. (af) SEM images of pure g-C3N4 and its WO3/g-C3N4 nanocomposite organic compounds with various amounts of tungsten oxide (0.001, 0.01, 0.05, 0.1, and 0.5 g of WO3, respectively).
Materials 15 02482 g002
Figure 3. (ac) Diffused reflectance optics UV-Vis (a), (α)2 (b) and (α)1/2 (c) versus the incident photon energy hυ of pure g-C3N4 and its WO3/g-C3N4, with various amounts of tungsten oxide (0.001, 0.01, 0.05, 0.1, and 0.5 g of WO3).
Figure 3. (ac) Diffused reflectance optics UV-Vis (a), (α)2 (b) and (α)1/2 (c) versus the incident photon energy hυ of pure g-C3N4 and its WO3/g-C3N4, with various amounts of tungsten oxide (0.001, 0.01, 0.05, 0.1, and 0.5 g of WO3).
Materials 15 02482 g003aMaterials 15 02482 g003b
Figure 4. The degradation (%) of MB (a,b) phenol for pure g-C3N4 and its WO3/g-C3N4 nanocomposites.
Figure 4. The degradation (%) of MB (a,b) phenol for pure g-C3N4 and its WO3/g-C3N4 nanocomposites.
Materials 15 02482 g004aMaterials 15 02482 g004b
Figure 5. (a,b) The kinetic degradation curves of MB (a) and (b) Phenol for pure g-C3N4 and its WO3/g-C3N4 nanocomposites.
Figure 5. (a,b) The kinetic degradation curves of MB (a) and (b) Phenol for pure g-C3N4 and its WO3/g-C3N4 nanocomposites.
Materials 15 02482 g005
Figure 6. Photocatalytic mechanism of WO3/g-C3N4 nanocomposites.
Figure 6. Photocatalytic mechanism of WO3/g-C3N4 nanocomposites.
Materials 15 02482 g006
Figure 7. The recycling process for pure g-C3N4 and its WO3/g-C3N4 nanocomposites in photodegradation of (a) MB, (b) phenol.
Figure 7. The recycling process for pure g-C3N4 and its WO3/g-C3N4 nanocomposites in photodegradation of (a) MB, (b) phenol.
Materials 15 02482 g007
Table 1. N4 and its WO3/g-C3N4 with various amounts of tungsten oxide (0.001, 0.01, 0.05, 0.1, and 0.5% WO3).
Table 1. N4 and its WO3/g-C3N4 with various amounts of tungsten oxide (0.001, 0.01, 0.05, 0.1, and 0.5% WO3).
SamplesPhasesMean Values
Crystallite Size,
(nm)
Dislocation Density,
(1/(nm)2)
Lattice
Strain
Pure g-C3N4Pure g-C3N439.176.51 × 10−48.850 × 10−4
0.001 g WO3-doped g-C3N4Pure g-C3N435.917.999 × 10−49.753 × 10−4
0.01 g WO3-doped g-C3N4Pure g-C3N435.937.993 × 10−49.749 × 10−4
0.05 g WO3-doped g-C3N4Pure g-C3N440.846.866 × 10−48.886 × 10−4
0.1 g WO3-doped g-C3N4Pure g-C3N444.839.336 × 10−49.667 × 10−4
0.5 g WO3-doped g-C3N4Phase 1: Pure g-C3N481.761.710 × 10−44.436 × 10−4
Phase 2: WO352.933.641 × 10−46.593 × 10−4
Table 2. Particle size and corresponding rate constants of pure g-C3N4 and its WO3/g-C3N4, with various amounts of tungsten oxide (0.001, 0.01, 0.05, 0.1, and 0.5% WO3).
Table 2. Particle size and corresponding rate constants of pure g-C3N4 and its WO3/g-C3N4, with various amounts of tungsten oxide (0.001, 0.01, 0.05, 0.1, and 0.5% WO3).
SamplesParticle Size, (µm)K, (min−1)
MBPhenol
Pure g-C3N41.650.00280.0053
0.001 g WO3/g-C3N41.210.00540.0064
0.01 g WO3/g-C3N41.260.00490.0092
0.05 g WO3/g-C3N41.30.01080.0194
0.1 g WO3/g-C3N41.50.00520.0172
0.5 g WO3/g-C3N41.510.00450.0145
Table 3. Photocatalytic comparison between the prepared WO3/g-C3N4 and various previous works.
Table 3. Photocatalytic comparison between the prepared WO3/g-C3N4 and various previous works.
PhotocatalystMethod of PreparationOrganic SolutionIrradiation TimeSource% DegradationRefs.
WO3/g-CNPyrolysisMB120 minVisible
Light lamp
80%Present work
Phenol90%
C-doped g-C3N4/WO3Hydrothermal impregnationTetracycline60 minUV-light
lamp
75%[33]
WO3/g-C3N4Calcination methodTetracycline120 minXenon lamp90.54%[34]
Ag/WO3/g-C3N4Calcination methodMB300 minUV-light
lamp
-----[35]
WO3/g-C3N4Thermal polymerizationMO120 minXenon lamp93%[36]
WO3/g-C3N4HydrothermalRhB90 minUV-light
lamp
91%[37]
WO3/g-C3N4/
Photo-Fenton
Calcination methodP-nitrophenol240 minXenon lamp91%[38]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hussien, M.S.A.; Bouzidi, A.; Abd-Rabboh, H.S.M.; Yahia, I.S.; Zahran, H.Y.; Abdel-wahab, M.S.; Alharbi, W.; Awwad, N.S.; Ibrahim, M.A. Fabrication and Characterization of Highly Efficient As-Synthesized WO3/Graphitic-C3N4 Nanocomposite for Photocatalytic Degradation of Organic Compounds. Materials 2022, 15, 2482. https://doi.org/10.3390/ma15072482

AMA Style

Hussien MSA, Bouzidi A, Abd-Rabboh HSM, Yahia IS, Zahran HY, Abdel-wahab MS, Alharbi W, Awwad NS, Ibrahim MA. Fabrication and Characterization of Highly Efficient As-Synthesized WO3/Graphitic-C3N4 Nanocomposite for Photocatalytic Degradation of Organic Compounds. Materials. 2022; 15(7):2482. https://doi.org/10.3390/ma15072482

Chicago/Turabian Style

Hussien, Mai S. A., Abdelfatteh Bouzidi, Hisham S. M. Abd-Rabboh, Ibrahim S. Yahia, Heba Y. Zahran, Mohamed Sh. Abdel-wahab, Walaa Alharbi, Nasser S. Awwad, and Medhat A. Ibrahim. 2022. "Fabrication and Characterization of Highly Efficient As-Synthesized WO3/Graphitic-C3N4 Nanocomposite for Photocatalytic Degradation of Organic Compounds" Materials 15, no. 7: 2482. https://doi.org/10.3390/ma15072482

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