ACS Publications. Most Trusted. Most Cited. Most Read
Ternary Photodegradable Nanocomposite (BiOBr/ZnO/WO3) for the Degradation of Phenol Pollutants: Optimization and Experimental Design
My Activity
  • Open Access
Article

Ternary Photodegradable Nanocomposite (BiOBr/ZnO/WO3) for the Degradation of Phenol Pollutants: Optimization and Experimental Design
Click to copy article linkArticle link copied!

  • Hager R. Ali
    Hager R. Ali
    Spectroscopic Division, Analysis and Evaluation Department, Egyptian Petroleum Research Institute (EPRI), Nasr City 11727 Cairo, Egypt
    More by Hager R. Ali
  • Eman A. Motawea*
    Eman A. Motawea
    Spectroscopic Division, Analysis and Evaluation Department, Egyptian Petroleum Research Institute (EPRI), Nasr City 11727 Cairo, Egypt
    *Email: [email protected]. Phone: 002-01227024388.
Open PDF

ACS Omega

Cite this: ACS Omega 2021, 6, 34, 22047–22064
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.1c02644
Published August 17, 2021

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

The degradation of environmental contaminants with photocatalysts has bright prospects for application in the control of pollution. In this study, BiOBr/ZnO/WO3 heterojunctions have been documented to be reliable visible-light photocatalysts for phenol deterioration. X-ray diffraction, X-ray photoelectron spectroscopy, Fourier-transform infrared spectroscopy, scanning electron microscopy, transmission electron microscopy, photoluminescence spectral analysis, electrochemical impedance spectroscopy (EIS), EIS Bode plots, linear sweep voltammetry, and UV–visible diffuse reflectance spectroscopy were employed to describe the heterojunction’s structure in addition to its optical features. The results revealed that the BiOBr/ZnO/WO3 ternary photocatalyst displayed more degradation activity in comparison to single-phase ZnO, WO3, or BiOBr, which is also higher than that of binary mixture photocatalysts with a phenol degradation efficiency of 90%. The influence of degradation variables, for instance, the potential of hydrogen (pH) and the initial organic contaminant content besides the heterojunction dose, on the deterioration efficiency was optimized using the response surface methodology. The degradation efficiency reached 95% under the optimal conditions of 0.08 g/0.03 L catalyst dose, a pH of 9, and an initial organic contaminant content of 10 mg L–1. However, the optimal phenol degradation efficiency of 39.37 mg g–1 was achieved under the conditions of 0.08 g/0.03 L catalyst dose, pH of 9, and 200 mg L–1 initial phenol concentration.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2021 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

Treatment of water contaminated with industrial effluents is a persistent global problem today. Several related bodies have raised concerns about the green technology and energy efficiency in wastewater treatment in recent years to reduce the current environmental problems. Phenol is present in the industrial synthesis of adhesive and antiseptic products as an intermediate. (1) An innate deterioration rate of monophenol can happen depending on a condition through various processes. In water bodies, the acceptable limit for phenol is 0.1–1.0 mg L–1. Owing to its toxicity, genotoxicity, and exceptionally high endocrine-disrupting potency, phenol remediation therapies are becoming problems despite the low concentration limit. (2) One of the best approaches is producing a photocatalyst and its use in process treatment. Some of the rare-earth semiconductors such as BiOBr and ZnO besides WO3 are used during photocatalytic processes. To start the photocatalytic reaction and ultimately degrade the contaminant, these semiconductors will use sunlight. (3) Recently, due to their peculiar properties and possible uses, BiO halogenides (BiOX, where X = Cl, Br, and I) were studied. Bismuth oxybromide is a positive-category semiconductor that has a band gap of 2.80 eV and an enticing photocatalyst for contaminant deterioration under solar light illumination. (4,5) Some researchers have reported that photocatalytic activity is strongly dependent on the size, diameter, and structure of catalysts. Nanophotocatalysts’ shape, scale, and structure have therefore been tuned to enhance their activity. (4,6) Recently, due to their high crystallinity and ultrathin thickness, the lamellar nanostructures have been gaining greater attention. It is well known that this structure could help in reducing the recombination possibilities of the illumination-created free charge carriers, so that they can easily move toward the surface to degrade the organic molecules. (5,7) Fabricating a blend of different p–n types through integrating different transistors with suitable band models is the most efficient way. Besides, an n-type may be generated by hybridizing ZnO with WO3, which could facilitate the isolation of illumination-produced carriers. Tungsten(VI) oxide (WO3) nanoparticles (NPs) have attracted particular interest among researchers in the domain of illumination catalysts owing to their heavy sorption in the presence of solar spectra (almost 500 nm), their sustainable physical and chemical properties, and their resistance toward photocorrosion. (8) Tungstic anhydride (WO3) has exceptional electrical illumination characteristics of a small band gap of 2.40–2.80 eV with a favorable hole dispersion width that may potentially use ∼12% solar light. (9−11) However, because of stagnant kinetics and low charge transfer at the transistor–electrolyte system, WO3 demonstrates a fair photocatalytic activity toward the deterioration of phenolic contaminants. (12,13) Several techniques have been employed to decorate WO3 to solve these problems, including metal implantation (14,15) and coupling with other semiconductors. (6,16) ZnO, additionally defined as the II–IV transistor, contains a broadband energy band gap (3.3 eV), high photosensitivity, and stability. (17,18) Also, at room temperature, ZnO possessed good heat sustainability (19) and is relatively inexpensive. The variables of a process can be developed, improved, and optimized by the response surface methodology (RSM), which is an assimilation of theoretical and numerical tools used for assessing the significance of deferential process factors in the presence of different correlations. RSM is a viable procedure for optimizing responses that are modeled by process parameters. (20) To our knowledge, no previous research has focused on the fabrication of a BiOBr/WO3/ZnO p–n heterojunction and its phenol degradation performance. We present a simple and cost-effective solvothermal synthesis technique for fabricating a unique hybrid BiOBr/WO3/ZnO p–n heterojunction. In the current work, we discuss the effects of p–n heterojunction formation on the visible-light photodegradation properties of phenol and its mechanisms.

Results and Discussion

Click to copy section linkSection link copied!

Characterization of NPs

X-ray Diffraction

To characterize the synthesized NPs, various techniques were used. X-ray diffraction (XRD) was used to describe the crystal structure and primary crystal scale. XRD analysis was used to analyze the BiOBr/WO3/ZnO phase and crystalline composition (Figure 1). XRD detected the pure WO3 as well. The sharp diffraction peaks may be well ascribed to the WO3 phase (JCPDS card no. 024-0747) with 2θ values of 24.35, 26.56, 33.23, 34.15, 35.50, 45.70, and 49.90° and a crystallinity of 79.03%. XRD peaks of the as-prepared ZnO (JCPDS card no. 01-081-8838) were observed at angles of 31.80, 34.46, 36.27, 47.50, 56.60, 62.89, 67.97, 69.11, 72.59, and 76.97o (2θ). The most common diffraction peaks of BiOBr were observed at 2θ angles of 25.16, 31.80, 32.17, 46.16, and 57.17° with a crystallinity of 59.35% and indexed to BiOBr according to JCPDS card no. 52-0084 with a crystallinity of 81.63%. The BiOBr/WO3 XRD pattern (JCPDS file no. 52-0084) and of BiOBr/ZnO (JCPDS file no. 09-0393) corresponded to JCPDS file no. 52-0084; ZnO- and WO3-related peaks were not identified due to the presence of low BiOBr-related ZnO and WO3 amounts. In the BiOBr/WO3/ZnO composites, the XRD pattern corresponded to JCPDS file no. 52-0084. This may be due to the great dispersion of ZnO and WO3 in BiOBr; the position of the peaks showed some changes suggesting the formation of a new compound. Besides, the presence of BiOBr limited the growth of ZnO and WO3 NPs.

Figure 1

Figure 1. XRD spectra of the synthesized WO3, ZnO, BiOBr, BiOBr/WO3, BiOBr/ZnO, and BiOBr/WO3/ZnO.

Fourier-Transform Infrared Spectroscopy

To evaluate the composition, Fourier-transform infrared (FTIR) spectroscopy was performed. Figure 2 presents the contrast of FTIR spectra for all the prepared photocatalysts; the FTIR spectrum of the synthesized BiOBr/WO3/ZnO composite shows a high peak of absorption at 3508 cm–1 owing to the stretching vibration of nonchemical bond association of OH groups, while a peak at 1637 cm–1 was observed for H–O–H bending vibrations. The peaks are observed due to the existence of CO2 at 2390 cm–1. For Zn–O, the most characteristic peak at 514 cm–1 is observed, while for the bending vibration of C–H stretching, an absorption peak at about 1392 cm–1 is attributed. The findings mentioned above are in line with the results of XRD. The peak at a lower frequency of 810.84 cm–1 is the product of W–O–W bond stretching vibrations. The presence of W–O–W/O–W–O (842 and 781 cm–1, respectively) and also a further peak at 507 cm–1 is due to the Bi-O bond that is observed in the FTIR spectra of binary and ternary nanocomposites. (21) These findings further show that the preparation of the ternary nanocomposite BiOBr/WO3/ZnO has been effectively done.

Figure 2

Figure 2. FTIR patterns of WO3, ZnO, BiOBr, BiOBr/WO3, BiOBr/ZnO, and BiOBr/WO3/ZnO.

X-ray Photoelectron Spectroscopy

The electronic states and the chemical composition of BiOBr/ZnO/WO3 were examined by X-ray photoelectron spectroscopy (XPS), and the results are given in Figure 3. It can be observed from the survey spectrum in Figure 3a that the characteristic peaks of Zn, Bi, Br, O, and W are present, indicating their coexistence on the surface of the BiOBr/ZnO/WO3 sample. As for the high-resolution Bi 4f spectrum in Figure 3b, two significant peaks are observed at 159 and 164.5 eV, corresponding to Bi 4f5/2 and Bi 4f7/2, respectively, due to the spin–orbital component of Bi3+ in BiOBr. (22) Moreover, as shown in Figure 3c, the XPS spectra of Br 3d consisted of Br 3d5/2 at 68.2 eV and Br 3d3/2 at 69.5 eV attributed to Br in the monovalent oxidation state. (22) Also, the high-resolution spectrum of W 4f depicted in Figure 3d shows that W 4f7/2 at 35.4 eV and W 4f 5/2 at 37.5 eV are ascribed to W6+ species. (23) From the Zn 2p XPS spectrum in Figure 3e, the fitted peaks at 1045 and 1022 eV can be observed, which were attributed to 2p1/2 and 2p3/2 of Zn2+, respectively, in ZnO. (24) The O 1s spectrum in Figure 3f could be resolved into four individual peaks, which indicated the presence of four different kinds of O species in the sample. The main peak at 530.4 eV was assigned to O2 ions that are in oxygen-deficient regions. The peak at 533.9 eV was attributed to the crystal lattice O atom in WO3. (23) Also, the peak at 531.9 eV was related to the crystal lattice O atom in BiOBr, (24) while the peak at 530.6 eV could be attributed to the O atom in ZnO. (24) The XPS analysis proves the existence of the BiOBr/ZnO/WO3 composite.

Figure 3

Figure 3. XPS spectrum of BiOBr/ZnO/WO3: (a) survey spectrum, (b) Bi 4f spectrum, (c) Br 3d spectrum, (d) W 4f spectrum, (e) Zn 2p spectrum, and (f) O 1s spectrum.

Optical Properties of Photocatalysts

UV–visible diffuse reflection spectroscopy (UV–vis DRS) and photoluminescence (PL) spectroscopy were utilized to identify the optical characteristics of pure ZnO, WO3, BiOBr, and binary and ternary composites (Figure 4). As shown in Figure 4A, the neat BiOBr, ZnO, and WO3 display absorption edges at 425, 420, and 450 nm, respectively, while the binary BiOBr/ZnO and BiOBr/WO3 exhibit absorption edges at 428 and 470 nm, respectively; additionally, BiOBr/ZnO/WO3 shows the highest visible-light absorption edge at 485 nm, confirming the existence of additional energy levels of metal ions in the forbidden gap of BiOBr. The band gap of the prepared photocatalysts was calculated according to the classical equation of Taucʼs plot (Figure 4B,C)
(1)

Figure 4

Figure 4. Optical properties of the prepared photocatalysts: (A) UV–vis DRS spectra, (B) Taucʼs plots [(αhv)2 vs Hv] for WO3 and ZnO photocatalysts, (C) Taucʼs plots [(αhv)1/2vs Hv] for BiOBr, BiOBr/ZnO, BiOBr/WO3, and BiOBr/ZnO/WO3, and (D) PL spectra.

It should be noted that the absorption coefficient per unit length (4,17) is expressed as α, Eg represents the band gap energy, h refers to Planck’s constant, v is the frequency of vibration, hv implies the incident photon energy, A refers to the proportional constant, and n is 0.50 for the indirect band-gap semiconductor and 2 for the direct band-gap semiconductor. ZnO and WO3 have been reported to be attributed, as shown in Figure 4B, to the n-type direct band-gap semiconductor (5) and BiOBr to the p-type indirect band-gap semiconductor (7) (Figure 4C). The band gaps of the synthesized ZnO and WO3 were estimated to be 3.20 and 2.85 eV, respectively, and that of BiOBr was 2.80 eV. By doping BiOBr with ZnO and WO3, the band gap was reduced to approximately 2.75 and 2.70 eV, respectively, while the band gap was reduced to approximately 2.60 eV for the BiOBr/WO3/ZnO composite. This confirms that BiOBr/WO3/ZnO exhibits the narrowest band gap, and this reduction in energy value is due to excessive absorption in the visible region due to the energy level existing in the forbidden gap through the doping of ZnO and WO3. Furthermore, the conduction band (CB) and valence band (VB) edges of ZnO, BiOBr, and WO3 that are depicted in Table 1 have been studied more as a result of their significant roles in assessing the heterojunction flowchart of photogenerated electrons and holes. WO3 (band gap: 2.85 eV), ZnO (band gap: 3.20 eV), and BiOBr (band gap: 2.80 eV) band-edge positions have been assessed according to the following equations
(2)
(3)
where the VB edge potential is expressed as EVB and X is the semiconductor electronegativity, which is the geometric mean of electronegativity for constituent atoms. The X values are determined as 6.18, 5.79 and 6.49 eV, respectively, for BiOBr, ZnO, and WO3. E0 is the free electron energy of the hydrogen scale (∼4.5 eV) and Eg is the semiconductor’s band gap energy. The EVB and ECB values of BiOBr are calculated as 3.16 and 0.20 eV, respectively, while those of WO3 are calculated as 3.365 and 0.615 eV and of ZnO are 2.89 and −0.31 eV, respectively.
Table 1. Values of Calculated ECB and EVB for ZnO, WO3, and BiOBr
semiconductor absolute electronegativityX (eV)ECB (eV)EVB (eV)Eg (eV)
ZnO5.790–0.312.893.2
BiOBr6.1800.23.162.8
WO36.490.6153.362.85
PL is an efficient tool for investigating the performance of charge carrier transfer, trapping, and separation of photogenerated electrons and holes in heterojunctions. (7)Figure 4D displays the PL spectra of all samples, where the main emission peak is located at 430 nm. The PL spectrum of composite photocatalysts is similar to that of pure BiOBr, while the intensity emission is reduced in the BiOBr/WO3/ZnO photocatalyst. Generally, the intensity of PL emission is derived from free charge carrier recombination, so the lower emission intensity of the BiOBr/WO3/ZnO heterojunction implies a lower electron and hole recombination rate under visible light irradiation. PL results prove that WO3 and ZnO are favorable for charge separation in BiOBr/WO3/ZnO photocatalysts as a result of the lower recombination rate of photogenerated charge carriers.

Photoelectrochemical Measurements

The charge generation, migration, and recombination processes in photocatalysts were investigated by electrochemical impedance spectroscopy (EIS Nyquist spectra); it is evident that the higher the electron and hole separation efficiency, the better the photocatalytic performance. Figure 5a demonstrates that compared to BiOBr, BiOBr/ZnO/WO3 has a much lower impedance, as evidenced by the smaller arc radius in the EIS Nyquist plots, implying that BiOBr/ZnO/WO3 has a lower interfacial layer resistance, beneficial to a faster interfacial transfer of electron–hole pairs and thus a more effective separation. For better analysis of the EIS data, the corresponding Bode phase plot is presented in Figure 5b; a characteristic of the Bode phase plot is its ability to identify the predominant electrical behavior of the system in a given frequency range. The maximum oscillation frequency (fmax) of the impedance semicircle of BiOBr/ZnO/WO3 is less than that of BiOBr. Thus, the corresponding lifetime of photoelectrons in BiOBr/ZnO/WO3 increases. This can be ascribed to the efficient separation of light-generated electrons and holes in BiOBr/ZnO/WO3. In other words, the electrons transferred from the CB of BiOBr to that of ZnO and WO3, while the holes were transported to the VB of BiOBr. Therefore, the recombination of electron–hole pairs was inhibited, and the lifetime of photoelectrons was prolonged. Figure 5c shows the linear sweep voltammetry (LSV) measurements under visible-light irradiation for the BiOBr and BiOBr/ZnO/WO3 electrodes. The photocurrent values of both electrodes increased with the upgrade of the applied potential. The photocurrent densities of BiOBr/ZnO/WO3 were much higher than that of bare BiOBr at the same applied potential, which further indicated that BiOBr/ZnO/WO3 has higher charge separation and migration.

Figure 5

Figure 5. Photoelectrochemical measurements: (a) Nyquist plots of EIS, (b) EIS Bode plots, and (c) LSV curves for BiOBr and BiOBr/ZnO/WO3.

Scanning Electron Microscopy, Transmission Electron Microscopy, and Energy-Dispersive X-ray spectroscopy and Electron Microscopy Studies

The size and morphologies of the prepared BiOBr/ZnO/WO3 nanocomposite were investigated using scanning electron microscopy (SEM) and transmission electron microscopy (TEM). For comparison, Figure 6A,E displays the SEM and TEM photos of ZnO NPs; the particles clearly show the spherical structural composition, also nearly spongy-like structural NPs were observed with an average size of 122–193 nm; however, Figure 6B,F shows disc-shaped of agglomerated WO3 plates and also confirms the orthorhombic shapes of different dimensions with an average size of 71.9–168 nm. Evidently, it is clear that the prepared BiOBr, as shown in Figure 6C,G, contained ultrathin nanocrystals of irregular-shaped nanoplates with an average width of 50–122 nm, whereas ZnO and WO3 can be observed deposited on the surface of flakes/plates of BiOBr, as shown in Figure 6D,H, and such a structure is expected to provide abundant heterojunction interfaces between BiOBr, ZnO, and WO3.

Figure 6

Figure 6. SEM images of the synthesized (A) ZnO NPs, (B) WO3NPs, (C) BiOBr NPs, and (D) BiOBr/ZnO/WO3 nanocomposites, TEM images of (E) ZnO NPs, (F) WO3NPs, (G) BiOBr NPs, and (H) BiOBr/ZnO/WO3, and (I) EDX spectra of BiOBr/ZnO/WO3.

Energy-dispersive X-ray spectroscopy (EDX) analysis was utilized to explore the surface chemical composition of ternary BiOBr/ZnO/WO3. As per Figure 6I, the constituents of Bi, Br, O, Zn, and W were revealed in the spectrum of BiOBr/ZnO/WO3, implying that the ternary nanocomposite was successfully formed.

Evaluation of Photocatalytic Activity

The efficiency of the prepared photocatalyst toward organic pollutant photocatalytic degradation was investigated by conducting phenol removal. Adsorptive and photocatalytic activities were assessed individually for evaluating and comparing the performance. To screen the best catalyst that can be included in the upcoming experiments and achieve the highest removal of monophenol under visible-light irradiation, the photocatalytic performance of ZnO, WO3, BiOBr, BiOBr/ZnO, BiOBr/WO3, and BiOBr/ZnO/WO3 photocatalysts was evaluated. To attain adsorption–desorption equilibrium, the phenol samples with photocatalysts were shaken in the dark for 1 h. Due to its significant role in the polluted water-treatment process, sorption kinetics was evaluated. At an initial concentration of 50 mg/L, time-dependent phenol adsorption/degradation on the prepared catalysts was evaluated and is presented in Figure 7A as a function of adsorption/degradation capacity and in Figure 7B as a function of adsorption/degradation efficiency. As a result, in the dark, all catalysts reached the adsorption equilibrium after 60 min, while in the presence of visible light, catalysts reached degradation equilibrium after 300 min. From Figure 7A, it can be observed that the maximum phenol degradation capacity reached 27 mg/g with BiOBr/ZnO/WO3 photocatalysts. In the dark, the phenol removal percent was negligible, as depicted in Figure 7B, and decreased in the order of BiOBr/ZnO/WO3 (25%) > BiOBr/WO3 (22%) > BiOBr/ZnO (18%) > BiOBr (17.6%) > ZnO (16%) > WO3 (14%). This may be due to the high surface acidity of ZnO and WO3 (29,30) that increases the positive charge on the surface of the catalyst and also the repulsion forces between the catalyst and the cationic adsorbents. However, the efficiency of phenol photodegradation under visible light enhanced and is in the order of ZnO (50%) < WO3 (60%) < BiOBr (66.5%) < BiOBr/ZnO (73%) < BiOBr/WO3 (81%) < BiOBr/ZnO/WO3 (90%), which may be attributed to ZnO and WO3 photocatalysts operating as an acceptor of the electron, which is helpful for the efficient isolation of electron–hole pairs generated in the system and prevent their recombination. (27)

Figure 7

Figure 7. (A) Degradation capacity of phenol vs time with different catalysts, (B) degradation percentage at an equilibrium time of 300 min, and (C) kinetics of the photodegradation of phenol using different photocatalysts at the initial concentration of 50 mg/L, dose of 0.05 g/0.03 L, and a pH of 6.

The kinetics of the photodegradation of phenol illustrated in Figure 7C implies that the lowest deterioration performance that may be due to the unsuitable band gap under visible-light-induced conditions was shown by pure ZnO and WO3. Under visible-light irradiation, these photocatalysts were hard to be simulated. (26) The deterioration activity of ZnO and WO3 was drastically improved by doping with BiOBr and it reached 90% for BiOBr/ZnO/WO3. We conclude according to the above results that BiOBr/ZnO/WO3 p–n heterojunctions display a high photocatalytic activity rather than the solitary phases of ZnO, WO3, and BiOBr, which is also higher than that of the two component mixture photocatalyst. Compared to an individual or binary components, the improvement of the deterioration activity of the BiOBr/ZnO/WO3 p–n heterojunctions could be related to the creation of heterojunctions through n-type ZnO, WO3, and p-type BiOBr, enabling the interface isolation of electron–hole pairs.
Pseudo-first-order and pseudo-second-order models were utilized to evaluate the kinetics of phenol sorption onto photocatalysts, and the values were calculated according to eqs 3 and 4, respectively, whose data are depicted in Table 2. The regression results predicted (R2) demonstrate that, in all cases, the pseudo-first-order model defines phenol sorption rather than the pseudo-second-order model, indicating that phenol degradation depends on the catalyst used. Therefore, the rate of phenol sorption process depending on the pseudo-first-order kinetic model increases in the order of WO3 (0.0014 min–1) < ZnO (0.0019 min–1) < BiOBr (0.0023 min–1) < BiOBr/ZnO (0.0024 min–1) < BiOBr/WO3 (0.0040 min–1) < BiOBr/ZnO/WO3 (0.0062 min–1). BiOBr/ZnO/WO3 was found to be the best catalyst with the best sorption capacity (7.5 mg/g) and a maximum sorption rate (0.0062 min–1) among all catalysts. Experimental data were applied to the pseudo-first-order (eq 3) to obtain more information on the kinetics of phenol deterioration catalyzed by different photocatalysts, where the rate constant K value generally provides an estimate about photocatalyst’s efficiency. (28) The dark and visible-light irradiation catalytic efficiencies of the prepared catalysts are illustrated in Figure 7C. Due to the high band gap of ZnO and WO3 (3.20 and 2.85 eV, respectively), their catalytic efficiencies were significantly limited under visible-light irradiation with nearly 50 and 60 percent, respectively, for monophenol deterioration after 5 h. In the case of BiOBr/ZnO/WO3, the efficiency of removal (R = 90%, k = 0.0065 min–1, and capacity = 27 mg/g) was significantly more than that of ZnO and WO3, indicating that the combination of BiOBr increased the catalytic activity. The p–n heterojunction semiconductor is a convenient system for relatively reliable charge collection and isolation. Typically, the generated electron–hole pairs could be rapidly isolated by the electric field in the charge carriers when the p–n heterojunction was irradiated by photons. Electrons were diverted to the CB of negative-type semiconductors, driven by an electric field, and holes were diverted to the VB of the positive-type transistors. (29) The Fermi level of positive-type BiOBr moved up when positive-type BiOBr and negative-type ZnO and WO3 were in contact, while that of n-type ZnO and WO3 moved down until the balance state was established. The entire energy band of positive-type BiOBr increased, compatible with the motion of the Fermi level, whereas that of negative-type ZnO and WO3 reduced. Consequently, a charging area was formed in functionalities from negative-type ZnO and WO3 to positive-type BiOBr. Both ZnO and WO3 may simulate to create electron–hole pairs during visible-light irradiation. BiOBrʼs photogenerated electrons in the CB would be transported to ZnO and WO3, whereas the holes of ZnO and WO3 in the VB would be transported to BiOBr. The built-in electrical field can endorse such transfers of photogenerated carriers. Therefore, as a consequence of the development of positive–negative heterojunction among positive-type BiOBr and negative-type ZnO and WO3 interfaces, the photogenerated electrons and holes split up successfully. This BiOBr/ZnO/WO3 p–n heterojunction can therefore create VB holes (ZnO and WO3) with a high oxidation activity and CB electrons (BiOBr) with a high reduction activity sequentially and effectively. Photocatalytic reactions may be involved in a greater amount of electrons on the BiOBr surface and holes on the ZnO and WO3 surface, leading to the understanding of the visible-light photocatalytic activity of the BiOBr/ZnO/WO3 p–n heterojunction. (30) From the above results, it can be concluded that BiOBr/ZnO/WO3 is an excellent catalyst for further experimental analysis of degradation.
Table 2. Kinetic Parameters and Correlation Coefficients for the Pseudo-First-Order and Pseudo-Second-Order Models
  Pseudo-first-order modelPseudo-second-order model
 catalystK1 (min–1)R2K2 (g mg–1 min–1)R2
dark studyZnO0.00190.9980.000040.729
 WO30.00140.9850.001900.282
 BiOBr0.00230.9750.002330.788
 BiOBr/ZnO0.00240.9730.003500. 882
 BiOBr/WO30.00400.9670.007910.946
 BiOBr/ZnO/WO30.00620.9770.000960.945
light studyZnO0.00250.9700.001970.945
 WO30.00150.9800.000770.883
 BiOBr0.00290.9600.001140.927
 BiOBr/ZnO0.00370.9900.001010.976
 BiOBr/WO30.00460.9900.000950.939
 BiOBr/ZnO/WO30.00650.9950.001030.972

Optimization Analysis and Design of Experiments

The central composite design, CCD, is a scientific and computational technique that is suitable for modeling tests, multivariate modeling with visual aspect measurements of deferent parameters. Herein, the impact of degradation variables, for instance, the potential of hydrogen (pH), initial monophenol concentration, and BiOBr/ZnO/WO3 dose, on the deterioration efficiency of phenol has been analyzed. The real benefit of CCD is to minimize the number of tests to be held on, also specifying the correlation of operational conditions in addition to maximizing the parameters of degradation.

Combined Influence of Initial Monophenol Concentration and pH on the Deterioration Activity

The initial reactant concentration is recognized as one of the significant parameters of photodegradation of organic compounds. Figure 8A shows how the photodegradation efficiency is affected by the combined effect of initial phenol concentration and solution pH. Evidently, with the increase of the initial phenol concentration from 10 to 200 mgL–1, the degradation efficiency reduced; the degradation efficiency reached 95 percent at an initial concentration of 10 mg L–1 of phenol and a pH of 9, which is ascribed to the existence of adequate surface area of the catalyst employed, and the reaction of phenol molecules with OH increased. (31) At an initial concentration of 200 mg L–1 and a pH of 9, the degradation efficiency dropped to 52.50 percent. Such related to the possibility of monophenol molecules interacting with OH declines; therefore, excessive oxidant radicals involving OH and O2 have been needed for improving the oxidation procedure and a lower efficiency of degradation is identified. (32) In the photocatalytic process, pH is the core parameter that influences the charging characteristics of the surface of the catalyst, followed by the effects of the sorption activity that influences the photocatalytic degradation. (27,36) As seen in Figure 8A, as the pH value increased, there is a clear increase in phenol photodegradation. At a pH of 9 with an initial monophenol concentration of 10 mg L–1, the optimum degradation value (95%) was obtained, although the minimal degradation value (42.80%) was achieved at a pH of 3 with an initial phenol concentration of 200 mg L–1. This is related to the greater H+ content that in the case of lower pH, it improves the functional protonation of the catalyst surface. In the meantime, a massive amount of H+ competes for the catalyst binding site with phenol. (28) The efficiency of phenol removal steadily increased once the concentration of H+ in the solution declined. Therefore, the favorite pH for phenol degradation was found to be 9. (33,34)

Figure 8

Figure 8. (A) 3D diagram of the combined effect of pH and the initial concentration of phenol for the degradation efficiency of phenol at a BiOBr/ZnO/WO3 dose of 0.08 g/0.03 L, (B) 3D diagram of the combined effect of pH and dose of BiOBr/ZnO/WO3 for the degradation efficiency of phenol at an initial phenol concentration of 10 mg L–1 and a contact time of 300 min, (C) proposed photocatalytic degradation pathway of phenol under visible light irradiation, and (D) total organic carbon (TOC) content in phenol suspension vs visible light irradiation time after degradation with BiOBr/ZnO/WO3 at the initial phenol concentration of 50 mg/L, catalyst dose of 0.08 g/0.03 L, and a pH of 9.

Combined Effect of BiOBr/ZnO/WO3 Dose and pH on the Degradation Efficiency

Photocatalytic studies with various BiOBr/ZnO/WO3 photocatalyst doses of 0.02, 0.05, and 0.08 g per 0.03 L were carried out. An increase in the photocatalytic dose showed a significant enhancement in the photocatalytic degradation efficiency (Figure 8B), reaching an optimum of 95% degradation at 0.08 g/0.03 L catalyst dose, the pH of 9, and an initial monophenol concentration of 10 mg L–1. The improvement of photocatalytic efficiency with the increase of the BiOBr/ZnO/WO3 photocatalyst dose may be attributed to the fact that the increase of the photocatalyst’s available surface area and the creation of more OH radicals at the active sites would be sufficient to degrade and mineralize the organic pollutant. (31,35,37)
The photocatalytic degradation pathway for phenol under visible-light irradiation in the presence of BiOBr/ZnO/WO3 is depicted in Figure 8C, showing the complete degradation process of phenol. The radical produced on the surface of the BiOBr/ZnO/WO3 photocatalyst reacts with phenol to produce intermediates, and with continuous photo-oxidation, the benzene ring can open, leading to the formation of an aliphatic compound that mineralized into carbon dioxide and water.

TOC Content

The degradation of the organic molecule to CO2 and H2O is the final goal of the photocatalytic process of organic pollutants, where the TOC content at a given irradiation time is reduced. The organic molecule breaks down into intermediate products, which are finally mineralized through the oxidative attack of the hydroxyl radical or by the direct attack of the photogenerated holes. Figure 8D implies that the initial content of TOC (38.2 mg/L) in the phenol suspension decreased gradually with visible light irradiation time. However, the remaining amount in the suspension after photocatalytic treatment of BiOBr/ZnO/WO3 for 300 min of visible light irradiation decreased to 4.58 mg/L with 88% degradation in the TOC content. These results emphasize the degradation of phenol intermediates to CO2 and H2O.

Analysis of Variance Study

Analysis of variance (ANOVA) study for the analysis of the quadratic surface and the conditions of the model is presented in Table 3. The ANOVA study covers the degrees of freedom (DF), that is, the amount of freedom to vary values at the end of the computation of the statistics. Fischer’s F-value test is the ratio between the mean square of the model and the residual error where higher F-values with a P-value (probability) are justified as the lowest, as applicable by a highly significant regression model. Also, the probability results greater than 0.05 imply a nonsignificant impact, whereas the results below 0.05 imply a significant impact. (37,38) The correlation between the practical results and the expected results was quantitatively evaluated with the coefficient of regression R2. Adjusted R2 is also a monitor of a fit’s perfection, correcting the R2 value in the model for the sample by utilizing the DF in its calculations if the adjusted R2 is much lower than R2; as a result, the model is not fitted. (36) The R2 results implied that the high BiOBr/ZnO/WO3 phenol degradation coefficient (R2 = 0.999) was close to the adjusted R2 values (0.998), indicating a significant correlation among the actual and anticipated values and showed that the quadratic mods utilized highly adequate values to achieve the expected results. Model F-values (36.94) also demonstrated that the statistical analysis was appropriate by evaluating probability values (0.0005) lower than 0.05, implying that the model is essential. A nonsignificant number of designs lacking fit (0.200) also demonstrated that the designs created were realistic. The ideal conditions for phenol degradation from the model are the catalyst dose of 0.08 g/0.03 L, pH of 9, and initial phenol concentration of 10 mg L–1 with a predicted 94.40 percent efficiency of phenol degradation. Although the initial efficiency of phenol degradation was 95%, which exemplifies the precision of the models, the optimal phenol degradation capacity of 39.37 mg g–1 was achieved under conditions of 0.08 g/0.03 L catalyst dose, a pH of 9, and 200 mg L–1 initial phenol concentration.
Table 3. ANOVA Study of the Response Surface for the Degradation of Phenol by BiOBr/ZnO/WO3
catalystsourceDFSSMSF-valueprob.> F
BiOBr/ZnO/WO3model104068.3406.8036.940.0005
 error51.81400.3628  
 lack of fit41.81400.4536 0.2000
 pure error100  
 total (model + error)154070.20   
The plots of expected results versus actual BiOBr/ZnO/WO3 phenol degradation results are shown in Figure 9A; the results were found to be well dispersed near the straight line, expecting an incredible link between the practical and expected output results. The accuracy of the designs was also evaluated and is presented in Figure 9B with the residual plot against the run order; the residuals were non-uniformly dispersed about ±2.50 for the adsorbent. This implied an effective fitting of the model with the practical data. The normal residual probability plot is a graphical description, a diagnostic tool to assess the deviation from the hypothesis whether errors are uniformly distributed or not, where the points will follow a straight line. However, with normal data, some scattering is anticipated. (23)Figure 9C shows the plots of normal residual probability, wherein we can clearly observe that the results were uniformly scattered as the points in the plot form a fairly straight line. For more discussion of the data obtained, Pareto graphical analysis was applied, as depicted in Figure 9D. According to the Pareto diagram, the percentage of efficacy for each phenol degradation variable is in the following order: quadratic effect of initial phenol concentration (23.45 percent) > initial phenol concentration (12.92 percent) > solution pH (12 percent) > catalyst dose used (5.40 percent) > solution pH (12 percent) > catalyst dose used (5.40 percent).

Figure 9

Figure 9. A) Predicted values vs actual values of phenol degradation, (B) studentized residuals vs run order of phenol degradation, (C) normal probability plot of regular residuals for phenol degradation, and (D) phenol Pareto graphic analysis.

T-Test

The T-test has been utilized for investigating the importance of the parameter for the correlation coefficient. A high t-value combined with a small probability value (p < 0.05) of the parameters implies that the corresponding model term is of high significance. (37) In Table 4, the t values, probability values in addition to coefficients are listed, providing all data achieved by Design-Expert software (V.11). The results indicate that the most important parameter is the pH, backed by the monophenol concentration. The coefficient of interaction between the pH and phenol concentration is also significant, followed by the correlation coefficient of the monophenol concentration with the dose of the catalyst.
Table 4. Estimated Effects and Coefficients
termcoefficientT ratioprob.
constant53.5933.3190.0009
A: pH12.006.72700.0214
B: catalyst dose6.7643.79100.0631
C: phenol concentration–11.525–6.67200.0217
AB0.2600.145710.8975
AC4.862–2.72500.0112
BC1.687–0.945410.0444
ABC2.572–1.4420.0286

Essential and Interaction Influences

Essential influences reflect the average variation for each variable between the maximum and small degrees. Unless a variable’s influence is positive, the efficiency of degradation increases and the factor shifts from low to high levels. (37)Figure 10A presents the main effects plot, where at a minimum degree of initial monophenol concentration (10 mg L–1), higher degrees of pH of 9, and a catalyst dose of 0.08 g, the maximum design was obtained. The negative value (C = −11.52) of the phenol initial concentration effect indicates that the quantity of deteriorated monophenol decreases in the case of the initial phenol concentration increasing from 10 to 200 mg L–1. The influences of pH and catalyst dose, fortunately, are positive, implying that the response increases with the increase of these two factors. Figure 10B shows an interaction plot; once the variation in the result from a low to a high level for a variable relies on the level of the second variable, the interaction is effective, that is, whenever the layers do not run parallel. (38,39) The interaction plot exhibits a significant interaction between various factors, where all the lines representing the impacts in the graph quadrants are not parallel. The interaction plot suggests a significant interaction between various factors, where all the lines representing the impacts in the graph quadrants are not parallel.

Figure 10

Figure 10. (A) Main effects of different parameters on phenol deterioration by BiOBr/ZnO/WO3 and (B) interaction effects of different parameters on phenol deterioration by BiOBr/ZnO/WO3.

Suggested Photocatalytic Mechanism

For revealing the photodegradation mechanism of phenol by BiOBr/WO3/ZnO, the formation of oxidative radicals was investigated by adding trapping agents to the phenol solution, such as benzoquinone (BQ) for scavenging the superoxide radical O2 and isopropanol for scavenging the hydroxyl radical OH. The results shown in Figure 11A indicate that the addition of the BQ trapping agent significantly suppressed the degradation of phenol to 31.60%, while the addition of the isopropanol trapping agent reduced the degradation of phenol to 44.80%, proving that the superoxide radical O2 and the hydroxyl radical OH play a decisive role in the degradation of phenol by BiOBr/WO3/ZnO.

Figure 11

Figure 11. (A) Effect of scavengers on the photodegradation efficiency of phenol, (B) suggested photocatalytic degradation mechanism of phenol with BiOBr/WO3/ZnO, (C) photocatalytic stability of BiOBr/WO3/ZnO toward phenol degradation, (D) photocatalytic system efficiency for the degradation of phenol and malachite green at the initial concentration, 50 mg/L; BiOBr/WO3/ZnO and TiO2 P25 dose, 0.05 g/0.03 L; and pH, 6, and (E) realistic application of refinery petroleum wastewater.

Based on the above findings, the proposed photocatalytic mechanism of phenol degradation using BiOBr/WO3/ZnO is illustrated in Figure 11B. ZnO, WO3, and BiOBr can be activated easily at simulated visible-light illumination which led to the generation of electrons and holes. When the CB electrons of BiOBr were transferred to the CBs of ZnO and WO3, these excited electrons showed a reduction behavior and react with oxygen O2 to produce the superoxide radical O2 and then simultaneously caused the transfer of VB holes of ZnO and WO3 to the VB of BiOBr. Positive holes can oxidize phenol directly or react with H2O generating hydroxyl radicals OH; generally, the degradation of phenol occurs by oxidation with O2 and OH radicals. The OH radical is the most efficient oxidizing species, and it can degrade organic contaminants on the photocatalyst surface. In such a way, it is possible to effectively separate photogenerated electrons and holes and to effectively impede the recombination of electron–hole pairs, which is one of the main elements that can improve the photocatalytic activity. Photocatalytic reactions can result in a greater quantity of electrons on the BiOBr surface and holes on the ZnO and WO3 surfaces, leading to the superior visible-light photocatalytic activity of the ternary photocatalyst BiOBr/ZnO/WO3.

Catalytic Stability Study

The recyclability and stability of the BiOBr/ZnO/WO3 photocatalyst toward the degradation of aqueous solutions of phenol were investigated. Figure 11C illustrates the reuse of BiOBr/ZnO/WO3. For five cycles after desorption with ethanol, the photocatalytic degradation efficiency of phenol after the first cycle was 95%, which then decreased to 90% in the fifth cycle. This indicates that the decrease of BiOBr/ZnO/WO3 photocatalytic activity is negligible and emphasizes the long-term stability of the prepared BiOBr/ZnO/WO3 photocatalyst.
Based on the degradation efficiency (%) and degradation capacity, a comparative study of the current work and various photocatalysts for the degradation of phenol was done and is discussed in Table 5. The findings listed in the table demonstrate that BiOBr/ZnO/WO3 is a promising photocatalyst compared to other photocatalysts. Therefore, BiOBr/ZnO/WO3 is considered a promising catalyst for the simultaneous degradation of phenol in aqueous solutions.
Table 5. Comparative Study of Photocatalytic Degradation Efficiency (%) and Degradation Capacity of Phenol with Various Photocatalysts
adsorbentsdegradation %degradation capacity, q(mg/g)references
N-doped meso-titania69.228 (2)
TiO2/Gr/30PW91.045.5 (35)
TiO2/rGO90.022.7 (36)
Ag3PO4–Ag2CO373.924.6 (37)
BiOBr/WO3/ZnO95.039.37current work

Investigation of Photocatalytic System Efficiency

The photocatalytic efficiency of the BiOBr/ZnO/WO3 ternary nanocomposite was investigated toward the degradation of malachite green in addition to phenol, and the results were compared with the efficiency of commercial TiO2 P25 nano-powder toward the photodegradation of phenol and malachite green under the same experimental conditions. The results depicted in Figure 11D indicate that the efficiency of BiOBr/ZnO/WO3 toward the degradation of malachite green (93.6%) is greater than that for the degradation of phenol (90%), while the commercial TiO2 P25 nano-powder has a lower efficiency toward the degradation of both malachite green and phenol that reached 74 and 51.55%, respectively, which emphasize that BiOBr/ZnO/WO3 has a superior photocatalytic activity than the commercial TiO2 P25 nano-powder.

Realistic Application of Refinery Petroleum Wastewater

The efficiency of ternary photodegradable BiOBr/ZnO/WO3 was investigated for the treatment of real petroleum refinery wastewater samples. The analytical characterization results of the collected refinery wastewater sample before treatment were as follows: pH, 7.9 ± 0.2; total dissolved solids (TDS), 2.48 ± 0.4 g/L; phenol, 35 ± 0.5 mg/L; chemical oxygen demand (COD), 750 ± 0.8 mg/L; and biological oxygen demand (BOD), 210 ± 0.6 mg/L.
The results depicted in Figure 11E imply that the phenol content of petroleum wastewater decreased drastically with irradiation time rather than the COD and BOD contents, where the removal efficiency of phenol reached 89% after 300 min of irradiation, while that of COD and BOD reached 80.2 and 42.6%, respectively. These results imply the effectiveness of BiOBr/ZnO/WO3 for the treatment of organic pollutants. Generally, these findings revealed that BiOBr/ZnO/WO3 has the potency to be utilized in the eco-friendly treatment of refinery wastewater with high photocatalytic activity.

Conclusions

Click to copy section linkSection link copied!

Herein, single-nanophase ZnO, WO3, and BiOBr were synthesized, in addition to their binary- and ternary-phase BiOBr/ZnO, BiOBr/WO3, and BiOBr/ZnO/WO3 p–n heterojunctions, by the solvothermal technique. Their catalytic efficiency for phenol degradation was assessed under visible-light irradiation. The results indicate that the ternary BiOBr/ZnO/WO3 p–n heterojunction exhibits the highest photocatalytic activity that reached the degradation equilibrium after 60 min with an efficiency of 90%. RSM was used for optimizing the impact of degradation factors. Under the optimum conditions of 0.08 g/0.03 L catalyst dose, the pH of 9, and an initial monophenol concentration of 10 mg L–1, the degradation efficiency reached 95%. Also, the maximum phenol degradation capacity (39.37 mg g–1) was achieved under the conditions of 0.08 g/0.03 L catalyst dose, pH of 9, and 200 mg L–1 initial phenol concentration. The radical scavenging test proved that the superoxide radical O2 and the hydroxyl radical OH play a decisive role in the degradation of phenol by BiOBr/WO3/ZnO.

Experimental Section

Click to copy section linkSection link copied!

Materials and Methods

All the reagents utilized in this paper were obtained from Shanghai Chemical Company and were of analytical purity, and demineralized water was utilized for all preparations.

Synthesis of WO3 NPs

Sodium tungstate dihydrate powder (Na2WO4·2H2O, 0.003 mol) was first mixed in demineralized water (20 mL). (40) The final pH of the starting liquid was changed in two phases. First, by magnetic stirring, the pH limit for the liquid was acidified to 1.58 by hydrochloric acid (HCl, 37%) and then a white precipitate was formed. Finally, the liquid was diluted to 40 mL and then the mixture was mixed with oxalic acid (H2C2O4, 0.022 mol) to change the end pH to 2. Sodium chloride (NaCl, 0.068 mol) was added after stirring for 30 min and then heated at 90 °C for 3 h. The as-deposited substrate was then rinsed after the reaction was completed; the prepared salt was subsequently washed with demineralized water repeatedly, then calcined at 500 °C, and dried in air.

Preparation of Nano ZnO

Nano ZnO was prepared as follows: (41) 1.48 g of Zn(CH3COO)2·2H2O (Sigma-Aldrich, India) was dissolved in 63 mL of absolute ethanol in a 250 mL Schott bottle and heated with continuous stirring at 60° C. 0.74 g of NaOH was also dissolved separately in 33 mL of absolute ethanol in a 100 mL Schott bottle under the same Zn(CH3COO)2·2H2O condition. After both solutions had been fully dissolved, NaOH was added slowly to Zn(CH3COO)2·2H2O with vigorous stirring at 60 °C. The solution mixture was left to stand for 3 h before the reaction was finished. A white precipitate (ZnO) was produced and collected within 10 min by centrifugation at 4000 rpm and washed twice with acetone and once with ultrapure water to eliminate all the impurities. The product obtained was then dried at room temperature and then calcined at 500 °C and ground to form a powder.

Synthesis of Nano BiOBr and BiOBr/ZnO/WO3 Ternary Photocatalysts

The hydrolysis solvothermal process method was used for the preparation of BiOBr nanocrystals. Generally, to 0.035 L of 1,2-ethanediol (EG), 2.43 g of bismuth nitrate pentahydrate was added followed by 0.595 g of KBr with vigorous stirring until a transparent solution was obtained. Then, at room temperature, 0.3 L of demineralized water was gradually dropped into a transparent solution, and plenty of precipitate was immediately obtained. The obtained precipitate was purified, rinsed repeatedly with demineralized water, and then allowed to dry at 80 °C. (42) Typically, for BiOBr/ZnO/WO3 synthesis, 2.43 g of bismuth nitrate pentahydrate was added to 0.035 L of 1,2-ethanediol and stirred until the mixture became clear. Then, with vigorous stirring, 0.595 g of KBr was added and continuously stirred until the mixture became transparent. Subsequently, stoichiometric quantities of ZnO and WO3 with proportions of 4BiOBr/1ZnO/1WO3 (0.607 g equal to 25 wt% of bismuth nitrate pentahydrate) were dissipated by ultrasonication in 0.30 L of deionized water and stepwise added to the transparent solution at room temperature. The resultant mixture was collected, rinsed with demineralized water, and dried at 80 °C in air. (43,44)

Equipment

A Shimadzu XRD-600 diffractometer operating at a voltage of 40 KV and a current of 20 mA was utilized to record the patterns of XRD at both low angles (2θ range, 0.5–10) and wide angles (2θ range, 10–80) with Ni-filtered Cu Kα radiation (λ = 1.54° A). An FTIR spectrophotometer (Perkin-Elmer 720) was used to perform FTIR analysis. The KBr technique was applied nearly in a quantitative manner for all samples under investigation since the sample weight and that of KBr were fixed constant. XPS spectra were collected on K-ALPHA (Thermo Fisher Scientific, USA) with monochromatic X-ray Al K-α radiation of −10–1350 eV and a spot size of 400 μm at a pressure of 10–9 mbar with a full-spectrum pass energy of 200 eV and at a narrow-spectrum of 50 eV. Analysis by SEM coupled with EDS (SEM–EDX) was done using a scanning electron microscope JEOL GSM 6510LV. A transmission electron microscope (JEOL2100, 200 kV) was utilized to observe the particle size and structure morphologies of the synthesized photocatalysts. The photoelectrochemical measurements were conducted with an OrigaFlex-Pack OGF01A/potentiostat–galvanostat (Origalys, France) system. A UV–vis spectrophotometer (Jasco V-507) coupled with a Shimadzu IRS-2200 was utilized to perform UV–vis DRS and for the detection of malachite green dye concentration at λmax = 617 nm. The PL spectral analysis was done using a PL spectrometer (HITACHI F7000). For the quantitative analysis of the reaction intermediates, TOC analysis was employed using a high-sensitivity Shimadzu apparatus (4 μg C L–1 to 25,000 mg C L–1), model TOC-VCSH.

High-Performance Liquid Chromatography Analytical Procedure

High-performance liquid chromatography (HPLC) was used to analyze the resulting solutions after the reaction time has ended. An Agilent 1200 series (Waldbronn, Germany) HPLC devise model outfitted with an auto-sampler and a photodiode array detector (wavelength set at 190–400 nm in the full UV scan range) was used to analyze the phenol limits. A constant flow level of 1.0 mL/min with a reversed column phase C8 (4.6 × 25 cm, 300 A°, and 5 μm) was used. For elution, the mobile phase consisting of a 40 percent acetonitrile and 60 percent water mixture was utilized. Samples of the solution were taken at different reaction times to evaluate the phenol concentration and then photocatalysts were filtered and collected at 5000 rpm in a centrifuge.

Photocatalytic Performance

The degradation of phenol was utilized to evaluate the photocatalytic efficiency of the synthesized photocatalysts. A batch system was used to perform all experimental runs, where 0.05 g of each prepared photocatalyst was put into 30 mL of phenol contaminants (50 mg/l) at ambient temperature (25 ± 2 °C); first, all experimental runs were shaken in the dark for 1 h to reach the adsorption–desorption equilibrium and then visible-light illumination (tungsten lamp 300 W, λ > 400 nm) was switched on for 300 min. Samples were collected at different times, and the residual phenol concentration was assessed using the HPLC instrument. The percentage of phenol removal (R %) and the capacity for adsorption (qt, mg g–1) were determined using eqs 4 and 5, respectively. (24,25)
(4)
(5)
where the sorption capacity of the sorbent at intervals t (mg/g) is represented as qt, the starting phenol concentration as Co (mg/L), the remaining phenol concentration at intervals t as Ct (mg/L), the catalyst mass as m (g), and the quantity of phenol solution as V (L).
Relying on the solution concentration, kinetic information was gained by studying the pseudo-first-order kinetic system (eq 6), while depending upon the sorption equilibrium capacity, the pseudo-second-order kinetic model (eq 7) was studied. (26,27)
(6)
(7)
where the first-order rate constant is given by K1 (min–1), the second-order rate constant by K2 (g mg–1 min–1), and the quantity of phenol uptake at equilibrium by qe.
The conversion of phenol intermediates is represented by TOC as the instant TOC and TOC0, the initial TOC.

Factorial Design and Optimization

Depending on the CCD, a simplified RSM has been included to optimize the effect of other variables such as the pH, catalyst dose, and starting phenol concentration on the degradation efficiency. The verity of parameters that indicate the optimum phenol degradation efficiency was analyzed using the CCD model. Each degradation factor has three elevated levels designated as +1, 0, and −1. In the model, the catalyst dose was within 0.02–0.08 g/30 mL, the pH ranged from 3 to 9, and the initial phenol concentration varied from 10 to 200 mg L–1. Design-Expert software (V.11) was used to analyze statistical data. All the optimized tests were shaken in the dark for 1 h until the sorption balance was achieved and then the tests were conducted under visible light for 300 min. The phenol degradation efficiency behavior was described by the following polynomial equation of second order
(8)
where Y is the response variable, βo, βi, βij, and βii are the regression coefficients for the intercept, linear effect, double interaction, and quadratic effects, respectively, and Xi and Xj are the independent variables. (45)

Realistic Application of Refinery Petroleum Wastewater

A real wastewater sample was collected from a refinery petroleum company in Egypt. The TDS, pH, and organic constituents of the sample were characterized using standard methods (e.g., American Society for Testing and Materials and American Public Health Association Guidelines, 20th edition) for the examination of water and wastewater. For the degradation treatment, 1 g/L of BiOBr/ZnO/WO3 nanocomposite was dispersed in 100 mL of collected wastewater sample and irradiated for 300 min under visible light. At the end of the degradation time, the treated wastewater samples were separated by filtration, and the treatment efficiency was determined based on the decrease of organic pollutants, including phenol, COD (mg/L), and BOD (mg/L) at different time intervals, where COD and BOD are representatives of organic pollutants in wastewater.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Author
  • Author
    • Hager R. Ali - Spectroscopic Division, Analysis and Evaluation Department, Egyptian Petroleum Research Institute (EPRI), Nasr City 11727 Cairo, Egypt
  • Author Contributions

    H.R.A. contributed equally.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

All authors gratefully thank the Egyptian Petroleum Research Institute staff for their encouragement and for providing analysis.

References

Click to copy section linkSection link copied!

This article references 45 other publications.

  1. 1
    Alnaizy, R.; Akgerman, A. Advanced oxidation of phenolic compounds. Adv. Environ. Res. 2000, 4, 233244,  DOI: 10.1016/S1093-0191(00)00024-1
  2. 2
    Ahmed, S.; Rasul, M. G.; Martens, W. N.; Brown, R.; Hashib, M. A. Heterogeneous photocatalytic degradation of phenols in wastewater: A review on current status and developments. Desalination 2010, 261, 318,  DOI: 10.1016/j.desal.2010.04.062
  3. 3
    Karunakaran, C.; Dhanalakshmi, R. Semiconductor-catalyzed degradation of phenols with sunlight. Sol. Energy Mater. Sol. Cells 2008, 92, 13151321,  DOI: 10.1016/j.solmat.2008.05.002
  4. 4
    Reynolds, D. C.; Look, D. C.; Jogai, B. Optically pumped ultraviolet lasing from ZnO. Solid State Commun. 1996, 99, 873875,  DOI: 10.1016/0038-1098(96)00340-7
  5. 5
    Janotti, A.; Van De Walle, C. G. Fundamentals of zinc oxide as a semiconductor. Rep. Prog. Phys. 2009, 72, 126501,  DOI: 10.1088/0034-4885/72/12/126501
  6. 6
    Wang, S. L.; Zhu, Y.; Luo, X. 2D WC/WO3Heterogeneous Hybrid for Photocatalytic Decomposition of Organic Compounds with Vis-NIR Light. Adv. Funct. Mater. 2018, 28, 1705357,  DOI: 10.1002/adfm.201705357
  7. 7
    Xia, J.; Di, J.; Yin, S. Solvothermal synthesis and enhanced visible-light photocatalytic decontamination of bisphenol A (BPA) by g-C3N4/BiOBr heterojunctions. Mater. Sci. Semicond. Process. 2014, 24, 96103,  DOI: 10.1016/j.mssp.2014.02.036
  8. 8
    Motawea, E. A.; Ali, H. R. Exploitation of Heterostructure Photocatalyst Impregnated Alginate Beads for Enhanced Removal of Methyl Orange under Solar Light. Petrol. Petrochem. Eng. J. 2018, 2, 110,  DOI: 10.23880/ppej-16000180
  9. 9
    Berger, S.; Tsuchiya, H.; Ghicov, A.; Schmuki, P. High photocurrent conversion efficiency in self-organized porous WO3. Appl. Phys. Lett. 2006, 88, 203119,  DOI: 10.1063/1.2206696
  10. 10
    Momeni, M. M.; Hakimian, M.; Kazempour, A. In-Situ Manganese Doping of TiO2 Nanostructures via Single-Step Electrochemical Anodizing of Titanium in an Electrolyte Containing Potassium Permanganate: A Good Visible-Light Photocatalyst. Ceram. Int. 2015, 41, 13692,  DOI: 10.1016/j.ceramint.2015.07.158
  11. 11
    Momeni, M. M.; Nazari, Z. Preparation of TiO2 and WO3-TiO2 nanotubes decorated with PbO nanoparticles by chemical bath deposition process: A stable and efficient photo catalyst. Ceram. Int. 2016, 42, 86918697,  DOI: 10.1016/j.ceramint.2016.02.103
  12. 12
    Zeng, Q.; Li, J.; Bai, J.; Li, X.; Xia, L.; Zhou, B. Preparation of vertically aligned WO3 nanoplate array films based on peroxotungstate reduction reaction and their excellent photoelectrocatalytic performance. Appl. Catal., B 2017, 202, 388396,  DOI: 10.1016/j.apcatb.2016.09.045
  13. 13
    Zeng, X.; Wang, Z.; Wang, G. Highly dispersed TiO2 nanocrystals and WO3 nanorods on reduced graphene oxide: Z-scheme photocatalysis system for accelerated photocatalytic water disinfection. Appl. Catal., B 2017, 218, 163173,  DOI: 10.1016/j.apcatb.2017.06.055
  14. 14
    Momeni, M. M.; Ghayeb, Y.; Davarzadeh, M. Electrochemical construction of different titania-tungsten trioxide nanotubular composite and their photocatalytic activity for pollutant degradation: a recyclable photocatalysts. J. Mater. Sci.: Mater. Electron. 2015, 26, 15601567,  DOI: 10.1007/s10854-014-2575-x
  15. 15
    Momeni, M. M.; Ghayeb, Y.; Ghonchegi, Z. Fabrication and characterization of copper doped TiO2 nanotube arrays by in situ electrochemical method as efficient visible-light photocatalyst. Ceram. Int. 2015, 41, 87358741,  DOI: 10.1016/j.ceramint.2015.03.094
  16. 16
    Lei, R.; Ni, H.; Chen, R.; Zhang, B.; Zhan, W.; Li, Y. Growth of Fe 2 O 3 /SnO 2 nanobelt arrays on iron foil for efficient photocatalytic degradation of methylene blue. Chem. Phys. Lett. 2017, 673, 16,  DOI: 10.1016/j.cplett.2017.01.052
  17. 17
    Tsay, C.-Y.; Fan, K.-S.; Chen, S.-H.; Tsai, C.-H. Preparation and characterization of ZnO transparent semiconductor thin films by sol-gel method. J. Alloys Compd. 2010, 495, 126130,  DOI: 10.1016/j.jallcom.2010.01.100
  18. 18
    Sakthivel, S.; Neppolian, B.; Shankar, M. V.; Arabindoo, B.; Palanichamy, M.; Murugesan, V. Solar photocatalytic degradation of azo dye: Comparison of photocatalytic efficiency of ZnO and TiO2. Sol. Energy Mater. Sol. Cells 2003, 77, 6582,  DOI: 10.1016/S0927-0248(02)00255-6
  19. 19
    Li, D.; Haneda, H.; Ohashi, N.; Saito, N.; Hishita, S. Morphological reform of ZnO particles induced by coupling with MOx (M=V,W,Ce) and the effects on photocatalytic activity. Thin Solid Films 2005, 486, 2023,  DOI: 10.1016/j.tsf.2004.11.237
  20. 20
    El-saied, H. A.-a.; Motawea, E. A.-T. Optimization and Adsorption Behavior of Nanostructured NiFe2O4/Poly AMPS Grafted Biopolymer. J. Polym. Environ. 2020, 28, 23352351,  DOI: 10.1007/s10924-020-01774-z
  21. 21
    Lv, J.; Dai, K.; Zhang, J.; Liu, Q.; Liang, C.; Zhu, G. Facile constructing novel 2D porous g-C3N4/BiOBr hybrid with enhanced visible-light-driven photocatalytic activity. Sep. Purif. Technol. 2017, 178, 617,  DOI: 10.1016/j.seppur.2017.01.019
  22. 22
    Huang, W.; Hua, X.; Zhao, Y. Enhancement of visible-light-driven photocatalytic performance of BiOBr nanosheets by Co2+ doping. J. Mater. Sci.: Mater. Electron. 2019, 30, 1496714976,  DOI: 10.1007/s10854-019-01869-x
  23. 23
    Khampuanbut, A.; Santalelat, S.; Pankiew, A. Visible-light-driven WO3/BiOBr heterojunction photocatalysts for oxidative coupling of amines to imines: Energy band alignment and mechanistic insight. J. Colloid Interface Sci. 2020, 560, 213224,  DOI: 10.1016/j.jcis.2019.10.057
  24. 24
    Yang, H.; Zhang, Q.; Chen, Y.; Huang, Y.; Yang, F.; Lu, Z. Ultrasonic-microwave synthesis of ZnO/BiOBr functionalized cotton fabrics with antibacterial and photocatalytic properties. Carbohydr. Polym. 2018, 201, 162171,  DOI: 10.1016/j.carbpol.2018.08.068
  25. 25
    Tae Kwon, Y.; Yong Song, K.; In Lee, W.; Jin Choi, G.; Rag Do, Y. Photocatalytic behavior of WO3-loaded TiO2 in an oxidation reaction. J. Catal. 2000, 191, 192199,  DOI: 10.1006/jcat.1999.2776
  26. 26
    Keller, V.; Bernhardt, P.; Garin, F. Photocatalytic oxidation of butyl acetate in vapor phase on TiO2, Pt/TiO2 and WO3/TiO2 catalysts. J. Catal. 2003, 215, 129138,  DOI: 10.1016/S0021-9517(03)00002-2
  27. 27
    El-Salamony, R. A.; Amdeha, E.; Badawy, N. A.; Ghoneim, S. A.; Al-Sabagh, A. M. Visible light sensitive activated carbon-metal oxide (TiO2, WO3, NiO, and SnO) nano-catalysts for photo-degradation of methylene blue: a comparative study. Toxicol. Environ. Chem. 2018, 100, 143156,  DOI: 10.1080/02772248.2018.1497634
  28. 28
    Amdeha, E.; Mohamed, R. S. A green synthesized recyclable ZnO/MIL-101(Fe) for Rhodamine B dye removal via adsorption and photo-degradation under UV and visible light irradiation. Environ. Technol. 2019, 42, 842859,  DOI: 10.1080/09593330.2019.1647290
  29. 29
    Wang, H.; Zhang, L.; Chen, Z. Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 52345244,  DOI: 10.1039/c4cs00126e
  30. 30
    Zhang, J.; Zhang, L.; Shen, X.; Xu, P.; Liu, J. Synthesis of BiOBr/WO3p-n heterojunctions with enhanced visible light photocatalytic activity. CrystEngComm 2016, 18, 38563865,  DOI: 10.1039/c6ce00824k
  31. 31
    Rafiee, E.; Noori, E.; Zinatizadeh, A. A.; Zanganeh, H. Photocatalytic degradation of phenol using a new developed TiO2/graphene/heteropoly acid nanocomposite: Synthesis, characterization and process optimization. RSC Adv. 2016, 6, 9655496562,  DOI: 10.1039/c6ra09897e
  32. 32
    Ali, M. H. H.; Al-qahtani, K. M.; El-sayed, S. M. Enhancing photodegradation of 2,4,6 trichlorophenol and organic pollutants in industrial effluents using nanocomposite of TiO2 doped with reduced graphene oxide. Egypt. J. Aquat. Res. 2019, 45, 321328,  DOI: 10.1016/j.ejar.2019.08.003
  33. 33
    Ali, H. R. Photocoupling of Ag3PO4-Ag2CO3 with molecularly imprinted polymer for enhanced removal of phenol under solar light: Application of taguchi method. Int. J. Appl. Eng. Res. 2017, 12, 33533359
  34. 34
    Abd El Salam, H. M.; Younis, S. A.; Ali, H. R.; Zaki, T. Statistical modeling and optimization of phenol adsorption from water by modified Cu3(BTC)2: Kinetic, isotherm, and thermodynamic analysis. Microporous Mesoporous Mater. 2017, 241, 210217,  DOI: 10.1016/j.micromeso.2016.12.010
  35. 35
    Rodriguez, A. F. R.; Coaquira, J. A. H.; Morales, M. A. Synthesis, characterization and magnetic properties of polymer-Fe3O4 nanocomposite. Spectrochim. Acta, Part A 2013, 100, 101103,  DOI: 10.1016/j.saa.2012.02.081
  36. 36
    Mäkelä, M. Experimental design and response surface methodology in energy applications: A tutorial review. Energy Convers. Manage. 2017, 151, 630640,  DOI: 10.1016/j.enconman.2017.09.021
  37. 37
    Tebbouche, L.; Hank, D.; Zeboudj, S.; Namane, A.; Hellal, A. Evaluation of the phenol biodegradation byAspergillus niger: application of full factorial design methodology. Desalin. Water Treat. 2016, 57, 61246130,  DOI: 10.1080/19443994.2015.1053991
  38. 38
    Choquette-Labbé, M.; Shewa, W.; Lalman, J.; Shanmugam, S. Photocatalytic degradation of phenol and phenol derivatives using a Nano-TiO2 catalyst: Integrating quantitative and qualitative factors using response surface methodology. Water (Switzerland) 2014, 6, 17851806,  DOI: 10.3390/w6061785
  39. 39
    Ocholi, O.; Menkiti, M.; Auta, M.; Ezemagu, I. Optimization of the operating parameters for the extractive synthesis of biolubricant from sesame seed oil via response surface methodology. Egypt. J. Pet. 2018, 27, 265275,  DOI: 10.1016/j.ejpe.2017.04.001
  40. 40
    Tehrani, F. S.; Ahmadian, H.; Aliannezhadi, M. Hydrothermal synthesis and characterization of WO3 nanostructures: Effect of reaction time. Mater. Res. Express 2020, 7, 015911,  DOI: 10.1088/2053-1591/ab66fc
  41. 41
    Shamhari, N. M.; Wee, B. S.; Chin, S. F.; Kok, K. Y. Synthesis and characterization of zinc oxide nanoparticles with small particle size distribution. Acta Chim. Slov. 2018, 65, 578585,  DOI: 10.17344/acsi.2018.4213
  42. 42
    Yuan, Q. L.; Zhang, Y.; Yin, H. Y.; Nie, Q. L.; Wu, W. W. Rapid, simple and low-cost fabrication of BiOBr ultrathin nanocrystals with enhanced visible light photocatalytic activity. J. Exp. Nanosci. 2016, 11, 359369,  DOI: 10.1080/17458080.2015.1065014
  43. 43
    Liu, X.; Luo, J.; Zhu, Y.; Yang, Y.; Yang, S. Removal of methylene blue from aqueous solutions by an adsorbent based on metal-organic framework and polyoxometalate. J. Alloys Compd. 2015, 648, 986993,  DOI: 10.1016/j.jallcom.2015.07.065
  44. 44
    Younis, S. A.; Motawea, E. A.; Moustafa, Y. M.; Lee, J.; Kim, K.-H. A strategy for the efficient removal of chlorophenols in petrochemical wastewater by organophilic and aminated silica@alginate microbeads: Taguchi optimization and isotherm modeling based on partition coefficient. J. Hazard. Mater. 2020, 397, 122792,  DOI: 10.1016/j.jhazmat.2020.122792
  45. 45
    Bakr, A.-S. A.; Moustafa, Y. M.; Motawea, E. A.; Yehia, M. M.; Khalil, M. M. H. Removal of ferrous ions from their aqueous solutions onto NiFe2O4-alginate composite beads. J. Environ. Chem. Eng. 2015, 3, 14861496,  DOI: 10.1016/j.jece.2015.05.020

Cited By

Click to copy section linkSection link copied!

This article is cited by 7 publications.

  1. Iftekhar Ahmad, Mohammad Muneer, Abdelrahman S. Khder, Saleh A. Ahmed. Novel Type-II Heterojunction Binary Composite (CdS/AgI) with Outstanding Visible Light-Driven Photocatalytic Performances toward Methyl Orange and Tetracycline Hydrochloride. ACS Omega 2023, 8 (25) , 22708-22720. https://doi.org/10.1021/acsomega.3c01517
  2. Malathi Arumugam, Ravindranadh Koutavarapu, Kamala-Kannan Seralathan, Supareak Praserthdam, Piyasan Praserthdam. Noble metals (Pd, Ag, Pt, and Au) doped bismuth oxybromide photocatalysts for improved visible light-driven catalytic activity for the degradation of phenol. Chemosphere 2023, 324 , 138368. https://doi.org/10.1016/j.chemosphere.2023.138368
  3. Xiaolong Wu, Jing Xu, Pengfei Zhu, Mei Liu, Ming Duan, Shasha Zhang. High performance visible light response of a Z-type Bi 2 WO 6 /BiOBr/RGO heterojunction photocatalyst for the degradation of norfloxacin. Dalton Transactions 2022, 51 (47) , 17994-18009. https://doi.org/10.1039/D2DT03038A
  4. Paymaneh Taghizadeh-Lendeh, Amir Hossein Mohsen Sarrafi, Afshar Alihosseini, Naeimeh Bahri-Laleh. Degradation of methyldiethanolamine and gas refinery effluent using a TiO 2 @WO 3 /ZnO photocatalyst: central composite design optimization. New Journal of Chemistry 2022, 46 (38) , 18146-18156. https://doi.org/10.1039/D2NJ03769F
  5. Masoomeh Ghorbani, Ali Reza Solaimany Nazar, Mehrdad Frahadian, Mohsen Khosravi. Facile synthesis of Z-scheme ZnO-nanorod @ BiOBr-nanosheet heterojunction as efficient visible-light responsive photocatalyst: The effect of electrolyte and scavengers. Journal of Photochemistry and Photobiology A: Chemistry 2022, 429 , 113930. https://doi.org/10.1016/j.jphotochem.2022.113930
  6. Meysam Habibi, Aziz Habibi-Yangjeh, Shima Rahim Pouran, Cailing Xu, Chundong Wang. Binary visible-light-triggered ZnO/Bi4O5Br2 photocatalysts with n-n heterojunction: Simple fabrication and impressively activation of peroxodisulfate ions for degradation of tetracycline. Surfaces and Interfaces 2022, 32 , 102147. https://doi.org/10.1016/j.surfin.2022.102147
  7. Shreya Modi, Virendra Kumar Yadav, Amel Gacem, Ismat H. Ali, Dhruv Dave, Samreen Heena Khan, Krishna Kumar Yadav, Sami-ullah Rather, Yongtae Ahn, Cao Truong Son, Byong-Hun Jeon. Recent and Emerging Trends in Remediation of Methylene Blue Dye from Wastewater by Using Zinc Oxide Nanoparticles. Water 2022, 14 (11) , 1749. https://doi.org/10.3390/w14111749

ACS Omega

Cite this: ACS Omega 2021, 6, 34, 22047–22064
Click to copy citationCitation copied!
https://doi.org/10.1021/acsomega.1c02644
Published August 17, 2021

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

1780

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. XRD spectra of the synthesized WO3, ZnO, BiOBr, BiOBr/WO3, BiOBr/ZnO, and BiOBr/WO3/ZnO.

    Figure 2

    Figure 2. FTIR patterns of WO3, ZnO, BiOBr, BiOBr/WO3, BiOBr/ZnO, and BiOBr/WO3/ZnO.

    Figure 3

    Figure 3. XPS spectrum of BiOBr/ZnO/WO3: (a) survey spectrum, (b) Bi 4f spectrum, (c) Br 3d spectrum, (d) W 4f spectrum, (e) Zn 2p spectrum, and (f) O 1s spectrum.

    Figure 4

    Figure 4. Optical properties of the prepared photocatalysts: (A) UV–vis DRS spectra, (B) Taucʼs plots [(αhv)2 vs Hv] for WO3 and ZnO photocatalysts, (C) Taucʼs plots [(αhv)1/2vs Hv] for BiOBr, BiOBr/ZnO, BiOBr/WO3, and BiOBr/ZnO/WO3, and (D) PL spectra.

    Figure 5

    Figure 5. Photoelectrochemical measurements: (a) Nyquist plots of EIS, (b) EIS Bode plots, and (c) LSV curves for BiOBr and BiOBr/ZnO/WO3.

    Figure 6

    Figure 6. SEM images of the synthesized (A) ZnO NPs, (B) WO3NPs, (C) BiOBr NPs, and (D) BiOBr/ZnO/WO3 nanocomposites, TEM images of (E) ZnO NPs, (F) WO3NPs, (G) BiOBr NPs, and (H) BiOBr/ZnO/WO3, and (I) EDX spectra of BiOBr/ZnO/WO3.

    Figure 7

    Figure 7. (A) Degradation capacity of phenol vs time with different catalysts, (B) degradation percentage at an equilibrium time of 300 min, and (C) kinetics of the photodegradation of phenol using different photocatalysts at the initial concentration of 50 mg/L, dose of 0.05 g/0.03 L, and a pH of 6.

    Figure 8

    Figure 8. (A) 3D diagram of the combined effect of pH and the initial concentration of phenol for the degradation efficiency of phenol at a BiOBr/ZnO/WO3 dose of 0.08 g/0.03 L, (B) 3D diagram of the combined effect of pH and dose of BiOBr/ZnO/WO3 for the degradation efficiency of phenol at an initial phenol concentration of 10 mg L–1 and a contact time of 300 min, (C) proposed photocatalytic degradation pathway of phenol under visible light irradiation, and (D) total organic carbon (TOC) content in phenol suspension vs visible light irradiation time after degradation with BiOBr/ZnO/WO3 at the initial phenol concentration of 50 mg/L, catalyst dose of 0.08 g/0.03 L, and a pH of 9.

    Figure 9

    Figure 9. A) Predicted values vs actual values of phenol degradation, (B) studentized residuals vs run order of phenol degradation, (C) normal probability plot of regular residuals for phenol degradation, and (D) phenol Pareto graphic analysis.

    Figure 10

    Figure 10. (A) Main effects of different parameters on phenol deterioration by BiOBr/ZnO/WO3 and (B) interaction effects of different parameters on phenol deterioration by BiOBr/ZnO/WO3.

    Figure 11

    Figure 11. (A) Effect of scavengers on the photodegradation efficiency of phenol, (B) suggested photocatalytic degradation mechanism of phenol with BiOBr/WO3/ZnO, (C) photocatalytic stability of BiOBr/WO3/ZnO toward phenol degradation, (D) photocatalytic system efficiency for the degradation of phenol and malachite green at the initial concentration, 50 mg/L; BiOBr/WO3/ZnO and TiO2 P25 dose, 0.05 g/0.03 L; and pH, 6, and (E) realistic application of refinery petroleum wastewater.

  • References


    This article references 45 other publications.

    1. 1
      Alnaizy, R.; Akgerman, A. Advanced oxidation of phenolic compounds. Adv. Environ. Res. 2000, 4, 233244,  DOI: 10.1016/S1093-0191(00)00024-1
    2. 2
      Ahmed, S.; Rasul, M. G.; Martens, W. N.; Brown, R.; Hashib, M. A. Heterogeneous photocatalytic degradation of phenols in wastewater: A review on current status and developments. Desalination 2010, 261, 318,  DOI: 10.1016/j.desal.2010.04.062
    3. 3
      Karunakaran, C.; Dhanalakshmi, R. Semiconductor-catalyzed degradation of phenols with sunlight. Sol. Energy Mater. Sol. Cells 2008, 92, 13151321,  DOI: 10.1016/j.solmat.2008.05.002
    4. 4
      Reynolds, D. C.; Look, D. C.; Jogai, B. Optically pumped ultraviolet lasing from ZnO. Solid State Commun. 1996, 99, 873875,  DOI: 10.1016/0038-1098(96)00340-7
    5. 5
      Janotti, A.; Van De Walle, C. G. Fundamentals of zinc oxide as a semiconductor. Rep. Prog. Phys. 2009, 72, 126501,  DOI: 10.1088/0034-4885/72/12/126501
    6. 6
      Wang, S. L.; Zhu, Y.; Luo, X. 2D WC/WO3Heterogeneous Hybrid for Photocatalytic Decomposition of Organic Compounds with Vis-NIR Light. Adv. Funct. Mater. 2018, 28, 1705357,  DOI: 10.1002/adfm.201705357
    7. 7
      Xia, J.; Di, J.; Yin, S. Solvothermal synthesis and enhanced visible-light photocatalytic decontamination of bisphenol A (BPA) by g-C3N4/BiOBr heterojunctions. Mater. Sci. Semicond. Process. 2014, 24, 96103,  DOI: 10.1016/j.mssp.2014.02.036
    8. 8
      Motawea, E. A.; Ali, H. R. Exploitation of Heterostructure Photocatalyst Impregnated Alginate Beads for Enhanced Removal of Methyl Orange under Solar Light. Petrol. Petrochem. Eng. J. 2018, 2, 110,  DOI: 10.23880/ppej-16000180
    9. 9
      Berger, S.; Tsuchiya, H.; Ghicov, A.; Schmuki, P. High photocurrent conversion efficiency in self-organized porous WO3. Appl. Phys. Lett. 2006, 88, 203119,  DOI: 10.1063/1.2206696
    10. 10
      Momeni, M. M.; Hakimian, M.; Kazempour, A. In-Situ Manganese Doping of TiO2 Nanostructures via Single-Step Electrochemical Anodizing of Titanium in an Electrolyte Containing Potassium Permanganate: A Good Visible-Light Photocatalyst. Ceram. Int. 2015, 41, 13692,  DOI: 10.1016/j.ceramint.2015.07.158
    11. 11
      Momeni, M. M.; Nazari, Z. Preparation of TiO2 and WO3-TiO2 nanotubes decorated with PbO nanoparticles by chemical bath deposition process: A stable and efficient photo catalyst. Ceram. Int. 2016, 42, 86918697,  DOI: 10.1016/j.ceramint.2016.02.103
    12. 12
      Zeng, Q.; Li, J.; Bai, J.; Li, X.; Xia, L.; Zhou, B. Preparation of vertically aligned WO3 nanoplate array films based on peroxotungstate reduction reaction and their excellent photoelectrocatalytic performance. Appl. Catal., B 2017, 202, 388396,  DOI: 10.1016/j.apcatb.2016.09.045
    13. 13
      Zeng, X.; Wang, Z.; Wang, G. Highly dispersed TiO2 nanocrystals and WO3 nanorods on reduced graphene oxide: Z-scheme photocatalysis system for accelerated photocatalytic water disinfection. Appl. Catal., B 2017, 218, 163173,  DOI: 10.1016/j.apcatb.2017.06.055
    14. 14
      Momeni, M. M.; Ghayeb, Y.; Davarzadeh, M. Electrochemical construction of different titania-tungsten trioxide nanotubular composite and their photocatalytic activity for pollutant degradation: a recyclable photocatalysts. J. Mater. Sci.: Mater. Electron. 2015, 26, 15601567,  DOI: 10.1007/s10854-014-2575-x
    15. 15
      Momeni, M. M.; Ghayeb, Y.; Ghonchegi, Z. Fabrication and characterization of copper doped TiO2 nanotube arrays by in situ electrochemical method as efficient visible-light photocatalyst. Ceram. Int. 2015, 41, 87358741,  DOI: 10.1016/j.ceramint.2015.03.094
    16. 16
      Lei, R.; Ni, H.; Chen, R.; Zhang, B.; Zhan, W.; Li, Y. Growth of Fe 2 O 3 /SnO 2 nanobelt arrays on iron foil for efficient photocatalytic degradation of methylene blue. Chem. Phys. Lett. 2017, 673, 16,  DOI: 10.1016/j.cplett.2017.01.052
    17. 17
      Tsay, C.-Y.; Fan, K.-S.; Chen, S.-H.; Tsai, C.-H. Preparation and characterization of ZnO transparent semiconductor thin films by sol-gel method. J. Alloys Compd. 2010, 495, 126130,  DOI: 10.1016/j.jallcom.2010.01.100
    18. 18
      Sakthivel, S.; Neppolian, B.; Shankar, M. V.; Arabindoo, B.; Palanichamy, M.; Murugesan, V. Solar photocatalytic degradation of azo dye: Comparison of photocatalytic efficiency of ZnO and TiO2. Sol. Energy Mater. Sol. Cells 2003, 77, 6582,  DOI: 10.1016/S0927-0248(02)00255-6
    19. 19
      Li, D.; Haneda, H.; Ohashi, N.; Saito, N.; Hishita, S. Morphological reform of ZnO particles induced by coupling with MOx (M=V,W,Ce) and the effects on photocatalytic activity. Thin Solid Films 2005, 486, 2023,  DOI: 10.1016/j.tsf.2004.11.237
    20. 20
      El-saied, H. A.-a.; Motawea, E. A.-T. Optimization and Adsorption Behavior of Nanostructured NiFe2O4/Poly AMPS Grafted Biopolymer. J. Polym. Environ. 2020, 28, 23352351,  DOI: 10.1007/s10924-020-01774-z
    21. 21
      Lv, J.; Dai, K.; Zhang, J.; Liu, Q.; Liang, C.; Zhu, G. Facile constructing novel 2D porous g-C3N4/BiOBr hybrid with enhanced visible-light-driven photocatalytic activity. Sep. Purif. Technol. 2017, 178, 617,  DOI: 10.1016/j.seppur.2017.01.019
    22. 22
      Huang, W.; Hua, X.; Zhao, Y. Enhancement of visible-light-driven photocatalytic performance of BiOBr nanosheets by Co2+ doping. J. Mater. Sci.: Mater. Electron. 2019, 30, 1496714976,  DOI: 10.1007/s10854-019-01869-x
    23. 23
      Khampuanbut, A.; Santalelat, S.; Pankiew, A. Visible-light-driven WO3/BiOBr heterojunction photocatalysts for oxidative coupling of amines to imines: Energy band alignment and mechanistic insight. J. Colloid Interface Sci. 2020, 560, 213224,  DOI: 10.1016/j.jcis.2019.10.057
    24. 24
      Yang, H.; Zhang, Q.; Chen, Y.; Huang, Y.; Yang, F.; Lu, Z. Ultrasonic-microwave synthesis of ZnO/BiOBr functionalized cotton fabrics with antibacterial and photocatalytic properties. Carbohydr. Polym. 2018, 201, 162171,  DOI: 10.1016/j.carbpol.2018.08.068
    25. 25
      Tae Kwon, Y.; Yong Song, K.; In Lee, W.; Jin Choi, G.; Rag Do, Y. Photocatalytic behavior of WO3-loaded TiO2 in an oxidation reaction. J. Catal. 2000, 191, 192199,  DOI: 10.1006/jcat.1999.2776
    26. 26
      Keller, V.; Bernhardt, P.; Garin, F. Photocatalytic oxidation of butyl acetate in vapor phase on TiO2, Pt/TiO2 and WO3/TiO2 catalysts. J. Catal. 2003, 215, 129138,  DOI: 10.1016/S0021-9517(03)00002-2
    27. 27
      El-Salamony, R. A.; Amdeha, E.; Badawy, N. A.; Ghoneim, S. A.; Al-Sabagh, A. M. Visible light sensitive activated carbon-metal oxide (TiO2, WO3, NiO, and SnO) nano-catalysts for photo-degradation of methylene blue: a comparative study. Toxicol. Environ. Chem. 2018, 100, 143156,  DOI: 10.1080/02772248.2018.1497634
    28. 28
      Amdeha, E.; Mohamed, R. S. A green synthesized recyclable ZnO/MIL-101(Fe) for Rhodamine B dye removal via adsorption and photo-degradation under UV and visible light irradiation. Environ. Technol. 2019, 42, 842859,  DOI: 10.1080/09593330.2019.1647290
    29. 29
      Wang, H.; Zhang, L.; Chen, Z. Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 52345244,  DOI: 10.1039/c4cs00126e
    30. 30
      Zhang, J.; Zhang, L.; Shen, X.; Xu, P.; Liu, J. Synthesis of BiOBr/WO3p-n heterojunctions with enhanced visible light photocatalytic activity. CrystEngComm 2016, 18, 38563865,  DOI: 10.1039/c6ce00824k
    31. 31
      Rafiee, E.; Noori, E.; Zinatizadeh, A. A.; Zanganeh, H. Photocatalytic degradation of phenol using a new developed TiO2/graphene/heteropoly acid nanocomposite: Synthesis, characterization and process optimization. RSC Adv. 2016, 6, 9655496562,  DOI: 10.1039/c6ra09897e
    32. 32
      Ali, M. H. H.; Al-qahtani, K. M.; El-sayed, S. M. Enhancing photodegradation of 2,4,6 trichlorophenol and organic pollutants in industrial effluents using nanocomposite of TiO2 doped with reduced graphene oxide. Egypt. J. Aquat. Res. 2019, 45, 321328,  DOI: 10.1016/j.ejar.2019.08.003
    33. 33
      Ali, H. R. Photocoupling of Ag3PO4-Ag2CO3 with molecularly imprinted polymer for enhanced removal of phenol under solar light: Application of taguchi method. Int. J. Appl. Eng. Res. 2017, 12, 33533359
    34. 34
      Abd El Salam, H. M.; Younis, S. A.; Ali, H. R.; Zaki, T. Statistical modeling and optimization of phenol adsorption from water by modified Cu3(BTC)2: Kinetic, isotherm, and thermodynamic analysis. Microporous Mesoporous Mater. 2017, 241, 210217,  DOI: 10.1016/j.micromeso.2016.12.010
    35. 35
      Rodriguez, A. F. R.; Coaquira, J. A. H.; Morales, M. A. Synthesis, characterization and magnetic properties of polymer-Fe3O4 nanocomposite. Spectrochim. Acta, Part A 2013, 100, 101103,  DOI: 10.1016/j.saa.2012.02.081
    36. 36
      Mäkelä, M. Experimental design and response surface methodology in energy applications: A tutorial review. Energy Convers. Manage. 2017, 151, 630640,  DOI: 10.1016/j.enconman.2017.09.021
    37. 37
      Tebbouche, L.; Hank, D.; Zeboudj, S.; Namane, A.; Hellal, A. Evaluation of the phenol biodegradation byAspergillus niger: application of full factorial design methodology. Desalin. Water Treat. 2016, 57, 61246130,  DOI: 10.1080/19443994.2015.1053991
    38. 38
      Choquette-Labbé, M.; Shewa, W.; Lalman, J.; Shanmugam, S. Photocatalytic degradation of phenol and phenol derivatives using a Nano-TiO2 catalyst: Integrating quantitative and qualitative factors using response surface methodology. Water (Switzerland) 2014, 6, 17851806,  DOI: 10.3390/w6061785
    39. 39
      Ocholi, O.; Menkiti, M.; Auta, M.; Ezemagu, I. Optimization of the operating parameters for the extractive synthesis of biolubricant from sesame seed oil via response surface methodology. Egypt. J. Pet. 2018, 27, 265275,  DOI: 10.1016/j.ejpe.2017.04.001
    40. 40
      Tehrani, F. S.; Ahmadian, H.; Aliannezhadi, M. Hydrothermal synthesis and characterization of WO3 nanostructures: Effect of reaction time. Mater. Res. Express 2020, 7, 015911,  DOI: 10.1088/2053-1591/ab66fc
    41. 41
      Shamhari, N. M.; Wee, B. S.; Chin, S. F.; Kok, K. Y. Synthesis and characterization of zinc oxide nanoparticles with small particle size distribution. Acta Chim. Slov. 2018, 65, 578585,  DOI: 10.17344/acsi.2018.4213
    42. 42
      Yuan, Q. L.; Zhang, Y.; Yin, H. Y.; Nie, Q. L.; Wu, W. W. Rapid, simple and low-cost fabrication of BiOBr ultrathin nanocrystals with enhanced visible light photocatalytic activity. J. Exp. Nanosci. 2016, 11, 359369,  DOI: 10.1080/17458080.2015.1065014
    43. 43
      Liu, X.; Luo, J.; Zhu, Y.; Yang, Y.; Yang, S. Removal of methylene blue from aqueous solutions by an adsorbent based on metal-organic framework and polyoxometalate. J. Alloys Compd. 2015, 648, 986993,  DOI: 10.1016/j.jallcom.2015.07.065
    44. 44
      Younis, S. A.; Motawea, E. A.; Moustafa, Y. M.; Lee, J.; Kim, K.-H. A strategy for the efficient removal of chlorophenols in petrochemical wastewater by organophilic and aminated silica@alginate microbeads: Taguchi optimization and isotherm modeling based on partition coefficient. J. Hazard. Mater. 2020, 397, 122792,  DOI: 10.1016/j.jhazmat.2020.122792
    45. 45
      Bakr, A.-S. A.; Moustafa, Y. M.; Motawea, E. A.; Yehia, M. M.; Khalil, M. M. H. Removal of ferrous ions from their aqueous solutions onto NiFe2O4-alginate composite beads. J. Environ. Chem. Eng. 2015, 3, 14861496,  DOI: 10.1016/j.jece.2015.05.020