Next Article in Journal
What Drives the Morphological Traits of Stress-Tolerant Plant Cynodon dactylon in a Riparian Zone of the Three Gorges Reservoir, China
Next Article in Special Issue
Surfactant-Capped Silver-Doped Calcium Oxide Nanocomposite: Efficient Sorbents for Rapid Lithium Uptake and Recovery from Aqueous Media
Previous Article in Journal
Assessing the Applicability of Mainstream Global Isoscapes for Predicting δ18O, δ2H, and d-excess in Precipitation across China
Previous Article in Special Issue
Synthesis of Ag-OMS Catalyst for Sunlight-Assisted Photodegradation of Crystal Violet Dye
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Photocatalytic Activity of the Bi2O3-NiO Heterojunction for the Degradation of Methyl Orange under Irradiation of Sunlight

1
Department of Physics, University of Sargodha, Sargodha 40100, Pakistan
2
Department of Chemistry, Government College University Faisalabad, Faisalabad 38000, Pakistan
3
Oil and Gas Engineering Department, Oil and Gas Engineering College, Basrah University for Oil and Gas, Basrah 61004, Iraq
4
Department of Civil Engineering, Government Polytechnic Mankeda, Agra 283102, Uttar Pradesh, India
5
Arnold and Marie Schwartz College of Pharmacy and Health Sciences, Long Island University, Brooklyn, NY 11201, USA
6
Department of Chemistry, University of Malakand, Chakdara 18800, Pakistan
7
Department of Chemistry, College of Science, King Khalid University, Abha 61413, Saudi Arabia
8
Department of Chemistry, Al Lieth University College, Umm Al-Qura University, Makkah 24382, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Water 2023, 15(18), 3182; https://doi.org/10.3390/w15183182
Submission received: 15 August 2023 / Revised: 23 August 2023 / Accepted: 5 September 2023 / Published: 6 September 2023

Abstract

:
Recently, the development of visible-light-responsive catalysts for the photodegradation of organic pollutants has captured the attention of researchers globally. The ineffectiveness and high costs associated with conventional methods and techniques used for the abatement of water pollution have forced researchers to develop effective and low-cost innovative techniques for this purpose. Photocatalysis is considered an effective protocol for this purpose. Therefore, this study was conducted for the development of the Bi2O3-NiO heterojunction as a visible-light-responsive photocatalyst for the degradation of methyl orange. Ni(NO3)2∙6H2O (Fluka) and Bi(NO3)3∙5H2O (Merck) were used as precursor materials for the synthesis of NiO-Bi2O3. After fabrication, the Bi2O3-NiO heterojunction was characterized using XRD, EDX, SEM, FTIR, and TGA techniques. Then, it was employed as a catalyst for the photodegradation of methyl orange under sunlight irradiation. The fabricated Bi2O3-NiO showed higher photocatalytic activity than Bi2O3 and NiO with 100, 67, and 46% degradation of methyl orange, respectively. The rate constant determined by the non-linear method of analysis for the photodegradation of MO in the presence of Bi2O3-NiO was 3.2-fold and 1.7-fold of the rate constant with NiO and Bi2O3, respectively. The higher photocatalytic activity of Bi2O3-NiO than of its individual components in the present study is also attributed to the separation and transfer of positive holes and electrons. The recycling of spent Bi2O3-NiO under similar experimental conditions exhibited the same photocatalytic activity suggesting the stability of the fabricated Bi2O3-NiO photocatalyst.

1. Introduction

The rapid expansion of the textile and chemical industries and the rapid increase in world population has resulted in a dramatic condition regarding the pollution of water reserves. Other sources such as science and technology, agriculture, and domestic activities also contribute to water pollution. However, the contribution to aqueous contamination by these various industries is very significant [1,2,3,4,5,6,7]. Various industries such as pharmaceutics, leather, textile, food, etc., are responsible for water pollution. The effluents of these industries transmit various organic and inorganic pollutants to the environment. Water pollution due to the release of organic dyes from different industries is the most dangerous type of pollution that causes a threat to living organisms and society. Aqueous pollution due to dyes is a critical issue worldwide. These dyes severely affect the environment due to their carcinogenic nature, high toxicity, and persistent nature [8,9,10,11,12]. Various traditional methods including physical, chemical, and biological methods have been attempted for the remediation of dyes-polluted water. However, these methods have some drawbacks such as high cost and ineffectiveness of treatment processes. The ineffectiveness and high costs associated with conventional methods and techniques used for the abatement of water pollution have forced researchers to develop effective and low-cost innovative techniques for this purpose. Photocatalysis has merged as a green, low-cost, and effective technique for the destruction of organic pollutants like dyes, pesticides, and pharmaceutics present in wastewater recently [13,14,15,16]. The photocatalytic technique is an effective protocol for the abatement of organic pollutants like dyes and pesticides. The photocatalytic technique is a chemical process that is based on the production of hydroxyl radicals that take part in the complete mineralization of organic molecules into simple inorganic molecules [17,18,19,20]. The photocatalytic protocol has the following advantages over conventional methods:
  • Direct conversion of organic pollutants molecules into simple inorganic molecules like water and carbon dioxide
  • Mineralization of a wide range of organic pollutants irrespective of selectivity
  • No production of hazardous side products.
The photocatalytic technique is based on the in situ production of excitons in the valance band and conduction band of semiconductors photocatalysts under ultraviolet or visible light. These excitons generate highly reactive OH radicals through subsequent reactions. These OH radicals then mineralize the pollutants species. Depending upon the bandgap energy of the semiconductors photocatalysts, both ultraviolet and visible light can be used in the photocatalytic techniques. The semiconductor metal oxides like TiO2, ZnO, SnO2, CeO2, CuO, MgO, perovskite metal oxides, etc., have been used as traditional photocatalysts for a long time [21,22,23,24,25]. However, the traditional photocatalysts are not successful; the ineffective degradation of pollutants is due to their wide band gap energies. The traditional photocatalysts can only be used under the irradiation of ultraviolet light. As visible light is abundantly available, due to the sun (44% of sunlight is in the visible range), the development of a photocatalyst that can be effectively employed for the destruction of pollutants under the irradiation of sunlight is essential [26,27,28]. Semiconductors with narrow band gap energy can be employed for this purpose because such types of semiconductors can absorb visible light [29]. Therefore, the bismuth oxide, Bi2O3, which is a narrow band gap semiconductor, has gained the interest of researchers recently. It has desirable characteristics such as appropriate band gap energy, stable structure, low cost, and an environmentally benign nature. It has been widely used in many fields such as for thin films, photovoltaic cells, fuel cells, and catalysis. The bismuth oxide exists in various forms, however, the α-Bi2O3 (band gap 2.85 eV) and β-Bi2O3 (band gap 2.58 eV) are more important for catalytic applications [30,31,32]. However, due to the low energy difference in the valence band and conduction band of Bi2O3, the photo-induced positive holes and electrons quickly recombine. As a result, the photocatalytic performance of Bi2O3 is badly affected. Therefore, researchers have attempted to reduce the recombination of positive holes and electrons by modification of the structure of Bi2O3. One of the attempted modifications is the construction of heterojunctions between Bi2O3 and other semiconductors. The construction of a heterojunction between two semiconductors produces an electric field at the interface. This electric field assists in the separation of photo-induced positive holes and electrons resulting in the enhancement of photocatalytic activities [33,34,35]. This study reports the enhancement in photocatalytic activity of Bi2O3 for the photodegradation of methyl orange by the construction of a heterojunction with NiO. NiO is a p-type wide band gap (3.6–4.0 eV) semiconductor with high electro-optical proficiency and chemical stability. It has been used in several applications such as gas sensing, solar cells, and catalysis [36,37,38,39,40]. The driving force for the construction of heterojunction between NiO and Bi2O3 was to develop a solar light-driven photocatalyst for the photodegradation of dyes. The Bi2O3-NiO heterojunction exhibits higher photocatalytic activity due to two reasons: (1) It harvests an extended portion of the UV-visible spectrum. (2) It suppresses the unwanted recombination of photo-induced positive holes and electrons due to its synergic effect.

2. Materials and Methods

2.1. Synthesis of Bi2O3-NiO

Analytical-grade chemicals were used in this study. All the chemicals and reagents were used as received, i.e., without further purification. Ni(NO3)2∙6H2O (Fluka) and Bi(NO3)3∙5H2O (Merck) were used as precursor materials for the synthesis of NiO-Bi2O3. Typically, equimolar quantities of both Ni(NO3)2∙6H2O and Bi(NO3)3∙5H2O were dissolved in 10 mL concentrated HNO3. Then, the mixed solution was diluted with 100 mL of distilled water. The resultant solution was hydrolyzed by dropwise addition of 1M NaOH solution under continuous stirring. The addition of NaOH to the mixed solution produced a gel. The gel formed was filtered, washed, and dried at 100 °C for 24 h. Finally, the substance formed was calcined first at 300 °C and then at 600 °C for 4 h at each temperature. Pristine NiO and Bi2O3 were also prepared in the same way. The prepared samples were kept in glass vials for further study.

2.2. Characterization of Bi2O3-NiO

The NiO-Bi2O3 was characterized using XRD, EDX, SEM, FTIR, and TGA techniques. The instruments used for the characterization of prepared Bi2O3, NiO, and Bi2O3–NiO samples are listed in Table 1.

2.3. Photocatalytic Activity

The photocatalytic activity of Bi2O3-NiO was explored through the photodegradation of methyl orange under sunlight. The photodegradation of methyl orange was studied in a Pyrex glass beaker. In the first step, blank experiments were performed. In the first blank experiment, the dye solution was stirred under sunlight for an estimation of photolysis. A sample was taken after 30 min and analyzed with a UV-visible spectrophotometer. There was no removal of methyl orange due to photolysis. The second blank experiment was performed for confirmation of the leaching of the catalyst. It was confirmed by stirring 0.1 g Bi2O3-NiO in 20 mL distilled water for one hour. After stirring, the Bi2O3-NiO was separated and the clear liquid was analyzed for the existence of Bi or Ni ions. The analysis confirmed that there was no leaching of the metal ions from Bi2O3-NiO. In the second step, an optimized amount of catalyst was suspended in a 50 mL solution of methyl orange and stirred in the dark to maintain adsorption equilibrium. The initial concentrations of the dye solution used were 50, 100, and 150 mg/L. A sample was taken and analyzed with a UV-visible spectrophotometer. About 10% removal of methyl orange was determined due to adsorption. Then, the beaker was placed in sunlight under continuous stirring in the third step. Samples were taken and analyzed with a UV-visible spectrophotometer.

3. Results and Discussion

3.1. Characterization

The XRD spectra of Bi2O3, NiO, and Bi2O3–NiO are given in Figure 1. The existence of sharp peaks in the XRD spectrum of Bi2O3 indicates the crystalline nature of the prepared sample. The XRD spectrum of Bi2O3 is dominated by diffraction peaks corresponding to planes of tetragonal α-Bi2O3 at 2θ 24.71° (121), 30.68° (012), 33.59° (−112), 46.84° (041), and 47.61° (−104) [41,42,43,44]. The XRD pattern of NiO also shows the crystallinity of the fabricated sample with diffraction peaks and hkl planes of cubic NiO at 2θ 37.31° (111), 43.32° (200), and 62.78° (220) according to PDF 01-089-5881 [45,46,47]. The XRD of the Bi2O3-NiO heterojunction shows the diffraction peaks of both pure Bi2O3 and NiO which confirms the successful fabrication of the heterojunction. Based on the comparison of the XRD of the fabricated Bi2O3-NiO heterojunction with the available literature, it was concluded that the heterojunction had been successfully fabricated with a tetragonal Bi2O3 phase and cubic NiO phase with space groups P-421c and Fm-3m, respectively. The 2θ values and hkl planes corresponding to Bi2O3 observed in the XRD pattern and Equation (1) were used for the calculation of the lattice parameters of the tetragonal structure. Equation (1) was applied to the observed data through the non-linear method using the “Solver” software (Microsoft 365) of MS Word. The lattice parameters of the tetragonal structure were calculated as a = b = 2.32 Å and c = 1.34 Å. Similarly, using Equation (2) and the observed XRD data of NiO, the lattice parameters for the cubic structure were calculated as a = b = c = 0.98 Å [42].
1 d h k l 2 = h 2 + k 2 a 2 + l 2 c 2
1 d h k l 2 = h 2 + k 2 + l 2 a 2
The stability of the fabricated samples was investigated by thermal gravimetric analysis (TGA). Figure 2 shows the results of the thermal gravimetric analyses. All three fabricated samples were quite stable over a wide range of temperatures. A minute weight loss (about 2%) observed in each sample is attributed to the evaporation of adsorbed moisture content.
The fabrication of the samples was further confirmed by elemental composition analyses using energy-dispersive X-ray (EDX) spectroscopy. Figure 3 shows the results of the energy-dispersive X-ray (EDX) analyses of the samples. The results confirmed that Bi2O3 comprised Bi and O. Similarly, the NiO comprised Ni and O only. The Bi2O3-NiO heterojunction is composed of Bi, Ni, and O. The demonstrated EDX results confirmed the successful fabrication and purity of the samples. A trace amount of C was also detected in the Bi2O3-NiO heterojunction which might be due to impurity in the precursor material or contamination during the synthesis process.
The morphology and particle size analyses were carried out using scanning electron microscopy (SEM). Figure 4 demonstrates the scanning electron micrographs of Bi2O3, NiO, and Bi2O3-NiO samples. Particles of both Bi2O3 and NiO are non-agglomerated and irregular in shape. The SEM of Bi2O3-NiO indicates that both Bi2O3 and NiO retain their identity and particle shape in the heterojunction.

3.2. Photocatalytic Activity

The comparison of the photocatalytic activity of Bi2O3, NiO, and Bi2O3-NiO was studied by performing separate photodegradation experiments using a 50 mL solution of methyl orange with a 100 mg/L concentration. The photodegradation profile of methyl orange dye in terms of Mt/Mo (Mt: concentration or absorbance at a different time interval, Mo: concentration or absorbance at zero time) is given in Figure 5. The data given in Figure 5 shows that the photodegradation of methyl orange dye observed with Bi2O3-NiO is significantly higher than the photodegradation observed with pristine Bi2O3 and NiO. In terms of the percent photodegradation of methyl orange, the photocatalytic activity of Bi2O3-NiO, Bi2O3, and NiO was observed as ~100%, 67%, and 46%, respectively. After vigilant analyses of the observed results, it was predicted that various factors contribute to the enhanced photocatalytic response of the Bi2O3-NiO composite.
The photocatalytic efficiency of a photocatalyst is significantly influenced by the separation, transfer, and recombination of the photo-induced positive holes and electrons. It is accepted that efficient separation and transfer of the positive holes and electrons suppresses the recombination of these charge carriers and ultimately causes a significant improvement in the photocatalytic activity of the photocatalyst. The higher photocatalytic activity of Bi2O3-NiO than of its individual components in the present study is also attributed to the separation and transfer of the positive holes and electrons [48,49,50].
Methyl orange is a colored substance that absorbs light in the wavelength range of 340–530 nm of the UV-visible spectrum. Similarly, Bi2O3, having a bandgap energy of 2.58 eV, also absorbs in the visible region at ~480 nm. Ordinary sunlight is composed of 10% ultraviolet light and 45% visible light in addition to other types of radiation. It shows that methyl orange masks the absorption region of Bi2O3. Therefore, the available region of sunlight to be absorbed by the photocatalyst is only 300–340 nm. Hence, it is predicted that the photocatalytic degradation of methyl orange in the present study takes place via an indirect excitation process, unlike the normal photocatalytic process in which the photocatalytic degradation takes place via direct excitation. Furthermore, the inhibition of the recombination of positive holes and electrons through a synergic effect is also expected due to the composite photocatalyst. The irradiation of the reaction mixture (methyl orange and Bi2O3-NiO) causes excitation of the methyl orange due to the absorption of light. The excited methyl orange then transfers the absorbed energy to the Bi2O3 component of the photocatalyst. The transfer of this energy promotes the electrons from the VB to CB of Bi2O3 resulting in the formation of the exciton (a pair of positive holes and electrons). The positive holes are then diffused to the VB of NiO. As a result, the recombination of the positive hole and electron is inhibited. Similarly, the absorption of ultraviolet light causes the production of exciton in NiO as well. The photo-induced electron in the CB of NiO then flows to the CB of Bi2O3 [51,52,53]. The positive holes and electrons react with water and oxygen, respectively, and ultimately produce OH radicals. The OH radicals then take part in the degradation of methyl orange molecules. The formation of positive holes and hydroxyl radicals was verified experimentally using EDTA and t-BuOH as scavengers for charge carriers, respectively. It was noted that the photocatalytic activity of Bi2O3-NiO decreased from 80% to 63 and 41% in the presence of EDTA and t-BuOH, respectively [54,55,56]. The proposed process can be summarized as follows (BN: Bi2O3-NiO). This is explained in Figure 6 as well.
M O + h ϑ V i s       M O *
M O * + B N      B h + , e N
B h + , e N B e N ( h + )
B N + h ϑ U V       B N ( h + , e )
  B N h + , e       B e N ( h + )
B N h + + H 2 O   O H
B e N + O 2   O H
M O ( a d s ) + O H D e g r a r a t i o n   p r o d u c t s
It is necessary to mention that the excitation of NiO through the transfer of energy from methyl orange through indirect photocatalysis is less likely, as the energy absorbed by methyl orange is not sufficient to promote an electron from the VB to CB of the NiO counterpart of the composite photocatalyst. The higher photocatalytic activity of Bi2O3-NiO than of its individual components verifies the proposed assumptions.
Based on the proposed mechanism, the rate of photodegradation of MO in the present study is expressed in the following equation (M: concentration of methyl orange, θM: fraction of catalyst surface covered by methyl orange).
r = d M d t = k   O H θ M
As the amount of catalyst is constant and the reaction mixture is open to the atmosphere and continuously irradiated, the rate of reaction therefore does not depend on the concentration of OH radicals. Hence, the rate expression becomes
r = d M d t = k   θ M
The surface of the catalyst covered by methyl orange is expressed in terms of Langmuir adsorption isotherm. Hence, the rate expression modifies as (K: Langmuir constant for adsorption of methyl orange, Ki: Langmuir constant for adsorption of degradation products, Pi: concentration of products):
r = d M d t = k   K M 1 + K M + K i P i
As
K M + K i [ P ] i = M 0 = C o n s t a n t
Therefore, the rate equation becomes
r = d M d t = k o b s M  
where
k o b s = k K 1 + M o
On integration of Equation (7)
ln M o M t = k o b s   t
On re-arrangement of Equation (9), we obtain
M t = M o   e k o b s t
The photodegradation data of methyl orange over Bi2O3, NiO, and Bi2O3-NiO were analyzed according to kinetics Equation (10) using the non-linear method of analysis. The Microsoft Excel Solver add-in was used for non-linear analyses. Figure 7 shows the treatment of the degradation data according to kinetics Equation (10). The rate constants (kobs) are given in Table 2. The rate constant for the photodegradation of MO in the presence of Bi2O3-NiO was 3.2-fold and 1.7-fold of the rate constant with NiO and Bi2O3, respectively.

3.3. Dependance of Photocatalytic Activity on Catalyst Dosage

The optimization of catalyst dosage is also an important parameter in catalysis because the catalytic activity and rate of reaction significantly depend on catalyst dosage. Therefore, photocatalytic reactions were conducted with various catalyst dosages of Bi2O3-NiO using a 50 mL (100 mg/L) solution of methyl orange. Figure 8 shows the photocatalytic activity of Bi2O3-NiO towards the degradation of methyl orange after 120 min of reaction. It is evident that although the photodegradation increased with the catalyst dosage, the increase was not continuous. The catalytic activity initially significantly increased up to 0.08 g of the catalyst and then there was a slight increase with further increase in the catalyst dosage. On the other hand, the rate of reaction (degradation per gram of catalyst per minute) initially increased up to 0.08 g of the catalyst and then decreased with further additions. Hence, 0.08 g was chosen as an optimum catalyst dosage in this investigation. It is obvious that the number of active sites increases with catalyst dosage, therefore the catalytic activity also increases with catalyst dosage. However, a higher dosage of the catalyst creates a hindrance to the penetration of photons due to scattering and turbidity, therefore, the catalytic activity decreased at the higher dosage [57,58].

3.4. Dependance of Photocatalytic Activity on Concentration of Methyl Orange

The initial concentration of dye significantly influences the efficiency of the photocatalytic process. Therefore, we investigated the dependence of the photocatalytic activity of Bi2O3-NiO on the concentration of methyl orange. This investigation was accomplished by performing independent photodegradation experiments using 0.08 g Bi2O3-NiO and 50 mL solution of methyl orange with concentrations of 50, 100, and 150 mg/L. The removal of dye due to adsorption was excluded by stirring the reaction mixture in the dark for 30 min. Figure 9 shows the obtained data. The data shows that the catalytic efficiency of Bi2O3-NiO is inversely proportional to the concentration of methyl orange. The catalytic efficiency of Bi2O3-NiO towards the photodegradation of methyl orange was 100, 80, and 65% after 60 min of reaction with 50, 100, and 150 mg/L solutions of methyl orange, respectively. The obtained degradation data was analyzed for kinetics analysis according to kinetics Equation (10). The rate constants were found as 0.0227, 0.0125, and 0.0082 per minute for the photodegradation of 50, 100, and 150 mg/L solutions of methyl orange, respectively. An increase in the initial concentration of methyl orange decreased the photocatalytic performance. Generally, the rate of a reaction is a directly proportional to the concentration of the reactant. However, in this study, we observed that the rate of reaction is inversely proportional to the concentration of methyl orange. It is because photons cannot penetrate the catalyst surface due to the intense color of the solution at higher concentrations. Furthermore, as the catalyst dose was kept fixed in all three experiments the concentration of hydroxyl radicals was therefore also expected to be the same in all experiments. The increase in the concentration of methyl orange causes a decrease in the number of hydroxyl radicals per molecule of methyl orange. Hence, the photocatalytic efficiency is inversely proportional to the initial concentration of methyl orange [59,60,61,62].

3.5. Comparison of the Present Photocatalyst with Reported Photocatalysts

Various Bi2O3-based photocatalysts have been reported for the photodegradation of dyes. Dhiman et al. [63] have reported the CoFe2O4@Bi2O3/NiO photocatalyst for degradation of ofloxacin under visible light. They reported a 92% degradation of ofloxacin in 90 min using 100 mL of a 10 mg/L solution of ofloxacin. Shahzad et al. [64] have reported an organosilica-supported Bi2O3 photocatalyst for the degradation of methylene blue and methyl orange dyes. They obtained about 90% degradation of these dyes using 30 mL of a 20 mg/L solution of each dye. In another study, Co3O4-Bi2O3 was used as a catalyst for the photodegradation of rhodamine B dye with 92% degradation efficiency using a 100 mg/L solution of the dye [54]. Similarly, a 94% degradation of MB and MO over the NiO-Bi2O3 photocatalyst has been reported recently [65]. Banoth et al. [66] have prepared novel BiFeO3 and BiFeO3-Fe2O3 photocatalysts for the photodegradation of dyes. They tested their catalysts for the photodegradation of methylene blue dye. Almost complete degradation of methylene blue was obtained using 40 mg of catalyst and a 10 mg/L (100 mL) solution of methylene blue dye under visible light with a reaction duration of 70 min. Poorsajadi et al. [67] have synthesized a CuO-Bi2O3 nanocomposite via a hydrothermal method. After characterization, they evaluated the photodegradation of methylene blue dye using the prepared nanocomposite as a photocatalyst. About 88% degradation of BM was obtained at an acidic pH using 0.2 g/L of CuO-Bi2O3 nanocomposite as a catalyst and a 10 mg/L solution of the dye. Zhang et al., [68] have studied the synthesis of Bi2O3@Zn-MOF composite via a hydrothermal method. They used the prepared composite for the degradation of rhodamine B dye. They found 97% degradation of the dye using a 50 mL (40 mL) solution of rhodamine B dye under visible light. In this study, we report Bi2O3-NiO as a photocatalyst for the degradation of methyl orange. The catalyst developed in this study is more effective for the degradation of dyes because the almost complete degradation of methyl orange was observed using a 100 mg/L (50 mL) solution of the dye under natural sunlight.

4. Conclusions

This study showed that photocatalytic activity can be improved with the formation of a heterojunction by coupling two suitable semiconductors. The enhanced photocatalytic performance of the heterojunction is attributed to an extended response in the wide region of the UV–visible spectrum. The heterojunction photocatalysts play an important role in the enhancement of drawbacks associated with conventional photocatalysts like poor light response and low efficiency. Hence, a 50 mL solution of methyl orange (100 mg/L) was completely degraded under sunlight irradiation in the presence of Bi2O3-NiO. In comparison, the photocatalytic activity of Bi2O3 and NiO was found to be 67 and 46%, respectively. The optimum catalyst dosage was found as 0.08g. The degradation data were analyzed according to first-order kinetics through the non-linear method of analysis using Solver software. The rate constant for the photodegradation of MO in the presence of Bi2O3-NiO was 3.2-fold and 1.7-fold of the rate constant with NiO and Bi2O3, respectively. The photocatalytic activity of Bi2O3-NiO was inversely related to the initial concentration of methyl orange. The catalytic efficiency of Bi2O3-NiO towards photodegradation of methyl orange was 100, 80, and 65% after 60 min of reaction with 50, 100, and 150 mg/L solutions of methyl orange, respectively.

Author Contributions

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

Funding

This research was funded by King Khalid University grant number RGP2/272/44.

Data Availability Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding the work through the group Research Project Funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ibrahim, A.A.; Salama, R.S.; El-Hakam, S.A.; Khder, A.S.; Ahmed, A.I. Synthesis of Sulfated Zirconium Supported MCM-41 Composite with High-Rate Adsorption of Methylene Blue and Excellent Heterogeneous Catalyst. Colloids Surf. A Physicochem. Eng. Asp. 2021, 616, 126361. [Google Scholar] [CrossRef]
  2. Alasri, T.M.; Ali, S.L.; Salama, R.S.; Alshorifi, F.T. Band-Structure Engineering of TiO2 Photocatalyst by AuSe Quantum Dots for Efficient Degradation of Malachite Green and Phenol. J. Inorg. Organomet. Polym. Mater. 2023, 33, 1729–1740. [Google Scholar] [CrossRef]
  3. Saleh, T.S.; Badawi, A.K.; Salama, R.S.; Mostafa, M.M.M. Design and Development of Novel Composites Containing Nickel Ferrites Supported on Activated Carbon Derived from Agricultural Wastes and Its Application in Water Remediation. Materials 2023, 16, 2170. [Google Scholar] [CrossRef]
  4. Leonel, L.V.; de Oliveira Reis, F.; de Assis Figueiredo, F.A.M.M.; Ferraz, T.M.; de Oliveira Maia Júnior, S.; Silva, P.C.; de Andrade, J.R. Light Intensity and Hydrogel Soil Amendment Differentially Affect Growth and Photosynthesis of Successional Tree Species. J. For. Res. 2023, 34, 257–268. [Google Scholar] [CrossRef]
  5. Wei, S.; Chen, T.; Hou, H.; Xu, Y. Recent Advances in Electrochemical Sterilization. J. Electroanal. Chem. 2023, 937, 117419. [Google Scholar] [CrossRef]
  6. Feng, X.; Sun, L.; Wang, W.; Zhao, Y.; Shi, J.W. Construction of CdS@ZnO Core–Shell Nanorod Arrays by Atomic Layer Deposition for Efficient Photoelectrochemical H2 Evolution. Sep. Purif. Technol. 2023, 324, 124520. [Google Scholar] [CrossRef]
  7. Liang, Y.; Li, J.; Xue, Y.; Tan, T.; Jiang, Z.; He, Y.; Shangguan, W.; Yang, J.; Pan, Y. Benzene Decomposition by Non-Thermal Plasma: A Detailed Mechanism Study by Synchrotron Radiation Photoionization Mass Spectrometry and Theoretical Calculations. J. Hazard. Mater. 2021, 420, 126584. [Google Scholar] [CrossRef] [PubMed]
  8. Nizam, N.U.M.; Hanafiah, M.M.; Mahmoudi, E.; Halim, A.A.; Mohammad, A.W. The Removal of Anionic and Cationic Dyes from an Aqueous Solution Using Biomass-Based Activated Carbon. Sci. Rep. 2021, 11, 8623. [Google Scholar] [CrossRef]
  9. Kamran, U.; Rhee, K.Y.; Lee, S.Y.; Park, S.J. Innovative Progress in Graphene Derivative-Based Composite Hybrid Membranes for the Removal of Contaminants in Wastewater: A Review. Chemosphere 2022, 306, 135590. [Google Scholar] [CrossRef]
  10. Rawal, J.; Kamran, U.; Park, M.; Pant, B.; Park, S.J. Nitrogen and Sulfur Co-Doped Graphene Quantum Dots Anchored TiO2 Nanocomposites for Enhanced Photocatalytic Activity. Catalysts 2022, 12, 548. [Google Scholar] [CrossRef]
  11. Kamran, U.; Lee, S.Y.; Rhee, K.Y.; Park, S.J. Rice Husk Valorization into Sustainable Ni@TiO2/Biochar Nanocomposite for Highly Selective Pb (II) Ions Removal from an Aqueous Media. Chemosphere 2023, 323, 138210. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, J.; Zhong, A.; Huang, G.; Yang, M.; Li, D.; Teng, M.; Han, D. Enhanced Efficiency with CDCA Co-Adsorption for Dye-Sensitized Solar Cells Based on Metallosalophen Complexes. Sol. Energy 2020, 209, 316–324. [Google Scholar] [CrossRef]
  13. Saeed, M.; Al-Saeed, F.A.; Altaf, M.; Alahmari, S.D.; Bokhari, T.H.; Khan, I.; Bajaber, M.A.; Ahmed, A.E. Synthesis of Visible-Light-Driven Ag2O-Co3O4 Z-Scheme Photocatalyst for Enhanced Photodegradation of Reactive Yellow Dye. New J. Chem. 2022, 13, 936. [Google Scholar] [CrossRef]
  14. Khan, I.; Luo, M.; Khan, S.; Asghar, H.; Saeed, M.; Khan, S.; Khan, A.; Humayun, M.; Guo, L.; Shi, B. Green Synthesis of SrO Bridged LaFeO3/g-C3N4 Nanocomposites for CO2 Conversion and Bisphenol A Degradation with New Insights into Mechanism. Environ. Res. 2022, 207, 112650. [Google Scholar] [CrossRef]
  15. Khan, I.; Luo, M.; Guo, L.; Khan, S.; Wang, C.; Khan, A.; Saeed, M.; Zaman, S.; Qi, K.; Liu, Q.L. Enhanced Visible-Light Photoactivities of Porous LaFeO3 by Synchronously Doping Ni2+ and Coupling TS-1 for CO2 reduction and 2,4,6-Trinitrophenol Degradation. Catal. Sci. Technol. 2021, 11, 6793–6803. [Google Scholar] [CrossRef]
  16. Saeed, M.; Muneer, M.; ul Haq, A.; Akram, N. Photocatalysis: An Effective Tool for Photodegradation of Dyes—A Review. Environ. Sci. Pollut. Res. 2022, 29, 293–311. [Google Scholar] [CrossRef]
  17. Yu, H.; Zhu, J.; Qiao, R.; Zhao, N.; Zhao, M.; Kong, L. Facile Preparation and Controllable Absorption of a Composite Based on PMo12/Ag Nanoparticles: Photodegradation Activity and Mechanism. ChemistrySelect 2022, 7, e202103668. [Google Scholar] [CrossRef]
  18. Zhao, Z.; Shi, X.; Shen, Z.; Gu, Y.; He, L.; Zhang, M.; Lu, N. Single-Atom Fe Nanozymes Coupling with Atomic Clusters as Superior Oxidase Mimics for Ratiometric Fluorescence Detection. Chem. Eng. J. 2023, 469, 143923. [Google Scholar] [CrossRef]
  19. Zheng, Y.; Liu, Y.; Guo, X.; Chen, Z.; Zhang, W.; Wang, Y.; Tang, X.; Zhang, Y.; Zhao, Y. Sulfur-Doped g-C3N4/RGO Porous Nanosheets for Highly Efficient Photocatalytic Degradation of Refractory Contaminants. J. Mater. Sci. Technol. 2020, 41, 117–126. [Google Scholar] [CrossRef]
  20. Pang, S.; Zhou, C.; Sun, Y.; Zhang, K.; Ye, W.; Zhao, X.; Cai, L.; Hui, B. Natural Wood-Derived Charcoal Embedded with Bimetallic Iron/Cobalt Sites to Promote Ciprofloxacin Degradation. J. Clean. Prod. 2023, 414, 137569. [Google Scholar] [CrossRef]
  21. Shimi, A.K.; Parvathiraj, C.; Kumari, S.; Dalal, J.; Kumar, V.; Wabaidur, S.M.; Alothman, Z. Green Synthesis of SrO Nanoparticles Using Leaf Extract of Albizia Julibrissin and Its Recyclable Photocatalytic Activity: An Eco-Friendly Approach for Treatment of Industrial Wastewater. Environ. Sci. Adv. 2022, 1, 849–861. [Google Scholar] [CrossRef]
  22. Kadian, N.; Kumari, R.; Panchal, A.; Dalal, J.; Padalia, D. Structural and Optical Properties of Gadolinium Doped-Magnetite Nano-Crystal for Photocatalytic Application. J. Alloys Compd. 2023, 960, 170811. [Google Scholar] [CrossRef]
  23. He, X.; Xie, S.; Xu, J.; Yin, X.-B.; Zhang, M. Reactive Template-Engaged Synthesis of NiSx/MoS2 Nanosheets Decorated on Hollow and Porous Carbon Microtubes with Optimal Electronic Modulation toward High-Performance Enzyme-like Performance. Inorg. Chem. 2023, 62, 8033–8042. [Google Scholar] [CrossRef] [PubMed]
  24. Xia, G.; Zheng, Y.; Sun, Z.; Xia, S.; Ni, Z.; Yao, J. Fabrication of ZnAl-LDH Mixed Metal-Oxide Composites for Photocatalytic Degradation of 4-Chlorophenol. Environ. Sci. Pollut. Res. 2022, 29, 39441–39450. [Google Scholar] [CrossRef]
  25. Lu, C.; Xu, C.; Sun, W.; Ren, R.; Qiao, J.; Wang, Z.; Sun, K.; Pan, G.; Cao, Y. Enhancing Catalytic Activity of CO2 Electrolysis by Building Efficient and Durable Heterostructure for Solid Oxide Electrolysis Cell Cathode. J. Power Sources 2023, 574, 233134. [Google Scholar] [CrossRef]
  26. Khan, I.; Sun, N.; Zhang, Z.; Li, Z.; Humayun, M.; Ali, S.; Qu, Y.; Jing, L. Improved Visible-Light Photoactivities of Porous LaFeO3 by Coupling with Nanosized Alkaline Earth Metal Oxides and Mechanism Insight. Catal. Sci. Technol. 2019, 9, 3149–3157. [Google Scholar] [CrossRef]
  27. Li, G.; Huang, S.; Li, K.; Zhu, N.; Zhao, B.; Zhong, Q.; Zhang, Z.; Ge, D.; Wang, D. Near-Infrared Responsive Z-Scheme Heterojunction with Strong Stability and Ultra-High Quantum Efficiency Constructed by Lanthanide-Doped Glass. Appl. Catal. B 2022, 311, 121363. [Google Scholar] [CrossRef]
  28. Khan, I.; Luo, M.; Guo, L.; Khan, S.; Shah, S.A.; Khan, I.; Khan, A.; Wang, C.; Ai, B.; Zaman, S. Synthesis of Phosphate-Bridged g-C3N4/LaFeO3 Nanosheets Z-Scheme Nanocomposites as Efficient Visible Photocatalysts for CO2 Reduction and Malachite Green Degradation. Appl. Catal. A Gen. 2022, 629, 118418. [Google Scholar] [CrossRef]
  29. Luo, J.; Han, H.; Wang, X.; Qiu, X.; Liu, B.; Lai, Y.; Chen, X.; Zhong, R.; Wang, L.; Wang, C. Single-Atom Nb Anchored on Graphitic Carbon Nitride for Boosting Electron Transfer towards Improved Photocatalytic Performance. Appl. Catal. B 2023, 328, 122495. [Google Scholar] [CrossRef]
  30. Adeel, M.; Saeed, M.; Khan, I.; Muneer, M.; Akram, N. Synthesis and Characterization of Co-ZnO and Evaluation of Its Photocatalytic Activity for Photodegradation of Methyl Orange. ACS Omega 2021, 6, 1426–1435. [Google Scholar] [CrossRef]
  31. Chen, M.; Yao, J.; Huang, Y.; Gong, H.; Chu, W. Enhanced Photocatalytic Degradation of Ciprofloxacin over Bi2O3/(BiO)2CO3 Heterojunctions: Efficiency, Kinetics, Pathways, Mechanisms and Toxicity Evaluation. Chem. Eng. J. 2018, 334, 453–461. [Google Scholar] [CrossRef]
  32. Song, C.; Wang, L.J.; Sun, S.M.; Wu, Y.; Xu, L.J.; Gan, L. Preparation of Visible-Light Photocatalysts of Bi2O3/Bi Embedded in Porous Carbon from Bi-Based Metal Organic Frameworks for Highly Efficient Rhodamine B Removal from Water. Xinxing Tan. Cailiao/New Carbon. Mater. 2020, 35, 609–618. [Google Scholar] [CrossRef]
  33. Bhaviya Raj, R.; Umadevi, M.; Parimaladevi, R. Synergy Photodegradation of Basic Dyes by Zno/Bi2o3 Nanocomposites under Visible Light Irradiation. Iran. J. Chem. Chem. Eng. 2021, 40, 1121–1131. [Google Scholar] [CrossRef]
  34. Kaur, G.; Sharma, S.; Bansal, P. Visible Light Responsive Heterostructured α-Bi2O3/Zno Doped β-Bi2O3 Photocatalyst for Remediation of Organic Pollutants. Desalination Water Treat. 2020, 201, 349–355. [Google Scholar] [CrossRef]
  35. Ansari, S.; Ansari, M.S.; Satsangee, S.P.; Jain, R. Bi2O3/ZnO Nanocomposite: Synthesis, Characterizations and Its Application in Electrochemical Detection of Balofloxacin as an Anti-Biotic Drug. J. Pharm. Anal. 2021, 11, 57–67. [Google Scholar] [CrossRef]
  36. Liu, L.; Li, Y.; Yuan, S.; Ge, M.; Ren, M.; Sun, C.; Zhou, Z. Nanosheet-Based NiO Microspheres: Controlled Solvothermal Synthesis and Lithium Storage Performances. J. Phys. Chem. C 2009, 114, 251–255. [Google Scholar] [CrossRef]
  37. Jegatha Christy, A.; Umadevi, M. Novel Combustion Method to Prepare Octahedral NiO Nanoparticles and Its Photocatalytic Activity. Mater. Res. Bull. 2013, 48, 4248–4254. [Google Scholar] [CrossRef]
  38. Wu, Z.; Wang, Y.; Sun, L.; Mao, Y.; Wang, M.; Lin, C. An Ultrasound-Assisted Deposition of NiO Nanoparticles on TiO2 Nanotube Arrays for Enhanced Photocatalytic Activity. J. Mater. Chem. A Mater. 2014, 2, 8223–8229. [Google Scholar] [CrossRef]
  39. Ji, Z.; Natu, G.; Wu, Y. Cyclometalated Ruthenium Sensitizers Bearing a Triphenylamino Group for P-Type NiO Dye-Sensitized Solar Cells. ACS Appl. Mater. Interfaces 2013, 5, 8641–8648. [Google Scholar] [CrossRef]
  40. Wan, X.; Yuan, M.; Tie, S.L.; Lan, S. Effects of Catalyst Characters on the Photocatalytic Activity and Process of NiO Nanoparticles in the Degradation of Methylene Blue. Appl. Surf. Sci. 2013, 277, 40–46. [Google Scholar] [CrossRef]
  41. Yi, S.; Yue, X.; Xu, D.; Liu, Z.; Zhao, F.; Wang, D.; Lin, Y. Study on Photogenerated Charge Transfer Properties and Enhanced Visible-Light Photocatalytic Activity of p-Type Bi2O3/n-Type ZnO Heterojunctions. New J. Chem. 2015, 39, 2917–2924. [Google Scholar] [CrossRef]
  42. Saeed, M.; Haq, A.U.; Muneer, M.; Ahmad, A.; Bokhari, T.H.; Sadiq, Q. Synthesis and Characterization of Bi2O3 and Ag-Bi2O3 and Evaluation of Their Photocatalytic Activities towards Photodegradation of Crystal Violet Dye. Phys. Scr. 2021, 96, 125707. [Google Scholar] [CrossRef]
  43. Balachandran, S.; Swaminathan, M. Facile Fabrication of Heterostructured Bi2O3-ZnO Photocatalyst and Its Enhanced Photocatalytic Activity. J. Phys. Chem. C 2012, 116, 26306–26312. [Google Scholar] [CrossRef]
  44. Jabeen Fatima, M.J.; Navaneeth, A.; Sindhu, S. Improved Carrier Mobility and Bandgap Tuning of Zinc Doped Bismuth Oxide. RSC Adv. 2015, 5, 2504–2510. [Google Scholar] [CrossRef]
  45. Song, L.; Zhang, S. A Simple Mechanical Mixing Method for Preparation of Visible-Light-Sensitive NiO–CaO Composite Photocatalysts with High Photocatalytic Activity. J. Hazard. Mater. 2010, 174, 563–566. [Google Scholar] [CrossRef]
  46. Davar, F.; Fereshteh, Z.; Salavati-Niasari, M. Nanoparticles Ni and NiO: Synthesis, Characterization and Magnetic Properties. J. Alloys Compd. 2009, 476, 797–801. [Google Scholar] [CrossRef]
  47. Vimal Kumar, M.; Gokul Raja, T.S.; Selvakumar, N.; Jeyasubramanaian, K. Synthesis and Characterization of NiO-ZnO Nanocomposite by a Cost Efficient Self-Combustion Technique. J. Achiev. Mater. Manuf. Eng. 2016, 79, 13–18. [Google Scholar] [CrossRef]
  48. Saeed, M.; Adeel, S.; Muneer, M.; ul Haq, A. Photo Catalysis: An Effective Tool for Treatment of Dyes Contaminated Wastewater. Bioremediat. Biotechnol. 2020, 2, 175–187. [Google Scholar] [CrossRef]
  49. Nisar, A.; Saeed, M.; Usman, M.; Muneer, M.; Adeel, M.; Khan, I.; Akhtar, J. Kinetic Modeling of ZnO-RGO Catalyzed Degradation of Methylene Blue. Int. J. Chem. Kinet. 2020, 52, 645–654. [Google Scholar] [CrossRef]
  50. Saeed, M.; Muneer, M.; Khosa, M.K.K.; Akram, N.; Khalid, S.; Adeel, M.; Nisar, A.; Sherazi, S. Azadirachta Indica Leaves Extract Assisted Green Synthesis of Ag-TiO2 for Degradation of Methylene Blue and Rhodamine B Dyes in Aqueous Medium. Green. Process. Synth. 2019, 8, 659–666. [Google Scholar] [CrossRef]
  51. Valian, M.; Beshkar, F.; Salavati-Niasari, M. Urchin-like Dy2CoMnO6 Double Perovskite Nanostructures: New Simple Fabrication and Investigation of Their Photocatalytic Properties. J. Mater. Sci. Mater. Electron. 2017, 28, 12440–12447. [Google Scholar] [CrossRef]
  52. Beshkar, F.; Zinatloo-Ajabshir, S.; Bagheri, S.; Salavati-Niasari, M. Novel Preparation of Highly Photocatalytically Active Copper Chromite Nanostructured Material via a Simple Hydrothermal Route. PLoS ONE 2017, 12, e0158549. [Google Scholar] [CrossRef] [PubMed]
  53. Deng, Y.; Chen, Y.; Chen, B.; Ma, J. Preparation, Characterization and Photocatalytic Activity of CuBi2O4/NaTaO3 Coupled Photocatalysts. J. Alloys Compd. 2013, 559, 116–122. [Google Scholar] [CrossRef]
  54. Saeed, M.; Alwadai, N.; ben Farhat, L.; Baig, A.; Nabgan, W.; Iqbal, M. Co3O4-Bi2O3 Heterojunction: An Effective Photocatalyst for Photodegradation of Rhodamine B Dye. Arab. J. Chem. 2022, 15, 103732. [Google Scholar] [CrossRef]
  55. Alshorifi, F.T.; Alswat, A.A.; Mannaa, M.A.; Alotaibi, M.T.; El-Bahy, S.M.; Salama, R.S. Facile and Green Synthesis of Silver Quantum Dots Immobilized onto a Polymeric CTS-PEO Blend for the Photocatalytic Degradation of p-Nitrophenol. ACS Omega 2021, 6, 30432–30441. [Google Scholar] [CrossRef]
  56. Rohilla, S.; Gupta, A.; Kumar, V.; Kumari, S.; Petru, M.; Amor, N.; Noman, M.T.; Dalal, J. Excellent Uv-light Triggered Photocatalytic Performance of Zno.Sio2 Nanocomposite for Water Pollutant Compound Methyl Orange Dye. Nanomaterials 2021, 11, 2548. [Google Scholar] [CrossRef]
  57. de Moraes, N.P.; Silva, F.N.; da Silva, M.L.C.P.; Campos, T.M.B.; Thim, G.P.; Rodrigues, L.A. Methylene Blue Photodegradation Employing Hexagonal Prism-Shaped Niobium Oxide as Heterogeneous Catalyst: Effect of Catalyst Dosage, Dye Concentration, and Radiation Source. Mater. Chem. Phys. 2018, 214, 95–106. [Google Scholar] [CrossRef]
  58. Mostafa, E.M.; Amdeha, E. Enhanced Photocatalytic Degradation of Malachite Green Dye by Highly Stable Visible-Light-Responsive Fe-Based Tri-Composite Photocatalysts. Environ. Sci. Pollut. Res. 2022, 29, 69861–69874. [Google Scholar] [CrossRef]
  59. Patil, S.B.; Ravishankar, T.N.; Lingaraju, K.; Raghu, G.K.; Nagaraju, G. Multiple Applications of Combustion Derived Nickel Oxide Nanoparticles. J. Mater. Sci. Mater. Electron. 2018, 29, 277–287. [Google Scholar] [CrossRef]
  60. Khairnar, S.D.; Shrivastava, V.S. Facile Synthesis of Nickel Oxide Nanoparticles for the Degradation of Methylene Blue and Rhodamine B Dye: A Comparative Study. J. Taibah Univ. Sci. 2019, 13, 1108–1118. [Google Scholar] [CrossRef]
  61. Alshorifi, F.T.; Ali, S.L.; Salama, R.S. Promotional Synergistic Effect of Cs–Au NPs on the Performance of Cs–Au/MgFe2O4 Catalysts in Catalysis 3,4-Dihydropyrimidin-2(1H)-Ones and Degradation of RhB Dye. J. Inorg. Organomet. Polym. Mater. 2022, 32, 3765–3776. [Google Scholar] [CrossRef]
  62. El-Hakam, S.A.; ALShorifi, F.T.; Salama, R.S.; Gamal, S.; El-Yazeed, W.S.A.; Ibrahim, A.A.; Ahmed, A.I. Application of Nanostructured Mesoporous Silica/Bismuth Vanadate Composite Catalysts for the Degradation of Methylene Blue and Brilliant Green. J. Mater. Res. Technol. 2022, 18, 1963–1976. [Google Scholar] [CrossRef]
  63. Dhiman, P.; Sharma, G.; Alodhayb, A.N.; Kumar, A.; Rana, G.; Sithole, T.; ALOthman, Z.A. Constructing a Visible-Active CoFe2O4@Bi2O3/NiO Nanoheterojunction as Magnetically Recoverable Photocatalyst with Boosted Ofloxacin Degradation Efficiency. Molecules 2022, 27, 8234. [Google Scholar] [CrossRef] [PubMed]
  64. Shahzad, K.; Fernandez-Garcia, J.; Khan, M.I.; Shanableh, A.; Khan, N.A.; ur Rehman, A. Formulation of Bismuth (Bi2O3) and Cerium Oxides (CeO2) Nanosheets for Boosted Visible Light Degradation of Methyl Orange and Methylene Blue Dyes in Water. Catalysts 2022, 12, 1197. [Google Scholar] [CrossRef]
  65. Masula, K.; Sreedhar, P.; Vijay Kumar, P.; Bhongiri, Y.; Pola, S.; Basude, M. Synthesis and Characterization of NiO–Bi2O3 Nanocomposite Material for Effective Photodegradation of the Dyes and Agricultural Soil Pollutants. Mater. Sci. Semicond. Process 2023, 160, 107432. [Google Scholar] [CrossRef]
  66. Banoth, P.; Narsaiah, B.P.; De Los Santos Valladares, L.; Kargin, J.; Kollu, P. Single-Phase BiFeO3 and BiFeO3-Fe2O3 Nanocomposite Photocatalysts for Photodegradation of Organic Dye Pollutants. Nanoscale Adv. 2023, 5, 2646–2656. [Google Scholar] [CrossRef]
  67. Poorsajadi, F.; Sayadi, M.H.; Hajiani, M.; Rezaei, M.R. Synthesis of CuO/Bi2O3 Nanocomposite for Efficient and Recycling Photodegradation of Methylene Blue Dye. Int. J. Environ. Anal. Chem. 2022, 102, 7165–7178. [Google Scholar] [CrossRef]
  68. Zhang, Q.; Wang, D.; Yu, R.; Luo, L.; Li, W.; Cheng, J.; Zhang, Y. Excellent Photocatalytic Degradation of Rhodamine B over Bi2O3 supported on Zn-MOF Nanocomposites under Visible Light. Green. Process. Synth. 2023, 12, 20228123. [Google Scholar] [CrossRef]
Figure 1. XRD analysis of Bi2O3, NiO, and Bi2O3–NiO samples.
Figure 1. XRD analysis of Bi2O3, NiO, and Bi2O3–NiO samples.
Water 15 03182 g001
Figure 2. TGA analysis of Bi2O3, NiO, and Bi2O3–NiO samples.
Figure 2. TGA analysis of Bi2O3, NiO, and Bi2O3–NiO samples.
Water 15 03182 g002
Figure 3. EDX analysis of Bi2O3, NiO, and Bi2O3–NiO samples.
Figure 3. EDX analysis of Bi2O3, NiO, and Bi2O3–NiO samples.
Water 15 03182 g003
Figure 4. SEM analysis of Bi2O3, NiO, and Bi2O3–NiO samples.
Figure 4. SEM analysis of Bi2O3, NiO, and Bi2O3–NiO samples.
Water 15 03182 g004
Figure 5. Comparison of photocatalytic activity of Bi2O3, NiO, and Bi2O3-NiO towards photodegradation of methyl orange.
Figure 5. Comparison of photocatalytic activity of Bi2O3, NiO, and Bi2O3-NiO towards photodegradation of methyl orange.
Water 15 03182 g005
Figure 6. Mechanism of photocatalytic process.
Figure 6. Mechanism of photocatalytic process.
Water 15 03182 g006
Figure 7. Kinetics analyses of degradation data of MO.
Figure 7. Kinetics analyses of degradation data of MO.
Water 15 03182 g007
Figure 8. Dependence of photocatalytic activity on catalyst dosage.
Figure 8. Dependence of photocatalytic activity on catalyst dosage.
Water 15 03182 g008
Figure 9. Dependance of photocatalytic activity on concentration of methyl orange.
Figure 9. Dependance of photocatalytic activity on concentration of methyl orange.
Water 15 03182 g009
Table 1. Instruments used for characterization of Bi2O3, NiO, and Bi2O3–NiO samples.
Table 1. Instruments used for characterization of Bi2O3, NiO, and Bi2O3–NiO samples.
NoCharacterization TechniqueInstrumentModel
1XRD analysisX-ray diffractometerJOEL-JDX-3532, Japan
2EDX analysisEDX spectrophotometerJSM5910, UK
3SEM analysisScanning electron microscopeJEOL-JSM 5910, Japan
4FTIR analysisIR spectrophotometerBruker VRTEX70, USA
5Thermal gravimetric analysisTGA analyzerPerkin Elmer 6300 TGA analyzer, USA
6Analysis of dye solutionUV-visible spectrophotometerHitachi U-2800, Japan
Table 2. Rate constants determined by the non-linear method of analysis.
Table 2. Rate constants determined by the non-linear method of analysis.
CatalystkobsR2
NiO0.00390.99
Bi2O30.00720.99
Bi2O3-NiO0.01250.94
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ashfaq, M.; Ali, A.; Abbood, N.K.; Panchal, S.; Akram, N.; Saeed, M.; Doshi, O.P.; Ali, F.; Muhammad, S.; Sameeh, M.Y.; et al. Enhanced Photocatalytic Activity of the Bi2O3-NiO Heterojunction for the Degradation of Methyl Orange under Irradiation of Sunlight. Water 2023, 15, 3182. https://doi.org/10.3390/w15183182

AMA Style

Ashfaq M, Ali A, Abbood NK, Panchal S, Akram N, Saeed M, Doshi OP, Ali F, Muhammad S, Sameeh MY, et al. Enhanced Photocatalytic Activity of the Bi2O3-NiO Heterojunction for the Degradation of Methyl Orange under Irradiation of Sunlight. Water. 2023; 15(18):3182. https://doi.org/10.3390/w15183182

Chicago/Turabian Style

Ashfaq, Muhammad, Akbar Ali, Nabeel K. Abbood, Sandeep Panchal, Nadia Akram, Muhammad Saeed, Ojas Prakashbhai Doshi, Faiz Ali, Shabbir Muhammad, Manal Y. Sameeh, and et al. 2023. "Enhanced Photocatalytic Activity of the Bi2O3-NiO Heterojunction for the Degradation of Methyl Orange under Irradiation of Sunlight" Water 15, no. 18: 3182. https://doi.org/10.3390/w15183182

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