Next Article in Journal
Highly Tuning of Sunlight-Photocatalytic Properties of SnO2 Nanocatalysts: Function of Gd/Fe Dopants
Previous Article in Journal
Constructing Polyphosphazene Microsphere-Supported Pd Nanocatalysts for Efficient Hydrogenation of Quinolines under Mild Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization of the Novel Heterojunction Photocatalyst Sm2NdSbO7/BiDyO3 for Efficient Photodegradation of Methyl Parathion

1
School of Physics, Changchun Normal University, Changchun 130032, China
2
State Key Laboratory of Pollution Control and Resource Reuse, School of the Environment, Nanjing University, Nanjing 210093, China
*
Author to whom correspondence should be addressed.
Catalysts 2024, 14(6), 346; https://doi.org/10.3390/catal14060346
Submission received: 2 May 2024 / Revised: 19 May 2024 / Accepted: 21 May 2024 / Published: 27 May 2024

Abstract

:
A new catalyst, Sm2NdSbO7, was synthesized for the first time by solid-phase sintering. The study utilized X-ray diffraction, transmission electron microscope energy dispersive X-ray spectroscopy, and X-ray photoelectron spectroscopy to examine the structural characteristics of monocrystal BiDyO3, monocrystal Sm2NdSbO7 and Sm2NdSbO7/BiDyO3 heterojunction photocatalysts (SBHP) prepared by solid-phase sintering. The Sm2NdSbO7 photocatalyst owned a pyrochlorite structure, belonged to the face-centered cubic crystal system, possessed a space group of Fd3m and a bandgap width of 2.750 eV. After 145 min of visible light irradiation (145-VLIRD), the removal rate (RMR) of methyl parathion (MP) or total organic carbon of SBHP was 100% or 97.58%, respectively. After 145-VLIRD, the photocatalytic degradation rates of SBHP to MP were 1.13 times, 1.20 times, and 2.43 times higher than those of the Sm2NdSbO7 photocatalyst, the BiDyO3 photocatalyst, and the nitrogen-doped TiO2 catalyst, respectively. The experimental results showed that SBHP had good photocatalytic activity. After four cycles of cyclic degradation experiments with SBHP, the elimination rates of MP were 98.76%, 97.44%, 96.32%, and 95.72%, respectively. The results showed that SBHP had good stability. Finally, the possible degradation pathways and degradation mechanisms of MP were speculated. In this study, we successfully developed a high-efficiency heterojunction catalyst which responded to visible light and possessed significant photocatalytic activity. The catalyst could be used in photocatalytic reaction system for eliminating the harmful organic pollutants from wastewater.

1. Introduction

Global population growth has put enormous pressure on the food system; thus, more pesticides are used to increase crop yields to meet food demand. Organophosphorus pesticides are mainly used for controlling plant diseases, insect pests, weeds, etc., and at the same time, food production could be increased. Organophosphorus pesticides have the advantages of economy, high efficiency and convenience of use. In agricultural production, the widespread use of organophosphorus pesticides led to different degrees of residues in crops, resulting in water pollution [1,2,3]. The residues of organophosphorus pesticides might cause acute poisoning. Therefore, it was urgent to find an effective method for removing pesticide residues from wastewater and purifying water resources [4].
Methyl parathion (MP), one of the most widely used organophosphorus pesticides in agriculture, could effectively control a variety of pests. MP was highly toxic and harmful to the nervous systems of humans and animals. In addition, MP was easily soluble in water; once the excess MP was dissolved in the underground water, it would cause serious environmental pollution. For the degradation of MP, the traditional methods mainly included chemical oxidation and biological degradation. However, there were several drawbacks associated with these approaches. For example, the chemical oxidation method was complex and required a large number of chemical reagents, and it also produced a large amount of waste liquid, which might lead to secondary pollution. Although the biodegradation method could effectively degrade organic matter, it required specific microbial populations, and the degradation rate was relatively slow [5,6,7,8]. Therefore, it was necessary to find a more effective degradation method for reducing the impact of its residues on the environment and humans.
Compared with other degradation methods for the degradation of MP, photocatalytic technology was a more promising choice. Photocatalysis is a chemical reaction process that uses light energy to stimulate catalysts. This technology has a wide range of applications in science and technology. Moreover, it could use solar energy, which is an environmentally friendly and energy-saving technology [9,10]. The principle of photocatalytic technology was mainly based on the photocatalytic activity of photocatalysts. When incident light with energy greater than the band gap width (Eg) was used to irradiate the photocatalysts, the electrons on the valence band of the photocatalysts were excited by the incident light and transited to the conduction band. As a result, negatively charged and highly reactive electrons formed on the conduction band, and correspondingly, positively charged holes could be generated on the valence band [11]. These photogenerated electrons and holes participated in the redox reactions of organic matter at the surface active sites and decomposed organic pollutants into harmless substances such as carbon dioxide and water, thereby achieving the degradation of organic pollutants. Therefore, the development of a photocatalyst with high degradation efficiency was the key to the treatment of MP, which is derived from pesticide wastewater (PW).
Traditional commercial metal oxide photocatalysts such as TiO2 and ZnO could only be stimulated by ultraviolet light or near-ultra-violet light radiation due to the band gap. These materials possessed other limitations, such as weak light scattering properties and the presence of point defects [12,13]. In order to effectively utilize the largest proportion of the solar spectrum in the visible light region (λ > 400 nm), it was feasible to construct a composite material system with a more complex construction compared with the single metal oxide [14,15]. Based on the previous reports, A2B2O7 and ABO3 compounds showed better photocatalytic performance under visible light irradiation (VLIRD) [16]. The A2B2O7 and ABO3 compounds were characterized by the substitution of metal ions to improve the photocatalytic efficiency while maintaining their structure. Xing et al. [17,18] prepared Bi2Sn2O7 and Y2Ti2O7, which showed good photocatalytic performance under VLIRD. Wang et al. studied the ABO3-type (A = Ca, Sr, Ba; B = Ti, Zr) catalyst, which indicated that the structure had the potential to improve performance by changing the structure [19].
In a prior investigation [20], it was determined that Bi2InNbO7 exhibited a pyrochlore structure and a face-centered cubic crystal system. The structure of Bi2InNbO7 had the capability to improve photocatalytic activity. We anticipated that replacing Sm3+ and Bi3+ in Bi2InNbO7, Nd3+ and In3+ in Bi2InNbO7, and Sb5+ and Nb5+ in Bi2InNbO7 might potentially boost the carrier concentration based on the study provided. We hypothesized that elevating the carrier concentration would yield an improved Sm2NdSbO7 photocatalyst with enhanced photocatalytic capabilities.
In the past, many researchers had studied various strategies for improving the photocatalytic efficiency of photocatalysts [21,22,23,24,25], such as proper surface engineering [26], ion doping [27], and the formation of semiconductor heterojunctions with metals [28,29,30,31,32,33,34] or other semiconductors. Among these strategies, the construction of semiconductor heterojunctions has received widespread attention owing to its perfect effect on improving photocatalytic activity [35,36,37,38,39,40,41,42]. In addition, the construction of heterostructural photocatalysts has shown great prospects for improving photocatalytic performance [43,44,45,46,47,48,49]. The interlaced band structure in the heterostructure formed a tight interface, resulting in the establishment of a strong internal electric field near the interface. This promoted the efficient transfer and separation of photocharged load charges [50,51]. Higher photocatalytic efficiency could also be obtained by constructing many different types of heterojunction catalysts using different methods [52]. Baaloudj and Nguyen-Tri et al. discussed the formation, morphological modification, doping, and hybridization processes of advanced engineering strategies for designing silicate-based photocatalysts, including heterojunctions. Each of these strategies provided important implications for our research [53]. We wanted to create a photocatalyst by constructing a heterojunction. The SBHP was synthesized for the first time, and its structural type was Z-scheme heterojunction. The SBHP was synthesized for the first time, and its structural type is Z-scheme heterojunction. The removal rate of the SBHP prepared by us reached 98% under VLIRD for 120 min, which was better than other reports. For example, the GO-Fe3O4/Bi2MoO6 catalyst and NiO/Bi2MoO6 catalyst prepared by Nasiripur, P et al. [54,55] achieved a removal rate of 95% for MP derived from PW under the same illumination time.
This study utilized X-ray diffraction (XRD), transmission electron microscope energy dispersive X-ray spectroscopy (TEM-EDS), and X-ray photoelectron spectroscopy (XPS) to examine the structural characteristics of monocrystal BiDyO3 and single-phase Sm2NdSbO7 created through high-temperature solid-phase sintering. The elimination effectiveness of microplastics was measured during very low input rate digestion using several materials such as Sm2NdSbO7, BiDyO3, N-doped TiO2 (N-T), and SBHP. This work is intended to create a novel heterostructural catalyst to break down MP produced from PW using VLIRD. To our knowledge, no work has synthesized the catalyst Sm2NdSbO7, and this is the first time that it has been used in heterojunction photocatalytic applications. This work introduced a unique approach by utilizing high-temperature solid-phase synthesis to create a new Sm2NdSbO7 nanocatalyst and SBHP for the first time. An efficient photocatalyst responsive to visible light was developed, demonstrating great activity in removing MP originating from PW. The SBHP demonstrated superior efficiency in degrading organic pollutants originating from PW while ensuring safety.

2. Findings and Analysis

2.1. XRD Analysis

Figure 1 shows that the catalyst was well crystalline with no heterophase present. Data analysis was conducted using the Rietveld analysis approach in the Materials Studio application. It could be inferred from Figure 2a and Figure 3a that Sm2NdSbO7 with a Rp factor of 3.21% and BiDyO3 with a Rp factor of 4.22% exhibited good crystallization. The final structural refinement of Sm2NdSbO7 confirmed agreement between observed and calculated intensities of the pyrochlore-type structure. The crystal system was cubic with space group Fd3m, resulting in a lattice parameter of 10.58142 angstroms. The crystal structure of BiDyO3 was revised using the Rietveld method based on the XRD data from Figure 3. The final refinement for BiDyO3 showed excellent agreement between the observed and estimated intensities. BiDyO3 possessed a fluorite-type structure with a face-centered cubic crystal system and belonged to the space group Fm3m. The lattice parameter a for BiDyO3 in the improved finding was 5.455 angstroms. Figure 2b displays the atomic arrangement of Sm2NdSbO7, whereas Figure 3b illustrates the atomic configurations of BiDyO3. Table 1 displays the atomic coordinates and structural characteristics of Sm2NdSbO7. Table 2 displays the atomic coordinates and structural characteristics of BiDyO3.
The x-coordinate of the O(1) atom in the pyrochlore-type A2B2O7 compound (with cube syngony and space group Fd3m) serves as an indicator of changes in the crystal structure. When the length of the six A−O(1) bonds equals the length of the two A−O(2) bonds, the x-coordinate is 0.375. Information on MO6 (M = Nd3+ and Sb5+) could be derived from the x value [56]. The crystal structure of Sm2NdSbO7 exhibited a clear deformation of the MO6 (M = Nd3+ and Sb5+) octahedral due to an x-value shift of 0.375. To prevent the recombination of photogenerated electrons with photoinduced holes during the PHDE of MP under VLIRD, charge separation was necessary. Scientific reports by Inoue and Kudo suggest that local distortion of the MO6 octahedron in photocatalysts like BaTi4O9 and Sr2M2O7 (M = Nb5+ and Ta5+) is crucial for inhibiting charge recombination and enhancing photocatalytic activity [57,58]. The deformation of the MO6 (M = Nd3+ and Sb5+) octahedron in the crystal structure of Sm2NdSbO7 could be utilized to improve photocatalytic activity. The structure of Sm2NdSbO7 is a three-dimensional mesh composed of MO6 octahedra (where M = Nd3+ and Sb5+) that share angles. Sm3+ ions connected MO6 octahedra (where M = Nd3+ and Sb5+) into chains. There were two different Sm−O bond lengths observed: six Sm−O(1) bonds were 2.842 Å long, while two Sm−O(2) bonds were 2.479 Å long. Six bond lengths of M−O (M = Nd3+ and Sb5+) were 2.195 Å each, whereas M−Sm (M = Nd3+ and Sb5+) bond lengths measured 3.565 Å. The bond angle in the Sm2NdSbO7 crystal structure, including Nd3+ and Sb5+ ions, was measured at 139.424°. The bond angle between Sm-M-Sm (M = Nd3+ and Sb5+) was measured at 135.721° in the crystal structure of Sm2NdSbO7. The bond angle in the Sm2NdSbO7 crystal structure involving Sm-M-O (M = Nd3+ and Sb5+) was measured at 133.680°. Research on luminescence qualities determined that an M-O-M bond angle closer to 180° resulted in a higher degree of delocalized excited states [59]. The study found that the angle between MO6 (M = Nd3+ and Sb5+) octahedra with common angles, such as the M-O-M bond angle in Sm2NdSbO7, significantly influenced the photocatalytic activity of Sm2NdSbO7. As the M-O-M bond angle approached 180°, the mobility of photogenerated electrons and holes increased. The movement of electrons and holes generated by light impacts the efficiency of photocatalysis by influencing the likelihood of these particles reaching the reaction site on the catalyst’s surface.

2.2. FTIR Analysis

Figure 4 shows distinct absorption peaks in the FTIR spectra of SBHP, Sm2NdSbO7, and BiDyO3. The peaks identified were related to Sm−O, Nd−O, Sb−O, Sb−O−Sb, Bi−O, and Dy−O bonds. The stretching vibration (SV) of Sm−O occurred at 520 cm−1 [60,61], whereas the SV of Nd−O was seen at 630 cm−1 [62]. The Bi−O bending vibration (BV) peak was observed at 428 cm−1 [63], and the Dy−O SV peak was found at 613 cm−1 [64]. The Sb−O BV was detected at 458 cm−1 and 583 cm−1 [65]. The BV of Sb−O−Sb was shown by peaks at 658 cm−1 [66]. The wide peaks found between 3431 cm−1 and 3576 cm−1 represent the SV of O−H groups from chemisorbed water molecules [67,68]. The peak at 1632 cm−1 represents the BV mode of the O−H groups [69]. The bands found between 1379 cm−1 and 1632 cm−1 were identified as vibrations of C−H bonds from adsorbed water [70,71].

2.3. Raman Analysis

Figure 5 shows the Raman spectra of SBHP, Sm2NdSbO7, and BiDyO3. The Raman spectra of Sm2NdSbO7 displayed distinctive modes such as the Ag internal Sm−O stretching modes at 562 cm−1 [72] and peaks at 363 cm−1 related to the SV of Nd−O bonds [73]. The peaks at 504 cm−1 and 622 cm−1 are likely associated with the BV of Sb−O and Sb−O−Sb [74,75]. The Raman spectra of BiDyO3 displayed a wide band at 131 cm−1 and 641 cm−1, corresponding to the SV of Bi−O and Dy−O, respectively [76,77]. Notably, the Raman spectra of SBHP exhibited strong peaks encompassing the distinct absorption peaks of both Sm2NdSbO7 and BiDyO3, including peaks at 134 cm−1, 357 cm−1, 503 cm−1, 558 cm−1, 636 cm−1 and 641 cm−1.

2.4. UV–Vis Diffuse Reflectance Spectra

Figure 6 displays the absorption spectra of SBHP, Sm2NdSbO7, and BiDyO3 samples. The absorption edges were observed at 551 nm, 450 nm and 602 nm in the visible range of the spectra, as shown in Figure 6. The band gap energies of crystalline semiconductors could be calculated by identifying the point where the photon energy (hν) axis intersects with the extrapolated line from the linear section of the absorption edge of the Kubelka–Munk function (1) [78].
1 R d ( h ν ) 2 2 R d ( h ν ) = α ( h ν ) S
where S was the scattering factor, Rd was the diffuse reflectance, and α represented the absorption coefficient of radiation.
The optical absorption near the band edge of the crystalline semiconductors obeyed Equation (2) [79,80]:
Ahν = A (hν − Eg)n
A, α, Eg, and ν symbolize the proportional constant, absorption coefficient, band gap, and light frequency, respectively. The variable “n” dictates the nature of the transition in a semiconductor inside this equation. Eg and n could be determined using the following steps: (1) Plotting the natural logarithm of (αhν) against the natural logarithm of (hν-Eg) using an estimated value of Eg; (2) determining the value of n based on the slope of this graph; (3) improving the value of Eg by graphing (αhν)1/n against hν and extending the plot until (αhν) 1/n = 0 [81,82]. The Eg values of SBHP, Sm2NdSbO7, and BiDyO3 were determined to be 2.250 eV, 2.750 eV, and 2.060 eV, respectively, using the approach described above. The approximate value of n was around 2, and the optical transition was classified as an indirect transition.

2.5. Performance Characterization of Sm2NdSbO7/BiDyO3 Heterojunction Photocatalysts

Figure 7 shows the XPS measurement spectra of SBHP, Sm2NdSbO7 and BiDyO3. The presence of Sm, Nd, Sb, Bi, Dy and O elements in SBHP could be clearly seen, which proved the successful preparation of heterojunction structure. An observed carbon signal was identified as adventitious hydrocarbon, used as a calibration.
It could be seen from Figure 8 that various element peaks with specific binding energies were obtained; demonstrably, the Sm 3d5/2 peak of samarium for Sm2NdSbO7 was situated at 1084.09 eV, and the Nd 3d5/2 peak of neodymium for Sm2NdSbO7 was situated at 982.62 eV. The Sb 4d5/2 peak of antimony for Sm2NdSbO7 was situated at 35.09 eV. The Bi 4d5/2 peak of the bismuth for BiDyO3 was situated at 442.44 eV. The Dy 4d5/2 peak of the dysprosium for BiDyO3 was situated at 158.95 eV. Comparison with the XPS map of SBHP found that these peaks showed a slight shift toward higher binding energy. These shifts confirmed the presence of a strong interfacial interaction between Sm2NdSbO7 and BiDyO3, potentially resulting from electron transfer and delocalization between these two components in heterojunction photocatalytic materials.
Figure 8f displays the separated O 1s spectra of SBHP and Sm2NdSbO7. The peaks seen at 529.97 eV and 529.71 eV were attributed to lattice oxygen. The peaks at 531.12 eV and 530.86 eV represented the signal coming from hydroxyl groups. The peaks at 532.29 eV and 532.03 eV were related to the signal of oxygen vacancies. The deconvoluted O 1s peaks in the SBHP sample showed noticeable alterations compared to their positions in the pure Sm2NdSbO7 sample. The shifts indicated more evidence of interfacial interactions between Sm2NdSbO7 and BiDyO3 species. The spin-orbit disassociation numerical value between Sb 3d5/2 and Sb 3d3/2 was consistently measured as 7.1 eV for both Sm2NdSbO7 and SBHP, confirming the sole occurrence of Sb5+ species [83,84].
The XPS spectrum of SBHP indicated the presence of Nd, Sm, Bi, Sb, Dy, and O components in the synthesized SBHP. The XPS examination results indicated that the oxidation states of Nd, Sm, Bi, Sb, Dy, and O ions were +3, +3, +3, +5, +3, and −2, respectively. The surface elemental analysis results indicated an average atomic ratio of Sm:Nd:Sb:Bi:Dy:O of 731:365:368:324:326:3593. The samples of SBHP have atomic ratios of 1.99:0.99:1.00 for Sm:Nd:Sb and 0.99:1.00 for Bi:Dy. The elevated oxygen concentration was a result of the significant oxygen adsorption on the surface of SBHP. No hetero-peaks were detected in the XPS peak of the SBHP, indicating the absence of additional components.
Figure 9 and Figure 10 show that the bigger hexagonal particles are from Sm2NdSbO7, while the smaller circular and quadrangular particles are from BiDyO3. Figure 9 and Figure 10 show that the Sm2NdSbO7 particles were enveloped by smaller BiDyO3 particles, and the two particles were closely connected, suggesting the effective synthesis of SBHP. Sm2NdSbO7 exhibited an octahedral shape. Figure 9 displays experimental data revealing that BiDyO3 exhibited a consistent spherical shape and even particle distribution. The particle size of BiDyO3 was approximately 446 nm, whereas the particle size of Sm2NdSbO7 was around 775 nm.
The TEM-EDS investigation results indicated the absence of impurity elements in the SBHP compound. The pure phase of Sm2NdSbO7 matched the XRD data presented in Figure 1. The SBHP was found to contain neodymium, samarium, antimony, bismuth, dysprosium, and oxygen based on the analysis of Figure 10 and Figure 11. The previous findings aligned with the XPS results of the SBHP, as depicted in Figure 7 and Figure 8.
The atomic ratio of Sm:Nd:Sb:Bi:Dy:O in the SBHP, as determined by the EDS spectra, was 853:385:386:305:331:7830, which matched the XPS results of the SBHP. The atomic ratio between Sm2NdSbO7 and BiDyO3 was around 65:53. Based on the results provided, we may infer that the SBHP exhibited good purity given our preparation conditions. The specific surface area of SBHP heterojunction catalyst is 6.28 m2·g−1, that of Sm2NdSbO7 is 5.19 m2·g−1, that of BiDyO3 is 4.15 m2·g−1, and that of N-T is 106.48 m2·g−1. This indicated that the catalytic activity and catalytic efficiency of SBHP were increased due to the increase of surface active sites.

2.6. Electrochemical Characterization

Electrochemical impedance mapping is a crucial technique for analyzing the movement of photogenerated electrons and photoholes in the prepared photocatalyst at the solid/electrolyte interface. Decreasing the arc radius increases the efficiency of photogenerated electron and photogenerated hole transport in the photocatalyst. Figure 12 illustrates that the arc radius diameter sequence is BiDyO3 > Sm2NdSbO7 > SBHP. The results indicated that the prepared SBHP demonstrated superior separation of photogenerated electrons and holes, as well as faster interfacial charge migration.

2.7. Optical Characterization

The influence of the sample on the separation of photogenerated charge carriers (PCC) was examined using electrochemical impedance spectroscopy (EIS) techniques. Figure 13 demonstrates that the SBHP exhibits the highest intensity of photocurrent response when compared to Sm2NdSbO7 and BiDyO3, as shown by the results. The enhanced photocurrent in SBHP resulted from the efficient diffusion of photoexcited electrons and the rapid transfer of photoexcited holes to the BiDyO3 surface. The event occurred because of the electric potential difference between the valence bands of Sm2NdSbO7 and BiDyO3 in the SBHP composite. The enhanced photocurrent response in SBHP indicated effective separation of PCC and extended durability in comparison to Sm2NdSbO7 and BiDyO3 in the photodegradation process. The observations clarified the improved photocatalytic efficiency achieved with SBHP [85,86].
UPS spectra were performed to determine the ionization potential of Sm2NdSbO7 and BiDyO3. UPS analysis could determine the valence band potentials of both p-type and n-type semiconductor materials. Figure 14 shows the initial (Ei) and maximum (Ecutoff) binding energies for the semiconductor samples. The measured numerical values were 1.378 eV and 19.965 eV for Sm2NdSbO7, and 0.749 eV and 20.471 eV for BiDyO3. The ionization potentials of Sm2NdSbO7 and BiDyO3 were determined to be 2.613 eV and 1.478 eV, respectively, using an excitation energy of approximately 21.2 eV [87,88]. The conduction band potentials of Sm2NdSbO7 and BiDyO3 were determined to be −0.137 eV and −0.582 eV, respectively. The results provide further support for the mechanistic paradigm described in this study.

2.8. Photocatalytic Activity

In the photocatalytic degradation experiment, the adsorption equilibrium experiment was carried out between −40 min and 0 min, and the adsorption equilibrium experiment was mainly to exclude the influence of photocatalyst adsorption on pollutant concentration. After the adsorption equilibrium was reached, it could be considered that the subsequent decrease in the concentration of pollutants was caused by photocatalytic degradation. N-doped TiO2 (N-T), a widely recognized visible-light-responsive photocatalyst, was employed as a benchmark to assess and compare the differences in photodegradation efficiency among various catalyst samples. In this study, we chose the organophosphorus pesticide methylparathion, and the potential surface charge present is a negative charge.
Thedata in Figure 15 indicated that using SBHP, the removal rate (RMR) of methyl parathion (MP) from pesticide wastewater (PW) approached 100% after 145 min of visible light irradiation (145-VLIRD). The reaction rate (RTR) was 2.87 × 10−9 mol·L−1·s−1, with a photon efficiency (PE) of 0.0603%. All subsequent studies adhered to the identical 145-VLIRD protocol. Using Sm2NdSbO7 resulted in an 88.24% RMR of MP, with a RTR of 2.54 × 10−9 mol·L−1·s−1 and a PE of 0.0533%. Using BiDyO3, the RMR of MP from PW was 83.16%, with a RTR of 2.39 × 10−9 mol·L−1·s−1 and a PE of 0.0502%. Using N-T as the photocatalyst, MP RMR was 41.16%, with a RTR of 1.18 × 10−9 mol·L−1·s−1 and a PE of 0.0248%. The results indicated that SBHP could enhance the photodegradation efficiency of MP. Among the other photocatalysts tested, Sm2NdSbO7 showed the highest efficiency, followed by BiDyO3 and N-T. BiDyO3 exhibited higher efficiency than N-T. The study demonstrated that SBHP had the highest visible photocatalytic activity when compared to Sm2NdSbO7, BiDyO3 and N-T. After 145-VLIRD, the RMR of MP by SBHP was 1.13, 1.20 and 2.43 times higher than that of Sm2NdSbO7, BiDyO3 and N-T photocatalysts, respectively.
The concentration curve (CTC) of MP gradually decreases with the increase of VLIRD time. It could be seen from Figure 16 that MP was degraded by SBHP, Sm2NdSbO7, BiDyO3 and N-T, respectively. The RMR of TOC, which was derived from PW, reached 97.58%, 84.15%, 79.06% and 40.52%, respectively, after 145-VLIRD. In summary, it could be seen from the above results that the RMR of TOC in the presence of SBHP with descending MP was higher than that of Sm2NdSbO7, BiDyO3, or N-T. The above results also showed that in the presence of Sm2NdSbO7, the RMR of TOC during the degradation of MP was much higher than that of BiDyO3 or N-T, which meant that compared with Sm2NdSbO7, BiDyO3, or N-T, SBHP has the greatest efficiency in the degradation of MP.
Figure 17a shows that the RMR of MP reached 98.76%, 97.44%, 96.32% and 95.72% after being irradiated with SBHP for 145-VLIRD following four cycle tests for MP degradation. Figure 17b shows that the RMR of TOC was 96.13%, 95.04%, 94.02%, or 93.11%. Figure 17c shows that when using SBHP for degradation, the RMR of MP was 99.3%, 100% and 98.5%, respectively, at pH values of 3, 7 and 11. This indicated that the pH value had no significant effect on the removal rate. Figure 17a–c show that the experimental findings demonstrated the SBHP’s significant stability.
Figure 17a–c show that after four cycles of deterioration testing under VLIRD, the RMR of MP decreased by 4.28% due to SBHP, the RMR of TOC decreased by 4.47%, and the effect of different pH values decreased by 0.7% and 1.5%. There was no significant difference in degradation efficiency, indicating that the photochemical structure stability of the photocatalyst was maintained.
According to Figure 18a,b, the kinetic constants k obtained from the kinetic curves (KC) of MP concentration and VLIRD time reached 0.02127 min−1 or 0.01001 min−1 or 0.00836 min−1 or 0.00836 min−1, respectively, using SBHP, Sm2NdSbO7, BiDyO3, or N-T. And the kinetic constant k of TOC concentration (KTOC) reached 0.02091 min−1 or 0.00868 min−1 or 0.00734 min−1 or 0.00285 min−1.
The KTOC degradation of MP was lower than the KC degradation of MP in the same catalyst, suggesting the occurrence of PHDE intermediate products in MP during the PHDE of MP under VLIRD. Compared to the other three photocatalysts, SBHP exhibited superior efficiency in degrading MP.
The results from Figure 19a,b show that in the four-cycle deterioration tests, the kinetic constant k obtained from the kinetic curve of MP concentration and the VLIRD time of SBHP ranged from 0.02127 min−1 to 0.01626 min−1. The kinetic constant k for the degradation test in four cycles using SBHP ranged from 0.01829 min−1 to 0.01439 min−1. The experimental data from Figure 18a,b and Figure 19a,b indicated that the PHDE of MP, produced from PW by SBHP under VLIRD, followed first-order reaction kinetics.

2.9. Mechanism Study

During the initial stage of the photocatalytic experiment, various radical trappers were individually introduced into the MP solution to identify the active substance involved in the degradation of MP. We employed isopropanol (IPA) with captured hydroxyl radicals (•OH), benzoquinone (BQ) with captured superoxide anions (•O2), and ethylenediaminetetraacetic acid (EDTA) with trapped holes (h+). The concentration of IPA, BQ, and EDTA was 0.15 mmol/L, and the dosage of each trapping agents was 1 mL. Figure 20 demonstrates that the addition of IPA, BQ, or EDTA to the MP solution resulted in a reduction of 63.80%, 47.64%, or 34.64% in the RMR of MP compared to the control group. It could be inferred that •OH, H+, and •O2 were all active radicals involved in the breakdown of MP. It could be seen from Figure 20 that the •OH in the MP solution plays a dominant role in degradation by SBHP under VLIRD. The experiment showed that hydroxyl radical was the most effective in oxidizing and eliminating MP derived from PW, compared to superoxide anion or hole.
EPR analysis was used to study the production of •O2 and •OH in the PHDE process [89,90,91]. We prepared a solution by combining 20 mg of the SBHP sample, 90 µL of DMPO (1 mol/L), and 1 mL of deionized water for the purpose of detecting the •OH radicals or •O2 radicals generated by SBHP in our experiment. Figure 21 demonstrates that after being exposed to visible light for 10 min, a four-line signal with a 1:2:2:1 intensity ratio, characteristic of the DMPO •OH signal, was identified. In addition, the EPR spectra showed a clear signature of DMPO •O2 with four conspicuous peaks of equal intensity ratio 1:1:1:1. This observation suggested the presence of superoxide radicals. The results indicated the simultaneous generation of •OH and •O2. The elevated relative intensity of the EPR signals suggested a higher production of these reactive radicals. The concentration of hydroxyl radicals surpassed the generation of superoxide radicals. The outcomes were consistent with the results of the radical-scavenger tests discussed previously.
A variety of effective characterization techniques were used to thoroughly understand the interfacial carrier dynamics and recombination. The PL spectra in Figure 22 show emission peaks at 470 nm for all samples. The PL emission of the sample may be caused by the recombination of the excitation associated with the defect through the excitation–excitation collision process. After the sample is excited, an electron-hole pair is generated, where the hole is in the valence band or defect-related position and the electron is in the conduction band or defect-related position, both of which emit photons during the recombination process [92,93], thus producing a light emission spectrum as shown in Figure 22. The SBHP has the lowest emission peak intensity, which means the lowest recombination rate of electrons and holes, indicating that the SBHP has the highest catalytic activity. A heterostructured sample could significantly improve the photocatalytic action on MP. Moreover, the study also demonstrated additional proof that SBHP has the most potent photocatalytic properties. Figure 23a–c displays the TRPL spectra of SBHP (τ1 = 1.652 ns, τ2 = 146.5 ns, τave = 81.29 ns), which presented a much higher PCC lifetime than Sm2NdSbO71 = 1.752 ns, τ2 = 111.5 ns, τave = 38.49 ns) and BiDyO31 = 1.61 ns, τ2 = 36.56 ns, τave = 7.6 ns), which indicated that SBHP was endowed with unbeatable photocatalytic efficacy over the individual Sm2NdSbO7 and BiDyO3.

2.10. Analysis of Possible Photocatalytic Degradation Mechanisms

Figure 24 illustrates that when the SBHP is exposed to visible light, both Sm2NdSbO7 and BiDyO3 could absorb the light and produce electron-hole pairs internally. It is composed of two staggered semiconductor photocatalysts, Sm2NdSbO7 and BiDyO3. After the photo-excitation of the two catalysts to generate electron holes, the photogenerated holes of BiDyO3 react with the photogenerated electrons of Sm2NdSbO7. The electron holes in the two catalysts are retained separately for the redox reaction. The use of this system could not only realize the spatial separation of redox sites but also ensure that the photocatalyst could maintain a suitable valence band position so as to maintain a strong redox reaction ability [94]. Consequently, generating more oxidized radicals like •OH or •O2 could enhance the degrading efficiency of MP. The conduction band potential of BiDyO3 was −0.582 eV, higher than the negative value of O2/•O2 (−0.33 eV vs. NHE). This suggested that electrons in the conduction band of BiDyO3 could react with oxygen to form O, leading to the degradation of MP (as illustrated in path 1 in Figure 24). The valence band potential of Sm2NdSbO7 was 2.613 eV, adjusted for OH/•OH (2.38 eV vs. NHE), suggesting that the valence band hole of Sm2NdSbO7 could oxidize H2O or OH to •OH to break down MP, as demonstrated in path 2. The light-induced holes in the valence band of BiDyO3 or Sm2NdSbO7 could directly oxidize MP and degrade it because of their potent oxidizing ability, as demonstrated by path 3. The superior photocatalytic performance of SBHP in degrading MP was primarily due to the effective separation of electrons and holes caused by SBHP. The blue dashed line and orange dashed line indicate the valance band and conduction band charged chemical potential of Sm2NdSbO7 and BiDyO3, respectively. The crimson lines mark the electrochemical potentials of O2/•O2 and OH/•OH. The shape mark 1 to 3 represent the three locations where degradation reactions occur.
The intermediates of the PHDE of MP were C6H5OH(NO2) (m/z = 139), C6H5OH (m/z = 94), (CH3O)3P(S) (m/z = 156), C6H4(OH)2 (m/z = 110), C6H3(OH)3 (m/z = 126), C6H4(NH2)OP(O)(OCH3)2 (m/z = 247), and (CH3O)2P(O)OH (m/z = 126). Based on the intermediates examined above, a PHDE pathway for MP was proposed. Figure 25 shows that both the oxidation reaction and the hydroxylation reaction of MP were realized during the PHDE. Eventually, MP was converted to small-molecule organic compounds that eventually combined with other organically active groups to convert to NO3, SO42−, CO2, and H2O.

3. Materials and Methods

3.1. Materials and Reagents

Sm2O3 (purity = 99.9%), Nd2O3 (purity = 99.9%), Sb2O5 (purity = 99.9%), Bi2O3 (purity = 99.9%), Dy2O3 (purity = 99.9%), Ethylenediaminetetracetic acid (EDTA, C10H16N2O8, purity = 99.5%), isopropyl alcohol (IPA, C3H8O, purity ≥ 99.7%), benzoquinone (BQ, C6H4O2, purity ≥ 98.0%), the above chemical reagents were all purchased from Sinopharm Group Chemical Reagent Co., Ltd. (Shanghai, China). Absolute ethanol (C2H5OH, purity ≥ 99.5%) was purchased from Aladdin Group Chemical Reagent Co., Ltd. (Shanghai, China). Parathion methyl (C8H10NO5PS, purity ≥ 98%) was gas chromatography grade and was purchased from Tianjin Bodi Chemical Co., Ltd., Tianjin, China, as the model material.

3.2. Preparation Method of Sm2NdSbO7

The novel photocatalyst Sm2NdSbO7 was produced using the solid-phase sintering process. Powders with a molar ratio of 2:1:1 of Sm2O3, Nd2O3, and Sb2O5 were dehydrated at 200 °C for 4 h prior to the synthesis process. The precursors of Sm2NdSbO7 were combined in the correct proportions, formed into tiny columns, and placed in an alumina crucible. The raw materials and the tiny columns were removed from the electric furnace after being calcined at 400 °C for 2 h. The blended materials were ground and then placed in an electric furnace. The calcination procedure was conducted at 1050 °C for 30 h in an electric furnace.

3.3. Preparation Method of BiDyO3

BiDyO3 was synthesized using a high-temperature solid-phase sintering technique. Because Bi2O3 performance is unstable at high temperatures, we decided to increase the amount of Bi2O3 by 120% after conducting 5 experiments. The materials with a molar ratio of 1.2:1 of Bi2O3 to Dy2O3 were well mixed and then milled in a ball mill until the particle size of the powder reached 1–2 µm. All powders were dehydrated at 200 °C for 4 h before synthesis. The powders were combined in an aluminum oxide crucible, formed into disks, and then sintered in an electric furnace at 750 °C for 10 h. The mixture was sintered in an electric furnace at 1050 °C for 12 h after being crushed and pressed. After complete grinding, a pure BiDyO3 catalyst was finally achieved.

3.4. Synthesis of N-Doped TiO2

The N-doped TiO2 (N-T) catalyst was prepared using the sol-gel method, with tetrabutyl titanate as a precursor and ethanol as a solvent. The process unfolded in the following manner: Solution A was produced by mixing 17 mL of tetrabutyl titanate with 40 mL of pure ethyl alcohol. Solution B was produced by mixing 40 mL of pure ethyl alcohol, 10 mL of glacial acetic acid, and 5 mL of double-distilled water. Solution A was gradually introduced into solution B with vigorous stirring, leading to the creation of a transparent colloidal suspension. Subsequently, an aqua ammonia solution with a N/Ti ratio of 8 mol% was introduced into the transparent colloidal suspension and agitated using a magnetic stirrer for 1 h. The xerogel was created during a 2-day aging procedure. The xerogel was crushed into fine particles and thereafter subjected to a temperature of 500 °C for a duration of 2 h. The powder was treated in an agate mortar and sifted using a shaker to obtain N-T particles.

3.5. Synthesis of Sm2NdSbO7/BiDyO3 Heterojunction Photocatalysts

(1)
Sm2NdSbO7 and BiDyO3, respectively, were weighed and evenly mixed, then added to the ball mill to be ground into a powder;
(2)
The above powder was taken to dry, pressed into sheets, put it into a high-temperature sintering furnace for sintering, raising the temperature from room temperature to 400 °C at a heating rate of 10 °C/min, and then maintained at 400 °C for 4 h. Then, the temperature was raised from 400 °C to 800 °C at a heating rate of 9 °C/min, and then maintained at 800 °C for 12 h. The temperature was then raised from 800 °C to 1250 °C at a heating rate of 8 °C/min, then maintained at 1250 °C for 30 h. Finally, the first sintered tablet sample was cooled from 1250 °C to room temperature at a cooling rate of 8 °C/min.
(3)
The above first sintering tablet was taken, crushed, pressed into a sheet, and then placed in a high-temperature sintering furnace for sintering. The temperature was raised from room temperature to 500 °C at a heating rate of 10 °C/min, and then maintained at 500 °C for 5 h. Then, the temperature was raised from 500 °C to 900 °C at a heating rate of 9 °C/min, and maintained at 900 °C for 15 h. The temperature was raised from 900 °C to 1300 °C at a heating rate of 8 °C/min, and then maintained at 1300 °C for 25 h. Finally, the second sintered tablet sample was obtained by cooling from 1300 °C to room temperature at a cooling rate of 7.5 °C/min.
(4)
The above second sintering tablet was taken and crushed again, pressed into a sheet, and placed in a high-temperature sintering furnace for sintering. The temperature was raised from room temperature to 550 °C at a heating rate of 10 °C/min, and then maintained at 550 °C for 4 h. Then, the temperature was raised from 550 °C to 950 °C at a heating rate of 9 °C/min, and it was then maintained at 950 °C for 16 h. The temperature was raised from 950 °C to 1350 °C at a heating rate of 8 °C/min, and then maintained at 1350 °C for 35 h. Finally, the third sintered tablet sample was obtained by cooling from 1350 °C at a cooling rate of 7 °C/min.
(5)
The above third sintered tablet sample was then taken and crushed to obtain the Sm2NdSbO7/BiDyO3 powder catalytic material.

3.6. Characterizations

The pristine crystals of the fabricated designs were examined using an XRD (XRD, Shimadzu, XRD-6000, Cu Kα radiation, Kyoto, Japan). The patterns, shape, and microstructure were analyzed using TEM models (JEM-F200, FEI Tecnai G2 F20, Waltham, MA, USA). The elemental composition was determined by EDS. The sample was analyzed for its diffuse reflectance spectrum using a UV–Vis DRS (Shimadzu, UV-3600, Kyoto, Japan). Analyzed the functional groups and chemical bonds using FTIR (WQF-530A, Beifen-Ruili Analytical Instrument Group Co., Ltd., Beijing, China). The chemical bond interactions were examined using a Raman spectrometer (INVIA0919-06, RENISHAW plx, Wotton-under-Edge, Gloucestershire, UK). Analyzed the surface chemical composition and states of the sample using an XPS (PHI 5000 VersaProbe, UlVAC-PHI, Maoqi City, Japan) equipped with an Al-kα X-ray source. The ionization potential of the valence band of the samples was measured using ultraviolet photoelectron spectroscopy (UPS) with an Escalab 250 xi instrument from Thermo Fisher Scientific in Waltham, MA, USA. The free radicals in the samples were identified using an electron paramagnetic resonance spectrometer (EPR, A300, Bruker Corporation, Karlsruhe, Germany).
The characteristics of PL and TRPL were analyzed using a fluorescence spectrophotometer (FLS980, Edinburgh Instruments Ltd., Edinburgh, UK).
Which were analyzed using double-exponential decay Equation (3) [95]:
I t = I 0 + A 1 e x p t τ 1 + A 2 e x p t τ 2
In the provided equation, τ1 and τ2 represent the first- and second-order decay times, respectively, and A1 and A2 are the weighting coefficients of each decay channel [96]. To determine the average photogenerated charge carrier lifetime (τave), the following Equation (4) was utilized [97]:
τ a v e = A 1 τ 1 2 + A 2 τ 2 2 / A 1 τ 1 + A 2 τ 2

3.7. Photoelectrochemical Experiments

Electrochemical impedance spectroscopy (EIS) and photocurrent (PC) measurements were performed using a CHI660D electrochemical station manufactured by Chenhua Instruments Co. in Shanghai, China, using a standard three-electrode configuration. The system comprised a working electrode, counter electrode, and reference electrode, which were catalysts: a platinum plate and a commercial Ag/AgCl electrode, respectively. A 0.5 mol·L−1 aqueous solution of Na2SO4 was used as an electrolyte. Photochemical experiments were carried out utilizing a 500 W xenon lamp with a UV cut-off filter serving as the visible light emitter. The electrode was produced using the following method: A 0.03 g sample and 0.01 g of chitosan were dissolved in 0.45 mL of dimethylformamide to form a uniform suspension solution after one hour of ultrasonic treatment. The solution was applied to a 10 mm × 20 mm indium tin oxide (ITO) conducting glass. The working electrode was dried at 80 °C for 10 min.

3.8. Experimental Setup and Procedure

The research was carried out in a photocatalytic reactor (CEL-LB70, China Education Au-Light Technology Co., Ltd., Beijing, China) at a controlled temperature of 20 °C with circulating cooling water. A 500 W xenon lamp with a 420 nm cut-off filter was utilized to replicate sunlight irradiation. There were 12 identical quartz tubes, each with 40 mL of reaction solution, totaling 480 mL for pesticide wastewater (PW). The concentration of Sm2NdSbO7, BiDyO3, or SBHP was 0.75 g·L−1, and the concentration of methyl parathion (MP) was 0.025 mmol·L−1. The concentration of MP refers to the residual quantity following biodegradation in real PW with a starting MP concentration of 1.0 mmol·L−1. 3 mL of suspension was intermittently removed during the reaction. The catalyst was filtered out using a 0.22 μm PES polyether sulfone filter membrane. The concentration of MP left in the solution was analyzed using Agilent 200 high-performance liquid chromatography (Agilent Technologies, Palo Alto, CA, USA) with a UV detector and a Zorbax 300SB-C18 column (4.6 mm × 150 mm, 5 μm). A mobile phase composed of a 50% mixture of CH3CN and distilled deionized water was used. The UV detection was calibrated specifically for a wavelength of 254 nm. 10 μL of the post-photodegradation MP solution was injected at a flow rate of 1 mL·min−1. Prior to subjecting the solution to visible light, the combination of photocatalyst and MP was stirred in darkness for 45 min to achieve an adsorption/desorption equilibrium between the components and ambient oxygen.
During exposure to visible light, the suspension is agitated at a speed of 500 rpm. Before the photocatalytic experiment, we first carried out the adsorption equilibrium experiment under the condition of avoiding light. The sample was added to the solution, and the magnetic stirrer was activated for stirring until the dye contaminants in the solution reached the adsorption equilibrium state on the surface of the sample. The experimental process took 40 min. The adsorption equilibrium experiment is mainly to exclude the effect of adsorption on the concentration of pollutants or degraded substances.
The mineralization experimental data of MP in the reaction solution were evaluated with a TOC analyzer (TOC-5000 A, Shimadzu Corporation, Kyoto, Japan). TOC concentration was measured during the PHDE of MP using potassium acid phthalate (KHC8H4O4) or anhydrous sodium carbonate as a reference reagent. Calibration solutions containing potassium acid phthalate with a known carbon concentration varying from 0 to 100 mg·L−1 were prepared. TOC concentration was measured by analyzing six samples, each consisting of 45 mL of reaction solution.
MP and its intermediate degradation products were identified and measured using liquid chromatography–mass spectrometry (LC-MS) with a Thermo Quest LCQ Duo instrument. A total of 20 microliters of the solution resulting from the photocatalytic process were automatically fed into the LC-MS apparatus. The mobile phase consisted of 60% methanol and 40% ultrapure water, with a flow rate of 0.2 mL·min−1. The mass spectrometry setup involved an electrospray ionization interface, a capillary temperature of 27 °C with a voltage of 19.00 V, a spray voltage of 5000 V, and a consistent sheath gas flow rate. The spectra were obtained using negative ion scan mode with a mass-to-charge ratio (m/z) range of 50 to 600.
The radiometer (Model FZ-A, Photoelectric Instrument Factory, Beijing Normal University, Beijing, China) detected an incoming photon flux Io of 4.76 × 10−6 Einstein·L−1·s−1 within the wavelength range of 400–700 nm. The incident photon flux on the photoreactor was altered by changing the distance between the photoreactor and the Xe arc lamp.
The photonic efficiency was calculated in accordance with the following Equation (5):
ϕ = R/Io
where ϕ was the photonic efficiency (%), R was the degradation rate of MP (mol·L−1·s−1), and Io was the incident photon flux (Einstein·L−1·s−1).

4. Conclusions

Sm2NdSbO7 was synthesized for the first time using high-temperature solid-state sintering, resulting in significant photocatalytic activity. Sm2NdSbO7/BiDyO3 heterojunction photocatalyst (SBHP) was produced for the first time using a straightforward solid-phase sintering method. Photophysical properties and photocatalytic performance were analyzed using various techniques, including X-ray diffractometer (XRD), transmission electron microscope (TEM), X-ray photoelectron spectrograph (XPS), Fourier transform infrared dpectrometer (FTIR), Raman spectrometer, fluorescence spectrophotometer, ultraviolet photoelectron spectroscopy (UPS), photocurrent (PC) test, photoluminescence (PL) spectroscopy, electron paramagnetic resonance (EPR) spectroscopy, and UV–Vis diffuse reflectance spectrophotometer (UV–Vis DRS). The results indicate that Sm2NdSbO7 is a single-phase material with a pyrochlore structure. It exhibits a cubic crystal system inside the Fd3m space group. Sm2NdSbO7 has lattice parameters of a = 10.98142 Å and a band gap of 2.750 eV. SBHP was demonstrated to be a successful photocatalyst for eliminating methyl parathion (MP) from wastewater. After 145 min of visible light exposure (145-VLIRD), the removal rates of MP and TOC were 100% and 97.58%, respectively. The relative metabolic rate of methyl palmitate by solid-phase high-pressure (SBHP) treatment was 1.13, 1.20, or 2.43 times more than that of Sm2NdSbO7, or BiDyO3 as a catalyst, or N-doped TiO2 as a catalyst, respectively, and it increased after 145-VLIRD. Thus, it may be inferred that SBHP could be a successful approach for treating wastewater contaminated with MP. The potential photodegradation mechanism of MP was hypothesized.

Author Contributions

Conceptualization, data curation, formal analysis, investigation, methodology, software, visualization, writing—original draft preparation, writing—review and editing, J.L. (Jingfei Luan); software, data curation, methodology, writing—original draft preparation and validation, J.L. (Jun Li); formal analysis, writing—original draft preparation, validation, investigation, Y.Y.; software, visualization, validation, B.N.; methodology, software, validation and investigation, L.H.; software, visualization and validation, Y.W.; methodology, software, validation and investigation, Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Free Exploring Key Item of Natural Science Fundation of Science and Technology Bureau of Jilin Province of China (Grant No. YDZJ202101ZYTS161).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Derbalah, A.; Chidya, R.; Jadoon, W.; Sakugawa, H. Temporal trends in organophosphorus pesticides use and concentrations in river water in Japan, and risk assessment. J. Environ. Sci. 2019, 79, 135–152. [Google Scholar] [CrossRef] [PubMed]
  2. John, E.M.; Shaike, J.M. Chlorpyrifos: Pollution and remediation. Environ. Chem. Lett. 2015, 13, 269–291. [Google Scholar] [CrossRef]
  3. Dahshan, H.; Megahed, A.M.; Abd-Elall, A.M.M.; Abd-El-Kader, M.A.G.; Nabawy, E.; Elbana, M.H. Monitoring of pesticides water pollution-The Egyptian River Nile. J. Environ. Health Sci. 2016, 14, 15. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, F.; Cao, X.L.; Tian, D.M.; Li, H.B. Synthesis of Coumarin-Pillararene as a Selective Fluorescent Probe for Methyl-Parathion. Chin. J. Chem. 2015, 33, 368–372. [Google Scholar] [CrossRef]
  5. Ji, X.Y.; Wang, Q.; Zhang, W.D.; Yin, F. Research advances in organophosphorus pesticide degradation: A review. Fresen Environ. Bull. 2016, 5, 1556–1561. [Google Scholar]
  6. Sirotkina, M.; Lyagin, I.; Efremenko, E. Hydrolysis of organophosphorus pesticides in soil: New opportunities with ecocompatible immobilized His6-OPH. Int. Biodeterior. Biodegrad. 2012, 68, 18–23. [Google Scholar] [CrossRef]
  7. Harrache, Z.; Abbas, M.; Aksil, T.; Trari, M. Thermodynamic and kinetics studies on adsorption of indigo carmine from aqueous solution by activated carbon. Microchem. J. 2019, 144, 180–189. [Google Scholar] [CrossRef]
  8. Li, M.X.; Wang, H.T.; Wu, S.J.; Li, F.T.; Zhi, P.D. Adsorption of hazardous dyes indigo carmine and acid red on nanofiber membranes. RSC Adv. 2012, 2, 900–907. [Google Scholar] [CrossRef]
  9. Namini, A.S.; Delbari, S.A.; Mousavi, M.; Ghasemi, J.B. Synthesis and characterization of novel ZnO/NiCr2O4 nanocomposite for water purification by degradation of tetracycline and phenol under visible light irradiation. Mater. Res. Bull. 2021, 139, 111247. [Google Scholar] [CrossRef]
  10. Mokhtari, F.; Tahmasebi, N. Hydrothermal synthesis of W-doped BiOCl nanoplates for photocatalytic degradation of rhodamine B under visible light. J. Phys. Chem. Solids 2021, 149, 109804. [Google Scholar] [CrossRef]
  11. Subhan, M.A.; Rifat, T.P.; Saha, P.C.; Alam, M.M.; Asiri, A.M.; Rahman, M.M.; Akter, S.; Raihan, T.; Azad, A.K.; Uddin, J. Enhanced visible light-mediated photocatalysis, antibacterial functions and fabrication of a 3-chlorophenol sensor based on ternary Ag2O·SrO·CaO. RSC Adv. 2020, 10, 11274–11291. [Google Scholar] [CrossRef] [PubMed]
  12. Chandiran, A.K.; Abdi-Jalebi, M.; Nazeeruddin, M.K.; Grätzel, M. Analysis of Electron Transfer Properties of ZnO and TiO2 Photoanodes for Dye-Sensitized Solar Cells. ACS Nano 2014, 8, 2261–2268. [Google Scholar] [CrossRef] [PubMed]
  13. Himmah, S.W.; Diantoro, M.; Astarini1, N.A.; Tiana1, S.K.G.; Nasikhudin; Hidayat, A.; Taufiq, A. Structural, morphological, optical, and electrical properties of TiO2/ZnO rods multilayer films as photoanode on dye-sensitized solar cells. J. Phys. Conf. Ser. 2021, 1816, 012095. [Google Scholar] [CrossRef]
  14. Volnistem, E.A.; Bini, R.D.; Silva, D.M.; Rosso, J.M.; Dias, G.S.; Cotica, L.F.; Santos, I.A. Intensifying the photocatalytic degradation of methylene blue by the formation of BiFeO3/Fe3O4 nanointerfaces. Ceram. Int. 2020, 46, 18768–18777. [Google Scholar] [CrossRef]
  15. Volnistem, E.A.; Bini, R.D.; Dias, G.S.; Cotica, L.F.; Santos, I.A. Photodegradation of methylene blue by mechano synthesized BiFeO3 submicron particles. Ferroelectrics 2018, 534, 190–198. [Google Scholar] [CrossRef]
  16. Zou, Z.G.; Ye, J.H.; Arakawa, H. Photocatalytic behavior of a new series of In0.8M0.2TaO4 (M = Ni, Cu, Fe) photocatalysts in aqueous solutions. Catal. Lett. 2001, 75, 209–213. [Google Scholar] [CrossRef]
  17. Xing, Y.L.; Yin, X.T.; Que, Q.H.; Que, W.X. Fabrication of Ag2O-Bi2Sn2O7 heterostructured nanoparticles for enhanced photocatalytic activity. J. Nanosci. Nanotechnol. 2018, 18, 4306–4310. [Google Scholar] [CrossRef] [PubMed]
  18. Higashi, M.; Abe, R.; Sayama, K.; Sugihara, H.; Abe, Y. Improvement of photocatalytic activity of titanate pyrochlore Y2Ti2O7 by addition of excess Y. Chem. Lett. 2005, 34, 1122–1123. [Google Scholar] [CrossRef]
  19. Wang, M.H.; Zhu, C.Z.; Hu, H.S.; Lin, S.J.; Zhou, L.F.; Wang, Z.Z.; Wang, H.; Zhang, T.M.; Li, Y.H.; Yang, D.Y.; et al. Correlation between radiation resistance and structural factors of ABO3-type perovskites. Nim. B 2023, 536, 88–96. [Google Scholar] [CrossRef]
  20. Luan, J.F.; Tan, W.C. Preparation, photophysical and photocatalytic property characterization of Sm2FeSbO7 during visible light irradiation. Chin. J. Inorg. Chem. 2018, 34, 1950–1965. [Google Scholar]
  21. Chen, D.; Ray, A.K. Removal of toxic metal ions from wastewater by semiconductor photocatalysis. Chem. Eng. Sci. 2001, 56, 1561–1570. [Google Scholar] [CrossRef]
  22. Ramos-Delgado, N.A.; Gracia-Pinilla, M.A.; Maya-Trevino, L.; Hinojosa-Reyes, L.; Guzman-Mar, J.; Hernández-Ramírez, A. Solar photocatalytic activity of TiO2 modified with WO3 on the degradation of an organophosphorus pesticide. J. Hazard. Mater. 2013, 263 Pt 1, 36–44. [Google Scholar] [CrossRef]
  23. Yu, J.; Wang, S.; Low, J.; Xiao, W. Enhanced photocatalytic performance of direct Z-scheme g-C3N4-TiO2 photocatalysts for the decomposition of formaldehyde in air. Phys. Chem. Chem. Phys. 2013, 15, 16883–16890. [Google Scholar] [CrossRef] [PubMed]
  24. Hu, S.Z.; Li, F.Y.; Fan, Z.P.; Gui, J.Z. The effect of H2-CCl4 mixture plasma treatment on TiO2 photocatalytic oxidation of aromatic air contaminants under both UV and visible light. Chem. Eng. J. 2014, 236, 285–292. [Google Scholar] [CrossRef]
  25. Zou, X.; Li, X.; Zhao, Q.; Liu, S. Synthesis of LaVO4/TiO2 heterojunction nanotubes by sol-gel coupled with hydrothermal method for photocatalytic air purification. J. Colloid. Interf. Sci. 2012, 383, 13–18. [Google Scholar] [CrossRef]
  26. Medina-Valtierra, J.; Frausto-Reyes, C.; Ramirez-Ortiz, J.; Camarillo-Martinez, G. Self-cleaning test of doped TiO2-coated glass plates under solar exposure. Ind. Eng. Chem. Res. 2009, 48, 598–606. [Google Scholar] [CrossRef]
  27. Hebeish, A.A.; Abdelhady, M.M.; Youssef, A.M. TiO2 nanowire and TiO2 nanowire doped Ag-PVP nanocomposite for antimicrobial and self-cleaning cotton textile. Carbohyd. Polym. 2013, 91, 549–559. [Google Scholar] [CrossRef]
  28. Khavar, A.H.C.; Moussavi, G.; Mahjoub, A.R.; Satari, M.; Abdolmaleki, P. Synthesis and visible-light photocatalytic activity of In, S-TiO2@rGO nanocomposite for degradation and detoxification of pesticide atrazine in water. Chem. Eng. J. 2018, 345, 300–311. [Google Scholar] [CrossRef]
  29. Zhang, Z.; Li, A.; Cao, S.W.; Bosman, M.; Li, S.; Xue, C. Direct evidence of plasmon enhancement on photocatalytic hydrogen generation over Au/Pt-decorated TiO2 nanofibers. Nanoscale 2014, 6, 5217–5222. [Google Scholar] [CrossRef]
  30. Zhang, N.; Shi, J.; Mao, S.S.; Guo, L. Co3O4 quantum dots: Reverse micelle synthesis and visible-light-driven photocatalytic overall water splitting. Chem. Commun. 2014, 50, 2002–2004. [Google Scholar] [CrossRef]
  31. Kohantorabi, M.; Moussavi, G.; Oulego, P.; Giannakis, S. Radical-based degradation of sulfamethoxazole via UVA/PMS-assisted photocatalysis, driven by magnetically separable Fe3O4@CeO2@BiOI nanospheres. Sep. Purif. Technol. 2021, 267, 118665. [Google Scholar] [CrossRef]
  32. Huang, J.G.; Zhao, X.G.; Zheng, M.Y.; Li, S.; Wang, Y.; Liu, X.J. Preparation of N-doped TiO2 by oxidizing TiN and its application on phenol degradation. Water Sci. Technol. 2013, 68, 934–939. [Google Scholar] [CrossRef]
  33. Mohammadi, K.; Sadeghi, M.; Azimirad, R. Facile synthesis of SrFe12O19 nanoparticles and its photocatalyst application. J. Mater. Sci. Mater. Electron. 2017, 28, 10042–10047. [Google Scholar] [CrossRef]
  34. Mohammadi, K.; Sadeghi, M.; Azimirad, R.; Ebrahimi, M. Barium hexaferrite nanoparticles: Morphology-controlled preparation, characterization and investigation of magnetic and photocatalytic properties. J. Mater. Sci. Mater. Electron. 2017, 28, 9983–9988. [Google Scholar] [CrossRef]
  35. Shieh, D.L.; Lin, Y.S.; Yeh, J.H.; Chen, S.C.; Lin, B.C.; Lin, J.L. N-doped, porous TiO2 with rutile phase and visible light sensitive photocatalytic activity. Chem. Commun. 2012, 48, 2528–2530. [Google Scholar] [CrossRef]
  36. Zhou, X.; Lu, J.; Jiang, J.; Li, X.; Lu, M.; Yuan, G.; Wang, Z.; Zheng, M.; Seo, H.J. Simple fabrication of N-doped mesoporous TiO2 nanorods with the enhanced visible light photocatalytic activity. Nanoscale Res. Lett. 2014, 9, 34. [Google Scholar] [CrossRef]
  37. Chen, X.; Burda, C. The electronic origin of the visible-light absorption properties of C-, N- and S-doped TiO2 nanomaterials. J. Am. Chem. Soc. 2008, 130, 5018–5019. [Google Scholar] [CrossRef]
  38. Li, H.; Zhang, X.; Huo, Y.; Zhu, J. Supercritical preparation of a highly active S-doped TiO2 photocatalyst for methylene blue mineralization. Environ. Sci. Technol. 2007, 41, 4410–4414. [Google Scholar] [CrossRef]
  39. Li, Z.; Ding, D.; Ning, C. P-Type hydrogen sensing with Al-and V-doped TiO2 nanostructures. Nanoscale Res. Lett. 2013, 8, 25. [Google Scholar] [CrossRef]
  40. Pan, L.; Wang, S.; Zou, J.J.; Huang, Z.F.; Wang, L.; Zhang, X. Ti3+-defected and V-doped TiO2 quantum dots loaded on MCM-41. Chem. Commun. 2014, 50, 988–990. [Google Scholar] [CrossRef]
  41. Zhang, Z.; Shao, C.; Zhang, L.; Li, X.; Liu, Y. Electrospun nanofibers of V-doped TiO2 with high photocatalytic activity. J. Colloid. Interface Sci. 2010, 351, 57–62. [Google Scholar] [CrossRef]
  42. Liu, G.; Han, C.; Pelaez, M.; Zhu, D.; Liao, S.; Likodimos, V.; Ioannidis, N.; Kontos, A.G.; Falaras, P.; Dunlop, P.S.; et al. Synthesis, characterization and photocatalytic evaluation of visible light activated C-doped TiO2 nanoparticles. Nanotechnology 2012, 23, 294003. [Google Scholar] [CrossRef]
  43. Yu, J.; Zhou, P.; Li, Q. New insight into the enhanced visible-light photocatalytic activities of B-, C- and B/C-doped anatase TiO2 by first-principles. Phys. Chem. Chem. Phys. 2013, 15, 12040–12047. [Google Scholar] [CrossRef]
  44. Zhang, W.J.; Ma, Z.; Du, L.; Li, H. Role of PEG4000 in sol-gel synthesis of Sm2Ti2O7 photocatalyst for enhanced activity. J. Alloy. Compd. 2017, 704, 26–31. [Google Scholar] [CrossRef]
  45. Wang, S.X.; Li, W.; Wang, S.; Jiang, J.M.; Chen, Z.H. Synthesis of well-defined hierarchical porous La2Zr2O7 monoliths via non-alkoxidesolegel process accompanied by phase separation. Micropor. Mesopor. Mat. 2016, 221, 32–39. [Google Scholar] [CrossRef]
  46. Zou, Z.; Ye, J.; Arakawa, H. Preparation, structural and photophysical properties of Bi2InNbO7 compound. J. Mater. Sci. 2000, 19, 1909–1911. [Google Scholar]
  47. Shin, H.; Byun, T.H. Effect of Pt loading onto ball-milled TiO2 on the visible-light photocatalytic activity to decompose rhodamine B. Res. J. Chem. Environ. 2013, 17, 41–46. [Google Scholar]
  48. Kanhere, P.; Shenai, P.; Chakraborty, S.; Ahuja, R.; Zheng, J.; Chen, Z. Mono- and co-doped NaTaO3 for visible light photocatalysis. Phys. Chem. Chem. Phys. 2014, 16, 16085–16094. [Google Scholar] [CrossRef]
  49. Daskalaki, V.M.; Antoniadou, M.; Li Puma, G.; Kondarides, D.I.; Lianos, P. Solar light-responsive Pt/CdS/TiO2 photocatalysts for hydrogen production and simultaneous degradation of inorganic or organic sacrificial agents in wastewater. Environ. Sci. Technol. 2010, 44, 7200–7205. [Google Scholar] [CrossRef]
  50. Xing, C.C.; Zhang, Y.; Liu, Y.P.; Wang, X.; Li, J.S.; Martinez-Alanis, P.R.; Spadaro, M.C.; Guardia, P.; Arbiol, J.; Llorca, J.; et al. Photodehydrogenation of ethanol over Cu2O/TiO2 heterostructures. Nanomaterials 2021, 11, 1399. [Google Scholar] [CrossRef]
  51. Stathi, P.; Solakidou, M.; Deligiannakis, Y. Lattice defects engineering in W-, Zr-doped BiVO4 by flame spray pyrolysis: Enhancing photocatalytic O2 evolution. Nanomaterials 2021, 11, 501. [Google Scholar] [CrossRef]
  52. Li, X.; Yan, X.; Lu, X.; Zuo, S.; Li, Z.; Yao, C.; Ni, C. Photo-assisted selective catalytic reduction of NO by Z-scheme natural clay based photocatalyst: Insight into the effect of graphene coupling. J. Catal. 2018, 357, 59–68. [Google Scholar] [CrossRef]
  53. Oussama, B.; Nhu-Nang, V.; Aymen, A.A.; Van, Q.L.; Phuong, N.T. Recent advances in designing and developing efficient sillenite-based materials for photocatalytic applications. Adv. Colloid Interface Sci. 2024, 327, 103136. [Google Scholar]
  54. Nasiripur, P.; Zangiabadi, M.; Baghersad, M.H. Visible light photocatalytic degradation of methyl parathion as chemical warfare agents simulant via GO-Fe3O4/Bi2MoO6 nanocomposite. J. Mol. Struct. 2021, 1243, 130875. [Google Scholar] [CrossRef]
  55. Zangiabadi, M.; Mehrabi, F.; Nasiripur, P.; Baghersad, M.H. Visible-light-driven photocatalytic degradation of methyl parathion as chemical warfare agent simulant by NiO/Bi2MoO6 heterojunction photocatalyst. J. Mol. Struct. 2022, 1256, 132472. [Google Scholar] [CrossRef]
  56. Nowak, M.; Kauch, B.; Szperlich, P. Determination of energy band gap of nanocrystalline SbSI. Rev. Sci. Instrum. 2009, 80, 046107. [Google Scholar] [CrossRef]
  57. Mohan, S.; Kaur, S.; Singh, D.P.; Kaur, P. Structural and luminescence properties of samarium doped lead alumino borate glasses. Opt. Mater. 2017, 73, 223–233. [Google Scholar] [CrossRef]
  58. Goh, K.H.; Haseeb, A.S.M.A.; Wong, Y.H. Effect of Oxidation Temperature on Physical and Electrical Properties of Sm2O3 Thin-Film Gate Oxide on Si Substrate. J. Electron. Mater. 2016, 45, 5302–5312. [Google Scholar] [CrossRef]
  59. Yuvakkumar, R.; Hong, S.I. Nd2O3: Novel synthesis and characterization. J. Sol-Gel Sci. Technol. 2015, 73, 511–517. [Google Scholar] [CrossRef]
  60. Bosca, M.; Pop, L.; Borodi, G.; Pascuta, P.; Culea, E. XRD and FTIR structural investigations of erbium-doped bismuth–lead–silver glasses and glass ceramics. J. Alloys Compd. 2009, 479, 579–582. [Google Scholar] [CrossRef]
  61. Pascuta, P.; Culea, E. FTIR spectroscopic study of some bismuth germanate glasses containing gadolinium ions. Mater. Lett. 2008, 62, 4127–4129. [Google Scholar] [CrossRef]
  62. Kaviyarasu, K.; Sajan, D.; Devarajan, P.A. A rapid and versatile method for solvothermal synthesis of Sb2O3 nanocrystals under mild conditions. Appl. Nanosci. 2012, 3, 529–533. [Google Scholar] [CrossRef]
  63. Rada, S.; Rus, L.; Rada, M.; Zagrai, M.; Culea, E.; Rusu, T. Compositional dependence of structure, optical and electrochemical properties of antimony(III) oxide doped lead glasses and vitroceramics. Ceram. Int. 2014, 40, 15711–15716. [Google Scholar] [CrossRef]
  64. Janani, B.; Okla, M.K.; Abdel-Maksoud, M.A.; AbdElgawad, H.; Thomas, A.M.; Raju, L.L.; Al-Qahtani, W.H.; Khan, S.S. CuO loaded ZnS nanoflower entrapped on PVA-chitosan matrix for boosted visible light photocatalysis for tetracycline degradation and anti-bacterial application. J. Environ. Manag. 2022, 306, 114396. [Google Scholar] [CrossRef]
  65. Cheng, T.; Gao, H.; Liu, G.; Pu, Z.; Wang, S.; Yi, Z.; Wu, X.; Yang, H. Preparation of core-shell heterojunction photocatalysts by coating CdS nanoparticles onto Bi4Ti3O12 hierarchical microspheres and their photocatalytic removal of organic pollutants and Cr(VI) ions. Colloids Surf. A Physicochem. Eng. Asp. 2022, 633, 127918. [Google Scholar] [CrossRef]
  66. Isari, A.A.; Hayati, F.; Kakavandi, B.; Rostami, M.; Motevassel, M.; Dehghanifard, E. N, Cu co-doped TiO2@functionalized SWCNT photocatalyst coupled with ultrasound and visible-light: An effective sono-photocatalysis process for pharmaceutical wastewaters treatment. Chem. Eng. J. 2020, 392, 123685. [Google Scholar] [CrossRef]
  67. Li, R.; Cai, M.; Xie, Z.; Zhang, Q.; Zeng, Y.; Liu, H.; Liu, G.; Lv, W. Construction of heterostructured CuFe2O4/g-C3N4 nanocomposite as an efficient visible light photocatalyst with peroxydisulfate for the organic oxidation. Appl. Catal. B-Environ. 2019, 244, 974–982. [Google Scholar] [CrossRef]
  68. Shao, B.; Liu, X.; Liu, Z.; Zeng, G.; Liang, Q.; Liang, C.; Cheng, Y.; Zhang, W.; Liu, Y.; Gong, S. A novel double Z-scheme photocatalyst Ag3PO4/Bi2S3/Bi2O3 with enhanced visible-light photocatalytic performance for antibiotic degradation. Chem. Eng. J. 2019, 368, 730–745. [Google Scholar] [CrossRef]
  69. Hadjiev, V.G.; Iliev, M.N.; Sasmal, K.; Sun, Y.Y.; Chu, C.W. Raman spectroscopy of RFeAsO (R=Sm, La). Phys. Rev. B 2008, 77, 220505. [Google Scholar]
  70. Bahareh, K.; Habibi, M.H. High photocatalytic activity of light-driven Fe2TiO5 nanoheterostructure toward degradation of antibiotic metronidazole. J. Ind. Eng. Chem. 2019, 80, 292–300. [Google Scholar] [CrossRef]
  71. Refat, M.S.; Elsabawy, K.M. Infrared spectra, Raman laser, XRD, DSC/TGA and SEM investigations on the preparations of selenium metal, (Sb2O3, Ga2O3, SnO and HgO) oxides and lead carbonate with pure grade using acetamide precursors. Bull. Mater. Sci. 2011, 34, 873–881. [Google Scholar] [CrossRef]
  72. Gilliam, S.J.; Jensen, J.O.; Banerjee, A.; Zeroka, D.; Kirkby, S.J.; Merrow, C.N. A theoretical and experimental study of Sb4O6: Vibrational analysis, infrared, and Raman spectra. Spectrochim. Acta. A Mol. Biomol. Spectrosc. 2004, 60, 425–434. [Google Scholar] [CrossRef]
  73. Chahine, A.; Et-tabirou, M.; Pascal, J.L. FTIR and Raman spectra of the Na2O–CuO–Bi2O3–P2O5 glasses. Mater. Lett. 2004, 58, 2776–2780. [Google Scholar] [CrossRef]
  74. Sharma, N.D.; Singh, J.; Vijay, A.; Samanta, K.; Pandey, S.D. Investigations of anharmonic effects via phonon mode variations in nanocrystalline Dy2O3, Gd2O3 and Y2O3. J. Raman Spectrosc. 2017, 48, 822–828. [Google Scholar] [CrossRef]
  75. Zhou, F.; Kang, K.; Maxisch, T.; Ceder, G.; Morgan, D. The electronic structure and band gap of LiFePO4 and LiMnPO4. Solid State Commun. 2004, 132, 181–186. [Google Scholar] [CrossRef]
  76. Tauc, J.; Grigorov, R.; Vancu, A. Optical properties and electronic structure of amorphous germanium. Phys. Status Solidi 1966, 15, 627–637. [Google Scholar] [CrossRef]
  77. Butler, M.A. Photoelectrolysis with YFeO3 electrodes. J. Appl. Phys. 1977, 48, 1914–1920. [Google Scholar] [CrossRef]
  78. Cui, B.Y.; Cui, H.T.; Li, Z.R.; Dong, H.Y.; Li, X.; Zhao, L.F.; Wang, J.W. Novel Bi3O5I2 hollow microsphere and its enhanced photocatalytic activity. Catalysts 2019, 9, 709. [Google Scholar] [CrossRef]
  79. Vallejo, W.; Cantillo, A.; Salazar, B.; Diaz-Uribe, C.; Ramos, W.; Romero, E.; Hurtado, M. Comparative study of ZnO thin films doped with transition metals (Cu and Co) for methylene blue photodegradation under visible irradiation. Catalysts 2020, 10, 528. [Google Scholar] [CrossRef]
  80. Liu, B.Y.; Du, J.Y.; Ke, G.L.; Jia, B.; Huang, Y.J.; He, H.C.; Zhou, Y.; Zou, Z.G. Boosting O2 Reduction and H2O Dehydrogenation Kinetics: Surface N-Hydroxymethylation of g-C3N4 Photocatalysts for the Efficient Production of H2O2. Adv. Funct. Mater. 2022, 32, 2111125. [Google Scholar] [CrossRef]
  81. Liu, C.; Feng, Y.; Han, Z.T.; Sun, Y.; Wang, X.Q.; Zhang, Q.F.; Zou, Z.G. Z-scheme N-doped K4Nb6O17/g-C3N4 heterojunction with superior visible-light-driven photocatalytic activity for organic pollutant removal and hydrogen production. Chin. J. Catal. 2021, 42, 164–174. [Google Scholar] [CrossRef]
  82. Ma, Q.; Kumar, R.K.; Xu, S.Y.; Koppens, F.H.L.; Song, J.C.W. Photocurrent as a multiphysics diagnostic of quantum materials. Nat. Rev. Phys. 2023, 5, 170–184. [Google Scholar] [CrossRef]
  83. Cheng, Y.X.; Ye, J.H.; Lai, L.; Fang, S.; Guo, D.Y. Ambipolarity Regulation of Deep-UV Photocurrent by Controlling Crystalline Phases in Ga2O3 Nanostructure for Switchable Logic Applications. Adv. Electron. Mater. 2023, 9, 2201216. [Google Scholar] [CrossRef]
  84. Annadi, A.; Gong, H. Success in both p-type and n-type of a novel transparent AgCuI alloy semiconductor system for homojunction devices. Appl. Mater. Today 2020, 20, 100703. [Google Scholar] [CrossRef]
  85. Cheng, T.; Gao, H.; Sun, X.; Xian, T.; Wang, S.; Yi, Z.; Liu, G.; Wang, X.; Yang, H. An excellent Z-scheme Ag2MoO4/Bi4Ti3O12 heterojunction photocatalyst: Construction strategy and application in environmental purification. Adv. Powder Technol. 2021, 32, 951–962. [Google Scholar] [CrossRef]
  86. Deonikar, V.G.; Patil, S.S.; Tamboli, M.S.; Ambekar, J.D.; Kulkarni, M.V.; Panmand, R.P.; Umarji, G.G.; Shinde, M.D.; Rane, S.B.; Munirathnam, N.R.; et al. Growth study of hierarchical Ag3PO4/LaCO3OH heterostructures. Phys. Chem. Chem. Phys. 2017, 19, 20541–20550. [Google Scholar] [CrossRef] [PubMed]
  87. Patil, S.S.; Tamboli, M.S.; Deonikar, V.G.; Umarji, G.G.; Ambekar, J.D.; Kulkarni, M.V.; Kolekar, S.S.; Kale, B.B.; Patil, D.R. Magnetically separable Ag3PO4/NiFe2O4 composites with enhanced photocatalytic activity. Dalton Trans. 2015, 44, 20426–20434. [Google Scholar] [CrossRef] [PubMed]
  88. Balakrishnan, G.; Velavan, R.; Mujasam Batoo, K.; Raslan, E.H. Microstructure, optical and photocatalytic properties of MgO nanoparticles. Results Phys. 2020, 16, 103013. [Google Scholar] [CrossRef]
  89. Ali, T.; Tripathi, P.; Azam, A.; Raza, W.; Ahmed, A.S.; Ahmed, A.; Muneer, M. Photocatalytic performance of Fe-doped TiO2nanoparticles under visible-light irradiation. Mater. Res. Express 2017, 4, 015022. [Google Scholar] [CrossRef]
  90. Zhuang, Y.; Luan, J. Improved photocatalytic property of peony-like InOOH for degrading norfloxacin. Chem. Eng. J. 2020, 382, 122770. [Google Scholar] [CrossRef]
  91. Chen, J.; Zhao, X.; Kim, S.G.; Park, N.G. Multifunctional Chemical Linker Imidazoleacetic Acid Hydrochloride for 21% Efficient and Stable Planar Perovskite Solar Cells. Adv. Mater. 2019, 31, 1902902. [Google Scholar] [CrossRef] [PubMed]
  92. Gao, Z.W.; Wang, Y.; Ouyang, D.; Liu, H.; Huang, Z.F.; Kim, J.; Choy, W.C.H. Triple Interface Passivation Strategy-Enabled Efficient and Stable Inverted Perovskite Solar Cells. Small Methods 2020, 4, 2000478. [Google Scholar] [CrossRef]
  93. Chen, J.; Kim, S.-G.; Ren, X.; Jung, H.S.; Park, N.-G. Effect of bidentate and tridentate additives on the photovoltaic performance and stability of perovskite solar cells. J. Mater. Chem. A 2019, 7, 4977–4987. [Google Scholar] [CrossRef]
  94. Jiang, L.B.; Yuan, X.Z.; Zeng, G.M.; Liang, J.; Chen, X.H.; Yu, H.B.; Wang, H.; Wu, Z.B.; Zhang, J.; Xiong, T. In-situ synthesis of direct solid-state dual Z-scheme WO3/g-C3N4/Bi2O3 photocatalyst for the degradation of refractory pollutant. Appl. Catal. B 2018, 227, 376–385. [Google Scholar] [CrossRef]
  95. Wang, J.H.; Zou, Z.G.; Ye, J.H. Synthesis, structure and photocatalytic property of a new hydrogen evolving photocatalyst Bi2InTaO7. Mater. Sci. Forum 2003, 423–425, 485–490. [Google Scholar] [CrossRef]
  96. Kohno, M.; Ogura, S.; Sato, K.; Inoue, Y. Properties of photocatalysts with tunnel structures: Formation of a surface lattice O− radical by the UV irradiation of BaTi4O9 with a pentagonal-prism tunnel structure. Chem. Phys. Lett. 1997, 267, 72–76. [Google Scholar] [CrossRef]
  97. Kudo, A.; Kato, H.; Nakagawa, S. Water Splitting into H2 and O2 on New Sr2M2O7 (M = Nb and Ta) Photocatalysts with layered perovskite structures: Factors affecting the photocatalytic activity. J. Phys. Chem. B 2000, 104, 571–575. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of as-prepared samples: (a) SBHP, (b) Sm2NdSbO7, and (c) BiDyO3.
Figure 1. XRD patterns of as-prepared samples: (a) SBHP, (b) Sm2NdSbO7, and (c) BiDyO3.
Catalysts 14 00346 g001
Figure 2. (a) XRD pattern and (b) the atomic structure (red atom: O, purple atom: Nd or Sb, green atom: Sm) of Sm2NdSbO7.
Figure 2. (a) XRD pattern and (b) the atomic structure (red atom: O, purple atom: Nd or Sb, green atom: Sm) of Sm2NdSbO7.
Catalysts 14 00346 g002
Figure 3. (a) XRD pattern and (b) the atomic structure (red atom: O, green atom: Bi, purple atom: Dy) of BiDyO3.
Figure 3. (a) XRD pattern and (b) the atomic structure (red atom: O, green atom: Bi, purple atom: Dy) of BiDyO3.
Catalysts 14 00346 g003
Figure 4. FTIR spectra of SBHP, Sm2NdSbO7, and BiDyO3.
Figure 4. FTIR spectra of SBHP, Sm2NdSbO7, and BiDyO3.
Catalysts 14 00346 g004
Figure 5. Raman spectra of (a) SBHP, (b) Sm2NdSbO7, and (c) BiDyO3.
Figure 5. Raman spectra of (a) SBHP, (b) Sm2NdSbO7, and (c) BiDyO3.
Catalysts 14 00346 g005
Figure 6. (a) UV–Vis diffuse reflectance spectra and (b) correlative diagram of (αhν) 1/2 and of SBHP, Sm2NdSbO7, and BiDyO3.
Figure 6. (a) UV–Vis diffuse reflectance spectra and (b) correlative diagram of (αhν) 1/2 and of SBHP, Sm2NdSbO7, and BiDyO3.
Catalysts 14 00346 g006
Figure 7. XPS spectrum of SBHP, Sm2NdSbO7, and BiDyO3.
Figure 7. XPS spectrum of SBHP, Sm2NdSbO7, and BiDyO3.
Catalysts 14 00346 g007
Figure 8. The corresponding high-resolution XPS spectra of (a) Sm 3d, (b) Nd 3d, (c) Sb 4d, (d) Bi 4d, (e) Dy 4d, and (f) O 1s of SBHP and Sm2NdSbO7.
Figure 8. The corresponding high-resolution XPS spectra of (a) Sm 3d, (b) Nd 3d, (c) Sb 4d, (d) Bi 4d, (e) Dy 4d, and (f) O 1s of SBHP and Sm2NdSbO7.
Catalysts 14 00346 g008aCatalysts 14 00346 g008b
Figure 9. TEM photograph of SBHP.
Figure 9. TEM photograph of SBHP.
Catalysts 14 00346 g009
Figure 10. (a) EDS elemental mapping of SBHP; (b) Sm, (c) Nd, (d) Sb, (e) Bi, (f) Dy, and (g) O.
Figure 10. (a) EDS elemental mapping of SBHP; (b) Sm, (c) Nd, (d) Sb, (e) Bi, (f) Dy, and (g) O.
Catalysts 14 00346 g010aCatalysts 14 00346 g010b
Figure 11. EDS spectrum of SBHP.
Figure 11. EDS spectrum of SBHP.
Catalysts 14 00346 g011
Figure 12. Nyquist impedance plots of SBHP, Sm2NdSbO7, and BiDyO3.
Figure 12. Nyquist impedance plots of SBHP, Sm2NdSbO7, and BiDyO3.
Catalysts 14 00346 g012
Figure 13. Transient photocurrent of SBHP, Sm2NdSbO7, and BiDyO3.
Figure 13. Transient photocurrent of SBHP, Sm2NdSbO7, and BiDyO3.
Catalysts 14 00346 g013
Figure 14. UPS spectra of (a) Sm2NdSbO7 and (b) BiDyO3 (the intersections of the black dash lines indicated by the black arrows indicated the onset (Ei) and cutoff (Ecutoff) binding energy).
Figure 14. UPS spectra of (a) Sm2NdSbO7 and (b) BiDyO3 (the intersections of the black dash lines indicated by the black arrows indicated the onset (Ei) and cutoff (Ecutoff) binding energy).
Catalysts 14 00346 g014
Figure 15. CTC of MP during PHDE with SBHP, Sm2NdSbO7, BiDyO3, N-T under VLIRD.
Figure 15. CTC of MP during PHDE with SBHP, Sm2NdSbO7, BiDyO3, N-T under VLIRD.
Catalysts 14 00346 g015
Figure 16. CTC of TOC during PHDE of MP, which is derived from PW with SBHP, Sm2NdSbO7, BiDyO3, or N-T under VLIRD.
Figure 16. CTC of TOC during PHDE of MP, which is derived from PW with SBHP, Sm2NdSbO7, BiDyO3, or N-T under VLIRD.
Catalysts 14 00346 g016
Figure 17. (a) CTC of MP during PHDE of MP, which is derived from PW with SBHP under VLIRD for four cycle degradation tests. (b) CTC of TOC during PHDE of MP, which is derived from PW with SBHP under VLIRD for four cycle degradation tests. (c) The effect of different pH values on the removal rate of MP.
Figure 17. (a) CTC of MP during PHDE of MP, which is derived from PW with SBHP under VLIRD for four cycle degradation tests. (b) CTC of TOC during PHDE of MP, which is derived from PW with SBHP under VLIRD for four cycle degradation tests. (c) The effect of different pH values on the removal rate of MP.
Catalysts 14 00346 g017
Figure 18. (a) Observed first-order kinetic plots for the PHDE of MP with SBHP or with Sm2NdSbO7 or with BiDyO3 or with N-T under VLIRD. (b) Observed first-order kinetic plots for TOC during PHDE of MP with SBHP or with Sm2NdSbO7 or with BiDyO3 or with N-T under VLIRD.
Figure 18. (a) Observed first-order kinetic plots for the PHDE of MP with SBHP or with Sm2NdSbO7 or with BiDyO3 or with N-T under VLIRD. (b) Observed first-order kinetic plots for TOC during PHDE of MP with SBHP or with Sm2NdSbO7 or with BiDyO3 or with N-T under VLIRD.
Catalysts 14 00346 g018
Figure 19. (a) Observed first-order kinetic plots for the PHDE of MP with SBHP under VLIRD for four cycle degradation tests. (b) The four-cycle degradation curve of SBHP for PHDE of TOC in MP under VLIRD was observed.
Figure 19. (a) Observed first-order kinetic plots for the PHDE of MP with SBHP under VLIRD for four cycle degradation tests. (b) The four-cycle degradation curve of SBHP for PHDE of TOC in MP under VLIRD was observed.
Catalysts 14 00346 g019
Figure 20. (a) RMR of MP with SBHP under VLIRD; (b) BQ, IPA, or EDTA on the removal efficiency of MP with SBHP under VLIRD.
Figure 20. (a) RMR of MP with SBHP under VLIRD; (b) BQ, IPA, or EDTA on the removal efficiency of MP with SBHP under VLIRD.
Catalysts 14 00346 g020
Figure 21. EPR spectrum for DMPO•O2 and DMPO•OH over SBHP.
Figure 21. EPR spectrum for DMPO•O2 and DMPO•OH over SBHP.
Catalysts 14 00346 g021
Figure 22. PL spectra of Sm2NdSbO7, BiDyO3, and SBHP.
Figure 22. PL spectra of Sm2NdSbO7, BiDyO3, and SBHP.
Catalysts 14 00346 g022
Figure 23. TRPL spectra of (a) Sm2NdSbO7, (b) BiDyO3, and (c) SBHP.
Figure 23. TRPL spectra of (a) Sm2NdSbO7, (b) BiDyO3, and (c) SBHP.
Catalysts 14 00346 g023
Figure 24. Possible PHDE mechanism of MP with SBHP under VLIRD.
Figure 24. Possible PHDE mechanism of MP with SBHP under VLIRD.
Catalysts 14 00346 g024
Figure 25. Suggested PHDE pathway scheme for MP under VLIRD with SBHP.
Figure 25. Suggested PHDE pathway scheme for MP under VLIRD with SBHP.
Catalysts 14 00346 g025
Table 1. Structural parameters of Sm2NdSbO7.
Table 1. Structural parameters of Sm2NdSbO7.
AtomxyzOccupation
Factor
Sm0001
Nd0.50.50.50.5
Sb0.50.50.50.5
O(1)−0.1850.1250.1251
O(2)0.1250.1250.1251
Table 2. Structural parameters of BiDyO3.
Table 2. Structural parameters of BiDyO3.
AtomxyzOccupation
Factor
Bi0.50.50.51
Dy0001
O0.25000.5
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

Luan, J.; Li, J.; Yao, Y.; Niu, B.; Hao, L.; Wang, Y.; Li, Z. Synthesis, Characterization of the Novel Heterojunction Photocatalyst Sm2NdSbO7/BiDyO3 for Efficient Photodegradation of Methyl Parathion. Catalysts 2024, 14, 346. https://doi.org/10.3390/catal14060346

AMA Style

Luan J, Li J, Yao Y, Niu B, Hao L, Wang Y, Li Z. Synthesis, Characterization of the Novel Heterojunction Photocatalyst Sm2NdSbO7/BiDyO3 for Efficient Photodegradation of Methyl Parathion. Catalysts. 2024; 14(6):346. https://doi.org/10.3390/catal14060346

Chicago/Turabian Style

Luan, Jingfei, Jun Li, Ye Yao, Bowen Niu, Liang Hao, Yichun Wang, and Zhe Li. 2024. "Synthesis, Characterization of the Novel Heterojunction Photocatalyst Sm2NdSbO7/BiDyO3 for Efficient Photodegradation of Methyl Parathion" Catalysts 14, no. 6: 346. https://doi.org/10.3390/catal14060346

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