Next Article in Journal
Nickel-Based Single-Atom Alloys for Methane Dehydrogenation and the Effect of Subsurface Carbon: First-Principles Investigations
Previous Article in Journal
Hydrogenation of Styrene-Butadiene Rubber Catalyzed by Tris(triisopropylphosphine)hydridorhodium(I)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Property Characterization of Eu2SmSbO7/ZnBiEuO4 Heterojunction Photocatalysts and Photocatalytic Degradation of Chlorpyrifos under Visible Light Irradiation

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(2), 144; https://doi.org/10.3390/catal14020144
Submission received: 26 November 2023 / Revised: 6 February 2024 / Accepted: 7 February 2024 / Published: 15 February 2024
(This article belongs to the Topic New Materials and Advanced Applications in Photocatalysis)

Abstract

:
Eu2SmSbO7 and ZnBiEuO4 were synthesized for the first time using the hydrothermal method. Eu2SmSbO7/ZnBiEuO4 heterojunction photocatalyst (EZHP) was synthesized for the first time using the solvothermal method. The crystal cell parameter of Eu2SmSbO7 was 10.5547 Å. The band gap width of Eu2SmSbO7 was measured and found to be 2.881 eV. The band gap width of ZnBiEuO4 was measured and found to be 2.571 eV. EZHP efficiently degraded the pesticide chlorpyrifos under visible light irradiation (VLID). After VLID of 160 min, the conversion rate of the chlorpyrifos concentration reached 100%, while the conversion rate of the total organic carbon (TOC) concentration was 98.02% using EZHP. After VLID of 160 min, the photocatalytic degradation conversion rates of chlorpyrifos using EZHP were 1.13 times, 1.19 times, and 2.84 times those using Eu2SmSbO7, ZnBiEuO4, and nitrogen-doped titanium dioxide (N-doped TiO2), respectively. The photocatalytic activity could be ranked as follows: EZHP > Eu2SmSbO7 > ZnBiEuO4 > N-doped TiO2. The conversion rates of chlorpyrifos were 98.16%, 97.03%, 96.03%, and 95.06% for four cycles of experiments after VLID of 160 min using EZHP. This indicated that EZHP was stable and could be reused. In addition, the experiments with the addition of capture agents demonstrated that the oxidation removal ability of three oxidation free radicals for degrading chlorpyrifos obeyed the following order: hydroxyl radical > superoxide anion > holes. This study examined the intermediates of chlorpyrifos during the photocatalytic degradation of chlorpyrifos, and a degradation path was proposed, at the same time, the degradation mechanism of chlorpyrifos was revealed. This study provides a scientific basis for the development of efficient heterojunction photocatalysts.

Graphical Abstract

1. Introduction

In recent years, there has been an increasing demand for food supplies due to the growing global population [1,2,3]. The use of pesticides is a valid method to increase crop yields [4,5,6,7]. Therefore, pesticides are widely used in agricultural production. However, usually only a small amount of the pesticide works, and most of the pesticide remains on the crop or in the soil. Pesticide residues on crops can be harmful to creatures, and other residues can cause serious pollution of the soil and even the underground water [8,9,10,11]. Chlorpyrifos, as one of the classical organophosphorus pesticides, has been widely used due to its low cost and effective ability to control pests, weeds, and diseases [12,13,14,15]. However, the disadvantages of chlorpyrifos are also obvious; for example, chlorpyrifos is an organic pollutant that is difficult to degrade; at the same time, chlorpyrifos also affects the neurological system, leading to various diseases [16,17,18,19,20]. Moreover, chlorpyrifos has adverse effects on the development of the brain and body, especially for children [21,22]. Therefore, elimination of chlorpyrifos derived from wastewater is a meaningful goal, considering the healthy living environment for humans.
After preliminary physical filtration, some methods have been invented for the degradation of small particle organic pollutants that are easily dissolved in water, such as chlorination, electrocatalysis, and anaerobic–aerobic biological treatment [23,24,25,26,27,28,29,30,31]. All the above methods have some drawbacks. For example, chlorination is hard to degrade completely. Moreover, electrocatalytic methods consume more energy, and, ultimately, the anaerobic–aerobic biological treatment requires complicated reaction conditions. Therefore, finding a method with higher degradation efficiency to solve the problem of organic pollutants is needed.
A promising method, photocatalytic technology, was reported to be an effective way of degrading organic pollutants [32,33,34,35,36,37,38,39]. Photocatalytic technology absorbs solar energy to degrade water, generating free radicals with significant oxidizing and reducing properties. The free radicals can degrade organic pollutants without producing secondary pollution. Photocatalytic technology has become a favored choice due to its advantages of high efficiency, non-toxicity, low cost, and environmental friendliness [40,41,42].
At the beginning, TiO2 became a popular research area in the field of photocatalysis due to its simple fabrication process, high photocatalytic activity, and excellent stability [43,44,45]. With TiO2 as a catalytic material, it was required that the incident photon energy must be higher than 3.2 eV, and only ultraviolet light would be able to satisfy this condition. At the same time, ZnO was also reported as the photocatalyst using ultraviolet light [46,47]. The energy of ultraviolet light in sunlight was too small compared with visible light (which occupies 43% of the sunlight spectrum). Based on the reports of the researchers, the construction of the composite would have a wider light absorption range and higher photocatalytic efficiency than that of single metal oxides and composite catalysts [48,49,50,51]. A2B2O7-type compounds and AB2O4-type compounds had been reported to have excellent catalytic performance under visible light irradiation (VLID). For example, Zou et al. found that Bi2InNbO7 had a good effect in decomposing water to produce hydrogen [52]. Based on our previous report, ZnBiYO4 also exhibited excellent photocatalytic properties under VLID [53]. It was speculated that Eu2SmSbO7 and ZnBiEuO4 could effectively degrade organic pollutants under VLID.
In the process of photochemical reactions, the photo-induced electrons and the photo-induced holes are generated after light irradiation on the surface of the photocatalytic material. The photo-induced electrons and the photo-induced holes participated in the subsequent reaction to generate superoxide anions and hydroxyl radicals. Superoxide radicals and hydroxyl radicals had strong redox capacity and therefore can degrade organic pollutants effectively. However, the photo-induced electrons and the photo-induced holes that do not participate in the reaction are recombined, affecting the generation of superoxide radicals and hydroxyl radicals, and thus affecting the photocatalytic efficiency. Therefore, the photocatalytic efficiency can be improved by inhibiting the recombination of the photo-induced electrons and the photo-induced holes [54,55]. The catalytic efficiency can be improved by constructing heterojunction structures to inhibit the complexation of photo-induced electrons and photo-induced holes [56,57]. For example, the heterojunction AgBr/BiPO4 that was synthesized by Xu H. et al. possessed higher catalytic activity than pure BiPO4, and the heterojunction Ag2MoO4/Bi4Ti3O12 by Cheng T. T. et al. had higher catalytic activity than pure Ag2MoO4 or pure Bi4Ti3O12 [58,59]. Thus, we synthesized the Eu2SmSbO7/ZnBiEuO4 heterojunction photocatalyst (EZHP) and found that EZHP had excellent catalytic activity under VLID. The catalyst Ag/TiO2 that was prepared by Fattah W. I. A. et al. achieved a conversion rate of chlorpyrifos of nearly 75% after light irradiation for 120 min [60]. The conversion rate of chlorpyrifos reached 84.5% using EZHP under VLID of 120 min in our manuscript. In conclusion, the EZHP that was prepared by Jingfei Luan showed higher photocatalytic activity compared with the Ag/TiO2 that was prepared by Fattah W. I. A. et al. The catalyst Cu/ZnO that was prepared by Pathania, D. et al. achieved a conversion rate of chlorpyrifos of 81% after solar light irradiation of 240 min [61]. The conversion rate of chlorpyrifos reached 100% using EZHP under VLID of 160 min in our manuscript. In summary, the EZHP that was prepared by Jingfei Luan showed higher photocatalytic activity compared with the Cu/ZnO that was prepared by Pathania, D. et al.
In this paper, the structural properties of pure-phase ZnBiEuO4 and single-phase Eu2SmSbO7 were analyzed using an X-ray diffractometer (XRD), a UV-Vis spectrophotometer, a Fourier transform infrared (FTIR) spectrometer, a Raman spectrometer, X-ray photoelectron spectroscopy (XPS), transmission electron microscopy (TEM), energy dispersive X-ray spectroscopy (EDS), and ultraviolet photoelectron spectroscopy (UPS). In addition, the conversion rates of chlorpyrifos under VLID were measured with pure-phase Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or Eu2SmSbO7/ZnBiEuO4 as a photocatalyst. The aim of this study was to prepare a novel heterojunction photocatalyst for the removal of chlorpyrifos from pesticide wastewater under VLID. Our innovative research involved the first synthesis and characterization of a novel Eu2SmSbO7 nanocatalyst, ZnBiEuO4, and EZHP. For the first time, EZHP was used to degrade chlorpyrifos and the conversion rates of chlorpyrifos under VLID were obtained. The results showed that EZHP had high photocatalytic activity and could effectively degrade chlorpyrifos. This paper will make a huge contribution to the structural construction of novel photocatalysts and to the study of pesticide degradation.

2. Result and Discussion

2.1. XRD Analysis

X-ray diffraction (XRD) uses the diffraction of X-rays in crystals to characterize the diffracted X-ray signals and obtain information about the structure of the crystalline material, the crystal size, and cell parameters. Figure 1 shows the XRD patterns of ZnBiEuO4, Eu2SmSbO7, and Eu2SmSbO7/ZnBiEuO4 heterojunction photocatalyst (EZHP). As can be seen from Figure 1, the diffraction peaks of Eu2SmSbO7 and ZnBiEuO4 could be detected in the XRD spectra of EZHP without any other diffraction peaks that belonged to the impure phase, which indicated that the EZHP was successfully synthesized.
Figure 2a displays the structural properties of Eu2SmSbO7 according to Rietveld analysis using the Materials Studio program. The results of Rietveld refinement for Eu2SmSbO7 showed that the observed intensities and calculated intensities of the pyrochlore-type structure were in fine agreement. Eu2SmSbO7 was a single-phase, cubic crystal system with a space group of Fd3m (modeled to include O atoms) [62]. The atomic structure of Eu2SmSbO7 is shown in Figure 2b, where the cell parameter a for Eu2SmSbO7 is 10.5547 Å. Table 1 presents the atomic coordinates and the structural parameters of Eu2SmSbO7. The full-profile structural refinement of Eu2SmSbO7 generated an unweighted R factor, RP = 2.14%.
The x-coordinate of the O(1) atom reflects the change in crystal structure of A2B2O7-type compounds. If the lengths of the A-O(1) bonds (of which there are six) are identical to the lengths of the A-O(2) bonds (two), it can be assumed that x is equal to 0.375 [63]. The value of x provides information about the octahedral distortion of MO6 (M = Sm3+ and Sb5+). Since the x value was shifted from x = 0.375, it was proved that there existed crystal structure distortions in Eu2SmSbO7 [63]. The photocatalytic degradation of chlorpyrifos under VLID required charge separation to prevent photo-induced electron and photo-induced hole recombination. Deformation of MO6 octahedra has been reported to prevent charge recombination and enhance photocatalytic performance in some photocatalysts such as BaTi4O9 and Sr2M2O7 (M = Nb5+ and Ta5+) [64,65]. Therefore, it was speculated that the deformation of MO6 (M = Sm3+ and Sb5+) octahedra in Eu2SmSbO7 could enhance the photocatalytic performance. Eu2SmSbO7 had a three-dimensional network structure which consisted of corner-sharing MO6 (M = Sm3+ and Sb5+) octahedra. Each Eu3+ ion was connected to two MO6 octahedra to form a chain. Eu-O bonds were found in two lengths, with the six Eu-O (1) bond lengths (4.376 Å) being longer than the two Eu-O (2) bond lengths (2.285 Å). The crystal structure of Eu2SmSbO7 had six M-O (1) (M = Sm3+ and Sb5+) bonds with lengths of 2.285 Å, six M-O (3) bonds with lengths of 2.285 Å, and six M-Eu (M = Sm3+ and Sb5+) bonds with lengths of 3.732 Å. The M-O-M (M = Sm3+ and Sb5+) bond angle in the Eu2SmSbO7 crystal structure was 109.47°, the Eu-M-Eu (M = Sm3+ and Sb5+) bond angle was 135.00°, and the Eu-M-O (M = Sm3+ and Sb5+) bond angle was 125.26°.
Many previous studies have indicated that the size of the bond angle affects the luminescent properties of the material [63,66]. When the bond angles of M-O-M (M = Sm3+ and Sb5+) were close to 180°, the produced photo-induced electrons and photo-induced holes were more easily accessible to the reaction sites on the catalyst surface. In addition, the larger the Eu-Sm-O bond angle or Eu-Sb-O bond angle of Eu2SmSbO7 was, the higher the photocatalytic activity was. The degradation of chlorpyrifos under VLID with Eu2SmSbO7 as a photocatalyst mainly depended on the crystal structure and electronic structure of Eu2SmSbO7.
Figure 3a displays the structural properties of ZnBiEuO4 according to the Rietveld refinement analysis using the Materials Studio version 2.2 software. ZnBiEuO4 had a tetragonal crystalline nature with the space group of I41/A [67]. The atomic structure of ZnBiEuO4 is shown in Figure 3b. It can be found from Figure 3b that the cell parameters for ZnBiEuO4 were a = b = 10.5572 Å and c = 10.0341 Å. The atomic coordinates and structural parameters of ZnBiEuO4 are listed in Table 2. The full-profile structural refinement of ZnBiEuO4 generated an unweighted R factor, RP = 6.62%.
The crystallite size (L) could be calculated using the Scherrer Formula (1) [68,69,70]:
L = K λ β c o s θ
K is the shape constant, normally taken as 0.9. λ is the X-ray wavelength (in this study, λ = 1.54056 Å), β is the half-width of the diffraction peak at the maximum height, and θ is the Bragg angle. Using the Scherrer formula, we could calculate the crystallite size of Eu2SmSbO7 (341 nm) and the crystallite size of ZnBiEuO4 (522 nm). The crystallite size of Eu2SmSbO7 or ZnBiEuO4 was 341 nm or 522 nm after the formation of the EZHP. The crystallite size of Eu2SmSbO7 and ZnBiEuO4 did not change after the formation of the EZHP.

2.2. FTIR Analysis

The absorption peaks of functional groups and chemical bonds can be obtained from FTIR spectroscopy to determine the chemical composition of a substance. Figure 4 displays the FTIR spectra of Eu2SmSbO7, EZHP, and ZnBiEuO4, including the characteristic absorption peaks associated with the Eu–O, Sm–O, Sb–O–Sb, Zn–O, and Bi–O bonds. The stretching vibrations of Eu–O or Sm–O bonds appeared at 590 cm−1or 508 cm−1, respectively [71,72]. The peaks at 728 cm−1 and 649 cm−1 were related to the bending vibrations of the Sb–O–Sb bond [73,74]. The bending vibration of Zn–O bond appeared at 468 cm−1 and the bending vibration of Bi–O bond was associated with the characteristic peak at 429 cm−1 [75,76]. The peak at 1631 cm−1 was related to the bending vibration of the O–H group [77]. The multiple bands that were observed from 1361 cm−1 to 1631 cm−1 corresponded to the vibrations of the C–H bonds [78].

2.3. Raman Analysis

Raman spectroscopy is based on the interaction of chemical bonds within a substance and can reflect the chemical structure of the substance. The Raman spectra of ZnBiEuO4, Eu2SmSbO7, and EZHP are presented in Figure 5. The characteristic peaks of Zn–O, Eu–O, or Bi–O which were found in the Raman spectrum of ZnBiEuO4 appeared at 278 cm−1, 438 cm−1, or 631 cm−1, respectively [79,80,81]. In the Raman spectrum of Eu2SmSbO7, the characteristic peaks of Eu–O appeared at 366 cm−1, 450 cm−1, and 707 cm−1; at the same time, the characteristic peaks of Sm–O appeared at 168 cm−1 [80,82]. The peaks at 224 cm−1, 281 cm−1, 472 cm−1, and 527 cm−1, which are displayed in Figure 5, belonged to Sb–O and Sb–O–Sb [83,84]. Furthermore, the Raman spectrum of the EZHP included different absorption peaks which were derived from Eu2SmSbO7 and ZnBiEuO4, including peaks at 163 cm−1, 225 cm−1, 275 cm−1, 288 cm−1, 367 cm−1, 440 cm−1, 454 cm−1, 471 cm−1, 536 cm−1, 625 cm−1, and 706 cm−1. The above results provide indirect evidence for our successful preparation of ZnBiEuO4, Eu2SmSbO7, and EZHP.

2.4. XPS Analysis

X-ray photoelectron spectroscopy can provide information on elemental composition, atomic valence, and elemental content. XPS analysis was carried out in order to study the chemical compositions and valence states of each element in Eu2SmSbO7, EZHP, and ZnBiEuO4. Figure 6 illustrates the XPS spectra of Eu2SmSbO7, EZHP, and ZnBiEuO4. From Figure 6 it can be seen that the prepared sample of Eu2SmSbO7, ZnBiEuO4, or EZHP was successfully synthesized. In addition, a carbon peak was observed and this carbon peak was attributed to adventitious hydrocarbon as a calibration reference.
Figure 7a–e shows the spectral peaks of Sm 3d, Sb 4d, Eu 4d, Zn 2p, and Bi 4f in Eu2SmSbO7, ZnBiEuO4, and EZHP. These peaks were located at 1084.25 eV (Sm 3d5/2), 34.53 eV (Sb 4d5/2), 135.91 eV (Eu 4d5/2), 135.83 eV (Eu 4d5/2), 1022.30 eV (Zn 2p3/2), 159.33 eV (Bi 4f7/2), and 164.70 eV (Bi 4f5/2), respectively. In EZHP, these peaks were slightly shifted towards higher binding energies, such as 1084.48 eV (Sm 3d5/2), 34.76 eV (Sb 4d5/2), 136.06 eV (Eu 4d5/2), 1022.45 eV (Zn 2p3/2), 159.48 eV (Bi 4f7/2), and 164.85 eV (Bi 4f5/2). This confirmed the existence of a strong interfacial interaction between Eu2SmSbO7 and ZnBiEuO4. The reason for this phenomenon could be the electron transfer and delocalization between Eu2SmSbO7 and ZnBiEuO4 in the heterojunction photocatalyst. Furthermore, in ZnBiEuO4 and EZHP, the spin-orbit separation value between Bi 4f7/2 and Bi 4f5/2 was determined to be 5.37 eV, which indicated that Bi3+ was exclusively present. Figure 7f presents the deconvolution O 1s spectrum of EZHP. The peak at 531.28 eV was assigned to O 1s. In addition, the peaks at 531.83 eV and 540.14 eV corresponded to Sb 3d5/2 and Sb 3d3/2, respectively.
The results of XPS analysis showed that the oxidation states of Eu, Sm, Sb, Zn, Bi, and O ions within Eu2SmSbO7, ZnBiEuO4, and EZHP were +3 [85], +3 [86], +5 [87,88,89], +2 [90], +3 [91,92], and −2 [93]. According to the results of XPS analysis, the surface elemental analyses result for EZHP indicated that the average atomic ratio of Eu:Sm:Sb:Zn:Bi:O was 934:314:301:323:337:7791. According to the results of XPS analysis, the surface atomic concentration ratio of Eu:Sm:Sb:Zn:Bi:O was 9.34%:3.14%:3.01%:3.23%:3.37%:77.91% for EZHP. In the EZHP sample, the atomic ratios of Eu:Sm:Sb and Zn:Bi:Eu were 2.06:1.04:1.00 and 1.03:1.08:1.00, respectively. Obviously, neither shoulders nor widening of the XPS peaks for Eu2SmSbO7 or ZnBiEuO4 were observed, which indicated the absence of any other phases.

2.5. UV-Vis Diffuse Reflectance Spectra

UV–visible diffuse reflectance is a commonly used spectral analysis technique that can be used to determine the bandgap width of a material. Figure 8a illustrates the absorption spectrum of Eu2SmSbO7. Figure 8b shows the plot of (αhν)2 and for Eu2SmSbO7. Figure 9a illustrates the absorption spectrum of ZnBiEuO4. Figure 9b shows the plot of (αhν)1/2 and for ZnBiEuO4. Figure 10a illustrates the absorption spectrum of EZHP. Figure 10b shows the plot of (αhν)1/2 and for EZHP. The absorption edge of the novel photocatalyst Eu2SmSbO7 was at 430 nm; at the same time, the absorption edge of EZHP was at 453 nm and the absorption edge of ZnBiEuO4 was at 482 nm. The three above catalysts had absorption edges in the visible range. Through the point of intersection of the (photon energy)-axis with the prolonging of the linear part of the absorption edge of the Kubelka–Monk function (2) (re-emission function), the bandgap energy of the semiconductor could be determined [94,95].
1 R d ( h ν ) 2 2 R d ( h ν ) = α ( h ν ) S
In above equations, S, Rd, and α represents the scattering quotiety, diffuse reflectance, and radiation absorption quotiety, respectively.
Photo-absorption near the band edge of the crystalloid semiconductor follows Equation (3) [96,97]:
α h ν = A ( h ν E g ) n
In this equation, A, α , Eg, and ν represent the proportional constant, absorption coefficient, band gap, and light frequency, respectively. The value of n controls the transition property of the semiconductor. The values of Eg and n could be identified by following these three steps: (1) draw a graph of l n ( α h ν ) with l n ( h ν E g ) and estimate the approximate value of Eg; (2) based on the rate of slope of the graph, speculate about the value of n; (3) draw a graph of ( α h ν ) 1 / n with h ν , refine the Eg values, and prolong the region of the curve that approximates a straight line until it i n t e r s e c t s t h e b a s e l i n e . The energy bandwidths of Eu2SmSbO7, EZHP, and ZnBiEuO4 could be estimated using the direct method (1240/transition wavelength λ).
The black dashed line in Figure 8b, Figure 9b, and Figure 10b is the extrapolation line of the linear part of the absorption edge; thus, the intersection of the black dashed line with the b a s e l i n e is the bandgap width [98]. According to the calculation that was conducted using above method, the Eg value of Eu2SmSbO7 was 2.881 eV and the Eg value of ZnBiEuO4 was 2.571 eV. Similarly, it could be concluded that the Eg value of EZHP was 2.737 eV. All of the three above catalysts had a visible light response characteristic. The n value of Eu2SmSbO7 was estimated to be 0.5, which indicated that the optical transition was directly allowed. The n value of ZnBiEuO4 or EZHP was estimated to be 2, which indicated that the optical transition was indirectly allowed.

2.6. Property Characterization of Eu2SmSbO7/ZnBiEuO4 Heterojunction Photocatalyst

Figure 11 shows the TEM image of Eu2SmSbO7. Figure 12 displays the TEM image of ZnBiEuO4. As can be seen from Figure 11 and Figure 12, the particle size of Eu2SmSbO7 was about 350 nm, while the particle size of ZnBiEuO4 was about 520 nm, which was about the same as that of the XRD analysis.
Figure 13 shows the TEM image of EZHP. Figure 14 displays EDS elemental maps of EZHP (Eu, Sm, Sb, and O for Eu2SmSbO7, and Zn, Bi, Eu, and O for ZnBiEuO4). Figure 15 illustrates the EDS spectrum of EZHP. From Figure 11, Figure 12, Figure 13 and Figure 14, it can be noticed that the larger particles belonged to ZnBiEuO4 while the smaller particles belonged to Eu2SmSbO7. ZnBiEuO4 particles were surrounded by smaller Eu2SmSbO7 particles, which indicated the successful synthesis of EZHP.
Based on Figure 14 and Figure 15, it could be proved that europium, samarium, antimony, zinc, bismuth, and oxygen elements were present in EZHP and there were no other impurities. It can be concluded that the EZHP that was synthesized in this study was of high purity. According to the EDS spectrum of EZHP in Figure 15, the surface atomic concentration ratio of Eu:Sm:Sb:Zn:Bi:O was 9.36%:2.94%:2.87%:3.29%:3.31%:78.23%, which was approximately the same as the results of XPS analysis.

2.7. Photocatalytic Activity

Figure 16a shows the concentration variation curves of chlorpyrifos during the photocatalytic degradation of chlorpyrifos under VLID using EZHP, Eu2SmSbO7, ZnBiEuO4, nitrogen-doped titanium dioxide (N-doped TiO2), or without a catalyst. N-doped TiO2 is a widely recognized visible light responsive photocatalyst; thus, we chose to use N-doped TiO2 for comparing the photocatalytic activity with that of other photocatalysts [99,100]. It can be obviously seen that the concentration of chlorpyrifos in pesticide wastewater gradually decreased with the extension in VLID time.
According to the above findings, it could be concluded that the descending order of photodegradation efficiency was as following: EZHP > Eu2SmSbO7 > ZnBiEuO4 > N-doped TiO2 under the same conditions. Comparing the conversion rates of chlorpyrifos after VLID of 160 min, it could be found that the conversion rate of chlorpyrifos using EZHP was 1.13 times higher than that with Eu2SmSbO7 as the photocatalyst, 1.19 times higher than that with ZnBiEuO4 as the photocatalyst, and 2.84 times higher than that with N-doped TiO2 as the photocatalyst.
Figure 16b presents the concentration variation curves of total organic carbon (TOC) during the photocatalytic degradation of chlorpyrifos within pesticide wastewater under VLID using EZHP, Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or without a catalyst. As can be seen from Figure 16b, the concentration of TOC decreased gradually with increasing VLID time.
By analyzing the data that are shown in Figure 16b, it was demonstrated that after VLID of 160 min, the conversion rates of TOC that was derived from pesticide wastewater were 98.02%, 84.04%, 80.12%, 30.23%, and 4.14% when chlorpyrifos was degraded using EZHP, Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or without catalyst, respectively. In other words, the descending order of the conversion rates of TOC during the degradation of chlorpyrifos was as follows: EZHP > Eu2SmSbO7 > ZnBiEuO4 > N-doped TiO2. The above results indicated that EZHP had a higher mineralization rate compared with Eu2SmSbO7, ZnBiEuO4, or N-doped TiO2 during the degradation of chlorpyrifos.
Figure 17 demonstrates the effect of different photocatalyst doses on the conversion rates when using EZHP to degrade chlorpyrifos under VLID. Beginning with the increase in the initial concentration of EZHP, the conversion rates of chlorpyrifos increased after VLID of 160 min. The conversion rate of chlorpyrifos reached 100% at an initial concentration of EZHP of 0.75 g/L. After that, the conversion rates of chlorpyrifos decreased with the increase in the initial concentration of EZHP. The reason for the decrease in the conversion rates of chlorpyrifos could be the aggregation of a high concentration of EZHP, which resulted in the decrease in active sites on the surface of the catalyst [101]. The above results indicated that the optimum conversion rate was achieved at an initial EZHP concentration of 0.75 g/L. Table 3 demonstrates the effect of different ratios of different catalysts to chlorpyrifos on the conversion rates of chlorpyrifos. From the data in Table 3, it can be seen that the optimum conversion rate of chlorpyrifos was achieved at a ratio of catalyst to chlorpyrifos of 66.96.
The concentration variation curves of chlorpyrifos during four cyclic degradation experiments using EZHP under VLID are shown in Figure 18a. It can be deduced from Figure 18a that after VLID of 160 min using EZHP, the conversion rates of chlorpyrifos reached 98.16%, 97.03%, 96.03%, or 95.06%, respectively. From the four cyclic degradation experiments, the corresponding photon efficiencies were 0.0687%, 0.0679%, 0.0672%, or 0.0666%, respectively. Figure 18b illustrates the concentration variation curves of TOC during the photocatalytic degradation of chlorpyrifos using EZHP under VLID. It can be concluded from Figure 18b that the conversion rates of TOC were 96.94%, 95.72%, 94.68%, or 93.66% using EZHP after VLID of 160 min. According to the four consecutive cyclic degradation experiments using EZHP, the conversion rates of chlorpyrifos decreased by 1.84%, 1.13%, 1%, or 0.97% after each cyclic degradation experiment, respectively. It can be concluded from Figure 18b that the conversion rates of TOC decreased by 1.08%, 1.22%, 1.04%, or 1.02% after each cyclic degradation experiment using EZHP under VLID. The above results demonstrated that EZHP had excellent structural stability and reusability.
To demonstrate the significance of our study in the field of photocatalytic degradation of chlorpyrifos, a comparative overview of relevant studies in this field is presented in Table 4. The data in Table 4 show that the photocatalytic activity of EZHP was better than that of other catalysts. These results indicate that EZHP can increase the conversion rates of chlorpyrifos and make a significant contribution to the field of photocatalysis. In conclusion, XRD and XPS analyses demonstrated the excellent structural stability of EZHP during the photocatalytic degradation of chlorpyrifos under VLID.
The first-order kinetic plots that correspond to chlorpyrifos concentration and VLID time during photocatalytic degradation of chlorpyrifos under VLID using EZHP, Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or without a catalyst, are shown in Figure 19a. Figure 19b presents the first-order kinetic curves that correspond to TOC concentration and VLID time during photocatalytic degradation of chlorpyrifos under VLID using EZHP, Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or without a catalyst. The kinetic constants were calculated from the equation ( l n C 0 C = k C t ) and ( l n T O C 0 T O C = k T O C t ) . In the equation, C0 is the initial saturation of chlorpyrifos, C is the reactive saturation of chlorpyrifos, TOC0 represents the initial saturation of total organic carbon, and TOC represents the reactive saturation of total organic carbon. It can be concluded from Figure 19a that the kinetic constant value KC that was derived from the dynamic curves for EZHP, Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or without a catalyst, was 0.0202 min−1, 0.0090 min−1, 0.0075 min−1, 0.0020 min−1, and 0.0003 min−1 under VLID, and its corresponding linear correlation coefficient (R2) was 0.8511, 0.8741, 0.8524, 0.9919, and 0.9773, respectively. Moreover, it can be deduced from Figure 19b that the kinetic constant value KTOC that originated from the dynamic curves for EZHP, Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or without a catalyst, was 0.0182 min−1, 0.0078 min−1, 0.0067 min−1, 0.0017 min−1, and 0.0002 min−1 under VLID, and its corresponding R2 was 0.8512, 0.8775, 0.8562, 0.9962, and 0.9745, respectively. According to the above studies, it was found that the mineralization efficiency during photocatalytic degradation of chlorpyrifos using EZHP was higher than that using Eu2SmSbO7 or ZnBiEuO4 under VLID. Furthermore, the KTOC value of the above four photocatalysts was lower than the KC value of the above four photocatalysts, which indicates that the intermediate products might be formed during the photocatalytic degradation of chlorpyrifos.
Figure 20a displays the first-order kinetic plots that correspond to chlorpyrifos concentration and VLID time during four cyclic degradation experiments for photocatalytic degradation of chlorpyrifos using EZHP. Figure 20b shows the first-order kinetic plots that correspond to TOC concentration and VLID time during four cyclic degradation experiments for photocatalytic degradation of chlorpyrifos using EZHP. Based on the analysis of the data in Figure 20a, the kinetic constant value KC that was derived from the dynamic curves using EZHP for the four cyclic degradation experiments was 0.0186 min−1, 0.0161 min−1, 0.0144 min−1, and 0.0131 min−1 under VLID, and its corresponding R2 was 0.8527, 0.8605, 0.8604, and 0.8562, respectively. It can be deduced from Figure 20b that the kinetic constant value KTOC that originated from the dynamic curves using EZHP for the four cyclic degradation experiments was 0.0158 min−1, 0.0137 min−1, 0.0125 min−1, and 0.0110 min−1 under VLID, and its corresponding R2 was 0.8608, 0.8517, 0.8509, and 0.8503, respectively. During the cycling experiments for degrading chlorpyrifos, we used the catalyst that was extracted from the previous degradation experiment of chlorpyrifos for the next degradation experiment of chlorpyrifos; as a result, KC and KTOC decreased with the increasing number of cycling experiments. According to the above results, it could be concluded that the photocatalytic degradation of chlorpyrifos that was derived from pesticide wastewater using EZHP followed the first-order reaction kinetics under VLID.
Figure 21 shows the XRD pattern of EZHP before and after the photocatalytic degradation reaction of chlorpyrifos. The crystal structure of EZHP before and after the photocatalytic degradation of chlorpyrifos under VLID did not change significantly, as can be seen in Figure 21. Figure 22 displays the XPS spectra of EZHP before and after the photocatalytic degradation reaction of chlorpyrifos. From Figure 22, it can be seen that the elemental composition and atomic valence of EZHP did not change remarkably before and after the photocatalytic degradation of chlorpyrifos under VLID.
Figure 23a shows the effect of different pH values on the conversion rates of chlorpyrifos during photocatalytic degradation of chlorpyrifos using EZHP under VLID. Analysis of the data in Figure 23a showed that the conversion rates of chlorpyrifos reached 99.2%, 100%, and 98.8% after VLID of 160 min at pH 3, 7, and 11, respectively. Therefore, it could be concluded that different pH values have little effect on the conversion rates of chlorpyrifos using EZHP under VLID.
Figure 23b displays the effect of different metal ions species on the conversion rates of chlorpyrifos during photocatalytic degradation of chlorpyrifos using EZHP under VLID. Ultrapure water containing Mg2+, Zn2+, or Ba2+ was introduced into the photocatalytic reaction system. It was found that the concentration of chlorpyrifos in pesticide wastewater decreased gradually with the increase in VLID time. In addition, by analyzing the data in Figure 23b, it could be seen that the conversion rate of chlorpyrifos using EZHP without the addition of metal ions after VLID of 160 min was 100%. Under the same conditions, the introduction of 1 mmol/L Mg2+, 1 mmol/L Zn2+, or 1 mmol/L Ba2+ metal ions resulted in the conversion rates of chlorpyrifos to 99.75%, 98.94%, and 99.37%, respectively. Overall, the addition of Mg2+, Zn2+, or Ba2+ metal ions did not significantly inhibit the photocatalytic degradation of chlorpyrifos using EZHP under VLID.
Figure 23c presents the effect of different anion species on the conversion rates of chlorpyrifos during photocatalytic degradation of chlorpyrifos using EZHP under VLID. Ultrapure water containing NO3, SO42−, CO32−, or Cl was introduced into the photocatalytic reaction system. It was found that the concentration of chlorpyrifos in pesticide wastewater decreased gradually with the increase of VLID time. Furthermore, by analyzing the data in Figure 23c, it could be seen that the conversion rate of chlorpyrifos using EZHP without the addition of anions after VLID of 160 min was 100%. Under the same conditions, the introduction of 1 mmol/L NO3, 1 mmol/L SO42−, 1 mmol/L CO32−, or 1 mmol/L Cl anions resulted in the conversion rates of chlorpyrifos of 96.56%, 98.13%, 73.44%, and 62.50%, respectively. The Cl ion has a stronger inhibitory effect on the degradation of chlorpyrifos. The Cl ion consumes •OH and photo-induced holes, leading to the reduction of free radicals and thus hindering the photocatalytic reaction. The inhibition of chlorpyrifos degradation by NO3 or SO42− ions was small. This is due to the fact that NO3 or SO42− only consume the photo-induced holes in the photocatalytic reaction system, and thus the inhibition is weaker.
Figure 24 shows the effect of the addition of isopropanol (IPA), benzoquinone (BQ), or ethylenediaminetetraacetic acid (EDTA) on the conversion rates of chlorpyrifos using EZHP under VLID. In order to identify the active species involved in the photocatalytic reaction during chlorpyrifos photocatalytic degradation, different radical scavengers, as described above, were introduced at the beginning of the photocatalytic experiments. IPA captured hydroxyl radicals (•OH), while BQ captured superoxide anions (•O2) and EDTA captured holes (h+). The volume of IPA, BQ, or EDTA added was 1 mL at a concentration of 0.15 mmol L−1. Compared with the control group, the conversion rates of chlorpyrifos in the presence of IPA, BQ, or EDTA decreased by 47.62%, 35.66%, and 22.97% after VLID of 160 min. Therefore, it could be concluded that •OH, h+, and •O2 acted as reactive radicals during the photocatalytic degradation of chlorpyrifos. The above experiments with the addition of scavengers showed that the oxidative removal capacities that were from the largest to the smallest for the three above oxidizing radicals during photocatalytic degrading of chlorpyrifos obeyed the following order: •OH > •O2 > h+. Thus, we could conclude that •OH possessed the strongest oxidative removal ability for photocatalytic degradation of chlorpyrifos within pesticide wastewater.
Photoluminescence (PL) spectra and time-resolved photoluminescence (TRPL) spectra can reflect the complexation rate of photo-induced electrons and photo-induced holes, and provide the electronic lifetime of the photocatalyst. Figure 25a displays the PL spectra of EZHP, Eu2SmSbO7, and ZnBiEuO4. Figure 25b–d show the TRPL spectrum of Eu2SmSbO7, ZnBiEuO4, and EZHP, respectively. The lower the relative intensities of the vertical coordinates corresponding to the PL spectra, the more difficult it was to combine the photo-induced electrons and photo-induced holes. The above results would lead to the prolongation of the survival life of photo-induced electrons and photo-induced holes; as a result, the numbers of •O2 and •OH would be increased. Ultimately, the photocatalytic activity was improved [107,108]. According to Figure 25a, it can be seen that the value of the relative intensity of the longitudinal coordinate of the PL spectrum corresponding to EZHP was lower than the value of the relative intensity of the longitudinal coordinate of the PL spectrogram corresponding to Eu2SmSbO7. At the same time, the value of the relative intensity of the longitudinal coordinate of the PL spectrum corresponding to Eu2SmSbO7 was lower than the value of the relative intensity of the longitudinal coordinate of the PL spectrum corresponding to ZnBiEuO4. Therefore, based on above results, it could be concluded that the photocatalytic activity was ranked from high to low as follows: EZHP > Eu2SmSbO7 > ZnBiEuO4. The above results suggest that the compounding rate of the photo-induced electrons and the photo-induced holes could be reduced by constructing a heterojunction structure. The TRPL spectrum of Eu2SmSbO7, ZnBiEuO4, and EZHP in Figure 25b–d was fitted by the double-exponential decay shown in Equation (4) [109]:
I ( t ) = I 0 + A 1   e x p ( t τ 1 ) + A 2   e x p ( t τ 2 )
In above equations, τ 1 or τ 2 denote the fast and slow attenuation components, respectively. It is usually assumed that the τ 1 component is attributed to non-radiative recombination involving defects or traps, while the τ 2 component corresponds to radiative recombination within the photocatalyst. In order to obtain the average electron lifetime ( τ a v e ), Equation (5) could be used as follows [110]:
τ a v e = ( A 1 τ 1 2 + A 2 τ 2 2 ) / ( A 1 τ 1 + A 2 τ 2 )
Table 5 lists the calculated lifetimes and the corresponding parameters. By comparing the data in Table 5, it was observed that the average electron lifetime of EZHP ( τ a v e = 3.0910 ns) was higher than the average electron lifetime of Eu2SmSbO7 ( τ a v e = 1.5094 ns) and the average electron lifetime of ZnBiEuO4 ( τ a v e = 1.1692 ns). The above results again indicate that EZHP had a better photocatalytic activity than Eu2SmSbO7 or ZnBiEuO4.
Electrochemical impedance spectroscopy (EIS) reflects the relationship between the impedance of the electrode interface and the frequency of the change in the applied voltage or current under light conditions. At the same time, the migration process of photo-induced electrons and photo-induced holes at the interfaces of Eu2SmSbO7 and ZnBiEuO4, which constituted the EZHP, could be understood based on the EIS. The radius of the arc in the Nyquist impedance diagram can be used to compare the compounding rates of photo-induced electrons and photo-induced holes in the catalysts; thus, the length of the survival lifetimes of the photo-induced electrons and photo-induced holes can be determined. Figure 26 illustrates the Nyquist impedance plot of EZHP, Eu2SmSbO7, and ZnBiEuO4. It should be known that the smaller the arc radius of the Nyquist impedance plots profile of the catalyst, the more difficult it was to combine the photo-induced electrons and photo-induced holes, which led to the longer survival life of the photo-induced electrons and photo-induced holes [67]. As can be seen from Figure 26, the descending order of the arc radius that was derived from the Nyquist impedance plot was as follows: ZnBiEuO4 > Eu2SmSbO7 > EZHP. The above results show that the prepared EZHP had a longer survival lifetime of photo-induced electrons and photo-induced holes and higher interfacial charge mobility; thus, EZHP had better photocatalytic activity.

2.8. Analysis of Possible Degradation Mechanisms

The conceivable photocatalytic degradation mechanism of chlorpyrifos using EZHP under VLID is illustrated in Figure 27. The electrochemical potential of the conduction band (CB) in semiconductor catalysts could be calculated according to Equation (6) [111]. According to Equation (7), the electrochemical potential of the valence band (VB) in the semiconductor catalyst could be calculated [111]. Equations (6) and (7) are as follows:
ECB = XEe − 0.5Eg
EVB = ECB + Eg
where Eg is the band gap of the semiconductor and Ee is the energy of the free electrons on the hydrogen scale (about 4.5 eV). X is the electronegativity of the semiconductor. X is the average of the ionization energy ( I ) and the electron affinity energy ( A ), expressed through Equation (8) [112]. The X of the photocatalyst could be calculated using Equation (9) [113].
X = ( I + A ) / 2
X = [ X ( A ) a X ( B ) b X ( C ) c ] 1 / ( a + b + c )
A, B, and C and a, b, and c are the corresponding element and atomic numbers in the semiconductor photocatalysts. The electronegativity of Eu2SmSbO7 and ZnBiEuO4 was calculated to be 5.745 and 4.667, respectively. Based on above equations, the VB potential and CB potential of Eu2SmSbO7 were estimated to be 2.685 eV and −0.196 eV, respectively; contemporaneously, the VB potential and CB potential of ZnBiEuO4 were estimated to be 1.452 eV and −1.119 eV, respectively. Eu2SmSbO7 and ZnBiEuO4 were both able to absorb visible light energy and produced electron–hole pairs when Eu2SmSbO7 or ZnBiEuO4 was exposed to VLID. The redox potential of the CB for ZnBiEuO4 (−1.119 eV) was more negative than the redox potential of the CB for Eu2SmSbO7 (−0.196 eV). At the same time, the redox potential position of the VB of Eu2SmSbO7 (2.685 eV) was more positive than the redox potential position of the VB for ZnBiEuO4 (1.452 eV). Therefore, the photo-induced electrons on the CB of ZnBiEuO4 could be transferred to the CB of Eu2SmSbO7, while the photo-induced holes on the VB of Eu2SmSbO7 could be transferred to the VB of ZnBiEuO4. Therefore, the composition of EZHP effectively increased the separation rate of photo-induced electrons and photo-induced holes; as a result, the above results improved the interfacial charge transfer efficiency [114]. More oxidizing radicals (•OH and •O2) participated during the photocatalytic reaction process of degrading chlorpyrifos, thereby improving the degradation efficiency.
In addition, as shown in pathway 1 of Figure 27, because the CB potential of ZnBiEuO4 (−1.119 eV) was more negative than the potential of O2/•O2 (−0.33 eV), the photo-induced electrons within the CB of ZnBiEuO4 could absorb oxygen to produce •O2. •O2 was used to degrade chlorpyrifos. Similarly, as illustrated in pathway 2 of Figure 27, because the VB potential of Eu2SmSbO7 (2.685 eV) was more positive than the potential of OH/•OH (2.38 eV), the photo-induced holes in the VB of Eu2SmSbO7 could oxidize H2O or OH into •OH, which could degrade chlorpyrifos. Finally, the photo-induced holes within the VB of ZnBiEuO4 or Eu2SmSbO7 could directly oxidize chlorpyrifos, as illustrated in Path 3 of Figure 27. In conclusion, the construction of EZHP could enhance the separation rate of the photo-induced electrons and the photo-induced holes; therefore, EZHP exhibited excellent photocatalytic performance during the process of degrading chlorpyrifos.
The basic principle of the ultraviolet photoelectron spectroscopy (UPS) is the photoelectric effect, with ultraviolet light as the excitation light source. The ionization potential of the photocatalyst and the electrochemical potential of the VB could be obtained by UPS. The UPS spectrum of Eu2SmSbO7 is illustrated in Figure 28a. It can be deduced from Figure 28a that the measured onset (Ei) energy and cut-off (Ecutoff) energy of Eu2SmSbO7 were 0.496 eV and 19.011 eV, respectively. Figure 28b presents the UPS spectrum of ZnBiEuO4. It can be deduced from Figure 28b that the onset (Ei) energy and cut-off (Ecutoff) energy of ZnBiEuO4 were measured to be 0.159 eV and 19.907 eV, respectively. The ionization potentials of Eu2SmSbO7 and ZnBiEuO4 could be calculated by decreasing the width of the UPS spectrum by the excitation energy, which was about 21.2 eV [115]. According to the above approach, the VB ionization potential of Eu2SmSbO7 and ZnBiEuO4 was calculated to be 2.685 eV and 1.452 eV, respectively, which was consistent with the VB potential that was derived from the above simulation calculation results.
The electrochemical potential of O2/•O2− was -0.33 eV and the absolute electrochemical potential value of OH/•OH was 2.38 eV. The ionization potentials of the VB for Eu2SmSbO7 and ZnBiEuO4 were calculated to be 2.685 eV and 1.452 eV, respectively. Thus, it can be deduced from Figure 28 that the photo-induced holes in the VB of Eu2SmSbO7 possessed the ability to oxidize H2O or OH to form •OH. In addition, the CB potential of Eu2SmSbO7 and ZnBiEuO4 was −0.196 eV and −1.119 eV, respectively, which indicated that the photo-induced electrons in the conduction band of ZnBiEuO4 could absorb O2 to form •O2.
Based on above studies, EZHP possessed the ability to produce •OH and •O2 under VLID, which could degrade chlorpyrifos efficiently within pesticide wastewater.
The intermediates produced during chlorpyrifos degradation by LC-MS were identified to investigate the mechanism of chlorpyrifos degradation. The intermediate products were identified as C5H2Cl3NO (m/z = 196), C5H2Cl3N (m/z = 180), C5H3Cl2NO (m/z = 162), C5H3Cl2N (m/z = 146), C5H5N (m/z = 79), C4H11O3PS (m/z = 170), C4H11O4P (m/z = 154), C2H7O4P (m/z = 126), and H3O4P (m/z = 97). Based on above-mentioned detected intermediate products, we could deduce the degradation pathway of chlorpyrifos. Figure 29 shows the degradation pathway of chlorpyrifos. From Figure 29, it can be seen that chlorpyrifos underwent oxidation and hydroxylation reactions during photocatalytic degradation. Eventually, the above results revealed that chlorpyrifos was converted to CO2, H2O, SO42−, NO3−, and PO43−.

3. Experimental Section

3.1. Materials and Reagents

Materials purchased from Sinopharm Group Chemical Reagent Co., Ltd. (Shanghai, China) includes ethylenediaminetetraacetic acid (EDTA, C10H16N2O8, purity = 99.5%), isopropanol (IPA, C3H8O, purity ≥ 99.7%), and P-benzoquinone (BQ, C6H4O2, purity ≥ 98.0%). Materials purchased from Aladdin Group Chemical Reagent Co., Ltd. (Shanghai, China) included Eu(NO3)3·6H2O (purity = 99.99%), Sm(NO3)3·6H2O (purity = 99.9%), SbCl5 (purity = 99.99%), Zn(NO3)2·6H2O (purity = 99.99%), Bi(NO3)3·5H2O (purity = 99.99%) tetrabutyl titanate (C16H36O4Ti, purity = 99.99%), ammonia (H5NO, purity ≥ 25%), glacial acetic acid (CH3CO2H, purity = 100%), absolute ethanol (C2H5OH, purity ≥ 99.5%), and chlorpyrifos (C9H11Cl3NO3PS, purity ≥ 99.0%). Ultrapure water (18.25 MU cm) was used during the experiment.

3.2. Synthesis of N-Doped TiO2

Nitrogen-doped titanium dioxide (N-doped TiO2) used in this research was synthesized by the sol–gel method using tetrabutyl titanate as the precursor and ethanol as the dissolvent. N-doped TiO2 was synthesized as follows:
In the first step, mixed Solution 1 (17 mL tetrabutyl titanate and 40 mL absolute ethanol) and mixed Solution 2 (5 mL double-distilled water, 10 mL glacial acetic acid, and 40 mL absolute ethanol) were prepared, and mixed Solution 1 was added to mixed Solution 2 by dropwise addition under continuous stirring. At this time, a transparent colloidal suspension was obtained. In the second step, ammonia (N/Ti ratio of 8 mol%) was added to the above solution to form a dry gel after 48 h of aging. Finally, the dry gel was ground into powder and calcined at 500 °C for 2 h and then ground again and passed through a vibrating sieve to obtain N-TiO2 powder.

3.3. Synthesis of Eu2SmSbO7/ZnBiEuO4 Heterojunction Photocatalyst

Eu2SmSbO7 and ZnBiEuO4 were prepared by the hydrothermal method. Quantities of 0.30 mol/L Eu(NO3)3·6H2O, 0.15 mol/L Sm(NO3)3·6H2O, and 0.15 mol/L SbCl5 were mixed and stirred for 20 h and then transferred to a Teflon-lined autoclave and heated at 200 °C for 15 h. Subsequently, the resulting powder was calcined in a tube furnace at 790 °C for 10 h under nitrogen to obtain Eu2SmSbO7 powder.
To prepare ZnBiEuO4, 0.15 mol/L Zn(NO3)2·6H2O, 0.15 mol/L Bi(NO3)3·5H2O, and 0.15 mol/L Eu(NO3)3·6H2O were mixed and stirred for 20 h. The solution was transferred to a Teflon-lined autoclave and heated at 200 °C for 15 h and then calcined in a tube furnace under nitrogen at 770 °C for 10 h.
In this study, the solvothermal method was used to synthesize EZHP. The synthesized Eu2SmSbO7 (1 mol) and ZnBiEuO4 (1 mol) were added to 300 mL of octanol (C8H18O) and dispersed for 1 h using an ultrasonic wave. Then, in order to make it easier for ZnBiEuO4 to adhere to the surface of Eu2SmSbO7, it was heated and refluxed at 150 v for 2 h under vigorous stirring. Finally, the product was collected by centrifugation after cooling to room temperature. The collected EZHP was washed several times with a hexane/ethanol mixture to purify the EZHP. The pure EZHP was dried at 60 °C for 6 h and stored.

3.4. Characterization

In order to analyze the properties such as the crystal structure of Eu2SmSbO7 and ZnBiEuO4, the tests were carried out using an X-ray diffractometer (XRD, Shimadzu, XRD-6000, Kyoto, Japan). The conformation and microstructure of EZHP were characterized using a transmission electron microscope (TEM, JEM—F200 FEI Tecnai G2 F20 FEI Talos F200s, Thermo Fisher Scientific Corporation, Waltham, MA, USA), and the micro-regional compositional elements of the samples were determined using energy dispersive spectroscopy (EDS). To obtain the energy band widths of Eu2SmSbO7, ZnBiEuO4, and EZHP, a UV–visible spectrophotometer (UV-Vis DRS, Shimadzu, UV-3600, Kyoto, Japan) was used. A Fourier transform infrared spectrometer (FTIR, WQF-530A, Beifen-Ruili Analytical Instrument (Group) Co., Ltd., Beijing, China) and laser Raman spectrometer (INVIA0919-06, Renishaw plc New Mills Wotton-under-Edge Gloucestershire GL12 8JR, London, United Kingdom) were used to analyze the chemical bonding of Eu2SmSbO7, ZnBiEuO4, and EZHP. The surface chemistry and valence states of Eu2SmSbO7, ZnBiEuO4, and EZHP were analyzed using an X-ray photoelectron spectrometer (XPS, PHI 5000 VersaProbe, UlVAC-PHI, Chigasaki, Japan) and an Al-kα X-ray source. The ionization potentials of Eu2SmSbO7 and ZnBiEuO4 were analyzed by Ultraviolet photoelectron spectroscopy (UPS, Escalab 250Xi, Thermo Fisher Scientific Corporation, Waltham, MA, USA).

3.5. Photoelectrochemical Experiments

The electrochemical impedance spectroscopy (EIS) of Eu2SmSbO7, ZnBiEuO4, and EZHP was experimentally derived from a CHI660D electrochemical bench (Shanghai Zhenhua Instrument Co., Ltd., Shanghai, China) equipped with standard three electrodes. The working electrodes were fabricated by mixing 0.03 g of photocatalyst (Eu2SmSbO7, ZnBiEuO4, or EZHP) and 0.01 g chitosan with 0.45 mL of dimethylformamide, and then ultrasonicated for 1 h. The processed solution was added dropwise to a conductive glass (10 mm × 20 mm) made with indium tin oxide (ITO) and then dried at 80 °C for 10 min. A platinum plate was used as the counter electrode and an Ag/AgCl electrode was used as the reference electrode. Na2SO4 aqueous solution (0.5 mol/L) was used as the electrolyte. A 500 W Xe lamp with a 420 nm cut-off filter was used as the experimental light source.

3.6. Experimental Setup and Procedure

The photocatalytic experiments were carried out at a constant temperature of 20 °C, and the visible light in the experiments was simulated using a xenon lamp (500 W) with a cut-off filter (420 nm) in a photochemistry chamber (CEL-LB70, China Education Au-Light Technology Co., Ltd., Beijing, China).
Twelve identical quartz tubes with a volume of 40 mL were used for the individual reaction solutions. The total reaction volume of the pesticide effluent was 480 mL. The amount of catalyst (Eu2SmSbO7 or ZnBiEuO4) was 0.75 g/L. The amount of EZHP was 0.35 g/L, 0.45 g/L 0.55 g/L, 0.65 g/L 0.75 g/L, 0.85 g/L, or 0.95 g/L. The initial concentration of chlorpyrifos was 0.032 mmol/L.
The concentration of chlorpyrifos in the solution was determined by liquid chromatography (Agilent Technologies, Palo Alto, CA, USA) after filtering out the catalyst from 3 mL of the suspension at the corresponding time according to the experimental requirements. The mobile phase consisted of one-half CH3CN and one-half distilled deionized water. The injection volume of the test solution was 10 μL, and the flow rate was 1 mL·min−1.
In order to establish a reaction system with an equilibrium of adsorption/desorption of the photocatalyst, chlorpyrifos, and atmospheric oxygen, the solution containing the photocatalyst and chlorpyrifos was magnetically stirred for 45 min under dark conditions before VLID. Stirring was conducted at 500 rpm/min during the desorption of chlorpyrifos under VLID.
In order to determine the concentration of TOC during the photocatalytic degradation of chlorpyrifos, the solution was analyzed using a TOC analyzer (TOC-5000 A, Shimadzu Corporation, Kyoto, Japan). Potassium acid phthalate (KHC8H4O4) solutions of known carbon concentration (0 to 100 mg/L) were prepared as standard reagents for comparison. The TOC concentration of six reaction solutions of 45 mL was measured.
A liquid chromatograph mass spectrometer (LC-MS, Thermo Quest LCQ Duo, Thermo Fisher Scientific Corporation, Waltham, MA, USA) was used to identify the intermediates produced during the photocatalytic degradation of chlorpyrifos. After the photocatalytic degradation of chlorpyrifos, 20 μL of the solution was automatically injected into the instrument for testing. The mobile phase consisted of 60% methanol and 40% ultrapure water at a flow rate of 0.2 mL/min. The relevant parameter conditions for the experiment were set to a capillary temperature of 27 °C, a voltage of 19.00 V, a spray voltage of 5000 V, and a constant sheath gas flow rate. The m/z range of the test spectrum was 50 to 600.
The incident photon flux Io was 4.76 × 10−6 Einstein L−1 s−1 under VLID measured with a radiometer (Model FZ-A, Photoelectric Instrument Factory Beijing Normal University, Beijing, China). The incident photon flux could be adjusted by varying the distance between the photoreactor and the xenon arc lamp.
The photonic efficiency could be calculated by Equation (10):
ϕ = R I o
In the above equations, ϕ, R, and Io represent the photonic efficiency (%), conversion rate of chlorpyrifos (mol L−1 s−1), and incident photon flux (Einstein L−1 s−1), respectively.

4. Conclusions

In this study, Eu2SmSbO7 with high photocatalytic activity was prepared via the hydrothermal method for the first time. The new photocatalyst, Eu2SmSbO7, was proven to be a pure phase with a pyrochlore structure, cubic crystal system, and Fd3m space group. Various characterization techniques, such as XRD analysis, FT-IR, Raman spectrometry, UV-Vis, XPS, and TEM-EDS, were utilized to detect the properties of the newly prepared photocatalysts. The lattice parameter and band gap of Eu2SmSbO7 were a = 10.5547 Å and 2.881 eV, respectively. EZHP was successfully prepared using the solvothermal method for the first time. The results demonstrated that EZHP was effective for removing chlorpyrifos from pesticide wastewater. After VLID of 160 min to degrade chlorpyrifos using EZHP, the conversion rate of chlorpyrifos reached 100% and the conversion rate of TOC was 98.02%. The kinetic constant value KC that was derived from the dynamic curves for EZHP was 0.0202 min−1. After VLID of 160 min to degrade chlorpyrifos using Eu2SmSbO7, the conversion rate of chlorpyrifos reached 88.16% and the conversion rate of TOC was 84.04%. The kinetic constant value KC that was derived from the dynamic curves for Eu2SmSbO7 was 0.0090 min−1. After VLID of 160 min, the conversion rates of chlorpyrifos using EZHP were 1.13 times, 1.19 times, and 2.84 times higher than the conversion rates of chlorpyrifos with Eu2SmSbO7, ZnBiEuO4, or N-doped TiO2 as the photocatalyst. Therefore, it can be concluded that the use of EZHP was an effective method for treating chlorpyrifos that was derived from pesticide wastewater. Finally, the possible photodegradation pathways of chlorpyrifos was hypothesized. Chlorpyrifos was degraded into inorganic compounds such as CO2, H2O, SO42−, NO3−, and PO43−.

Author Contributions

J.L. (Jingfei Luan): conceptualization, data curation, formal analysis, investigation, methodology, software, visualization, writing—original draft preparation, writing—review and editing, validation. Y.W.: software, data curation, methodology, writing—original draft preparation, validation. Y.Y.: formal analysis, investigation, writing—original draft preparation, validation, investigation. L.H.: software, visualization, validation. J.L. (Jun Li): formal analysis, methodology, writing—original draft preparation. Y.C.: software, validation, data curation. 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 the Natural Science Foundation of the Science and Technology Development Plan Project of Jilin Province, China (Grant No. YDZJ202101ZYTS161).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Soria-Lopez, A.; Garcia-Perez, P.; Carpena, M.; Garcia-Oliveira, P.; Otero, P.; Fraga-Corral, M.; Cao, H.; Prieto, M.A.; Simal-Gandara, J. Challenges for future food systems: From the Green Revolution to food supply chains with a special focus on sustainability. Food Front. 2023, 4, 9–20. [Google Scholar] [CrossRef]
  2. Fukase, E.; Martin, W. Economic growth, convergence, and world food demand and supply. World. Dev. 2020, 132, 104954. [Google Scholar] [CrossRef]
  3. Webb, R.; Buratini, J. Global Challenges for the 21st Century: The Role and Strategy of the Agri-Food Sector. Anim. Reprod. 2016, 13, 133–142. [Google Scholar] [CrossRef]
  4. Liu, Y.B.; Pan, X.B.; Li, J.S. A 1961–2010 record of fertilizer use, pesticide application and cereal yields: A review. Agron. Sustain. Dev. 2015, 35, 83–93. [Google Scholar] [CrossRef]
  5. Delcour, I.; Spanoghe, P.; Uyttendaele, M. Literature review: Impact of climate change on pesticide use. Food Res. Int. 2015, 68, 7–15. [Google Scholar] [CrossRef]
  6. Popp, J.; Peto, K.; Nagy, J. Pesticide productivity and food security. A review. Agron. Sustain. Dev. 2013, 33, 243–255. [Google Scholar] [CrossRef]
  7. Washuck, N.; Hanson, M.; Prosser, R. Yield to the data: Some perspective on crop productivity and pesticides. Pest Manag. Sci. 2022, 78, 1765–1771. [Google Scholar] [CrossRef]
  8. Rodrigues, E.T.; Lopes, I.; Pardal, M.A. Occurrence, fate and effects of azoxystrobin in aquatic ecosystems: A review. Environ. Int. 2013, 53, 18–28. [Google Scholar] [CrossRef]
  9. Wahab, S.; Muzammil, K.; Nasir, N.; Khan, M.S.; Ahmad, M.F.; Khalid, M.; Ahmad, W.; Dawria, A.; Reddy, L.K.V.; Busayli, A.M. Advancement and New Trends in Analysis of Pesticide Residues in Food: A Comprehensive Review. Plants 2022, 11, 1106. [Google Scholar] [CrossRef]
  10. Wan, M.T. Ecological risk of pesticide residues in the British Columbia environment: 1973–2012. J. Environ. Sci. Health B 2013, 48, 344–363. [Google Scholar] [CrossRef]
  11. El-Sheikh, E.S.A.; Ramadan, M.M.; El-Sobki, A.E.; Shalaby, A.A.; McCoy, M.R.; Hamed, I.A.; Ashour, M.B.; Hammock, B.D. Pesticide Residues in Vegetables and Fruits from Farmer Markets and Associated Dietary Risks. Molecules 2022, 27, 72. [Google Scholar] [CrossRef] [PubMed]
  12. Jyoti, J.L.; Shelton, A.M.; Taylor, A.G. Film-coating seeds with chlorpyrifos for germination and control of cabbage maggot (Diptera:Anthomyiidae) on cabbage transplants. J. Entomol. Sci. 2003, 38, 553–565. [Google Scholar] [CrossRef]
  13. Bundy, C.S.; Lewis, B. Impact of chlorpyrifos for pink bollworm control on secondary pests and beneficials. Southwest. Entomol. 2005, 30, 105–111. [Google Scholar]
  14. Goh, W.L.; Yiu, P.H.; Wong, S.C.; Rajan, A. Safe use of chlorpyrifos for insect pest management in leaf mustard (Brassica juncea L. Coss.). J. Food Agric. Env. 2011, 9, 1064–1066. [Google Scholar]
  15. Su, H.H.; Jiang, L.L.; Wang, H.T.; Yang, T.Z.; Harvey-Samuel, T.; Zhou, Q.X.; Lü, Y.B.; Yang, Y.Z. Sublethal effects of Bt toxin and chlorpyrifos on various Spodoptera exigua populations. Entomol. Exp. Appl. 2015, 157, 214–219. [Google Scholar] [CrossRef]
  16. Hites, R.A. The Rise and Fall of Chlorpyrifos in the United States. Environ. Sci. Technol. 2021, 55, 1354–1358. [Google Scholar] [CrossRef]
  17. John, E.M.; Shaike, J.M. Chlorpyrifos: Pollution and remediation. Environ. Chem. Lett. 2015, 13, 269–291. [Google Scholar] [CrossRef]
  18. Wang, L.; Liu, Z.; Zhang, J.J.; Wu, Y.H.; Sun, H.W. Chlorpyrifos exposure in farmers and urban adults: Metabolic characteristic, exposure estimation, and potential effect of oxidative damage. Environ. Res. 2016, 149, 164–170. [Google Scholar] [CrossRef]
  19. Rahman, H.U.U.; Asghar, W.; Nazir, W.; Sandhu, M.A.; Ahmed, A.; Khalid, N. A comprehensive review on chlorpyrifos toxicity with special reference to endocrine disruption: Evidence of mechanisms, exposures and mitigation strategies. Sci. Total Environ. 2021, 755, 142649. [Google Scholar] [CrossRef]
  20. Tudi, M.; Yang, L.S.; Wang, L.; Lv, J.; Gu, L.J.; Li, H.R.; Peng, W.; Yu, Q.M.; Ruan, H.D.; Li, Q.; et al. Environmental and Human Health Hazards from Chlorpyrifos, Pymetrozine and Avermectin Application in China under a Climate Change Scenario: A Comprehensive Review. Agriculture 2023, 13, 1683. [Google Scholar] [CrossRef]
  21. Guo, J.Q.; Zhang, J.M.; Wu, C.H.; Lv, S.L.; Lu, D.S.; Qi, X.J.; Jiang, S.; Feng, C.; Yu, H.X.; Liang, W.J.; et al. Associations of prenatal and childhood chlorpyrifos exposure with Neurodevelopment of 3-year-old children. Environ. Pollut. 2019, 251, 538–546. [Google Scholar] [CrossRef]
  22. Rauh, V.A.; Garfinkel, R.; Perera, F.P.; Andrews, H.F.; Hoepner, L.; Barr, D.B.; Whitehead, R.; Tang, D.; Whyatt, R.W. Impact of prenatal chlorpyrifos exposure on neurodevelopment in the first 3 years of life among inner-city children. Pediatrics 2006, 118, E1845–E1859. [Google Scholar] [CrossRef]
  23. Lee, J.E.; Kim, M.K.; Lee, J.Y.; Lee, Y.M.; Zoh, K.D. Degradation kinetics and pathway of 1H-benzotriazole during UV/chlorination process. Chem. Eng. J. 2019, 359, 1502–1508. [Google Scholar] [CrossRef]
  24. Mansor, N.A.; Tay, K.S. Degradation of 5,5-diphenylhydantoin by chlorination and UV/chlorination: Kinetics, transformation by-products, and toxicity assessment. Environ. Sci. Pollut. R 2017, 24, 22361–22370. [Google Scholar] [CrossRef] [PubMed]
  25. Qin, L.; Lin, Y.L.; Xu, B.; Hu, C.Y.; Tian, F.X.; Zhang, T.Y.; Zhu, W.Q.; Huang, H.; Gao, N.Y. Kinetic models and pathways of ronidazole degradation by chlorination, UV irradiation and UV/chlorine processes. Water Res. 2014, 65, 271–281. [Google Scholar] [CrossRef] [PubMed]
  26. Yang, S.M.; Feng, Y.; Gao, D.; Wang, X.W.; Suo, N.; Yu, Y.Z.; Zhang, S.B. Electrocatalysis degradation of tetracycline in a three-dimensional aeration electrocatalysis reactor (3D-AER) with a flotation-tailings particle electrode (FPE): Physicochemical properties, influencing factors and the degradation mechanism. J. Hazard. Mater. 2021, 407, 18. [Google Scholar] [CrossRef]
  27. Wang, Y.L.; Cai, N.C.; Huo, Y.D.; Chen, H. Electrochemical oxidation for the degradation of aniline on the SnO2/Ti anode. Acta Phys. Chim. Sin. 2001, 17, 609–613. [Google Scholar] [CrossRef]
  28. Zhou, M.H.; Wu, Z.C.; Wang, D.H. A novel electrocatalysis method for organic pollutants degradation. Chin. Chem. Lett. 2001, 12, 929–932. [Google Scholar]
  29. Zhang, M.H.; Liu, G.H.; Song, K.; Wang, Z.Y.; Zhao, Q.L.; Li, S.J.; Ye, Z.F. Biological treatment of 2,4,6-trinitrotoluene (TNT) red water by immobilized anaerobic-aerobic microbial filters. Chem. Eng. J. 2015, 259, 876–884. [Google Scholar] [CrossRef]
  30. Ji, Q.H.; Tabassum, S.; Yu, G.X.; Chu, C.F.; Zhang, Z.J. Determination of biological removal of recalcitrant organic contaminants in coal gasification waste water. Environ. Technol. 2015, 36, 2815–2824. [Google Scholar] [CrossRef]
  31. Jiang, Y.; Wang, H.Y.; Shang, Y.; Yang, K. Simultaneous removal of aniline, nitrogen and phosphorus in aniline-containing wastewater treatment by using sequencing batch reactor. Bioresour. Technol. 2016, 207, 422–429. [Google Scholar] [CrossRef]
  32. Goudarzi, M.; Abdulhusain, Z.H.; Salavati-Niasari, M. Low-cost and eco-friendly synthesis of Mn-doped T2WO4 nanostructures for efficient visible light photocatalytic degradation of antibiotics in water. Sol. Energy 2023, 262, 111912. [Google Scholar] [CrossRef]
  33. Mehdizadeh, P.; Jamdar, M.; Mahdi, M.A.; Abdulsahib, W.K.; Jasim, L.S.; Yousefi, S.R.; Salavati-Niasari, M. Rapid microwave fabrication of new nanocomposites based on Tb-Co-O nanostructures and their application as photocatalysts under UV/Visible light for removal of organic pollutants in water. Arab. J. Chem. 2023, 16, 104579. [Google Scholar] [CrossRef]
  34. Hafeez, A.; Shezad, N.; Javed, F.; Fazal, T.; Rehman, M.S.U.; Rehman, F. Synergetic effect of packed-bed corona-DBD plasma micro-reactor and photocatalysis for organic pollutant degradation. Sep. Purif. Technol. 2021, 269, 118728. [Google Scholar] [CrossRef]
  35. Mmelesi, O.K.; Masunga, N.; Kuvarega, A.; Nkambule, T.T.; Mamba, B.B.; Kefeni, K.K. Cobalt ferrite nanoparticles and nanocomposites: Photocatalytic, antimicrobial activity and toxicity in water treatment. Mat. Sci. Semicon. Proc. 2021, 123, 105523. [Google Scholar] [CrossRef]
  36. Masunga, N.; Mamba, B.B.; Kefeni, K.K. Trace samarium doped graphitic carbon nitride photocatalytic activity toward metanil yellow dye degradation under visible light irradiation. Colloid. Surface A 2020, 602, 125107. [Google Scholar] [CrossRef]
  37. Rostami, M.; Jourshabani, M.; Davar, F.; Ziarani, G.M.; Badiei, A. Black TiO2-g-C3N4 heterojunction coupled with MOF (ZIF-67)-derived Co3S4/nitrogen-doped carbon hollow nanobox. J. Am. Ceram. Soc. 2023, 107, 719–735. [Google Scholar] [CrossRef]
  38. Ye, K.H.; Li, Y.; Yang, H.; Li, M.Y.; Huang, Y.C.; Zhang, S.Q.; Ji, H.B. An ultrathin carbon layer activated CeO2 heterojunction nanorods for photocatalytic degradation of organic pollutants. Appl. Catal. B Environ. 2019, 259, 118085. [Google Scholar] [CrossRef]
  39. Yang, X.F.; Qin, J.L.; Jiang, Y.; Chen, K.M.; Yan, X.H.; Zhang, D.; Li, R.; Tang, H. Fabrication of P25/Ag3PO4/graphene oxide heterostructures for enhanced solar photocatalytic degradation of organic pollutants and bacteria. Appl. Catal. B-Environ. 2015, 166, 231–240. [Google Scholar] [CrossRef]
  40. Koe, W.S.; Lee, J.W.; Chong, W.C.; Pang, Y.L.; Sim, L.C. An overview of photocatalytic degradation: Photocatalysts, mechanisms, and development of photocatalytic membrane. Environ. Sci. Pollut. R 2020, 27, 2522–2565. [Google Scholar] [CrossRef] [PubMed]
  41. Wen, X.J.; Shen, C.H.; Fei, Z.H.; Fang, D.; Liu, Z.T.; Dai, J.T.; Niu, C.G. Recent developments on AgI based heterojunction photocatalytic systems in photocatalytic application. Chem. Eng. J. 2020, 383, 21. [Google Scholar] [CrossRef]
  42. Luan, J.F.; Niu, B.W.; Ma, B.B.; Yang, G.M.; Liu, W.L. Preparation and Property Characterization of In2YSbO7/BiSnSbO6 Heterojunction Photocatalyst toward Photocatalytic Degradation of Indigo Carmine within Dye Wastewater under Visible-Light Irradiation. Materials 2022, 15, 6648. [Google Scholar] [CrossRef] [PubMed]
  43. Soleimani, M.; Ghasemi, J.B.; Ziarani, G.M.; Karimi-Maleh, H.; Badiei, A. Photocatalytic degradation of organic pollutants, viral and bacterial pathogens using titania nanoparticles. Inorg. Chem. Commun. 2021, 130, 108688. [Google Scholar] [CrossRef]
  44. Zhang, Q.H.; Gao, L.; Guo, J.K. Photocatalytic activity of nanosized TiO2. J. Inorg. Mater. 2000, 15, 556–560. [Google Scholar]
  45. Wold, A. Photocatalytic Properties of TIO2. Chem. Mater. 1993, 5, 280–283. [Google Scholar] [CrossRef]
  46. Yeber, M.C.; Rodríguez, J.; Freer, J.; Durán, N.; Mansilla, H.D. Photocatalytic degradation of cellulose bleaching effluent by supported TiO2 and ZnO. Chemosphere 2000, 41, 1193–1197. [Google Scholar] [CrossRef] [PubMed]
  47. Villasenor, J.; Reyes, P.; Pecchi, G. Photodegradation of pentachlorophenol on ZnO. J. Chem. Technol. Biotechnol. 1998, 72, 105–110. [Google Scholar] [CrossRef]
  48. Dong, W.X.; Bao, Q.F.; Gu, X.Y.; Peng, G.; Zhao, X.G. C3N4/CaTi2O5 Composite: Synthesis and Photocatalytic Properties. Chinese. J.Org. Chem. 2017, 33, 292–298. [Google Scholar] [CrossRef]
  49. Li, J.H.; Han, M.S.; Guo, Y.; Wang, F.; Sun, C. Fabrication of FeVO4/Fe2TiO5 composite catalyst and photocatalytic removal of norfloxacin. Chem. Eng. J. 2016, 298, 300–308. [Google Scholar] [CrossRef]
  50. Issarapanacheewin, S.; Wetchakun, K.; Phanichphant, S.; Kangwansupamonkon, W.; Wetchakun, N. A novel CeO2/Bi2WO6 composite with highly enhanced photocatalytic activity. Mater. Lett. 2015, 156, 28–31. [Google Scholar] [CrossRef]
  51. Hunge, Y.M.; Yadav, A.A.; Kang, S.W. Photocatalytic Degradation of Eriochrome Black-T Using BaWO4/MoS2 Composite. Catalysts 2022, 12, 11. [Google Scholar] [CrossRef]
  52. Zou, Z.G.; Ye, J.H.; Abe, R.; Arakawa, H. Photocatalytic decomposition of water with Bi2InNbO7. Catal. Lett. 2000, 68, 235–239. [Google Scholar] [CrossRef]
  53. Cui, Y.B.; Luan, J.F. Synthesis, crystal structure, photodegradation kinetics and photocatalytic activity of novel photocatalyst ZnBiYO4. J. Environ. Sci. 2015, 29, 51–61. [Google Scholar] [CrossRef]
  54. Ajmal, A.; Majeed, I.; Malik, R.N.; Idriss, H.; Nadeem, M.A. Principles and mechanisms of photocatalytic dye degradation on TiO2 based photocatalysts: A comparative overview. RSC. Adv. 2014, 4, 37003–37026. [Google Scholar] [CrossRef]
  55. Tu, W.G.; Zhou, Y.; Zou, Z.G. Versatile Graphene-Promoting Photocatalytic Performance of Semiconductors: Basic Principles, Synthesis, Solar Energy Conversion, and Environmental Applications. Adv. Funct. Mater. 2013, 23, 4996–5008. [Google Scholar] [CrossRef]
  56. Wang, S.B.; Wang, F.F.; Su, Z.M.; Wang, X.N.; Han, Y.C.; Zhang, L.; Xiang, J.; Du, W.; Tang, N. Controllable Fabrication of Heterogeneous p-TiO2 QDs@g-C3N4 p-n Junction for Efficient Photocatalysis. Catalysts 2019, 9, 439. [Google Scholar] [CrossRef]
  57. Wang, W.Z.; Huang, X.W.; Wu, S.; Zhou, Y.X.; Wang, L.J.; Shi, H.L.; Liang, Y.J.; Zou, B. Preparation of p-n junction Cu2O/BiVO4 heterogeneous nanostructures with enhanced visible-light photocatalytic activity. Appl. Catal. B-Environ. 2013, 134, 293–301. [Google Scholar] [CrossRef]
  58. Xu, H.; Xu, Y.G.; Li, H.M.; Xia, J.X.; Xiong, J.; Yin, S.; Huang, C.J.; Wan, H.L. Synthesis, characterization and photocatalytic property of AgBr/BiPO4 heterojunction photocatalyst. Dalton Trans. 2012, 41, 3387–3394. [Google Scholar] [CrossRef]
  59. Cheng, T.T.; Gao, H.J.; Sun, X.F.; Xian, T.; Wang, S.F.; Yi, Z.; Liu, G.R.; Wang, X.X.; Yang, H. An excellent Z-scheme Ag2MoO4/Bi4Ti3O12heterojunction photocatalyst: Construction strategy and application in environmental purification. Adv. Powder Technol. 2021, 32, 951–962. [Google Scholar] [CrossRef]
  60. Fattah, W.I.A.; Gobara, M.M.; El-Hotaby, W.; Mostafa, S.F.M.; Ali, G.W. Coating stainless steel plates with Ag/TiO2 for chlorpyrifos decontamination. Mater. Res. Express. 2016, 3, 055009. [Google Scholar] [CrossRef]
  61. Pathania, D.; Sharma, A.; Kumar, S.; Srivastava, A.K.; Kumar, A.; Singh, L. Bio-synthesized Cu-ZnO hetro-nanostructure for catalytic degradation of organophosphate chlorpyrifos under solar illumination. Chemosphere 2021, 277, 130315. [Google Scholar] [CrossRef]
  62. Luan, J.F.; Wei, Z.J.; Niu, B.W.; Yang, G.M.; Huang, C.S.; Ma, B.B.; Liu, W.L. Synthesis, Property Characterization and Photocatalytic Activity of the Ag3PO4/Gd2BiTaO7 Heterojunction Catalyst under Visible Light Irradiation. Catalysts 2022, 12, 22. [Google Scholar] [CrossRef]
  63. Wang, J.H.; Zou, Z.G.; Ye, J.H. Synthesis, structure and photocatalytic property of a new hydrogen evolving photocatalyst Bi2InTaO7. In Functionally Graded Materials Vii; Pan, W., Gong, J., Zhang, L., Chen, L., Eds.; Trans Tech Publications Ltd.: Zurich-Uetikon, Switzerland, 2003; Volume 423–424, pp. 485–490. [Google Scholar]
  64. 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]
  65. 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]
  66. Wiegel, M.; Middel, W.; Blasse, G. Influence of ns(2) ions on the luminescence of niobates and tantalates. J. Mater. Chem. 1995, 5, 981–983. [Google Scholar] [CrossRef]
  67. Luan, J.F.; Ma, B.B.; Yao, Y.; Liu, W.L.; Niu, B.W.; Yang, G.M.; Wei, Z.J. Synthesis, Performance Measurement of Bi2SmSbO7/ZnBiYO4 Heterojunction Photocatalyst and Photocatalytic Degradation of Direct Orange within Dye Wastewater under Visible Light Irradiation. Materials 2022, 15, 3986. [Google Scholar] [CrossRef] [PubMed]
  68. Monshi, A.; Foroughi, M.R.; Monshi, M.R. Modified Scherrer Equation to Estimate More Accurately Nano-Crystallite Size Using XRD. World J. Nano Sci. Eng. 2012, 2, 154–160. [Google Scholar] [CrossRef]
  69. Nasiri, S.; Rabiei, M.; Palevicius, A.; Janusas, G.; Vilkauskas, A.; Nutalapati, V.; Monshi, A. Modified Scherrer equation to calculate crystal size by XRD with high accuracy, examples Fe2O3, TiO2 and V2O5. Nano Trends 2023, 3, 100015. [Google Scholar] [CrossRef]
  70. Burton, A.W.; Ong, K.; Rea, T.; Chan, I.Y. On the estimation of average crystallite size of zeolites from the Scherrer equation: A critical evaluation of its application to zeolites with one-dimensional pore systems. Microporous Mesoporous Mater. 2009, 117, 75–90. [Google Scholar] [CrossRef]
  71. Shaishta, N.; Khan, W.U.; Mane, S.K.B.; Hayat, A.; Zhou, D.D.; Khan, J.; Mehmood, N.; Inamdar, H.K.; Manjunatha, G. Red-emitting CaSc2O4:Eu3+ phosphor for NUV-based warm white LEDs: Structural elucidation and Hirshfeld surface analysis. Int. J. Energy Res. 2020, 44, 8328–8339. [Google Scholar] [CrossRef]
  72. Goh, K.H.; Haseeb, 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]
  73. Kaviyarasu, K.; Sajan, D.; Devarajan, P.A. A rapid and versatile method for solvothermal synthesis of Sb2O3 nanocrystals under mild conditions. Appl. Nanosci. 2013, 3, 529–533. [Google Scholar] [CrossRef]
  74. 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]
  75. Achehboune, M.; Khenfouch, M.; Boukhoubza, I.; Leontie, L.; Doroftei, C.; Carlescu, A.; Bulai, G.; Mothudi, B.; Zorkani, I.; Jorio, A. Microstructural, FTIR and Raman spectroscopic study of Rare earth doped ZnO nanostructures. In Proceedings of the Nanosmat Mediterrane Conference, Fac Sci Rabat, Rabat, Morocco, 2 May–4 June 2009; pp. 319–323. [Google Scholar]
  76. 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]
  77. 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, 16. [Google Scholar] [CrossRef]
  78. Karimi, B.; 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]
  79. Berg, R.W.; Thorup, N. The reaction between ZnO and molten Na2S2O7 or K2S2O7 forming Na2Zn(SO4)2 or K2Zn(SO4)2, studied by Raman spectroscopy and X-ray diffraction. Inorg. Chem. 2005, 44, 3485–3493. [Google Scholar] [CrossRef] [PubMed]
  80. Tsaryuk, V.I.; Zhuravlev, K.P.; Szostak, R.; Vologzhanina, A.V. Structure, Luminescence, and Raman Spectroscopy of Europium and Terbium Dipivaloylmethanates and Other β-Diketonates with 2,2′-Bipyridine. J. Struct. Chem. 2020, 61, 1026–1037. [Google Scholar] [CrossRef]
  81. 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]
  82. Hadjiev, V.G.; Iliev, M.N.; Sasmal, K.; Sun, Y.Y.; Chu, C.W. Raman spectroscopy of R FeAsO (R = Sm, La). Phys. Rev. B 2008, 77, 3. [Google Scholar] [CrossRef]
  83. 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]
  84. 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 Part A Mol. Biomol. Spectrosc. 2004, 60, 425–434. [Google Scholar] [CrossRef] [PubMed]
  85. Luan, J.; Hao, L.; Yao, Y.; Wang, Y.; Yang, G.; Li, J. Preparation and Property Characterization of Sm2EuSbO7/ZnBiSbO5 Heterojunction Photocatalyst for Photodegradation of Parathion Methyl under Visible Light Irradiation. Molecules 2023, 28, 7722. [Google Scholar] [CrossRef] [PubMed]
  86. Uwamino, Y.; Ishizuka, T.; Yamatera, H. X-ray photoelectron-spectroscopy of rare-earth compounds. J. Electron Spectrosc. Relat. Phenom. 1984, 34, 67–78. [Google Scholar] [CrossRef]
  87. Izquierdo, R.; Sacher, E.; Yelon, A. X-ray photoelectron-spectra of antimony oxides. Appl. Surf. Sci. 1989, 40, 175–177. [Google Scholar] [CrossRef]
  88. Birchall, T.; Connor, J.A.; Hillier, I.H. High-energy photoelectron-spectroscopy of some antimony compounds. J. Chem. Soc. Dalton Trans. 1975, 2003–2006. [Google Scholar] [CrossRef]
  89. Benvenutti, E.V.; Gushikem, Y.; Vasquez, A.; Decastro, S.C.; Zaldivar, G.A.P. X-ray photoelectron-spectroscopy and mossbauer-spectroscopy study of iron(iii) and antimony(v) oxides grafted onto a silica-gel surface. J. Chem. Soc. Chem. Commun. 1991, 1325–1327. [Google Scholar] [CrossRef]
  90. Dake, L.S.; Baer, D.R.; Zachara, J.M. Auger parameter measurements of zinc-compounds relevant to zinc transport in the environment. Surf. Interface Anal. 1989, 14, 71–75. [Google Scholar] [CrossRef]
  91. Morgan, W.E.; Stec, W.J.; Van Wazer, J.R. Inner-orbital binding-energy shifts of antimony and bismuth compounds. Inorg. Chem. 1973, 12, 953–955. [Google Scholar] [CrossRef]
  92. Ettema, A.; Haas, C. AN X-ray photoemission spectroscopy study of interlayer charge-transfer in some misfit layer compounds. J. Phys.-Condens. Matter 1993, 5, 3817–3826. [Google Scholar] [CrossRef]
  93. Wagner, C.D.; Zatko, D.A.; Raymond, R.H. Use of the oxygen kll auger lines in identification of surface chemical-states by electron-spectroscopy for chemical-analysis. Anal. Chem. 1980, 52, 1445–1451. [Google Scholar] [CrossRef]
  94. Nowak, M.; Kauch, B.; Szperlich, P. Determination of energy band gap of nanocrystalline SbSI using diffuse reflectance spectroscopy. Rev. Sci. Instrum. 2009, 80, 3. [Google Scholar] [CrossRef]
  95. Zhou, F.; Kang, K.S.; 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]
  96. Butler, M.A.; Ginley, D.S.; Eibschutz, M. Photoelectrolysis with YFeO3 electrodes. J. Appl. Phys. 1977, 48, 3070–3072. [Google Scholar] [CrossRef]
  97. Tauc, J.; Grigorovici, R.; Vancu, A.J.p.s.s. Optical properties and electronic structure of amorphous germanium. Phys. Status Solidi. 1966, 15, 627–637. [Google Scholar] [CrossRef]
  98. Makula, P.; Pacia, M.; Macyk, W. How To Correctly Determine the Band Gap Energy of Modified Semiconductor Photocatalysts Based on UV-Vis Spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [Google Scholar] [CrossRef] [PubMed]
  99. Cheng, X.W.; Yu, X.J.; Xing, Z.P.; Yang, L.S. Synthesis and characterization of N-doped TiO2 and its enhanced visible-light photocatalytic activity. Arab. J. Chem. 2016, 9, S1706–S1711. [Google Scholar] [CrossRef]
  100. Ananpattarachai, J.; Kajitvichyanukul, P.; Seraphin, S. Visible light absorption ability and photocatalytic oxidation activity of various interstitial N-doped TiO2 prepared from different nitrogen dopants. J. Hazard. Mater. 2009, 168, 253–261. [Google Scholar] [CrossRef] [PubMed]
  101. Kanmoni, V.G.G.; Daniel, S.; Raj, G.A.G. Photocatalytic degradation of chlorpyrifos in aqueous suspensions using nanocrystals of ZnO and TiO2. React. Kinet. Mech. Catal. 2012, 106, 325–339. [Google Scholar] [CrossRef]
  102. Poonam, V.; Poonam, D.D. Photocatalytic degradability of insecticide Chlorpyrifos over UV irradiated Titanium dioxide in aqueous phase. Int. J. Environ. Sci. 2012, 3, 743–755. [Google Scholar]
  103. Baharvand, A.; Ali, R.; Yusof, A.M.; Ibrahim, A.N.; Chandren, S.; Nur, H. Preparation of Anatase Hollow TiO2 Spheres and Their Photocatalytic Activity in the Photodegradation of Chlorpyrifos. J. Chin. Chem. Soc. 2014, 61, 1211–1216. [Google Scholar] [CrossRef]
  104. Majhi, D.; Bhoi, Y.P.; Samal, P.K.; Mishra, B.G. Morphology controlled synthesis and photocatalytic study of novel CuS-Bi2O2CO3 heterojunction system for chlorpyrifos degradation under visible light illumination. Appl. Surf. Sci. 2018, 455, 891–902. [Google Scholar] [CrossRef]
  105. Nekooie, R.; Shamspur, T.; Mostafavi, A. Novel CuO/TiO2/PANI nanocomposite: Preparation and photocatalytic investigation for chlorpyrifos degradation in water under visible light irradiation. J. Photochem. Photobiol. A Chem. 2021, 407, 113038. [Google Scholar] [CrossRef]
  106. Esfandian, H.; Cherati, M.R.; Khatirian, M. Electrochemical behavior and photocatalytic performance of chlorpyrifos pesticide decontamination using Ni-doped ZnO-TiO2 nanocomposite. Inorg. Chem. Commun. 2024, 159, 111750. [Google Scholar] [CrossRef]
  107. Balakrishnan, G.; Velavan, R.; Batoo, K.M.; Raslan, E.H. Microstructure, optical and photocatalytic properties of MgO nanoparticles. Results. Phys. 2020, 16, 4. [Google Scholar] [CrossRef]
  108. Ali, T.; Tripathi, P.; Azam, A.; Raza, W.; Ahmed, A.S.; Ahmed, A.; Muneer, M. Photocatalytic performance of Fe-doped TiO2 nanoparticles under visible-light irradiation. Mater. Res. Express. 2017, 4, 12. [Google Scholar] [CrossRef]
  109. 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]
  110. 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]
  111. 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-Environ. 2018, 227, 376–385. [Google Scholar] [CrossRef]
  112. Franco-Perez, M.; Gazquez, J.L. Electronegativities of Pauling and Mulliken in Density Functional Theory. J. Phys. Chem. A 2019, 123, 10065–10071. [Google Scholar] [CrossRef]
  113. Mousavi, M.; Habibi-Yangjeh, A.; Abitorabi, M. Fabrication of novel magnetically separable nanocomposites using graphitic carbon nitride, silver phosphate and silver chloride and their applications in photocatalytic removal of different pollutants using visible-light irradiation. J. Colloid Interface Sci. 2016, 480, 218–231. [Google Scholar] [CrossRef] [PubMed]
  114. Cao, W.; Jiang, C.Y.; Chen, C.; Zhou, H.F.; Wang, Y.P. A novel Z-scheme CdS/Bi4O5Br2 heterostructure with mechanism analysis: Enhanced photocatalytic performance. J. Alloys Compd. 2021, 861, 12. [Google Scholar] [CrossRef]
  115. Xu, S.; Gong, S.Q.; Jiang, H.; Shi, P.H.; Fan, J.C.; Xu, Q.J.; Min, Y.L. Z-scheme heterojunction through interface engineering for broad spectrum photocatalytic water splitting. Appl. Catal. B-Environ. 2020, 267, 12. [Google Scholar] [CrossRef]
Figure 1. The XRD patterns of (a) ZnBiEuO4, (b) Eu2SmSbO7, (c) EZHP (red circle: the crystal plane of ZnBiEuO4; blue diamond: the crystal plane of Eu2SmSbO7).
Figure 1. The XRD patterns of (a) ZnBiEuO4, (b) Eu2SmSbO7, (c) EZHP (red circle: the crystal plane of ZnBiEuO4; blue diamond: the crystal plane of Eu2SmSbO7).
Catalysts 14 00144 g001
Figure 2. (a) XRD and Rietveld refinement of Eu2SmSbO7: Simulated XRD data (blue solid line), experimental XRD data (red dotted line), difference between experimental XRD data and simulated XRD data (black solid line), and observed diffraction peaks (green perpendicular lines). (b) Atom construction of Eu2SmSbO7. (Red atom: O, green atom: Eu, purple atom: Sm or Sb).
Figure 2. (a) XRD and Rietveld refinement of Eu2SmSbO7: Simulated XRD data (blue solid line), experimental XRD data (red dotted line), difference between experimental XRD data and simulated XRD data (black solid line), and observed diffraction peaks (green perpendicular lines). (b) Atom construction of Eu2SmSbO7. (Red atom: O, green atom: Eu, purple atom: Sm or Sb).
Catalysts 14 00144 g002
Figure 3. (a) XRD and Rietveld refinement of ZnBiEuO4: simulated XRD data (blue solid line), experimental XRD data (red dotted line), difference between experimental and simulated XRD data (black solid line), and observed diffraction peaks (green perpendicular lines). (b) Atom construction of ZnBiEuO4. (Red atom: O, purple atom: Zn, green atomy: Bi or Eu.).
Figure 3. (a) XRD and Rietveld refinement of ZnBiEuO4: simulated XRD data (blue solid line), experimental XRD data (red dotted line), difference between experimental and simulated XRD data (black solid line), and observed diffraction peaks (green perpendicular lines). (b) Atom construction of ZnBiEuO4. (Red atom: O, purple atom: Zn, green atomy: Bi or Eu.).
Catalysts 14 00144 g003
Figure 4. FTIR spectra of Eu2SmSbO7, EZHP, and ZnBiEuO4.
Figure 4. FTIR spectra of Eu2SmSbO7, EZHP, and ZnBiEuO4.
Catalysts 14 00144 g004
Figure 5. Raman spectra of (a) ZnBiEuO4, (b) Eu2SmSbO7, and (c) EZHP.
Figure 5. Raman spectra of (a) ZnBiEuO4, (b) Eu2SmSbO7, and (c) EZHP.
Catalysts 14 00144 g005
Figure 6. XPS spectra of Eu2SmSbO7, EZHP, and ZnBiEuO4.
Figure 6. XPS spectra of Eu2SmSbO7, EZHP, and ZnBiEuO4.
Catalysts 14 00144 g006
Figure 7. XPS spectrum of (a) Sm3+, (b) Sb5+, (c) Eu3+, (d) Zn2+, and (e) Bi3+ for EZHP, Eu2SmSbO7, and ZnBiEuO4, and (f) XPS spectrum of O2− in EZHP.
Figure 7. XPS spectrum of (a) Sm3+, (b) Sb5+, (c) Eu3+, (d) Zn2+, and (e) Bi3+ for EZHP, Eu2SmSbO7, and ZnBiEuO4, and (f) XPS spectrum of O2− in EZHP.
Catalysts 14 00144 g007aCatalysts 14 00144 g007b
Figure 8. (a) The UV-Vis diffuse reflectance spectrum of Eu2SmSbO7; (b) plots of (αhν) 2 and for Eu2SmSbO7 (black dashed line: the linear fit of the correlative diagram; red dashed line: base line).
Figure 8. (a) The UV-Vis diffuse reflectance spectrum of Eu2SmSbO7; (b) plots of (αhν) 2 and for Eu2SmSbO7 (black dashed line: the linear fit of the correlative diagram; red dashed line: base line).
Catalysts 14 00144 g008
Figure 9. (a) The UV-Vis diffuse reflectance spectrum of ZnBiEuO4; (b) plots of (αhν) 1/2 and for ZnBiEuO4 (black dashed line: the linear fit of the correlative diagram; red dashed line: base line).
Figure 9. (a) The UV-Vis diffuse reflectance spectrum of ZnBiEuO4; (b) plots of (αhν) 1/2 and for ZnBiEuO4 (black dashed line: the linear fit of the correlative diagram; red dashed line: base line).
Catalysts 14 00144 g009
Figure 10. (a) The UV-Vis diffuse reflectance spectrum of EZHP; (b) plots of (αhν) 1/2 and for EZHP (black dashed line: the linear fit of the correlative diagram; red dashed line: base line).
Figure 10. (a) The UV-Vis diffuse reflectance spectrum of EZHP; (b) plots of (αhν) 1/2 and for EZHP (black dashed line: the linear fit of the correlative diagram; red dashed line: base line).
Catalysts 14 00144 g010
Figure 11. TEM image of Eu2SmSbO7.
Figure 11. TEM image of Eu2SmSbO7.
Catalysts 14 00144 g011
Figure 12. TEM image of ZnBiEuO4.
Figure 12. TEM image of ZnBiEuO4.
Catalysts 14 00144 g012
Figure 13. TEM image of EZHP.
Figure 13. TEM image of EZHP.
Catalysts 14 00144 g013
Figure 14. EDS elemental mapping of EZHP (Eu, Sm, Sb, and O from Eu2SmSbO7, and Zn, Bi, Sb, and O from ZnBiEuO4).
Figure 14. EDS elemental mapping of EZHP (Eu, Sm, Sb, and O from Eu2SmSbO7, and Zn, Bi, Sb, and O from ZnBiEuO4).
Catalysts 14 00144 g014
Figure 15. The EDS spectrum of EZHP.
Figure 15. The EDS spectrum of EZHP.
Catalysts 14 00144 g015
Figure 16. Concentration variation curves of (a) chlorpyrifos and (b) TOC during the photodegradation of chlorpyrifos using EZHP, Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or without a catalyst under VLID.The conversion rates of chlorpyrifos could be calculated by ( 1 C C 0 ) × 100%, where C represents the instantaneous concentration of chlorpyrifos and C0 represents the initial concentration of chlorpyrifos. For a better comparison in all experiments, the VLID time was set to 160 min. By analyzing the data in Figure 16a, it can be seen that the conversion rate of chlorpyrifos during degradation of chlorpyrifos using EZHP was 100% after VLID of 160 min, with the reaction rate of 3.33 × 10−9 mol·L−1·s−1 and a photonic efficiency of 0.0700%. When Eu2SmSbO7 was used as the photocatalyst, the conversion rate of chlorpyrifos was 88.16%, with a reaction rate of 2.94 × 10−9 mol·L−1·s−1 and a photonic efficiency of 0.0617%. The conversion rate of chlorpyrifos that was derived from pesticide wastewater with ZnBiEuO4 as the photocatalyst was 84.03%; at the same time, the reaction rate of 2.80 × 10−9 mol·L−1·s−1 and the photonic efficiency of 0.0588% were achieved. When N-doped TiO2 was used as the photocatalyst, the conversion rate of chlorpyrifos was 35.25%, with a reaction rate of 1.18 × 10−9 mol·L−1·s−1 and a photonic efficiency of 0.0247%. Finally, when chlorpyrifos was photodegraded directly under VLID without the addition of any catalysts, the conversion rate of chlorpyrifos was 4.72%, with a reaction rate of 0.16 × 10−9 mol·L−1·s−1 and a photonic efficiency of 0.0033%.
Figure 16. Concentration variation curves of (a) chlorpyrifos and (b) TOC during the photodegradation of chlorpyrifos using EZHP, Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or without a catalyst under VLID.The conversion rates of chlorpyrifos could be calculated by ( 1 C C 0 ) × 100%, where C represents the instantaneous concentration of chlorpyrifos and C0 represents the initial concentration of chlorpyrifos. For a better comparison in all experiments, the VLID time was set to 160 min. By analyzing the data in Figure 16a, it can be seen that the conversion rate of chlorpyrifos during degradation of chlorpyrifos using EZHP was 100% after VLID of 160 min, with the reaction rate of 3.33 × 10−9 mol·L−1·s−1 and a photonic efficiency of 0.0700%. When Eu2SmSbO7 was used as the photocatalyst, the conversion rate of chlorpyrifos was 88.16%, with a reaction rate of 2.94 × 10−9 mol·L−1·s−1 and a photonic efficiency of 0.0617%. The conversion rate of chlorpyrifos that was derived from pesticide wastewater with ZnBiEuO4 as the photocatalyst was 84.03%; at the same time, the reaction rate of 2.80 × 10−9 mol·L−1·s−1 and the photonic efficiency of 0.0588% were achieved. When N-doped TiO2 was used as the photocatalyst, the conversion rate of chlorpyrifos was 35.25%, with a reaction rate of 1.18 × 10−9 mol·L−1·s−1 and a photonic efficiency of 0.0247%. Finally, when chlorpyrifos was photodegraded directly under VLID without the addition of any catalysts, the conversion rate of chlorpyrifos was 4.72%, with a reaction rate of 0.16 × 10−9 mol·L−1·s−1 and a photonic efficiency of 0.0033%.
Catalysts 14 00144 g016
Figure 17. Effect of different photocatalyst doses on degradation of chlorpyrifos using EZHP under VLID.
Figure 17. Effect of different photocatalyst doses on degradation of chlorpyrifos using EZHP under VLID.
Catalysts 14 00144 g017
Figure 18. Concentration variation curves of (a) chlorpyrifos and (b) TOC during four cyclic degradation tests of chlorpyrifos that was derived from pesticide wastewater under VLID using EZHP.
Figure 18. Concentration variation curves of (a) chlorpyrifos and (b) TOC during four cyclic degradation tests of chlorpyrifos that was derived from pesticide wastewater under VLID using EZHP.
Catalysts 14 00144 g018
Figure 19. First-order kinetic plots of (a) chlorpyrifos and (b) TOC that were observed during photodegradation of chlorpyrifos under VLID using EZHP, Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or without a catalyst.
Figure 19. First-order kinetic plots of (a) chlorpyrifos and (b) TOC that were observed during photodegradation of chlorpyrifos under VLID using EZHP, Eu2SmSbO7, ZnBiEuO4, N-doped TiO2, or without a catalyst.
Catalysts 14 00144 g019
Figure 20. First-order kinetic plots of (a) chlorpyrifos and (b) TOC that were observed during four cyclic photodegradation experiments for chlorpyrifos under VLID using EZHP.
Figure 20. First-order kinetic plots of (a) chlorpyrifos and (b) TOC that were observed during four cyclic photodegradation experiments for chlorpyrifos under VLID using EZHP.
Catalysts 14 00144 g020
Figure 21. The XRD pattern of EZHP before photocatalytic degradation reaction of chlorpyrifos (red line), and the XRD pattern of EZHP after photocatalytic degradation reaction of chlorpyrifos (blue line).
Figure 21. The XRD pattern of EZHP before photocatalytic degradation reaction of chlorpyrifos (red line), and the XRD pattern of EZHP after photocatalytic degradation reaction of chlorpyrifos (blue line).
Catalysts 14 00144 g021
Figure 22. The XPS spectrum of EZHP before photocatalytic degradation reaction of chlorpyrifos (red line), and the XPS spectrum of EZHP after photocatalytic degradation reaction of chlorpyrifos (blue line).
Figure 22. The XPS spectrum of EZHP before photocatalytic degradation reaction of chlorpyrifos (red line), and the XPS spectrum of EZHP after photocatalytic degradation reaction of chlorpyrifos (blue line).
Catalysts 14 00144 g022
Figure 23. The effect of different (a) pH values, (b) metal ion species, and (c) anion species on the degradation of chlorpyrifos using EZHP under VLID.
Figure 23. The effect of different (a) pH values, (b) metal ion species, and (c) anion species on the degradation of chlorpyrifos using EZHP under VLID.
Catalysts 14 00144 g023
Figure 24. (a) Effect of different scavengers such as BQ, IPA, or EDTA on removal efficiency of chlorpyrifos using EZHP under VLID; (b) effect of different scavengers on chlorpyrifos conversion rates during photocatalytic degradation of chlorpyrifos using EZHP under VLID.
Figure 24. (a) Effect of different scavengers such as BQ, IPA, or EDTA on removal efficiency of chlorpyrifos using EZHP under VLID; (b) effect of different scavengers on chlorpyrifos conversion rates during photocatalytic degradation of chlorpyrifos using EZHP under VLID.
Catalysts 14 00144 g024
Figure 25. (a) PL spectra of EZHP, Eu2SmSbO7, and ZnBiEuO4, and TRPL spectra of (b) Eu2SmSbO7, (c) ZnBiEuO4, (d) and EZHP (red line: curves generated by nonlinear curve fitting).
Figure 25. (a) PL spectra of EZHP, Eu2SmSbO7, and ZnBiEuO4, and TRPL spectra of (b) Eu2SmSbO7, (c) ZnBiEuO4, (d) and EZHP (red line: curves generated by nonlinear curve fitting).
Catalysts 14 00144 g025
Figure 26. Nyquist impedance plots of EZHP, Eu2SmSbO7 photocatalyst, and ZnBiEuO4 photocatalyst.
Figure 26. Nyquist impedance plots of EZHP, Eu2SmSbO7 photocatalyst, and ZnBiEuO4 photocatalyst.
Catalysts 14 00144 g026
Figure 27. Possible photocatalytic degradation mechanism of degradation using EZHP under VLID (1, 2, and 3: the three degradation pathways of chlorpyrifos).
Figure 27. Possible photocatalytic degradation mechanism of degradation using EZHP under VLID (1, 2, and 3: the three degradation pathways of chlorpyrifos).
Catalysts 14 00144 g027
Figure 28. Ultraviolet photoelectron spectrum (UPS) of (a) Eu2SmSbO7 and (b) ZnBiEuO4.
Figure 28. Ultraviolet photoelectron spectrum (UPS) of (a) Eu2SmSbO7 and (b) ZnBiEuO4.
Catalysts 14 00144 g028
Figure 29. Suggested photocatalytic pathway scheme for the degradation of chlorpyrifos under VLID with EZ heterojunction as photocatalyst.
Figure 29. Suggested photocatalytic pathway scheme for the degradation of chlorpyrifos under VLID with EZ heterojunction as photocatalyst.
Catalysts 14 00144 g029
Table 1. Structural parameters of Eu2SmSbO7.
Table 1. Structural parameters of Eu2SmSbO7.
AtomxyzOccupation Factor
Eu0001
Sm0.50.50.50.5
Sb0.50.50.50.5
O(1)−0.1850.1250.1251
O(2)0.1250.1250.1251
Table 2. Structural parameters of ZnBiEuO4.
Table 2. Structural parameters of ZnBiEuO4.
AtomxyzOccupation Factor
Zn000.51
Bi0001
Eu0001
O0.757310.140130.081881
Table 3. The effect of different ratios of catalyst to chlorpyrifos on conversion rates of chlorpyrifos.
Table 3. The effect of different ratios of catalyst to chlorpyrifos on conversion rates of chlorpyrifos.
PhotocatalystRatio of Catalyst to ChlorpyrifosConversion Rates of Chlorpyrifos (%)
Eu2SmSbO7/ZnBiEuO431.2572.56
40.1879.26
49.1182.52
58.0493.75
66.96100
75.8999.87
84.8296.54
Table 4. Comparison of the photocatalytic activity of EZHP with that of other reported photocatalysts during photocatalytic degradation of chlorpyrifos.
Table 4. Comparison of the photocatalytic activity of EZHP with that of other reported photocatalysts during photocatalytic degradation of chlorpyrifos.
PhotocatalystRadiationIrradiation Time (min)PesticideConversion Rates (%)Ref.
TiO2/H2O2UV300Chlorpyrifos70[102]
Hollow TiO2UV180Chlorpyrifos75.21[103]
Cu/ZnOSolar light240Chlorpyrifos81[61]
TiO2/H2O2Solar light300Chlorpyrifos83[102]
CuS/Bi2O2CO3Visible light180Chlorpyrifos>95[104]
CuO/TiO2Visible light90Chlorpyrifos60[105]
Ni-doped ZnO/TiO2Visible light140Chlorpyrifos75.5[106]
Eu2SmSbO7Visible light160Chlorpyrifos88.16This study
EZHPVisible light160Chlorpyrifos100This study
Table 5. Fitted results of TRPL curves of Eu2SmSbO7, ZnBiEuO4, and EZHP.
Table 5. Fitted results of TRPL curves of Eu2SmSbO7, ZnBiEuO4, and EZHP.
Eu2SmSbO7ZnBiEuO4Eu2SmSbO7/ZnBiEuO4
A 1 3.1295 × 10102.4041 × 101339.4378
τ 1 (ns)1.50941.16927.6479
A 2 0.04552.2526811.5324
τ 2 (ns)140.808411.10172.3793
τ a v e (ns)1.50941.16923.0910
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.; Wang, Y.; Yao, Y.; Hao, L.; Li, J.; Cao, Y. Preparation and Property Characterization of Eu2SmSbO7/ZnBiEuO4 Heterojunction Photocatalysts and Photocatalytic Degradation of Chlorpyrifos under Visible Light Irradiation. Catalysts 2024, 14, 144. https://doi.org/10.3390/catal14020144

AMA Style

Luan J, Wang Y, Yao Y, Hao L, Li J, Cao Y. Preparation and Property Characterization of Eu2SmSbO7/ZnBiEuO4 Heterojunction Photocatalysts and Photocatalytic Degradation of Chlorpyrifos under Visible Light Irradiation. Catalysts. 2024; 14(2):144. https://doi.org/10.3390/catal14020144

Chicago/Turabian Style

Luan, Jingfei, Yichun Wang, Ye Yao, Liang Hao, Jun Li, and Yu Cao. 2024. "Preparation and Property Characterization of Eu2SmSbO7/ZnBiEuO4 Heterojunction Photocatalysts and Photocatalytic Degradation of Chlorpyrifos under Visible Light Irradiation" Catalysts 14, no. 2: 144. https://doi.org/10.3390/catal14020144

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