Next Article in Journal
Batch and Flow Nitroaldol Synthesis Catalysed by Granulicella tundricola Hydroxynitrile Lyase Immobilised on Celite R-633
Next Article in Special Issue
Innovative and Practical Trends in the Photocatalytic Degradation of Environmental Pollutants
Previous Article in Journal
Study on Catalytic Water Oxidation Properties of Polynuclear Manganese Containing Polyoxometalates
Previous Article in Special Issue
Synthesis, Property Characterization and Photocatalytic Activity of the Ag3PO4/Gd2BiTaO7 Heterojunction Catalyst under Visible Light Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation, Property Characterization of Gd2YSbO7/ZnBiNbO5 Heterojunction Photocatalyst for Photocatalytic Degradation of Benzotriazole 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 2022, 12(2), 159; https://doi.org/10.3390/catal12020159
Submission received: 31 December 2021 / Revised: 17 January 2022 / Accepted: 24 January 2022 / Published: 27 January 2022

Abstract

:
The Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst was synthesized for the first time by the facile in situ precipitation method. The structural properties of a Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst were characterized by X-ray diffractometer, scanning electron microscope-X ray energy dispersive spectra, X-ray photoelectron spectrograph and UV-Vis diffuse reflectance spectrophotometer. The band gap energy (BGE) of Gd2YSbO7 or ZnBiNbO5 was found to be 2.396 eV or 2.696 eV, respectively. The photocatalytic property of Gd2YSbO7 or ZnBiNbO5 or Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst (GZHP) was reported. After a visible-light irradiation of 145 minutes (VLI-145 min), the removal rate (RER) of benzotriazole reached 99.05%, 82.45%, 78.23% or 47.30% with Gd2YSbO7/ZnBiNbO5 heterojunction (GZH), Gd2YSbO7, ZnBiNbO5 or N-doped TiO2 (NTO) as photocatalyst. In addition, the kinetic constant k, derived from the dynamic curve toward benzotriazole concentration and visible light irradiation time with GZH as a photocatalyst, reached 0.0213 min−1. Compared with Gd2YSbO7 or ZnBiNbO5 or NTO, GZHP showed maximal photocatalytic activity (PHA) for the photocatalytic degradation of benzotriazole under visible-light irradiation. The RER of total organic carbon during the photocatalytic degradation of benzotriazole reached 90.18%, 74.35%, 70.73% or 42.15% with GZH as a photocatalyst or with Gd2YSbO7, ZnBiNbO5 or NTO as a photocatalyst after VLI-145 min. Moreover, the kinetic constant k, which came from the dynamic curve toward total organic carbon concentration and visible light irradiation time with GZH as a photocatalyst, reached 0.0110 min−1. Based on above results, GZHP showed the maximal mineralization percentage ratio when GZHP degraded benzotriazole. The results showed that hydroxyl radicals was the main oxidation radical during the degradation of benzotriazole. The photocatalytic degradation of benzotriazole with GZH as a photocatalyst conformed to the first-order reaction kinetics. Our research aimed to improve the photocatalytic properties of the single photocatalyst.

1. Introduction

In recent years, environmental pollution issue has attracted more attention, especially sewage pollution problems, which are closely related to our daily life and health. Benzotriazole was one of most important synthetic compounds in plastic additives, which were used in large quantities in building materials and cooling and heating fluids, paints and dishwashing detergents, metal processing, aircraft deicing and anti-icing, and polymer stabilization [1,2]. Therefore, the environmental and health risk caused by benzotriazole had gradually attracted increasing attention all over the world. Furthermore, benzotriazole was highly water-solubility, and a great deal of residual benzotriazole had been measured in environmental samples [3,4,5,6,7,8]. Large amounts of residual pollutants in the water might result in different degrees of benzotriazole bioaccumulation in an aquatic food web, and might finally be found in the human body. Due to its high water-solubility, low octanol water distribution coefficients, low vapor pressure and fair resistance to biodegradation [9,10,11] characters, the removal of benzotriazole had become more difficult.
Some traditional methods, such as electrochemistry [4,5,6,7], biodegradation [3,4,12], adsorption [8,9,13], and flocculation precipitation [9,11], had been utilized to remove the pollutant in wastewater, such as benzotriazole, alkylphenol, antioxidants, etc.; however, none of these methods had been good enough for use in a wide range of applications. Photocatalysis technology [14,15,16] was widely used in water-cleaning treatment and sewerage treatment due to its advantages of environmental protection, high treatment efficiency and low cost [17,18]. Photocatalyst technology could use sunlight as the energy to realize the activation of photocatalyst instead of using expensive oxidants, which made photocatalyst technology cheap, clean, attractive and promising from the perspective of energy [19]. Photocatalysts could purify sewage by decomposing organic pollutants to produce non-oxidation radicals. After use, the catalyst could be regenerated and recycled, which greatly reduced the cost of the photocatalyst.
The conventional photocatalysts were metal oxides such as TiO2 and ZnO [20,21], considering the economic problems. However, due to its wide band gap, anatase TiO2 could only absorb 5% of the solar energy, which corresponded to the ultraviolet light; thus, the rest of the optical energy was wasted. While exploring new photocatalysts, the abundant endeavors of previous scholars established the realization of the degradation of pollutants during visible light irradiation (VLIR) as a crucial goal for researchers. An important breakthrough was achieved as Zou reported that Ni-doped InTaO4 compound had the ability to respond to the wavelength of visible-light regions in 2001. Zou proved that developing visible-light-responsive photocatalysts using the rest of the sunlight (43%) was possible. Latterly, photocatalysts with a formula of AB2O5 and A2B2O7 were also reported. It was reported that Kapoor et al. [22] investigated the decomposition of gaseous acetaldehyde under UV light irradiation by MgTi2O5 with a pseudo-brookite structure type. Mixed metal oxide MgTi2O5 decomposed acetaldehyde under UV light irradiation. Zhang et al. [23] investigated the photocatalytic degradation (PHD) of methyl orange by La2Ce2O7 powder. Ladan et al. [24] investigated the photocatalytic performance of the nanostructured Bi2Sn2O7 in the degradation of the pollutant malachite green under solar light conditions.
Many effective methods had been used to improve the efficiency of the photocatalysts, such as the ion-doping method, heterojunction construction [25,26,27,28,29,30] and photosensitization [31,32]. Among them, the construction of composites was an active research field in the field of photocatalysis [29,30,31,32,33,34,35]. The function of a single photocatalyst could be improved by using the composite material system [36,37,38,39,40], which could possess a longer carrier life, higher photocatalytic performance and higher chemical stability [41,42,43,44,45,46,47]. In our previous work [48], we found that Sm2FeSbO7 had a pyrochlore type structure. As a photocatalyst under VLIR, its structural modification seemed to have the potential to improve the photocatalytic activity (PHA). Based on the above analysis, we could assume that substituting Sm3+ with Gd3+, and substituting Fe3+ with Y3+ in Sm2FeSbO7, might increase the carrier concentration. The results showed that the electrical transmission and photophysical properties of the new Gd2YSbO7 (GYS) compound had been changed and improved, and the compound Gd2YSbO7 might have advanced photocatalytic properties.
In the photocatalysis process, the heterojunction catalyst showed excellent properties [49,50,51]. The heterojunction construction had a positive effect on the redox performance and electron transfer rate of the catalyst, which caused a high separative efficiency of photoinduced electron and photoinduced hole within two semiconductors [52]. Yu et al. [53] constructed the Bi/BiOCl heterojunction photocatalyst, which exhibited an extraordinarily high activity in the photodegradation of rhodamine B dye and persistent organic pollutants. Wang et al. [54] showed that the Ag2O/Ag3PO4 heterostructure had a high degradation rate for methyl orange and phenol under VLIR in the photocatalytic test. The above results showed that the construction of a photocatalyst with a heterojunction could significantly improve the redox performance of the photocatalyst [55] and improve the overall selectivity and activity of the reaction. In conclusion, the heterojunction construction of a catalyst was a promising application direction. Therefore, the Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst (GZHP) was designed and synthetized.
In this paper, an X-ray diffraction (XRD), scanning electron microscope-X ray energy dispersive spectra (SEM-EDS), X-ray photoelectron spectrograph (XPS), and UV-Vis spectrophotometer were utilized to analyze the structural properties of pure phase Gd2YSbO7 and single-phase ZnBiNbO5 prepared by the kinetic control method and high-temperature, solid-phase sintering method. In addition, the removal rate of benzotriazole under VLIR with Gd2YSbO7 as a photocatalyst, or ZnBiNbO5 as a photocatalyst or N-doped TiO2 (NTO) as a photocatalyst or Gd2YSbO7/ZnBiNbO5 heterojunction (GZH) as a photocatalyst, was detected. The purpose of this study was to prepare a new heterojunction photocatalyst for removing benzotriazole from wastewater. The innovative research content of this paper was to synthesize a new type of Gd2YSbO7 nanocatalyst, using the high-temperature, solid-phase sintering method, for the first time. A visible light response photocatalyst with a high PHA was obtained, which could effectively remove benzotriazole. The removal of organic pollutants within dye wastewater with GZH as a photocatalyst was more efficient.

2. Result and Discussion

2.1. XRD Analysis

The structure of the as-prepared compound ZnBiNbO5 is confirmed by XRD, as shown in Figure S1, with the JCPDS card of Y2Ti2O7. The structure of the as-prepared product Gd2YSbO7 was also examined by XRD and Materials Studio program was used to obtain the collected data based on Rietveld analysis method. It could be concluded that Gd2YSbO7 or ZnBiNbO5 was a single phase and the lattice parameter of the photocatalyst Gd2YSbO7 or ZnBiNbO5 was 10.653(6) Å or 10.498(7) Å. At the same time, O atoms were considered in the refinement model; the good agreement between the observed and calculated intensities showed that Gd2YSbO7 or ZnBiNbO5 crystalized in the cubic crystal with the space group Fd3m. Figure S2 shows the atomic structure of ZnBiNbO5. The atomic coordinates and structural parameters of Gd2YSbO7 and ZnBiNbO5 are shown in Table S1 and Table S2, respectively.
It was known that the x coordinate of the O (1) atom could be considered as an index of the change in the crystal structure on the pyrochlore-type A2B2O7 compounds (Cubic, space group Fd3m) and was equal to 0.375 when the six A-O (1) bond lengths were the same as that of the two A-O (2) bond lengths [56]. Therefore, information on the distortion of the MO6 (M = Y3+ and Sb5+) octahedra could be gained from the x value [56]. The x value was shifted from x = 0.375 [56]; thus, the distortion of the MO6 (M = Y3+ and Sb5+) octahedra clearly existed in the crystal structure of Gd2YSbO7. Charge separation was required for the PHD of benzotriazole under VLIR for the sake of preventing recombination of the photoinduced electrons and photoinduced holes. Based on the reports of Inoue and Kudo, the local distortion of the MO6 octahedra, obtained from some photocatalysts such as BaTi4O9 and Sr2M2O7 (M = Nb5+ and Ta5+), was very important for preventing the charge recombination and contributed to the improvement in the PHA [57,58]. Therefore, the distortion of the MO6 (M = Y3+ and Sb5+) octahedra in the crystal structure of Gd2YSbO7 could also be considered to be useful for enhancing the PHA. Gd2YSbO7 consisted of a three-dimensional network structure of corner-sharing MO6 (M = Y3+ and Sb5+) octahedra. The MO6 (M = Y3+ and Sb5+) octahedra were connected into chains by Gd3+ ion. Two kinds of Gd-O bond lengths coexisted: the six Gd-O (1) bond lengths (2.491 Å) were clearly longer than the two Gd-O (2) bond lengths (2.230 Å). The six M-O (1) (M = Y3+ and Sb5+) bond lengths was 2.020 Å, and the M-Gd (M = Y3+ and Sb5+) bond length was 3.642 Å. The M-O-M (M = Y3+ and Sb5+) bond angles were 128.640° in the crystal structure of Gd2YSbO7. The Gd-M-Gd (M = Y3+ and Sb5+) bond angles were 135.00° in the crystal structure of Gd2YSbO7. The Gd-M-O (M = Y3+ and Sb5+) bond angles were 139.203° in the crystal structure of Gd2YSbO7. Based on the previous reports on the luminescent properties, when the M-O-M bond angle was closer to 180°, this helped the excited state become delocalized [56]. This showed that the angles between the corner-sharing MO6 (M = Y3+ and Sb5+) octahedra, for example, the M-O-M bond angles for Gd2YSbO7, were important, affecting the PHA of Gd2YSbO7. The closer the M-O-M bond angles were to 180°, the larger the mobilities of the photoinduced electrons and photoinduced holes [56]. The mobilities of the photoinduced electrons and photoinduced holes affected the PHA because they affected the probability of electrons and holes reaching reaction sites on the catalyst surface [56].
In addition, the Gd–Y–O bond angle or Gd–Sb–O bond angle of Gd2YSbO7 was larger, which surely helped to increase the PHA of Gd2YSbO7. As in Gd2YSbO7, Gd was a 5d-block rare earth metal element, Y was a 4d-block metal element, and Sb was a 5p-block metal element. According to the above analysis, the effect of degrading benzotriazole under VLIR with Gd2YSbO7 as a photocatalyst could mainly be attributed to the crystalline structure and electronic structure.

2.2. UV-Vis Diffuse Reflectance Spectra

The absorption spectra of the Gd2YSbO7 sample and ZnBiNbO5 sample are shown in Figure S3. The absorption edge of this new photocatalyst Gd2YSbO7 was found to be at 475 nm, which was at the visible region of the spectrum. The band gap energies of the crystalline semiconductors could be determined by the intersection point between photon energy axis and the line extrapolated from the linear portion of the absorption edge of the so-called Kubelka–Munk Equation (1) (known as the re-emission equation) [59,60].
[ 1 R d ( h ν ) ] 2 2 R d ( h ν ) = α ( h ν ) S
where S was the scattering factor, and Rd was the diffuse reflectance, and α represented the absorption coefficient of radiation.
The optical absorption near the band edge of the crystalline semiconductors obeyed Equation (2) [61,62]:
αhν = A (Eg)n
Here, A, α, Eg and ν represent proportional constant, absorption coefficient, band gap and light frequency, respectively. Within this equation, n controls the transition property of the semiconductor. Following the above equation, Eg and n could be determined: (1) plotting ln (αhν) versus ln (Eg) assuming an approximate value of Eg; (2) deducing the value of n in accordance with the slope in this graph; (3) refining the value of Eg by plotting (αhν)1/n versus and extrapolating the plot to (αhν)1/n = 0. First, the direct method (1240/transition wavelength λ) was used to estimate the band gap Eg of Gd2YSbO7 or ZnBiNbO5; as a result, the Eg of Gd2YSbO7 or ZnBiNbO5 was estimated to be 2.226 eV (1240/557) or 2.562 eV (1240/484). Secondly, an indirect method (Equation (2)) was utilized to measure the precise band gap width of Gd2YSbO7 or ZnBiNbO5 using 2.226 eV or 2.562 eV within Equation (2). Based on the above steps, the values of Eg for Gd2YSbO7 or ZnBiNbO5 were calculated to be 2.396 eV or 2.696 eV. The value of n was about 0.5, which meant that the optical transition for Gd2YSbO7 was directly allowed. The value of n was about 2, which meant that the optical transition for ZnBiNbO5 was indirectly allowed, as shown in Figure S3.
The band gap energy of Gd2YSbO7 or ZnBiNbO5 was 2.396 eV or 2.696 eV, the band gap energy of Bi3O5I2 was 2.02 eV [63], and the band gap energy of Co-doped ZnO was 2.39 eV [64]. The band-gap energies of all three compounds were less than 2.70 eV, which meant that the above four catalysts owned visible-light response characteristics and possess tremendous potential for displaying high catalytic activity under VLIR.

2.3. Property Characterization of Gd2YSbO7/ZnBiNbO5 Heterojunction Photocatalyst

To obtain the surface chemical compositions and valence states of each element of GZHP, the X-ray photoelectron spectrum (XPS) was performed. Figure 1 presents the XPS survey spectrum of GZHP and Figure 2 shows the XPS spectra of O2−, Gd3+, Y3+, Zn2+, Bi3+, Nb5+ and Sb5+ which derive from GZHP. Based on the full XPS spectrum, the synthesized GZHP included the elements of Gd, Y, Zn, Bi, Nb, Sb and O. According to the XPS analysis results, the oxidation state of Gd, Y, Zn, Bi, Nb, Sb or O ion was +3, +3, +2, +3, +5, +5 or −2, respectively. Based on the above analysis results, that the chemical formula of the novel compound was Gd2YSbO7 and ZnBiNbO5. In Figure 2, various elemental peaks with specific binding energies were obtained according to the peak-differentiating and -imitating (fitting procedure) results [65,66]. The O1s peak of the O element was located at 530.05 eV. The Gd4d5/2 peak of the Gd element was located at 142.10 eV. The Bi4f5/2 and Bi4f7/2 peaks of Bi element were located at 164.20 eV and 159.07 eV. The Sb3d3/2 peak or Sb3d5/2 peak of the Sb element were located at 539.98 eV or 531.50 eV. The Y3d3/2 and Y3d5/2 peaks of the Y element were located at 158.55 eV and 157.02 eV. The Nb3d3/2 and Nb3d5/2 peaks of the Nb element were located at 209.40 eV and 206.60 eV. The Zn2p1/2 peak of the Zn element were located at 1044.30 eV and the Zn2p3/2 peak of the Zn element was located at 1021.65 eV. In brief, Figure 1 and Figure 2 revealed the presence of gadolinium (Gd4d), yttrium (Y3p and Y3d), antimony (Sb3d and Sb4d), bismuth (Bi5d and Bi4f), zinc (Zn2p), niobium (Nb3d) and oxygen (O1s) within the prepared sample. The surface elemental analysis results displayed that the average atomic ratio of Gd:Y:Sb:Zn:Bi:Nb:O was 1146:592:579:586:612: 571:5914. The atomic ratio of Gd:Y:Sb or Zn:Bi:Nb in the sample of GZHP was 1.98:1.02:1.00 or 0.96:1.00:0.93, respectively. The atomic ratio of Gd2YSbO7:ZnBiNbO5 was close to 581:590. The reason that the value of oxygen was high could be due to the amount of oxygen adsorbed on the surface of GZHP. Apparently, the XPS reflections did not show shoulders or widening, which was a strong proof that Gd2YSbO7/ZnBiNbO5 was a pure phase without any other phases.
Figure 3 and Figure 4 show SEM images and the EDS elemental mapping of GZHP (Gd, Y, Sb, O from Gd2YSbO7 and Zn, Bi, Nb, O from ZnBiNbO5). Figure 5 shows the EDS spectra of GZHP. Figure 3 and Figure 4 show that the larger tetragonal particles belonged to Gd2YSbO7 and the smaller circular shaped particles belonged to ZnBiNbO5. As could be seen from Figure 3 and Figure 4, Gd2YSbO7 particles were surrounded by the smaller ZnBiNbO5 particles, and these two particles were tightly combined, which indicated the successful synthesis of GZHP. Gd2YSbO7 possessed a rhombic-dodecahedron-like morphology. It was established that the structural growth in the photocatalyst was controlled by the different surface energies of the crystallite facets. Previously, researchers found the order of increasing surface energy for different facets of silver phosphate (111) < silver phosphate (100) < silver phosphate (110). Among them, the silver phosphate (110) facet showed a higher surface energy than the (111) facet, which is the reason that irregularly shaped silver phosphate structures accumulate along the (110) crystallographic direction, leading to the formation of a rhombic-dodecahedron-like morphology for silver phosphate [67,68]. Therefore, the reason that Gd2YSbO7 has a rhombic-dodecahedron-like morphology could be explained by above explanation. The experimental results shown in Figure 3 indicated that ZnBiNbO5 owned a regular, sphere, snowflake-like morphology and a homogeneous particle distribution. The particle size of ZnBiNbO5 was measured to be around 750 nm, while the larger particle size of Gd2YSbO7 was measured to be about 2300 nm.
The SEM-EDS analysis results revealed that there were no other impure elements in the GZHP compound. At the same time, the pure phase of ZnBiNbO5 was consistent with the X-ray diffraction analysis result displayed in Figure S1. It could be concluded from Figure 4 and Figure 5 that the gadolinium element, yttrium element, antimony element, bismuth element, zinc element, niobium element and oxygen element are contained within GZHP. The above results are consistent with the XPS results of GZHP, shown in Figure 1 and Figure 2. In accordance with the EDS spectrum of GZHP, the atomic ratio of Gd:Y:Sb:Zn:Bi:Nb:O was 1213:547:587:611:572:590:5880, which is also consistent with the XPS results of GZHP. The atomic ratio of Gd2YSbO7:ZnBiNbO5 was close to 587:591. The more accurate stoichiometric ratio of the final product GZHP was summarized as 1168:1181 for Gd2YSbO7:ZnBiNbO5. According to the above results, we could conclude that GZHP was of a high purity under our preparation conditions.

2.4. Photocatalytic Activity

Figure 6 presents the concentration variation (COV) curves of benzotriazole during PHD of benzotriazole with GZH as a photocatalyst, with Gd2YSbO7 as a photocatalyst, with ZnBiNbO5 as a photocatalyst or with NTO as a photocatalyst under VLIR. Figure S4 shows the COV curves of benzotriazole during photocatalytic degradation of benzotriazole in the absence of photocatalyst under VLIR. Figure 6 shows that the removal rate (RER) of benzotriazole was 5.18%, 5.45%, 4.95% or 26.98% with GZH as a catalyst, Gd2YSbO7 as a catalyst, ZnBiNbO5 as a catalyst or with NTO as catalyst under dark conditions for 45 min. Moreover, Figure S4 also shows that the RER of benzotriazole was 2.03% in the absence of a photocatalyst. Figure 6 shows that the concentration of benzotriazole within chemical industry wastewater gradually decreased with an increase in VLIR time. The results obtained from Figure 6 indicated that the removal rate of benzotriazole within chemical industry wastewater reached 99.05%, the reaction rate was 4.55 × 10−9 mol·L−1·s−1 and the photonic efficiency was 0.0956% with GZH as a catalyst after visible light irradiation for 145 min (VLI-145 min). GZHP shows an outstanding removal efficiency in degrading benzotriazole compared with the normal photocatalysts, such as S-TiO2, N-TiO2, C-TiO2 or CdS-TiO2, after VLI-145 min, as shown in Table S3. All the other experiments followed the same VLI-145 min. When Gd2YSbO7, which possessed the specific surface area of 4.12 m2/g and was prepared with the sintering temperature of 1320 °C for 50 h, was used as photocatalyst, the removal rate of benzotriazole reached 82.45%, the reaction rate was 3.79 × 10−9 mol·L−1·s−1 and the photonic efficiency was 0.0796%. When the sintering temperature was set as 950 °C for 50 h, the RER of benzotriazole reached 51.42% with mixed oxides as a catalyst after VLI-145 min. When the sintering temperature was set as 1450 °C for 50 h, the RER of benzotriazole reached 76.21% with Gd2YSbO7, which owned the specific surface area of 3.07 m2/g as a catalyst after VLI-145 min. The removal rate of benzotriazole within chemical industry wastewater reached 78.23%, the reaction rate was 3.597 × 10−9 mol·L−1·s−1 and the photonic efficiency was 0.0754% with ZnBiNbO5 as a photocatalyst. The removal rate of benzotriazole reached 47.30%, the reaction rate was 2.17 × 10−9 mol·L−1·s−1 and the photonic efficiency was 0.0456% with NTO as a photocatalyst. Moreover, the above results suggest that the photodegradation efficiency of benzotriazole in the presence of GZHP was the highest; meanwhile, the photodegradation efficiency of benzotriazole with Gd2YSbO7 as a photocatalyst was higher than that with ZnBiNbO5 as a photocatalyst or with NTO as a photocatalyst. Ultimately, the photodegradation efficiency of benzotriazole with ZnBiNbO5 as a photocatalyst was higher than that with NTO as a photocatalyst, indicating that the visible light PHA of GZHP was maximal compared with that of Gd2YSbO7 or ZnBiNbO5 or NTO. The above results show that the removal rate of benzotriazole with GZH as a catalyst was 1.201 times or 1.266 times or 2.094 times higher than that with Gd2YSbO7 or ZnBiNbO5 as photocatalysts, or with NTO as a photocatalyst, after VLI-145 min.
Figure 7 shows the COV curves of total organic carbon during the PHD of benzotriazole in chemical industry wastewater with GZH, Gd2YSbO7, ZnBiNbO5 or NTO as a photocatalyst under VLIR. The concentration of benzotriazole gradually decreased with an increase in VLIR time. As can be seen in Figure 7, the removal rate of total organic carbon within chemical industry wastewater reached 90.18%, 74.35%, 70.73% and 42.15%, respectively, after VLI-145 min when GZHP, Gd2YSbO7, ZnBiNbO5 and NTO were used to degrade benzotriazole. In conclusion, we can observe from the above results that the removal rate of total organic carbon during degrading benzotriazole in the presence of GZHP was higher than that in the presence of Gd2YSbO7 or ZnBiNbO5 or NTO. The above results also indicate that the removal rate of total organic carbon during the degradation of benzotriazole in the presence of Gd2YSbO7 was much higher than that in the presence of ZnBiNbO5 or NTO, which meant that GZHP showed the maximal mineralization percentage ratio when GZHP degraded benzotriazole compared with Gd2YSbO7 or ZnBiNbO5 or NTO.
Figure 8 presents the COV curves of benzotriazole during PHD with Gd2YSbO7/ZnBiNbO5 heterojunction as a photocatalyst under VLIR for three cycle degradation tests. As seen in Figure 8, the removal rate of benzotriazole reached 97.25%, 96.08% or 94.90%, respectively, after VLI-145 min with GZH as a photocatalyst by accomplishing three cycle tests for degrading benzotriazole. Figure 9 shows the COV curves of total organic carbon during the PHD of benzotriazole with GZH as a catalyst under VLIR for three cycle degradation tests. From Figure 9, we can see that the removal rates of total organic carbon reach 88.19%, 87.02% or 85.27%, respectively. The experimental results obtained from Figure 8 and Figure 9 show that the GZHP has a high stability.
Figure S5 shows the observed first-order kinetic plots for the PHD of benzotriazole with GZH, Gd2YSbO7, ZnBiNbO5 or NTO as a photocatalyst under VLIR. Based on Figure S5, the kinetic constant k, which was obtained from the dynamic curve toward benzotriazole concentration, and VLIR time with Gd2YSbO7/ZnBiNbO5 heterojunction, Gd2YSbO7, ZnBiNbO5 or NTO as a photocatalyst reached 0.0213 min−1, 0.0095 min−1, 0.0077 min−1 or 0.0028 min−1, respectively. The kinetic constant k, which came from the dynamic curve toward total organic carbon concentration, reached 0.0110 min−1, 0.0075 min−1, 0.0065 min−1 or 0.0023 min−1 with GZH, Gd2YSbO7, ZnBiNbO5 or NTO as a photocatalyst. The fact that the KTOC value for degrading benzotriazole was lower than the value of KC for degrading benzotriazole using the same catalyst illustrated that the intermediate photodegradation products of benzotriazole probably appeared during the PHD of benzotriazole under VLIR. Simultaneously, compared with the other three photocatalysts, GZHP presented a higher mineralization efficiency for benzotriazole degradation.
Figure S6 shows the observed first-order kinetic plots for the PHD of benzotriazole with Gd2YSbO7/ZnBiNbO5 heterojunction as a photocatalyst under VLIR for three cycle degradation tests. Based on the results from Figure S6, the kinetic constant k, which was derived from the dynamic curve toward benzotriazole concentration and VLIR time with GZH as a catalyst for three cycle degradation tests, reached 0.0172 min−1, 0.0152 min−1 or 0.0138 min−1. The kinetic constant k, which came from the dynamic curve toward total organic carbon concentration and VLIR time with GZH as a catalyst for three cycle degradation tests, reached 0.0099 min−1, 0.0089 min−1 or 0.0080 min−1. According to the experimental results from Figures S5 and S6, the PHD of benzotriazole in chemical industry wastewater with GZH as a catalyst under VLIR conformed to the first-order reaction kinetics.
From Figure S6, the removal rate of benzotriazole is shown to decrease by 4.15% with GZH as a catalyst under VLIR after three cycle degradation tests, and the removal rate of total organic carbon decreased by 4.91%. No significant difference could be seen for the degradation efficiency in the above three cycle experiments, and the photocatalyst structure of GZHP was stable.
Figure 10 presents the effect of different radical scavengers, such as benzoquinone (BQ) or isopropanol (IPA) or ethylenediamine tetraacetic acid (EDTA), on the removal rate of benzotriazole with GZH as a catalyst under VLIR. To determine the mechanism used to degrade benzotriazole, different radical scavengers were separately added to the benzotriazole solution at the beginning of the photocatalytic experiment. Isopropanol (IPA) was used to capture hydroxyl radicals (OH); Benzoquinone (BQ) was utilized to capture superoxide anions (O2); Ethylenediaminetetraacetic acid (EDTA) was used to capture holes (h+). The scheming IPA, BQ or EDTA concentration was 0.15 mmol L−1, and the amount of IPA, BQ or EDTA added was 1 mL. According to Figure 10, while the IPA, BQ or EDTA were added to the benzotriazole solution, the removal rate of benzotriazole decreased by 68.16%, 59.41% or 36.75%, respectively, compared with the removal rate of benzotriazole that was derived from the control group. Thus, it could be concluded that OH, h+ and O2 were all active radicals during the benzotriazole degradation process. Figure 10 shows that OH in the benzotriazole solution played a leading role when benzotriazole was degraded with GZH as a photocatalyst under VLIR. By adding a capture agent, the hydroxyl radicals was shown to possess a maximal oxidation removal ability for removing benzotriazole in chemical industry wastewater compared with superoxide anion or holes. The oxidation removal ability for degrading benzotriazole, which ranged from high to low among the three oxidation radicals, was as follows: hydroxyl radicals > superoxide anions > holes.
Nyquist impedance plot measurement was another important characterization, which showed the photogenerated electron and photogenerated hole migration process of the prepared photocatalysts at solid/electrolyte interfaces. The smaller the arc radius, the higher the transportation efficiency of the photocatalysts. Figure S7 shows the corresponding Nyquist impedance plots of the prepared GZHP or Gd2YSbO7 photocatalyst or ZnBiNbO5 photocatalyst. It could easily be seen that the diameter of the arc radius was in the order ZnBiNbO5 > Gd2YSbO7 > GZHP. The above results indicate that the prepared GZHP presented a more efficient separation of the photogenerated electron and photogenerated hole, and a faster interfacial charge migration ability.

2.5. Possible Degradation Mechanism Analysis

Figure 11 shows the possible PHD mechanism of benzotriazole with GZH as a catalyst under VLIR. The potentials of the valence band (VB) and conductor band (CB) for the semiconductor catalyst could be calculated in accordance with Equation (3) and Equation (4), as follows [69]:
ECB = XEe − 0.5Eg
EVB = ECB + Eg
where Eg presents the band gap of the semiconductor, X presents the electronegativity of the semiconductor, and Ee presents the energy of free electrons on the hydrogen scale, which was about 4.5 eV. According to the above equations, the VB potential or the CB potential for ZnBiNbO5 were estimated to be 2.749 eV or 0.053 eV, respectively. For Gd2YSbO7, the VB potential or the CB potential were estimated to be 1.745 eV and −0.651 eV, respectively. Both ZnBiNbO5 and Gd2YSbO7 could absorb visible light and internally generated electron–hole pairs when the GZHP was irradiated by visible light. Since the redox potential position of CB of Gd2YSbO7 (−0.651 eV) was more negative than that of ZnBiNbO5 (0.053 eV), the electrons photoinduced on the CB of Gd2YSbO7 could transfer to the CB of ZnBiNbO5. Additionally, the redox potential position of VB of ZnBiNbO5 (2.749 eV) was more positive than that of Gd2YSbO7 (1.745 eV), and the photoinduced holes on the VB of ZnBiNbO5 could transfer to the VB of Gd2YSbO7. Therefore, the coupling of ZnBiNbO5 and Gd2YSbO7 to produce GZHP could effectively reduce the recombination rate of photoinduced electrons and photoinduced holes, and subsequently decrease the internal resistance, and prolong the lifetime of photo-induced electrons and photo-induced holes, as well as enhancing the interfacial charge transfer efficiency [70]. As a result, more oxidative radicals, such as OH or O2, could be produced to improve the degradation efficiency of benzotriazole. Moreover, the CB potential of Gd2YSbO7 was −0.651 eV, which is more negative than that of O2/O2 (−0.33 V), indicating that the electrons within the CB of Gd2YSbO7 could absorb oxygen to produce O2, which could degrade benzotriazole (as shown as path 1 in Figure 11). At the same time, the VB potential of ZnBiNbO5 was 2.749 eV, which was more positive than that of OH/OH (2.38 V), indicating that the holes in the VB of ZnBiNbO5 could oxidize H2O or OH into OH to degrade benzotriazole, as shown as path 2. As shown in path 3, the benzotriazole could be oxidized and degraded directly because of the strong oxidizing ability of the holes that were photoinduced in the VB of Gd2YSbO7 or ZnBiNbO5. In sum, the excellent PHA of GZHP toward benzotriazole degradation was mainly credited to the high efficiency of the electron–hole separation induced by GZHP.
To study the degradation mechanism of benzotriazole, the intermediate products that were generated during the degradation process of benzotriazole were detected by LC–MS. The intermediate products obtained during the PHD of benzotriazole were identified as C6H5N3O (m/z =136), C7H7N3 (m/z = 134), C5H5N3O2 (m/z = 138), C5H5N3O4 (m/z = 172), C6H5N3O4 (m/z =184), C4H3N3O4 (m/z = 158), C6H5N3O2 (m/z = 152), C4H5N3O2 (m/z = 128) and C7H7N3O (m/z = 150). According to the above-detected intermediate products, the PHD pathway of benzotriazole was proposed. Figure 12 presents the suggested PHD pathway scheme of benzotriazole under VLIR with Gd2YSbO7/ZnBiNbO5 heterojunction as a photocatalyst. Figure 12 shows that the oxidation reaction and methylation reaction of the triazole ring were realized during the PHD process of benzotriazole. As expected, benzotriazole was converted into small-molecule organic compounds; therefore, the new small-molecule products could combine with other organic active groups and finally turn into CO2 and water.

3. Experimental Section

3.1. Materials and Reagents

Ethylenediaminetetraacetic acid (EDTA, C10H16N2O8, purity = 99.5%) and isopropyl alcohol (IPA, C3H8O, purity ≥ 99.7%) were analytical grade. P-benzoquinone (BQ, C6H4O2, purity ≥ 98.0%) was chemical grade. Moreover, the above chemical reagents were all purchased from Sinopharm Group Chemical Reagent Co., Ltd. (Beijing, China). Absolute ethanol (C2H5OH, purity ≥ 99.5%) conformed to American Chemical Society Specifications and was purchased from Aladdin Group Chemical Reagent Co., Ltd. (Beijing, China). Benzotriazole (C6H5N3, purity ≥ 98%) was gas chromatography grade and was purchased from Tianjin Bodi Chemical Co., Ltd., Tianjin, China as the model material. Ultra-pure water (18.25 MU cm) was utilized throughout this work.

3.2. Preparation Method of Gd2YSbO7

Gd2YSbO7 was synthesized by solid-phase sintering method at a high temperature using Gd2O3 (99.99%), Y2O3 (99.99%) and Sb2O5 (99.99%) (Sinopharm Group Chemical Reagent Co., Ltd, Shanghai, China). In order to confirm the molar ratio of the raw oxides, the volatility of the raw oxide was tested using the following procedures: 1 g Gd2O3, 1 g Y2O3 and 1 g Sb2O5 were separately weighed and ground for 3 h in an agate mortar, and the above three raw oxides were separately placed in the alumina crucible (Shenyang Crucible Co., LTD, Shenyang, China) and then calcined at 1320 °C for 20 h in a high-temperature electric furnace (KSL 1700X, Hefei Kejing Materials Technology CO., LTD, Hefei, China). After cooling to room temperature, the three raw oxides were weighed again. As Gd2O3, Y2O3 or Sb2O5 still weighed 1 g after high-temperature sintering, the above three raw oxides were not volatile at 1320 °C. Therefore, a final molar ratio of n(Gd2O3):n(Y2O3):n(Sb2O5) = 2:1:1 was determined. Before the experiment, all powders were dried at 200 °C for 4 h. The precursors were stoichiometrically (n(Gd2O3):n(Y2O3):n(Sb2O5) = 2:1:1) mixed, then pressed into small columns, placed into an alumina crucible and calcined at 400 °C for 2 h. When naturally cooled to room temperature, the raw materials and the small columns were taken out of the electric furnace. The mixed materials should be ground and, subsequently, these mixed materials should be put into the electric furnace. Ultimately, the calcination process was separately carried out at 1320 °C for 50 h in an electric furnace. To clarify the preparation technology, a schematic representation of the synthesis process of the Gd2YSbO7 is shown in Figure S8a.

3.3. Preparation Method of ZnBiNbO5

ZnBiNbO5 was synthesized by the solid-phase sintering method at a high temperature using ZnO (99.99%), Bi2O3 (99.99%) and Nb2O5 (99.99%) (Sinopharm Group Chemical Reagent Co., Ltd., Shanghai, China). In order to confirm the molar ratio of the raw oxides, the volatility of the raw oxides was tested using the following procedures: 0.02 mol ZnO, 0.01 mol Bi2O3 and 0.01 mol Nb2O5 were separately weighed and were ground for 3 h in an agate mortar; the above three oxides were separately placed in an alumina crucible (Shenyang Crucible Co., LTD, Shenyang, China) and then calcined at 1100 °C for 25 h in a high-temperature electric furnace (KSL 1700X, Hefei Kejing Materials Technology CO., LTD, Hefei, China). After naturally cooling to room temperature, the quality of ZnO or Bi2O3 or Nb2O5 was remeasured and converted to a mole mass again. The results showed that 0.02 mol ZnO and 0.01 mol Nb2O5 remained unchanged. However, only 0.008 mol Bi2O3 was left, which meant that, when the temperature was 1100 °C, the Bi2O3 would be volatile. After five failed experiments, we concluded that the mole mass of 20% for Bi2O3 volatilized; as a result, the actual mole mass dosage of Bi2O3 should be 120% of the planned dosage. Therefore, a final molar ratio of n(ZnO):n(Bi2O3):n(Nb2O5) = 2:1.2:1 was determined.
These fully mixed materials were stoichiometrically (n(ZnO):n(Bi2O3):n(Nb2O5) = 2:1.2:1) quantified and put into a ball mill to obtain the final partial size of 1–2 µm. To ensure purity, all the powders were dried at 200 °C for 4 h before synthesis. Then, these powders were mixed in an aluminum oxide crucible after pressing into disks and, finally, these powders were sintered in an electric furnace at 400 °C for 8 h. After crushing and pressing, the mixture was sintered again in an electric furnace at 1100 °C for 25 h. Finally, pure ZnBiNbO5 catalyst was obtained after total grinding. To clarify the above preparation technology, a schematic representation of the synthesis process of the ZnBiNbO5 is shown in Figure S8b.

3.4. Synthesis of N-Doped TiO2

Nitrogen-doped titania (NTO) catalyst was prepared by sol-gel method; as a result, the precursor and solvent were tetrabutyl titanate and ethanol, respectively. Then, the following steps were carried out: (1) Solution A was made with 17 mL tetrabutyl titanate and 40 mL absolute ethyl alcohol; (2) Solution B consisted of 40 mL absolute ethyl alcohol, 10 mL glacial acetic acid and 5 mL double distilled water; (3) Under magnetic stirring condition, solution A was added drop by drop into solution until a transparent colloidal suspension was formed; (4) Aqua ammonia (N/Ti proportion was 8 mol%) was put into colloidal suspension and the magnetic stirring condition was maintained for 1 h. Next, the xerogel was generated after being aged for 2 days. The xerogel was ground into powder, which was calcined at 500 °C for 2 h; subsequently, the above powder was ground in agate mortar and screened by shaker to obtain NTO powders.

3.5. Synthesis of Gd2YSbO7/ZnBiNbO5 Heterojunction Photocatalyst

First, 0.30 mol/L Gd (NO3)3·6H2O, 0.15 mol/L Y(NO3)3·6H2Oand 0.15 mol/L SbCl5 were mixed and stirred for 20 h. The aforementioned solution was transferred into a Teflflon-lined autoclave, and then the temperature was raised to 200 °C for 15 h. Subsequently, the obtained powder was calcined at 800 °C for 10 h in a tube furnace at a rate of 8 °C/min under an atmosphere of N2. Gd2YSbO7 powder was finally obtained. Secondly, 0.30 mol/L Zn (NO3)2·6H2O, 0.15 mol/L Bi(NO3)3·5H2O and 0.15 mol/L NbCl5 were mixed and stirred for 20 h. This solution was transferred into a Teflflon-lined autoclave and heated at 200 °C for 15 h. Subsequently, the obtained powder was calcined at 790 °C for 10 h in a tube furnace at a rate of 8 °C min under an atmosphere of N2. ZnBiNbO5 powder was finally obtained.
A facile solvothermal method [71] was utilized to synthesize the new GZHP in this paper. GZHP was prepared by mixing 400 mg of Gd2YSbO7 with 400 mg of ZnBiNbO5 in 200 mL of octanol (C8H18O); then, the ultrasonic bath was used to disperse for 1 h. The mixture was then heated and refluxed at 140 °C for 2 h under vigorous stirring conditions to improve attachment of ZnBiNbO5 to the surface of Gd2YSbO7 nanoparticles to form GZHP. After being cooled to room temperature, the products were collected by centrifugation and then the mixture of n-hexane/ethanol was utilized to wash the sample several times. At last, the purified powders were dried at 60 °C for 6 h. Finally, GZHP was successfully prepared. To clarify this preparation technology, a schematic representation of the synthesis process of the Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst is shown in Figure S8c.

3.6. Characterizations

The pure crystals of the prepared samples were examined by the powder X-ray diffractometer (XRD, Shimadzu, XRD-6000, Cu Kɑ radiation, λ = 1.5418 Å, sampling step length of 0.02°, preset time of 0.3 s step−1, 2 theta of 10–80°). The scanning speed of the XRD instrument was 70°, which took 70 min; thus, each degree took 60 s. Each degree was divided into 50 steps, and each step took 1.2 s. The morphology and microstructure of the prepared samples were characterized using scanning electron microscope (SEM, FEI, Quanta 250) and the elementary composition derived from the samples prepared above prepared was captured by energy-dispersive spectroscopy (EDS). SEM minimized the amount of sample preparation required, and the low vacuum and ESEM abilities enabled charge-free imaging and analysis. Quanta’s patented “through-the-lens” differential pressure vacuum technology enabled EDS and EBSD analysis of samples under high- and low-vacuum conditions for enhanced analytical abilities. Stable high beam current (up to 2 μA) enabled fast and accurate analysis. The surface of the sample was plated with gold, and the catalyst powder was evenly coated on the disk to complete the surface imaging and obtain surface morphology and composition information for the sample. A diffuse reflectance spectrum was obtained for the sample prepared above using an UV-Vis spectrophotometer (UV-Vis DRS, Shimadzu, UV-3600). Shimadzu UV-3600 could extend the measurement range from the most common single-beam spectrophotometer to a deep UV-VIS-NIR spectrophotometer for UV and NIR regions. Photomultiplier tubes were used in the ultraviolet and visible regions; InGaAs and PbS detectors were used in the near-infrared region. The InGaAs detector covered the weak range of photomultiplier tubes and PbS detectors, ensuring high sensitivity over the entire measurement range. The surface chemical composition and states of the prepared sample were analyzed by X-ray photoelectron spectrograph (XPS, UlVAC-PHI, PHI 5000 VersaProbe) with an Al-kα X-ray source. The PHI 5000 VersaProbe was a multifunctional analytical instrument for PHI Scanning XPS Microprobe Devices in PHI’s industry-leading ZL Scanning X-Ray Technology and ZL Dual Beam Neutralization Technology. Scanning focused X-ray technology was the core technology of VersaProbe. Monochromatic, scanning, and focused X-ray sources allowed for a better analysis of microscopic and macroscopic regions, and all X-ray settings (including those with a beam spot diameter of less than 10 µm) could be used for spectral analysis, depth profiling and image analysis.

3.7. Photoelectrochemical Experiments

The electrochemical impedance spectroscopy (EIS) experiment was performed by a CHI660D electrochemical station (Chenhua Instruments Co. Shanghai, China) with a standard three-electrode. The three-electrode consisted of a working electrode (as-prepared catalysts), counter-electrode (platinum plate) and reference electrode (Ag/AgCl electrode). The electrolyte was Na2SO4 aqueous solution (0.5 mol/L). The light source for the experiment was a 500 W Xe lamp with a UV cut-off filter. The preparation of the working electrode was as follows: we pkaced the sample (0.03 g) and chitosan (0.01 g) in dimethylformamide (0.45 mL), and then we used ultrasonic treatment for an hour to obtain a uniform suspension solution. Subsequently, they were trickled on indium tin oxide (ITO)-conducting glass (10 mm × 20 mm). At last, the working electrode was dried at 80 °C for 10 min.

3.8. Experimental Setup and Procedure

All the experiments were carried out in a photocatalytic reactor (XPA-7, Xujiang Electromechanical Plant, Nanjing, China) and the temperature of the reaction system was 20 °C, controlled by circulating cooling water. Simulated sunlight irradiation was provided by a 500 W xenon with a 420 nm cut-off filter. For the same 12 quartz tubes, the volume of a single reaction solution was 40 mL, and the total reaction volume for chemical industry wastewater was 480 mL. The dosage of Gd2YSbO7 or ZnBiNbO5 or GZHP was 0.75 g/L; the concentration of benzotriazole was 0.040 mmol/L. The concentration of benzotriazole was the residual concentration after biodegradation for practical chemical industry wastewater, which contained a benzotriazole concentration of 1.2 mmol/L. During the reaction, 3 mL suspension was periodically withdrawn, and 27 mL suspension was taken to measure the concentration of residual benzotriazole after VLI-145 min; subsequently, the filtration (0.22 μm PES polyether sulfone filter membrane) was used to remove the catalyst. Finally, the residual concentration of benzotriazole in solution was determined by Agilent 200 high-performance liquid chromatography (Agilent Technologies, Palo Alto, CA, USA) with an UV detector and a Zorbax 300SB-C18 column (4.6 mm × 150 mm, 5 um). The mobile phase consisted of 50% CH3CN and 50% distilled deionized water. The UV detection wavelengths were located at 254 nm. The injection volume of post-photodegradation benzotriazole solution was 10 uL, and the flow rate was 1 mL·min−1. Prior to VLIR, we maintained magnetic stirring conditions in darkness for 45 min to guarantee the suspension (contained the photocatalyst and benzotriazole) in adsorption/desorption equilibrium among the photocatalyst, benzotriazole and atmospheric oxygen. During VLIR, the suspension was stirred at 500 rpm.
The mineralization experimental data for benzotriazole within the reaction solution were measured using a total organic carbon analyzer (TOC-5000 A, Shimadzu Corporation, Kyoto, Japan). To detect the concentration of total organic carbon during the PHD of benzotriazole, potassium acid phthalate (KHC8H4O4) or an anhydrous sodium carbonate was utilized as a standard reagent. Standard solutions of potassium acid phthalate with a known carbon concentration (in the range of 0–100 mg/L) were prepared for calibration purposes. Nine samples, which contained 68 mL reaction solution, were used to measure TOC concentration after VLI-145 min.
The identification and measurement of benzotriazole and its intermediate degradation products were carried out using liquid chromatography–mass spectrometry (LC-MS, Thermo Quest LCQ Duo, Thermo Fisher Scientific Corporation, Waltham, MA, USA. Beta Basic-C18 HPLC column: 150 × 2.1 mm, ID of 5 μm, Thermo Fisher Scientific Corporation, Waltham, MA, USA.). Here, 20 μL of solution, obtained after the photocatalytic reaction, was automatically injected into the LC–MS system. The mobile phase contained 60% methanol and 40% ultrapure water, and the flow rate was 0.2 mL/min. MS conditions included an electrospray ionization interface, capillary temperature of 27 °C with a voltage of 19.00 V, spray voltage of 5000 V and a constant sheath gas flow rate. The spectrum was acquired in the negative ion scan mode and the m/z ranged from 50 to 600.
The incident photon flux Io, which was tested by a radiometer (Model FZ-A, Photoelectric Instrument Factory Beijing Normal University, China), was determined as 4.76 × 10−6 Einstein L−1 s−1 under VLIR (wavelength range of 400–700 nm). The incident photon flux on the photoreactor was changed by adjusting the distance between the photoreactor and the Xe arc lamp.
The photonic efficiency was calculated in accordance with the following Equation (5):
ϕ = R/Io
where ϕ was the photonic efficiency (%), R was the degradation rate of benzotriazole (mol L−1 s−1), and Io was the incident photon flux (Einstein L−1 s−1).

4. Conclusions

For the first time, an efficient photocatalyst Gd2YSbO7 was successfully synthetized by the solid-phase sintering method at a high temperature. Furthermore, considering the superiority of the heterojunction construction, GZHP was synthesized by a facile solvothermal method. The photophysical properties and the photocatalytic properties of Gd2YSbO7 and GZHP were investigated via SEM, XRD, UV-Vis DRS and XPS tests. Based on the related tests and the analysis of the results, the prepared Gd2YSbO7 was pure phase and crystallized in a pyrochlore structure with the space group Fd3m. The lattice parameter and the band gap of Gd2YSbO7 were a = 10.653(6) Å and 2.396 eV, respectively. GZHP was proved to be an effective photocatalyst for removing benzotriazole in wastewater; after VLI-145 min, the RER of benzotriazole and total organic carbon reached 99.05% and 90.18%, respectively. The RER of benzotriazole with GZH as a photocatalyst was 1.201 times or 1.266 times or 2.094 times higher than that with Gd2YSbO7, ZnBiNbO5 or NTO as a photocatalyst after VLI-145 min. Furthermore, the kinetic constant k reached 0.0110 min−1. Therefore, it could be concluded that using a GZH as photocatalyst might be a potent method for treating wastewater polluted by benzotriazole. A degradation mechanism for benzotriazole was proposed, and nine intermediate products were detected during the degradation process of benzotriazole; as a result, the oxidation reaction and methylation reactions of the triazole ring were realized during the PHD process of benzotriazole. Finally, conceivable photodegradation pathways for the degradation of benzotriazole were presented.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12020159/s1, Figure S1: X-ray powder diffraction patterns of ZnBiNbO5 and JCPDS of Y2Ti2O7; Figure S2: Atomic structure of ZnBiNbO5. (Red atom: O, cyan atom: Zn, purple atom: Bi or Nb.); Table S1. Structural parameters of Gd2YSbO7 prepared by solid state reaction method. Table S2: Structural parameters of ZnBiNbO5 prepared by solid state reaction method; Figure S3. (a) UV-Vis diffuse reflectance spectrum of ZnBiNbO5 sample; (b) Plot of (αhν)1/2 versus for ZnBiNbO5; (c) UV-Vis diffuse reflectance spectrum of Gd2YSbO7 sample; (d) Plot of (αhν)1/2 versus for Gd2YSbO7 sample; Figure S4. Concentration variation curves of benzotriazole during photocatalytic degradation of benzotriazole in the absence of photocatalyst under VLIR.; Table S3: The removal rate of benzotriazole with Gd2YSbO7/ZnBiNbO5 heterojunction, S-TiO2, N-TiO2, C-TiO2 or CdS-TiO2 as photocatalyst; Figure S5: (a) Observed first-order kinetic plots for the photocatalytic degradation of benzotriazole with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst or with Gd2YSbO7 as photocatalyst or with ZnBiNbO5 as photocatalyst or with NTO as photocatalyst under VLIR; (b) Observed first-order kinetic plots for total organic carbon during photocatalytic degradation of benzotriazole in chemical industry wastewater with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst or with Gd2YSbO7 as photocatalyst or with ZnBiNbO5 as photocatalyst or with NTO as photocatalyst under VLIR.; Figure S6. Observed first-order kinetic plots for the photocatalytic degradation of benzotriazole with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR for three cycle degradation tests; Figure S7: Nyquist impedance plots of Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst or Gd2YSbO7 photocatalyst or ZnBiNbO5 photocatalyst; Figure S8: (a) Schematic representation of the synthesis process of the Gd2YSbO7; (b) Schematic representation of the synthesis process of the ZnBiNbO5; (c) Schematic representation of the synthesis process of the Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst.

Author Contributions

Conceptualization, Y.Y. and J.L.; methodology, Y.Y. and J.L.; software, Y.Y. and J.L.; validation, Y.Y. and J.L.; formal analysis, Y.Y. and J.L.; investigation, Y.Y. and J.L.; resources, Y.Y. and J.L.; data curation, Y.Y. and J.L.; writing—original draft preparation, Y.Y. and J.L.; writing—review and editing, Y.Y. and J.L.; visualization, Y.Y. and J.L.; supervision, Y.Y. and J.L.; project administration, Y.Y. and J.L.; funding acquisition, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Scientific and Technical Innovation Leading Personnel and Team Foundation for Middle-aged and Young Scientist of Science and Technology Bureau of Jilin Province of China (Grant No. 20200301033RQ), by the Free Exploring Key Item of Natural Science Fundation of Science and Technology Bureau of Jilin Province of China (Grant No. YDZJ202101ZYTS161), by the Industrial Technology Research and Development Fund of Jilin Province Capital Development Fund on Budget in 2021 of Jilin Province Development and Reform Commission of China (Grant No. 2021C037-1), by the Project Funded for Innovational and Enterprising Talents of Department of Human Resource and Social Security of Jilin Province of China (Grant No. 2020033), by the Scientific Research Planning Project of the Education Department of Jilin Province of China (Grant No. JJKH20200832KJ), by the Scientific Research Planning Project of the Education Department of Jilin Province of China (Grant No. JJKH20210885KJ), by Natural Science Foundation of Changchun Normal University (Grant No. [2019]13), by the Scientific Research Initating Foundation for Advanced Doctor of Changchun Normal University.

Acknowledgments

Thanks to Dongxu Yu for his contribution to the XPS data analysis. Thanks to Zhijie Wei, Bowen Niu, Bingbing Ma and Wenlu Liu for their contribution to the data analysis and making a drawing.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huntscha, S.; Hofstetter, T.B.; Schymanski, E.L.; Spahr, S.; Hollender, J. Biotransformation of benzotriazoles: Insights from transformation product identification and compound-specific isotope analysis. Environ. Sci. Technol. 2014, 48, 4435–4443. [Google Scholar] [CrossRef] [PubMed]
  2. Luan, J.; Huang, P.Q. Photophysical and photocatalytic properties of BiSnSbO6 under visible light irradiation. Materials 2018, 11, 491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Voutsa, D.; Hartmann, P.; Schaffner, C.; Giger, W. Benzotriazoles, alkylphenols and bisphenol A in bunicipal wastewaters and in the Glatt River, Switzerland. Environ. Sci. Pollut. Res. 2006, 13, 333–341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Liu, R.; Ruan, T.; Wang, T.; Song, S.; Guo, F.; Jiang, G. Determination of nine benzotriazole UV stabilizers in environmental water samples by automated on-line solid phase extraction coupled with high-performance liquid chromatography−tandem mass spectrometry. Talanta 2014, 120, 158–166. [Google Scholar] [CrossRef] [Green Version]
  5. Lai, H.; Ying, G.; Ma, Y.; Chen, Z.; Chen, F.; Liu, Y. Occurrence and dissipation of benzotriazoles and benzotriazole ultraviolet stabilizers in biosolid-amended soils. Environ. Toxicol. Chem. 2014, 33, 761–767. [Google Scholar] [CrossRef]
  6. Jia, Y.; Bakken, L.R.; Breedveld, G.D.; Aagaard, P.; Frostegård, Å. Organic compounds that reach subsoil may threaten groundwater quality: Effect of benzotriazole on degradation kinetics and microbial community composition. Soil Biol. Biochem. 2006, 38, 2543–2556. [Google Scholar] [CrossRef]
  7. Nakata, H.; Murata, S.; Filatreau, J. Occurrence and concentrations of benzotriazole UV stabilizers in marine organisms and sediments from the Ariake Sea, Japan. Environ. Sci. Technol. 2009, 43, 6920–6926. [Google Scholar] [CrossRef]
  8. Kim, J.; Ramaswamy, B.R.; Chang, K.; Isobe, T.; Tanabe, S. Multiresidue analytical method for the determination of antimicrobials, preservatives, benzotriazole UV stabilizers, flame retardants and plasticizers in fish using ultra high performance liquid chromatography coupled with tandem mass spectrometry. J. Chromatogr. A 2011, 1218, 3511–3520. [Google Scholar] [CrossRef]
  9. Hem, L.J.; Hartnik, T.; Roseth, R.; Breedveld, G.D. Photochemical degradation of benzotriazole. J. Environ. Sci. Health A 2003, 38, 471–481. [Google Scholar] [CrossRef]
  10. Hollingsworth, J.; Sierra-Alvarez, R.; Zhou, M.; Ogden, K.L.; Field, J.A. Anaerobic biodegradability and methanogenic toxicity of key constituents in copper chemical mechanical planarization effluents of the semiconductor industry. Chemosphere 2005, 59, 1219–1228. [Google Scholar] [CrossRef]
  11. Xu, J.; Li, L.; Guo, C.S.; Zhang, Y.; Wang, S.F. Removal of benzotriazole from solution by BiOBr photocatalysis under simulated solar irradiation. Chem. Eng. J. 2013, 221, 230–237. [Google Scholar] [CrossRef]
  12. Zhe, L.; Shirley, A.S.; Thomas, E.P.; Amila, O.D.S. Occurrence and fate of substituted diphenylamine antioxidants and benzotriazole UV stabilizers in various Canadian wastewater treatment processes. Water Res. 2017, 124, 158–166. [Google Scholar]
  13. Zhe, L.; Amila, O.D.S.; Thomas, E.P.; Cyril, J.C.; Gerald, R.T.; Mark, R.S.; Derek, C.G.M. Distribution, partitioning and bioaccumulation of substituted diphenylamine antioxidants and benzotriazole UV stabilizers in an Urban Creek in Canada. Environ. Sci. Technol. 2016, 50, 9089–9097. [Google Scholar]
  14. Javaid, R.; Qazi, U.Y. Catalytic oxidation process for the degradation of synthetic dyes: An overview. Inter. J. Env. Res. Pub. Heal. 2019, 16, 2066. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Aarthi, T.; Narahari, P.; Madras, G. Photocatalytic degradation of azure and sudan dyes using nano TiO2. J. Hazard. Mater. 2007, 149, 725–734. [Google Scholar] [CrossRef]
  16. Rahman, M.U.; Qazi, U.Y.; Hussain, T.; Nadeem, N.; Zahid, M.; Bhatti, H.N.; Shahid, I. Solar driven photocatalytic degradation potential of novel graphitic carbon nitride based nano zero-valent iron doped bismuth ferrite ternary composite. Opt. Mater. 2021, 120, 111408. [Google Scholar] [CrossRef]
  17. Tahmasebi, N.; Esmaeilpour, H.; Movahedifard, F.; Hakimyfard, A.; Moayeri, H. Fabrication of a novel MoO3/Cs3PMo12O40 composite for photocatalytic decolorization of rhodamine B. Mat. Sci. Semicon. Proc. 2021, 131, 105876. [Google Scholar] [CrossRef]
  18. Mokhtari, F.; Tahmasebi, N. Hydrothermal synthesis of W-doped BiOCl nanoplates for photocatalytic degradation of rhodamine B under visible light. J. Phys. Chem. Solids 2021, 149, 109804. [Google Scholar] [CrossRef]
  19. Subhan, M.A.; Rifat, T.P.; Saha, P.C.; Alam, M.M.; Asiri, A.M.; Rahman, M.M.; Akter, S.; Raihan, T.; Azad, A.K.; Uddin, J. Enhanced visible light-mediated photocatalysis, antibacterial functions and fabrication of a 3-chlorophenol sensor based on ternary Ag2O·SrO·CaO. RSC Adv. 2020, 10, 11274–11291. [Google Scholar] [CrossRef] [Green Version]
  20. Rizzo, L.; Della Sala, A.; Fiorentino, A.; Li Puma, G. Disinfection of urban wastewater by solar driven and UV lamp—TiO2 photocatalysis: Effect on a multi drug resistant Escherichia coli strain. Water Res. 2014, 53, 145–152. [Google Scholar] [CrossRef]
  21. Sakai, T.; Da Loves, A.; Okada, T.; Mishima, S. Titania/CnTAB nanoskeleton as adsorbent and photocatalyst for removal of alkylphenols dissolved in water. J. Hazard. Mater. 2013, 248, 487–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Kapoor, P.N.; Uma, S.; Rodriguez, S.; Klabunde, K.J. Aerogel processing of MTi2O5 (M = Mg, Mn, Fe, Co, Zn, Sn) compositions using single source precursors: Synthesis, characterization and photocatalytic behavior. J. Mol. Catal. A Chem. 2005, 229, 145–150. [Google Scholar] [CrossRef]
  23. Zhang, H.S.; Wen, Z.Y.; Zhao, Y.D.; Li, G.; Li, Z.J. Preparation, characterization of A2Ce2O7 (A = La and Gd) and their photo-catalytic properties. Energy Environ. Focus. 2015, 4, 324–329. [Google Scholar]
  24. Ladan, R.; Alireza, H.; Mohammad, S. High efficient solar light photocatalytic degradation of malachite green by solid state synthesized Bi2Sn2O7 and Bi2MxSn2O7 (M = Y3+, Eu3+, Gd3+ and Yb3+) nanomaterials. J. Nanoanal. 2020, 7, 1–17. [Google Scholar]
  25. Hebeish, A.A.; Abdelhady, M.M.; Youssef, A.M. TiO2 nanowire and TiO2 nanowire doped Ag-PVP nanocomposite for antimicrobial and self-cleaning cotton textile. Carbohyd. Polym. 2013, 91, 549–559. [Google Scholar] [CrossRef] [PubMed]
  26. Chen, J.; Wu, X.J.; Yin, L.; Li, B.; Hong, X.; Fan, Z.; Chen, B.; Xue, C.; Zhang, H. One-pot synthesis of CdS nanocrystals hybridized with single-layer transition-metal dichalcogenide nanosheets for efficient photocatalytic hydrogen evolution. Angew. ChemInt. Ed. Engl. 2015, 54, 1210–1214. [Google Scholar]
  27. Zhang, Z.; Li, A.; Cao, S.W.; Bosman, M.; Li, S.; Xue, C. Direct evidence of plasmon enhancement on photocatalytic hydrogen generation over Au/Pt-decorated TiO2 nanofibers. Nanoscale 2014, 6, 5217–5222. [Google Scholar] [CrossRef]
  28. Zhang, N.; Shi, J.; Mao, S.S.; Guo, L. Co3O4 quantum dots: Reverse micelle synthesis and visible-light-driven photocatalytic overall water splitting. Chem. Commun. 2014, 50, 2002–2004. [Google Scholar] [CrossRef]
  29. Wang, X.; Xu, Q.; Li, M.; Shen, S.; Wang, Y.; Feng, Z.; Shi, J.; Han, H.; Li, C. Photocatalytic overall water splitting promoted by an alpha-beta phase junction on Ga2O3. Angew. Chem. Int. 2012, 51, 13089–13092. [Google Scholar] [CrossRef]
  30. Huang, J.G.; Zhao, X.G.; Zheng, M.Y.; Li, S.; Wang, Y.; Liu, X.J. Preparation of N-doped TiO2 by oxidizing TiN and its application on phenol degradation. Water Sci. Technol. 2013, 68, 934–939. [Google Scholar] [CrossRef]
  31. Mohammadi, K.; Sadeghi, M.; Azimirad, R. Facile synthesis of SrFe12O19 nanoparticles and its photocatalyst application. J. Mater. Sci. Mater. EL 2017, 28, 10042–10047. [Google Scholar] [CrossRef]
  32. Mohammadi, K.; Sadeghi, M.; Azimirad, R.; Ebrahimi, M. Barium hexaferrite nanoparticles: Morphology-controlled preparation, characterization and investigation of magnetic and photocatalytic properties. J. Mater. Sci. Mater. EL 2017, 28, 9983–9988. [Google Scholar] [CrossRef]
  33. Shieh, D.L.; Lin, Y.S.; Yeh, J.H.; Chen, S.C.; Lin, B.C.; Lin, J.L. N-doped, porous TiO2 with rutile phase and visible light sensitive photocatalytic activity. Chem. Commun. 2012, 48, 2528–2530. [Google Scholar] [CrossRef] [PubMed]
  34. Zhou, X.; Lu, J.; Jiang, J.; Li, X.; Lu, M.; Yuan, G.; Wang, Z.; Zheng, M.; Seo, H.J. Simple fabrication of N-doped mesoporous TiO2 nanorods with the enhanced visible light photocatalytic activity. Nanoscale Res. Lett. 2014, 9, 34. [Google Scholar] [CrossRef] [Green Version]
  35. Chen, X.; Burda, C. The electronic origin of the visible-light absorption properties of C-, N- and S-doped TiO2 nanomaterials. J. Am. Chem. Soc. 2008, 130, 5018–5019. [Google Scholar] [CrossRef]
  36. Li, H.; Zhang, X.; Huo, Y.; Zhu, J. Supercritical preparation of a highly active S-doped TiO2 photocatalyst for methylene blue mineralization. Environ. Sci. Technol. 2007, 41, 4410–4414. [Google Scholar] [CrossRef]
  37. Li, Z.; Ding, D.; Ning, C. P-Type hydrogen sensing with Al-and V-doped TiO2 nanostructures. Nanoscale Res. Lett. 2013, 8, 25. [Google Scholar] [CrossRef] [Green Version]
  38. Pan, L.; Wang, S.; Zou, J.J.; Huang, Z.F.; Wang, L.; Zhang, X. Ti3+-defected and V-doped TiO2 quantum dots loaded on MCM-41. Chem. Commun. 2014, 50, 988–990. [Google Scholar] [CrossRef]
  39. Zhang, Z.; Shao, C.; Zhang, L.; Li, X.; Liu, Y. Electrospun nanofibers of V-doped TiO2 with high photocatalytic activity. J. Colloid. Interf. Sci. 2010, 351, 57–62. [Google Scholar] [CrossRef]
  40. Liu, G.; Han, C.; Pelaez, M.; Zhu, D.; Liao, S.; Likodimos, V.; Ioannidis, N.; Kontos, A.G.; Falaras, P.; Dunlop, P.S.; et al. Synthesis, characterization and photocatalytic evaluation of visible light activated C-doped TiO2 nanoparticles. Nanotechnology 2012, 23, 294003. [Google Scholar] [CrossRef]
  41. Yu, J.; Zhou, P.; Li, Q. New insight into the enhanced visible-light photocatalytic activities of B-, C- and B/C-doped anatase TiO2 by first-principles. Phys. Chem. Chem. Phys. 2013, 15, 12040–12047. [Google Scholar] [CrossRef]
  42. Zhang, W.J.; Ma, Z.; Du, L.; Li, H. Role of PEG4000 in sol-gel synthesis of Sm2Ti2O7 photocatalyst for enhanced activity. J. Alloy. Compd. 2017, 704, 26–31. [Google Scholar] [CrossRef]
  43. Wang, S.X.; Li, W.; Wang, S.; Jiang, J.M.; Chen, Z.H. Synthesis of well-defined hierarchical porous La2Zr2O7 monoliths via non-alkoxidesolegel process accompanied by phase separation. Micropor. Mesopor. Mat. 2016, 221, 32–39. [Google Scholar] [CrossRef]
  44. Zou, Z.; Ye, J.; Arakawa, H. Preparation, structural and photophysical properties of Bi2InNbO7 compound. J. Mater. Sci. 2000, 19, 1909–1911. [Google Scholar]
  45. Shin, H.; Byun, T.H. Effect of Pt loading onto ball-milled TiO2 on the visible-light photocatalytic activity to decompose rhodamine B. Res. J. Chem. Environ. 2013, 17, 41–46. [Google Scholar]
  46. Kanhere, P.; Shenai, P.; Chakraborty, S.; Ahuja, R.; Zheng, J.; Chen, Z. Mono- and co-doped NaTaO3 for visible light photocatalysis. Phys. Chem. Chem. Phys. 2014, 16, 16085–16094. [Google Scholar] [CrossRef] [Green Version]
  47. Daskalaki, V.M.; Antoniadou, M.; Li Puma, G.; Kondarides, D.I.; Lianos, P. Solar light-responsive Pt/CdS/TiO2 photocatalysts for hydrogen production and simultaneous degradation of inorganic or organic sacrificial agents in wastewater. Environ. Sci. Technol. 2010, 44, 7200–7205. [Google Scholar] [CrossRef]
  48. Luan, J.F.; Zhao, W.; Feng, J.W.; Cai, H.L.; Zheng, Z.; Pan, B.C.; Wu, X.S.; Zou, Z.G.; Li, Y.M. Structural, photophysical and photocatalytic properties of novel Bi2AlVO7. J. Hazard. Mater. 2009, 164, 781–789. [Google Scholar] [CrossRef]
  49. Xing, C.C.; Zhang, Y.; Liu, Y.P.; Wang, X.; Li, J.S.; Martinez-Alanis, P.R.; Spadaro, M.C.; Guardia, P.; Arbiol, J.; Llorca, J.; et al. Photodehydrogenation of ethanol over Cu2O/TiO2 heterostructures. Nanomaterials 2021, 11, 1399. [Google Scholar] [CrossRef]
  50. Stathi, P.; Solakidou, M.; Deligiannakis, Y. Lattice defects engineering in W-, Zr-doped BiVO4 by flame spray pyrolysis: Enhancing photocatalytic O2 evolution. Nanomaterials 2021, 11, 501. [Google Scholar] [CrossRef]
  51. Thomas, A.M.; Peter, J.; Nagappan, S.; Mohan, A.; Ha, C.S. Dual stimuli-responsive copper nanoparticles decorated SBA-15: A highly efficient catalyst for the oxidation of alcohols in water. Nanomaterials 2020, 10, 2051. [Google Scholar] [CrossRef]
  52. Zou, Y.; Shi, J.W.; Ma, D.; Fan, Z.; Lu, L.; Niu, C. In situ synthesis of C-doped TiO2@g-C3N4 core-shell hollow nanospheres with enhanced visible-light photocatalytic activity for H2 evolution. Chem. Eng. J. 2017, 322, 435–444. [Google Scholar] [CrossRef] [Green Version]
  53. Yu, Y.; Cao, C.Y.; Liu, H.; Li, P.; Wei, F.F.; Jiang, Y.; Song, W.G. A Bi/BiOCl heterojunction photocatalyst with enhanced electron-hole separation and excellent visible light photodegrading activity. J. Mater. Chem. A 2014, 2, 1677–1681. [Google Scholar] [CrossRef]
  54. Wang, P.Q.; Bai, Y.; Luo, P.Y.; Liu, J.Y. Ag2O/Ag3PO4 heterostructures: Highly efficient and stable visible-light-induced photocatalyst for degradation of methyl orange and phenol. Micro Nano Lett. 2013, 8, 340–344. [Google Scholar] [CrossRef]
  55. Tang, Y.; Tao, Y.; Zhou, T.; Yang, B.; Wang, Q.; Zhu, Z.; Xie, A.; Luo, S.; Yao, C.; Li, X. Direct Z-scheme La1-xCexMnO3 catalyst for photothermal degradation of toluene. Environ. Sci. Pollut. R 2019, 26, 36832–36844. [Google Scholar] [CrossRef]
  56. Wang, J.H.; Zou, Z.G.; Ye, J.H. Synthesis, structure and photocatalytic property of a new hydrogen evolving photocatalyst Bi2InTaO7. Mater. Sci. Forum. 2003, 423, 485–490. [Google Scholar] [CrossRef]
  57. 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]
  58. 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]
  59. Nowak, M.; Kauch, B.; Szperlich, P. Determination of energy band gap of nanocrystalline SbSI. Rev. Sci. Instrum. 2009, 80, 046107. [Google Scholar] [CrossRef]
  60. Zhou, F.; Kang, K.; Maxisch, T.; Ceder, G.; Morgan, D. The electronic structure and band gap of LiFePO4 and LiMnPO4. Solid State Commun. 2004, 132, 181–186. [Google Scholar] [CrossRef] [Green Version]
  61. Tauc, J.; Grigorov, R.; Vancu, A. Optical properties and electronic structure of amorphous germanium. Phys. Status Solidi. 1966, 15, 627–637. [Google Scholar] [CrossRef]
  62. Butler, M.A. Photoelectrolysis with YFeO3 electrodes. J. Appl. Phys. 1977, 48, 1914–1920. [Google Scholar] [CrossRef]
  63. Cui, B.Y.; Cui, H.T.; Li, Z.R.; Dong, H.Y.; Li, X.; Zhao, L.F.; Wang, J.W. Novel Bi3O5I2 hollow microsphere and its enhanced photocatalytic activity. Catalysts 2019, 9, 709. [Google Scholar] [CrossRef] [Green Version]
  64. Vallejo, W.; Cantillo, A.; Salazar, B.; Diaz-Uribe, C.; Ramos, W.; Romero, E.; Hurtado, M. Comparative study of ZnO thin films doped with transition metals (Cu and Co) for methylene blue photodegradation under visible irradiation. Catalysts. 2020, 10, 528. [Google Scholar] [CrossRef]
  65. Qazi, U.Y.; Javaid, R.; Zahid, M.; Tahir, N.; Afzal, A.; Lin, X.M. Bimetallic NiCo-NiCoO2 nano-heterostructures embedded on copper foam as a self-supported bifunctional electrode for water oxidation and hydrogen production in alkaline media. Int. J. Hydrogen Energ. 2021, 46, 18936–18948. [Google Scholar] [CrossRef]
  66. Qazi, U.Y.; Javaid, R.; Tahir, N.; Jamil, A.; Afzal, A. Design of advanced self-supported electrode by surface modification of copper foam with transition metals for efficient hydrogen evolution reaction. Int. J. Hydrogen Energ. 2020, 45, 33396–33406. [Google Scholar] [CrossRef]
  67. Deonikar, V.G.; Patil, S.S.; Tamboli, M.S.; Ambekar, J.D.; Kulkarni, M.V.; Panmand, R.P.; Umarji, G.G.; Shinde, M.D.; Rane, S.B.; Munirathnam, N.R.; et al. Growth study of hierarchical Ag3PO4/LaCO3OH heterostructures. Phys. Chem. Chem. Phys. 2017, 19, 20541–20550. [Google Scholar] [CrossRef]
  68. Patil, S.S.; Tamboli, M.S.; Deonikar, V.G.; Umarji, G.G.; Ambekar, J.D.; Kulkarni, M.V.; Kolekar, S.S.; Kale, B.B.; Patil, D.R. Magnetically separable Ag3PO4/NiFe2O4 composites with enhanced photocatalytic activity. Dalton Trans. 2015, 44, 20426–20434. [Google Scholar] [CrossRef]
  69. Jiang, L.B.; Yuan, X.Z.; Zeng, G.M.; Liang, J.; Chen, X.H.; Yu, H.B.; Wang, H.; Wu, Z.B.; Zhang, J.; Xiong, T. In-situ synthesis of direct solid-state dual Z-scheme WO3/g-C3N4/Bi2O3 photocatalyst for the degradation of refractory pollutant. Appl. Catal. B 2018, 227, 376–385. [Google Scholar] [CrossRef]
  70. 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, 158554. [Google Scholar] [CrossRef]
  71. Zhuang, Y.; Zhou, M.; Gu, J.; Li, X. Spectrophotometric and high performance liquid chromatographic methods for sensitive determination of bisphenol A. Spectrochim. Acta A 2013, 122, 153–157. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XPS survey spectrum of the GZHP.
Figure 1. XPS survey spectrum of the GZHP.
Catalysts 12 00159 g001
Figure 2. (a) XPS spectra of O2− and Sb5+ derived from the GZHP; (b) XPS spectra of Gd3+ derived from the GZHP; (c) XPS spectra of Y3+ and Bi3+ derived from the GZHP; (d) XPS spectra of Zn2+ derived from the GZHP; (e) XPS spectra of Nb5+ derived from the GZHP.
Figure 2. (a) XPS spectra of O2− and Sb5+ derived from the GZHP; (b) XPS spectra of Gd3+ derived from the GZHP; (c) XPS spectra of Y3+ and Bi3+ derived from the GZHP; (d) XPS spectra of Zn2+ derived from the GZHP; (e) XPS spectra of Nb5+ derived from the GZHP.
Catalysts 12 00159 g002
Figure 3. SEM photograph of Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst.
Figure 3. SEM photograph of Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst.
Catalysts 12 00159 g003
Figure 4. EDS elemental mapping of Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst (Gd, Y, Sb, O from Gd2YSbO7 and Zn, Bi, Nb, O from ZnBiNbO5).
Figure 4. EDS elemental mapping of Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst (Gd, Y, Sb, O from Gd2YSbO7 and Zn, Bi, Nb, O from ZnBiNbO5).
Catalysts 12 00159 g004
Figure 5. EDS spectrum of Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst.
Figure 5. EDS spectrum of Gd2YSbO7/ZnBiNbO5 heterojunction photocatalyst.
Catalysts 12 00159 g005
Figure 6. Concentration variation curves of benzotriazole during photocatalytic degradation of benzotriazole with GZH or Gd2YSbO7 as photocatalysts, or with ZnBiNbO5 or NTO as photocatalysts under VLIR.
Figure 6. Concentration variation curves of benzotriazole during photocatalytic degradation of benzotriazole with GZH or Gd2YSbO7 as photocatalysts, or with ZnBiNbO5 or NTO as photocatalysts under VLIR.
Catalysts 12 00159 g006
Figure 7. Concentration variation curves of total organic carbon during the photocatalytic degradation of benzotriazole in chemical industry wastewater with Gd2YSbO7/ZnBiNbO5 heterojunction, Gd2YSbO7, ZnBiNbO5 or NTO as a photocatalyst under VLIR.
Figure 7. Concentration variation curves of total organic carbon during the photocatalytic degradation of benzotriazole in chemical industry wastewater with Gd2YSbO7/ZnBiNbO5 heterojunction, Gd2YSbO7, ZnBiNbO5 or NTO as a photocatalyst under VLIR.
Catalysts 12 00159 g007
Figure 8. Concentration variation curves of benzotriazole during the photocatalytic degradation of benzotriazole in chemical industry wastewater with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR for three cycle degradation tests.
Figure 8. Concentration variation curves of benzotriazole during the photocatalytic degradation of benzotriazole in chemical industry wastewater with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR for three cycle degradation tests.
Catalysts 12 00159 g008
Figure 9. Concentration variation curves of total organic carbon during photocatalytic degradation of benzotriazole in chemical industry wastewater with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR for three cycle degradation tests.
Figure 9. Concentration variation curves of total organic carbon during photocatalytic degradation of benzotriazole in chemical industry wastewater with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR for three cycle degradation tests.
Catalysts 12 00159 g009
Figure 10. (a) Effect of different radical scavengers on the removal rate of benzotriazole with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR; (b) Effect of different radical scavengers such as benzoquinone (BQ), isopropanol (IPA) or ethylenediamine tetra-acetic acid (EDTA) on the removal efficiency of benzotriazole with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR.
Figure 10. (a) Effect of different radical scavengers on the removal rate of benzotriazole with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR; (b) Effect of different radical scavengers such as benzoquinone (BQ), isopropanol (IPA) or ethylenediamine tetra-acetic acid (EDTA) on the removal efficiency of benzotriazole with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR.
Catalysts 12 00159 g010
Figure 11. Possible photocatalytic degradation mechanism of benzotriazole with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR.
Figure 11. Possible photocatalytic degradation mechanism of benzotriazole with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst under VLIR.
Catalysts 12 00159 g011
Figure 12. Suggested photocatalytic degradation pathway scheme for benzotriazole under VLIR with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst.
Figure 12. Suggested photocatalytic degradation pathway scheme for benzotriazole under VLIR with Gd2YSbO7/ZnBiNbO5 heterojunction as photocatalyst.
Catalysts 12 00159 g012
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yao, Y.; Luan, J. Preparation, Property Characterization of Gd2YSbO7/ZnBiNbO5 Heterojunction Photocatalyst for Photocatalytic Degradation of Benzotriazole under Visible Light Irradiation. Catalysts 2022, 12, 159. https://doi.org/10.3390/catal12020159

AMA Style

Yao Y, Luan J. Preparation, Property Characterization of Gd2YSbO7/ZnBiNbO5 Heterojunction Photocatalyst for Photocatalytic Degradation of Benzotriazole under Visible Light Irradiation. Catalysts. 2022; 12(2):159. https://doi.org/10.3390/catal12020159

Chicago/Turabian Style

Yao, Ye, and Jingfei Luan. 2022. "Preparation, Property Characterization of Gd2YSbO7/ZnBiNbO5 Heterojunction Photocatalyst for Photocatalytic Degradation of Benzotriazole under Visible Light Irradiation" Catalysts 12, no. 2: 159. https://doi.org/10.3390/catal12020159

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