Next Article in Journal
Extraction and Analytical Techniques for Pharmaceuticals and Personal Care Products in Sediments: A Critical Review Towards Environmental Sustainability
Previous Article in Journal
Research on the Integration and Application of Industrial Architectural Heritage Information Under the Concept of Sustainability: A Case Study of the Architecture Building at Inner Mongolia University of Technology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Metronidazole Degradation via Visible Light-Driven Z-Scheme BiTmDySbO7/BiEuO3 Heterojunction Photocatalyst

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.
Sustainability 2025, 17(22), 10024; https://doi.org/10.3390/su172210024
Submission received: 19 September 2025 / Revised: 26 October 2025 / Accepted: 6 November 2025 / Published: 10 November 2025

Abstract

This study presented the successful synthesis of a visible light responsive Z-scheme BiTmDySbO7/BiEuO3 heterojunction photocatalyst (BBHP) via the hydrothermal method, exhibiting outstanding removal efficiency for degrading the metronidazole (MNZ) in wastewater. The BBHP exhibited exceptional photocatalytic activity during the degradation process of the MNZ which was a widely detected pharmaceutical pollutant in aquatic environments. The key to the high photocatalytic activity of the BBHP was the formation of a Z-scheme photogenerated carrier transport channel which existed between BiTmDySbO7 and BiEuO3 within the heterojunction structure. This innovative structural design was experimentally confirmed for enhancing the separation efficiency of the photogenerated charge carriers significantly, thereby, the efficient photocatalytic activity of the BBHP was promoted. After visible light irradiation for 130 min, the BBHP achieved a removal efficiency of 99.56% for degrading MNZ and a mineralization rate of 98.11% for removing the total organic carbon (TOC) concentration. In contrast to a single photocatalyst, the removal rate of the MNZ by using the BBHP was 1.14 times that by using the BiEuO3, 1.26 times that by using the BiTmDySbO7, and 2.65 times that by using the nitrogen-doped TiO2 (N-T) under visible light irradiation. The mineralization rate for removing the TOC concentration during the degradation process of the MNZ by using the BBHP was 1.17 times that by using the BiEuO3, 1.29 times that by using the BiTmDySbO7, and 2.86 times that by using the N-T under visible light irradiation. The photocatalytic degradation process of the MNZ by using the BBHP followed first-order kinetics model, concurrently, a dynamics rate constant of 0.0345 min−1 was obtained. Furthermore, the BBHP demonstrated excellent stability and durability in accordance with multiple cyclic degradation experiments. According to the capturing radicals experiments and the electron paramagnetic resonance test experiments, it was determined that the hydroxyl radicals (•OH) and the superoxide anions (•O2) played key role during the photocatalytic degradation process of the MNZ by using the BBHP under visible light irradiation. Finally, the intermediate products that were produced during the degradation process of the MNZ were analyzed by using liquid chromatography-mass spectrometer, as a result, a potential degradation pathway for the MNZ was proposed. Overall, this study could provide valuable references for future research on composite photocatalysts and effectively maintain the safety and sustainable utilization of water resource.

1. Introduction

The rapid expansion of industries such as chemical engineering, animal husbandry, and medical care has unfortunately led to a significant increase in antibiotic pollution, which is derived from water systems [1,2,3,4]. These pollutants posed a serious threat to human health and the ecological systems, primarily because of their chemical stability, diverse origins, and collective tendency within the food chain [5,6]. Metronidazole (MNZ), a nitroimidazole antibiotic that was widely used in the treatment of infections caused by anaerobic bacteria and protozoa, exemplified this challenge due to its potent bactericidal properties [7,8]. Despite its therapeutic benefits, MNZ was structurally stable, highly water-soluble, and difficult to be degraded by using traditional methods, often leading to its persistence in wastewater which was discharged from pharmaceutical facilities [9]. Even at low concentrations, MNZ has been linked to potential human carcinogenicity and could contribute to the rise in the resistance genes that derive from antibiotics in organisms [10]. Therefore, effective methods for degrading MNZ in wastewater are urgently needed to protect water resources and ecological systems.
Traditional approaches for treating wastewater containing MNZ, such as the biological degradation approach, the chemical oxidation approach, and the physical adsorption approach, often cause some drawbacks [11]. The aerobic biological degradation approach or the anaerobic biological degradation approach often suffers from slow reaction rates and could be greatly affected by the environment [12,13]. While chemical oxidation approach could be effective and typically involved high costs for chemical reagents, thus chemical oxidation approach might cause secondary pollution [14,15]. The physical adsorption approach was mainly used in the pretreatment stage, and it also suffered from inconsistent treatment outcomes [16]. These limitations highlighted a pressing need for the development of more solid and valid degradation methods that could overcome the shortcomings of current technologies.
The emergence of the high-level semiconductor photocatalytic technology offered a compelling pathway for sustainable water purification. This photocatalytic technology utilized light energy for driving chemical reactions, and the principle was primarily based on solid band theory [17,18]. When a semiconductor photocatalyst (SP) such as titanium dioxide (TiO2) was exposed to light irradiation with incident light energy higher than the band gap width of the SP, electrons would transit from the valence band (VB) of the SP to the conduction band (CB) of the SP. Accordingly, the photoinduced electrons stayed in the CB of the SP; concurrently, the holes would be left in the VB of the SP [19,20]. These photogenerated electrons and photoinduced holes then separated and migrated within or across the surface of the SP [21]. The electrons, acting as reducing agents, could convert oxygen into superoxide radicals [22,23]. Conversely, the holes that exhibited oxidizing properties could react with water or hydroxyl to produce highly reactive hydroxyl radicals. Ultimately, the generated radicals reacted with pollutants, breaking them down into innocuous chemical compounds such as carbon dioxide (CO2) and water (H2O) [24].
However, there was a fundamental challenge with single-component SP in balancing the utilization of light energy and the light stability. For instance, TiO2, which obtained widespread use, possessed a wide indirect bandgap value of approximately 3.2 eV [25], which significantly limited its effective utilization of the solar energy, resulting in a photon quantum efficiency of less than 4% [26,27,28]. Consequently, the practical applications often necessitated supplementary strong ultraviolet light sources for boosting their photocatalytic performance. In order to enhance solar utilization and improve photostability, developing efficient visible light-responsive photocatalysts was imperative for real-world deployment.
Recent research has found that the A2B2O7 compounds with perovskite structures exhibited excellent photocatalytic activity under visible light irradiation (VLI) [29,30,31]. For instance, Zhang et al. demonstrated that Sm2Ti2O7, which owns the pyrochlore structure, displayed excellent photocatalytic activity [32]. Similarly, Naceur et al. achieved a removal rate of 90.16% for degrading rhodamine B within 60 min by using Bi1.56Sb1.48Co0.96O7, which was synthesized via the solid-state reaction method [33]. On the basis of previous findings, we aimed to select more advantageous elements for synthesizing the A2B2O7 compounds. Murcia–López et al. found that the Bi-doped TiO2 catalyst possessed a higher efficiency for degrading phenol than the pure phase TiO2 catalyst under the condition of the same illumination time [34]. M. Mezyen et al. doped 5% Dy in SnO2, which resulted in the removal rate of 90.99% for degrading methylene blue within three hours under sunlight irradiation [35], consequently, Dy-doped SnO2 showed higher photocatalytic activity for degrading methylene blue compared with pure phase SnO2. Rao et al. observed that Tm-doped TiO2 exhibited improved degradation efficiency of acetaldehyde or o-xylene compared with the original pure phase TiO2 [36]. Gandelman et al. observed that Sb-doped TiO2 showed significant improvement in the photocatalytic degradation efficiency of acetaminophen compared to pure phase TiO2 [37]. The above-mentioned evidence led us to propose the BiTmDySbO7 photocatalyst as a novel candidate for a visible light-responsive photocatalyst. Ran et al. successfully prepared Eu doped TiO2. Compared with the TiO2, Eu doped TiO2 exhibited higher separation efficiency of the photogenerated electrons and the photogenerated holes [38].
In order to further enhance the photocatalytic activity of the visible light-responsive catalysts, the strategy of constructing the heterojunction photocatalysts has become a widely adopted and effective approach [39,40,41]. The heterojunction photocatalysts were formed by combining two or more photocatalysts with different band structures to create a heterostructure photocatalyst. Upon light irradiation on the surface of the heterojunction photocatalyst, the differences in the CB positions of different photocatalysts and the VB positions of different photocatalysts led to the redistribution of electrons and holes at the interface, aiming to improve the photocatalytic efficiency of different photocatalysts [42,43,44]. For example, Wang et al. reported that the degradation rate of metronidazole by using the Zn3In2S6/Bi2MoO6 heterojunction photocatalyst was 2.17 times that by using the single Zn3In2S6 photocatalyst [45]. Wang et al. constructed a ZnSnO3/CuBi2O4 heterojunction photocatalyst and found that the degradation rate of tetracycline by using ZnSnO3/CuBi2O4 heterojunction photocatalyst was 3.1 times higher than that by using the single CuBi2O4 [46].
Therefore, based on the appropriate band structure, we successfully prepared a novel Z-Scheme BiTmDySbO7/BiEuO3 heterojunction photocatalyst (BBHP) via the solvothermal method. The BBHP was designed for the efficient degradation of MNZ in wastewater. The BBHP exhibited excellent stability and sustainability; concurrently, the BBHP reduced long-term operational costs. Although the initial preparation cost of the BBHP was high, the BBHP demonstrated strong stability and reusability, thereby, the BBHP lowered the overall cost in light of a long-term application perspective. Recycling experiments for degrading the MNZ indicated that after five consecutive degradation cycles, the BBHP maintained a removal efficiency of 94.71% for degrading the MNZ and a total organic carbon mineralization efficiency of 93.39%, simultaneously, the photocatalytic activity of the BBHP did not decay significantly. Compared with adsorption materials, which required frequent replacement or oxidation technologies, which consumed large amounts of chemical reagents, the BBHP operated without the need for additional reagents; meanwhile, the BBHP was driven solely under the condition of visible light irradiation, as a result, using the BBHP could save electrical energy and thermal energy. Furthermore, the BBHP was reusable and could offer better cost-effectiveness for long-term use. Additionally, the BBHP resolved the performance limitations of traditional treatment technologies; correspondingly, the BBHP also resolved the performance limitations of the single-component catalyst. Ultimately, the BBHP was beneficial for maintaining the sustainable development of the ecological environment. We confirmed the strong potential of BBHP as a visible light response catalyst by comprehensively analyzing the crystal structure, morphological characteristics, and optical properties of BBHP. Our investigation thoroughly explored BBHP’s photocatalytic activity, stability, and the underlying mechanism for degrading MNZ. This work highlighted a promising pathway toward developing highly effective and durable solutions for complex pollutant removal.

2. Materials and Methods

2.1. Materials and Reagents

The following chemicals were purchased from Aladdin Reagent Co., Ltd. in Shanghai, China: bismuth (III) nitrate pentahydrate (Bi(NO3)3·5H2O, purity 99.99%), europium (III) nitrate hexahydrate (Eu(NO3)3·6H2O, purity 99.99%), thulium (III) nitrate hexahydrate (Tm(NO3)3·6H2O, purity 99.99%), dysprosium (III) nitrate hexahydrate (Dy(NO3)3·6H2O, purity 99.99%), antimony (V) pentachloride (SbCl5, purity 99.99%), ethanol (C2H5OH, purity ≥ 99.5%), ethylenediaminetetraacetic acid (EDTA, C10H16N2O8, purity ≥ 99.5%), isopropanol (IPA, C3H8O, purity ≥ 99.7%), p-benzoquinone (BQ, C6H4O2, purity ≥ 98.0%), metronidazole (C6H9N3O3, purity ≥ 99.8%), ultrapure water (18.25 MU-cm), Polyvinylpyrrolidone (PVP, (C6H6NO)n, purity ≥ 99.9%).

2.2. Preparation of BiEuO3

BiEuO3 photocatalyst was synthesized by the hydrothermal method. Bi(NO3)3·5H2O and Eu(NO3)3·6H2O were dissolved in deionized water with a stoichiometric ratio (Bi:Eu = 1:1). Next, the above mixed solution was stirred for 30 min to ensure complete dissolution. Subsequently, 2 mol/L sodium hydroxide solution was slowly added to adjust the pH value to 12. The resulting precursor solution was filled to 70% of the volume of the hydrothermal reactor, which was then well sealed. The sealed reactor was transferred to an oven and maintained at 180 °C for 1380 min. Following the reaction, the reactor was allowed to cool naturally up to room temperature. The precipitate was then collected and subjected to multiple washing cycles with deionized water and ethanol to remove any impurities. Finally, the thoroughly washed product was dried at 70 °C, followed by calcination in a muffle furnace at 600 °C for 4 h to enhance its crystallinity.

2.3. Preparation of BiTmDySbO7

BiTmDySbO7 photocatalyst was synthesized by the hydrothermal method. The precursor solution for the experiment was prepared using Bi(NO3)3·5H2O, Tm(NO3)3·6H2O, Dy(NO3)3·6H2O, SbCl5 and sodium hydroxide solution as raw materials. Subsequently, the reaction vessel was transferred to an oven and maintained at 200 °C for 1450 min. After the reaction, the product was collected, thoroughly washed, and dried at 70 °C. Ultimately, the dried product underwent calcination in a muffle furnace at 650 °C for 4 h to enhance its crystallinity.

2.4. Preparation of Nitrogen-Doped TiO2

The complete production process of the nitrogen-doped TiO2 was provided in the Supplementary Materials (Section S1).

2.5. Preparation of BiTmDySbO7/BiEuO3 Heterojunction Photocatalyst

The BiTmDySbO7/BiEuO3 heterojunction photocatalyst was prepared using a solvothermal method. Firstly, the prepared BiTmDySbO7 powder and BiEuO3 powder were mixed in a mass ratio of 1:1. Subsequently, ultrasonic dispersion was employed to treat the mixed powder and 0.5 g of polyvinylpyrrolidone in 30 mL of ethylene glycol for a duration of 30 min. The resulting well-dispersed mixed solution was shifted to a 50 mL high-pressure reaction vessel, which was sealed, and placed in an oven to react at 170 °C for 8 h. Following the reaction, the vessel was allowed to cool up to room temperature, and the product was centrifuged (8000 rpm, 10 min). Subsequently, the precipitate was alternately washed with deionized water and ethanol 3 times. Eventually, the precipitate was dried at 60 °C for 6 h to obtain the BiTmDySbO7/BiEuO3 heterojunction catalyst.

2.6. Feature Description

The specific feature description content was included in the Supplementary Materials (Section S2).

2.7. Photoelectrochemical Experiment

Electrochemical impedance spectroscopy (EIS) tests were conducted using a standard three-electrode system. It was composed of a platinum sheet counter electrode, an Ag/AgCl reference electrode, and a prepared glassy carbon working electrode.
Prior to the test, a suspension of the photocatalyst was formed by dispersing 2 mg of the synthesized photocatalyst in a small sample tube, which contained 200 μL of ethanol and 5 μL of Nafion reagent. The mixture underwent ultrasonic dispersion for 90 min to ensure thorough and homogeneous mixing. In the meantime, a 500 mL mixed solution, which contained 0.05 mol of potassium ferricyanide, 0.1 mol of potassium ferrocyanide, and 0.1 mol of potassium chloride, was prepared as the electrolyte for the test. The glassy carbon working electrode was carefully polished with polishing powder and deionized water to ensure optimal conductivity. Subsequently, 25 μL of the prepared catalyst suspension was precisely pipetted onto the conductive surface of the cleaned glassy carbon electrode and allowed to air dry for 25 min. The EIS measurements were then performed to quantify the interfacial charge transfer resistance of the photocatalyst. This analysis provided further insights into the electrochemical performance of the BBHP.

2.8. Experimental Setup and Procedure

The photocatalyst solution of BiTmDySbO7, BiEuO3, or BBHP was prepared at a concentration of 0.4 g/L and added to 480 mL of metronidazole (MNZ) solution for testing the adsorption ability and photodegradation activities. All photodegradation experiments were conducted under VLI and provided by a 500-watt xenon lamp, which was equipped with a 420-nanometer cutoff filter (Changtuo Co., Ltd., Beijing, China).
During the entire experiment, 5 mL samples were collected from the suspension every 15 min. The samples were centrifuged at 7000 rpm for 15 min to separate the catalyst. The concentration of MNZ in the separated liquid was detected by using a liquid chromatography-mass spectrometer (Thermo Fisher Scientific, Waltham, MA, USA, Thermo Quest LCQ Duo), which was equipped with a Beta Basic-C18 liquid chromatography column. A 10 μL aliquot of the supernatant was analyzed chromatographically at a flow rate of 0.5 mL/min for MNZ quantification.
In order to assess the mineralization efficiency during the degradation process of MNZ, total organic carbon (TOC) analysis was performed by using a Shimadzu Corporation (Kyoto, Japan) TOC analyzer. Potassium phthalate (KHC8H4O4) aqueous solution was used as the standard reference material for calibrating the TOC measurement. A calibration curve with a carbon concentration range of 0 mg/L to 100 mg/L was established by using potassium phthalate to ensure the accuracy and reliability of the TOC measurement.
For the purpose of characterizing the intermediate products of MNZ during the photocatalytic reaction process, a liquid chromatography mass spectrometry (LC-MS) was utilized. Following the photocatalytic reaction, 25 μL of the reaction solution was infused into the LC-MS system. A Beta Basic-C18 liquid chromatography column was used. Simultaneously, the mobile phase was methanol and water (60:40, v/v). The mass spectrometer scanning range was set from 50 m/z to 500 m/z. The wideband activation function was enabled to induce fragmentation of the intermediate products, thereby generating more structural information, which was crucial for the accurate identification of their chemical structures.

3. Results and Discussion

3.1. Characterization of Photocatalysts

3.1.1. X-Ray Diffraction (XRD) Pattern Analysis

Figure 1 shows the XRD patterns of BBHP, BiTmDySbO7, and BiEuO3. According to Figure 1, BiTmDySbO7 exhibited distinct diffraction peak at 29.27°, 33.93°, 48.72°, 57.81°, 60.79°, 71.19°, 79.07°, 81.45° or 91.26°, which corresponded to the crystal plane of (222), (400), (440), (622), (444), (800), (662), (840) or (844), respectively; meanwhile, BiEuO3 exhibited distinct diffraction peak at 28.18°, 32.16°, 46.71°, 55.53°, 58.41°, 68.44°, 75.38°, 77.88°, 86.95° or 93.76°, which corresponded to the crystal plane of (111), (200), (220), (311), (222), (400), (311), (420), (422) or (511). The diffraction peaks of BiTmDySbO7 and the diffraction peaks of BiEuO3 were all present in the diffraction peaks of BBHP, confirming that BiTmDySbO7 and BiEuO3 had successfully combined to form a heterojunction structure.
The weak impurity peak, which appeared at 30° in the XRD pattern of the BiTmDySbO7 derived from the sample, corresponds to the carrier glass sheet in Figure 1 and Figure 2a. The main raw material for the glass sheet was SiO2; thus, the weak impurity peak that appeared at 30° in the XRD pattern of the BiTmDySbO7 was the SiO2 peak. The Rietveld analysis of the XRD data for BiTmDySbO7 or BiEuO3 was conducted by using Materials Studio 2020 software. Figure 2a shows the Rietveld refinement results of BiTmDySbO7 according to the XRD data of BiTmDySbO7. Figure 2b displays the Rietveld refinement results of BiEuO3 according to the XRD data of BiEuO3. As shown in Figure 2a,b, the Rp value for BiTmDySbO7 or BiEuO3 was 7.91% or 4.87%, respectively, indicating that there was excellent agreement between the theoretical data and experimental data. The above results further confirmed that the BiTmDySbO7 or the BiEuO3 was a pure single-phase photocatalyst. The Rietveld refinement results of BiTmDySbO7 and BiEuO3 confirmed that BiTmDySbO7 possessed a cubic perovskite structure and BiEuO3 owned a face-centered cubic fluorite structure. Additionally, the space group of BiTmDySbO3 was Fd3m, and BiTmDySbO3 owned a lattice constant of 10.557 Å, concurrently, the lattice constant of BiEuO3 was 5.489 Å, and the space group of BiEuO3 was Fd3m. Atomic structure models, which are constructed based on these space groups, lattice constants, and refined atomic coordinates, are presented in Figure 2c,d for BiTmDySbO7 and BiEuO3. Table S1 provides the specific atomic coordinates and structural parameters of BiTmDySbO7. Table S2 [47] provides the specific atomic coordinates and structural parameters of BiEuO3.
Further examination of the BiTmDySbO7 crystal structure revealed the presence of the distorted MO6 octahedra (M = Dy3+ and Sb5+) [48,49]. This was due to the existence of two different A–O bond lengths (A = Bi3+ and Tm3+). One was the longer A–O (1) bond with a bond length of 2.286 Å, and the other was the shorter A–O (2) bond with a bond length of 1.988 Å. Additionally, the M–O–M (M = Dy3+ and Sb5+) bond angle within the BiTmDySbO7 crystal structure was measured to be 139.624°. The unique structure of BiTmDySbO7 resulted in a lower recombination rate of photogenerated charge carriers. Thus, the photocatalytic activity of BiTmDySbO7 was enhanced [50,51].

3.1.2. Fourier Transform Infrared Spectroscopy Analysis

Fourier transform infrared (FTIR) spectra were trialed to define the functional groups and chemical bonds that were present in BiTmDySbO7, BiEuO3, and BBHP. The FTIR spectra of BBHP, BiTmDySbO7, and BiEuO3 are shown in Figure 3. The analysis result, as shown in Figure 3, revealed the presence of characteristic vibrational modes corresponding to Bi–O, Dy–O, Tm–O, Eu–O, Sb–O, or Sb–O–Sb bond.
Specifically, bending vibrations which were attributed to the Bi–O bond were observed at 439 cm−1 and 485 cm−1 [52,53]. Stretching vibration of Dy–O, Eu–O, or Tm–O occurred at 516 cm−1, 595 cm−1, or 600 cm−1 [54,55,56]. In addition, the bending vibration of Sb–O or Sb–O–Sb occurred at 661 cm−1 or 745 cm−1 [57,58]. The broad peak at 3463 cm−1 corresponded to the stretching vibration of the O–H group, which derived from chemically adsorbed water molecules [59]. In the FTIR spectrum of the BiTmDySbO7, BiEuO3, or BBHP, differences were observed in OH- groups. The difference was primarily attributed to the varying hydrophilicity of the BBHP, BiTmDySbO7, or BiEuO3 and the distinct water adsorption capacities of the BBHP, BiTmDySbO7, or BiEuO3. The bending oscillation of the surface H–O–H group was related to the absorption peak at 1606 cm−1 [60]. The spectral band around 1361 cm−1 was due to the vibration of the C–H bond [61]. Additionally, variations which appeared in the C–H bonds were also noted in the FTIR spectrum of the BiTmDySbO7, BiEuO3, or BBHP. However, the absence of the C–H bonds was not a key characteristic for distinguishing the BiTmDySbO7, BiEuO3, and BBHP. The slight difference that derived from the C–H bond was related to the organic adsorbate, which appeared on the sample surface of the BiTmDySbO7, BiEuO3, or BBHP.

3.1.3. Raman Spectroscopy Analysis

The Raman spectra of BBHP, BiTmDySbO7, and BiEuO3 are presented in Figure 4. For BiTmDySbO7, seven distinct characteristic peaks were observed according to Figure 4. The vibrational modes of the Bi–O, Tm–O, and Dy–O bonds were identified from the characteristic peaks at 176 cm−1, 390 cm−1, and 727 cm−1 [62,63,64,65]. The stretching vibration of the Sb–O bond gave rise to the peak at 455 cm−1 or 501 cm−1 [66]. Furthermore, the Raman bands at 238 cm−1 and 295 cm−1 were characteristic of the Sb–O–Sb stretching vibration [66,67].
The Raman spectrum of BiEuO3 exhibited two major characteristic peaks. The Bi–O bond stretching vibration manifested as a peak at 135 cm−1, whereas the Eu–O bond stretching vibration gave rise to the peak at 627 cm−1 [68,69]. Moreover, the Raman spectrum of BBHP showed characteristic peaks at 135 cm−1, 176 cm−1, 238 cm−1, 298 cm−1, 390 cm−1, 455 cm−1, 501 cm−1, 627 cm−1, and 727 cm−1. These peaks were in perfect agreement with those peaks that existed in the Raman spectra of BiTmDySbO7 and BiEuO3. The above analysis result once again confirmed the successful preparation of BBHP.

3.1.4. Morphology and Elemental Analysis

Figure 5a shows the morphology image of the BBHP using transmission electron microscopy (TEM). The TEM morphology analysis results, which were obtained according to Figure 5a, indicated a close contact between the BiEuO3 particle and the BiTmDySbO7 particle. The TEM morphology image of the BBHP indicated the presence of a small number of the BiEuO3 particles, which were not attached to the BiTmDySbO7 particles or were free-standing. However, the TEM morphology analysis image of the BBHP was derived only from a specific localized area of the BBHP. The distribution of the BiEuO3 particles, which appeared in Figure 5a, was somewhat stochastic and did not fully represent the overall distribution of the BiEuO3 particles that existed in the BBHP. Based on the comprehensive characterization results, the BiEuO3 particles are generally uniformly distributed on the surface of the BiTmDySbO7. Figure 5b displays high-resolution TEM (HRTEM) images of BiTmDySbO7 and BiEuO3. The high-resolution TEM image of BiTmDySbO7 or BiEuO3 provided detailed insights into the crystal lattice. It revealed a lattice fringe spacing of 0.305 nm for BiTmDySbO7, which was consistent with the (222) crystal plane index of BiTmDySbO7. The crystal lattice fringe of BiEuO3 with a value of 0.316 nm was consistent with the (111) crystal plane index. Figure 5c manifests the TEM morphology image of the energy dispersive spectroscopy (EDS) measurement region, which belonged to BBHP. Figure 5d displays the EDS composition content distribution image of BBHP. Figure 5e is the EDS facial scanning element distribution image of BBHP. Figure 5c–e proved that the Bi element, Tm element, Dy element, Sb element, Eu element, and O element were uniformly distributed within BBHP. As shown in Figure 5d, the system peak, which was found from 8.5 KeV to 9.0 KeV, came from the copper mesh of the specimen stage, which was used for providing electroconductivity of the photocatalyst sample. The peak energy of the copper element was approximately 8.90 keV. Elemental surface analysis result indicated that the average atomic ratio of Bi, Tm, Dy, Sb, Eu, and O was approximately 823:411:412:411:413:4521, which was consistent with the result determined by spot scan of EDS. The atomic ratio of Bi, Tm, Dy, and Sb was calculated to be 0.98:0.99:1.01:1.00 in the BBHP sample. This precisely aligned with the proposed chemical formula and provided strong evidence for the successful and stoichiometric preparation of the BBHP.

3.1.5. X-Ray Photoelectron Spectroscopy Analysis

Figure 6a shows the full scan X-ray photoelectron spectroscopy (XPS) spectrum of BBHP. The Bi, Tm, Dy, Sb, Eu, or O were all identified in the full-scan XPS spectrum of BBHP, revealing the successful incorporation of BiEuO3 and BiTmDySbO7. The high-resolution XPS spectra of Bi 4f, Tm 4d, Dy 4d, Eu 3d, Sb 3d, and O 1s were presented in Figure 6b–f. The Bi 4f spectrum of BiTmDySbO7 exhibited two peaks at 169.58 and 164.96 eV in Figure 6b, corresponding to the Bi 4f7/2 and Bi 4f5/2 states. This was consistent with the +3 oxidation state of Bi [70,71]. A characteristic Bi 4f doublet was identified in BiEuO3, with the 4f7/2 and 4f5/2 located at 158.81 eV and 164.20 eV. In BBHP, the Bi 4f7/2 peak and Bi 4f5/2 peak shifted to 159.16 eV and 164.48 eV. The Tm 4d5/2 peak in BiTmDySbO7 was located at 177.46 eV according to Figure 6c. However, in BBHP, the Tm 4d5/2 peak shifted to 176.56 eV. Similarly, the Sb 3d5/2 peak and Sb 3d3/2 peak in BiTmDySbO7 were at 532.62 eV and 540.14 eV, respectively, while in BBHP, the Sb 3d5/2 peak and Sb 3d3/2 peak shifted to 531.89 eV and 539.49 eV according to Figure 6d. The Eu 3d3/2 peak of BiEuO3 was at 1165.23 eV, and the Eu 3d5/2 peak was at 1135.81 eV in accordance with Figure 6e. In BBHP, the Eu 3d3/2 peak and Eu 3d5/2 peak shifted to 1165.05 eV and 1135.47 eV. Table S3 shows the XPS peak position values of various elements that derive from the BiTmDySbO7 or the BBHP. Table S4 displays the XPS peak position values of various elements that originate from the BiEuO3 or the BBHP. From the analysis of the peak positions for different elements in the BiTmDySbO7, BiEuO3, and BBHP, it could be seen that the Bi 4f peak, Tm 4d peak, Dy 4d peak, and Sb 3d peak in BBHP all shifted to the right direction compared to those in BiTmDySbO7. The Bi 4f peak and Eu 3d peak in BBHP shifted to the left direction compared with those in BiEuO3. These systematic shifts within peak positions were confirmatory for the formation of complex electronic interactions within the BBHP, which were crucial for effectively enhancing the catalytic performance of BBHP [72].
Further insight into the chemical environment of oxygen was gained from the deconvoluted O 1s XPS spectra of BBHP, BiTmDySbO7, and BiEuO3 in accordance with Figure 6f. The BBHP, BiTmDySbO7, and BiEuO3 exhibited three distinct oxygen species. For BBHP, the peak, which is located at 529.32 eV, corresponds to the lattice oxygen. The peak, which is located at 530.06 eV, corresponds to the adsorbed oxygen. The peak, which is located at 530.91 eV, was consistent with the surface hydroxyl group [73,74]. Similarly, the peak located at 530.35 eV, 531.13 eV, or 531.77 eV in BiTmDySbO7 and the peak located at 529.93 eV, 531.05 eV, or 531.74 eV in BiEuO3 also corresponded to the above three oxygen species. The secondary phases were not detected within the XPS peaks of BBHP, BiTmDySbO7, and BiEuO3. These XPS results indirectly confirmed the existence of a heterostructure in the BBHP and further verified the chemical interaction between BiTmDySbO7 and BiEuO3.

3.1.6. Optical Properties Analysis

Figure 7a shows the ultraviolet and visible diffuse reflection spectra of BBHP, BiTmDySbO7, and BiEuO3. The optical absorption characteristic of BBHP, BiTmDySbO7, or BiEuO3 was investigated by using the ultraviolet and visible spectrophotometer. The analysis result of the diffuse reflection spectra, which were derived from Figure 7a, revealed that the absorption onset point of BBHP, BiTmDySbO7, or BiEuO3 was approximately at 570 nm, 445 nm, or 640 nm. The transition ranges of BBHP, BiTmDySbO7, and BiEuO3 were located within the visible light spectrum. Figure S1 shows the ultraviolet and visible diffuse reflection spectrum of the N-T. In contrast, the absorption onset of the N-T was observed at approximately 450 nm. These results indicated that the N-T exhibited a very limited capacity for visible light absorption.
Figure 7b displays the corresponding graphs of (αhν)1/2 and for BBHP, BiTmDySbO7, and BiEuO3. Combined with Equation (1), the band gap energy of BBHP, BiTmDySbO7, or BiEuO3 was calculated to preliminarily determine the visible light absorption capability of BBHP, BiTmDySbO7, or BiEuO3 [75,76,77]. The calculation results are shown in Figure 7b. The band gap energy of BBHP was 2.170 eV. Meanwhile, the band gap energy of BiTmDySbO7 or BiEuO3 was 2.784 eV or 1.933 eV. This indicated that the BBHP possessed a strong absorption capacity in the visible light region.
α h ν 1 2 = A h ν E g ,
In Equation (1), A was the absorbance factor, α was the absorption coefficient, Eg was the band gap energy, and ν was the photon energy.
Figure 7c shows the photoluminescence (PL) spectra of BBHP, BiTmDySbO7, and BiEuO3. Figure 7c revealed the recombination rate of the photoinduced electrons and the photoinduced holes (PEPH), and clarified the Difference in Photocatalytic activity for BBHP, BiTmDySbO7, and BiEuO3 [78,79]. A higher PL intensity corresponded to a higher recombination rate of PEPH, while a lower PL intensity indicated a lower recombination rate of PEPH; thus, a more efficient separation efficiency of PEPH was obtained when the PL intensity was low [80,81,82]. However, solely relying on the PL spectrum of the BBHP, BiTmDySbO7, or BiEuO3 for inferring that the BBHP possessed the highest separation efficiency of the PEPH during the photocatalytic degradation process of the MNZ was somewhat lacking in evidence. According to the results which were obtained by using multiple experimental methods, it was demonstrated that the BBHP possessed the highest separation efficiency of the PEPH. The PL experiment, TRPL experiment, photocurrent intensity variation experiment, and electrochemical impedance experiment collectively demonstrated that the separation efficiency of the PEPH by using the BBHP was superior to that by using BiTmDySbO7 or BiEuO3. As observed in Figure 7c, BBHP, BiTmDySbO7, and BiEuO3 exhibited the highest PL intensity in the wavelength range, which was from 460 nm to 480 nm. However, a significant difference in PL intensity of BBHP, BiTmDySbO7, and BiEuO3 was observed; therefore, the PL intensity of BBHP was lower than that of BiTmDySbO7 or BiEuO3. The above observation led to the conclusion that the BBHP owned a significantly accelerated separation rate of PEPH, which directly contributed to the strong photocatalytic performance of PEPH.
The time-resolved photoluminescence (TRPL) spectra of BBHP, BiEuO3, and BiTmDySbO7 are presented in Figure 7d–f. The average lifetime of the photoinduced electrons derived from BBHP, BiEuO3, or BiTmDySbO7 was calculated based on the data gained from Equation (2) [83,84]. Table S5 exhibits the time constant components and average fluorescence lifetime of BiEuO3, BiTmDySbO7, and BBHP.
The average lifetime, derived from BiTmDySbO7 or BiEuO3, was 2.201 ns or 4.771 ns. The average lifetime, derived from BBHP, was 9.485 ns. The lifetime of the photoinduced electrons derived from BBHP was extended compared with that derived from BiTmDySbO7 or BiEuO3. The above phenomenon provided strong evidence that PEPH, which originated from the BBHP, was effectively separated. The high separation efficiency of PEPH was a key factor for enhancing the photocatalytic efficiency of BBHP.
τ a v e = A 1 τ 1 2 + A 2 τ 2 2 / A 1 τ 1 + A 2 τ 2 ,
Figure 8a shows the effect of the photocurrent density on VLI time with BBHP, BiEuO3, or BiTmDySbO7 as the photoelectrode. The photocurrent response was a direct indicator of the density of PEPH [85,86]. The results achieved from Figure 8a revealed that the photocurrent density with BBHP as the photoelectrode was significantly higher than that with BiTmDySbO7 or BiEuO3 as the photoelectrode. The above observation was consistent with the enhanced separation efficiency of PEPH, which originated from the PL spectra results and TRPL spectra results.
Figure 8b displays the EIS of BiTmDySbO7, BiEuO3, and BBHP. In accordance with Figure 8b, the Nyquist plot analysis result of the EIS for BiTmDySbO7, BiEuO3, or BBHP was further consistent with the conclusion derived from the above PL spectra results, the above TRPL spectra results, or the above photocurrent density variation results. The smaller the radius which was observed from the Nyquist plot was, the smaller the resistance of charge transfer was [83]. The radius of the EIS for BBHP was smaller than that for BiTmDySbO7 or BiEuO3, indicating that the charge transfer resistance was smaller and the separation efficiency of PEPH is higher when using BBHP.

3.2. Examination of Photocatalytic Efficiency

Figure 9a shows the curves of Ct/C0 versus VLI time for the photocatalytic degradation of the MNZ by using BBHP, BiTmDySbO7, BiEuO3, or nitrogen-doped TiO2 (N-T). Ct meant the concentration of the MNZ at a specific VLI time; correspondingly, C0 referred to the initial concentration of the MNZ. Prior to the photodegradation of the MNZ, the MNZ solution was stirred in the dark for 45 min to establish the adsorption–desorption equilibrium. Figure S4 displays the relationship curve between the irradiation light wavelength and absorbance for the photocatalytic degradation of the MNZ by using the BBHP under different conditions of the visible light irradiation time. As shown in Figure S4, the maximal absorption wavelength of the MNZ was found to be 320 nm. Simultaneously, it could be found from Figure S4 that the absorbance intensity that corresponded to the MNZ concentration decreased gradually with the extension of the VLI time. The experimental results, which were exhibited in Figure S4, were consistent with the experimental results which is shown in Figure 9a. As shown in Figure 9a, the concentration of the MNZ remained essentially unchanged without the presence of a photocatalyst, indicating the resistance to natural photodegradation of the MNZ. In contrast, the newly synthesized BBHP demonstrated the most significant degradation efficiency of the MNZ compared with BiTmDySbO7, BiEuO3, or N-T. Figure 9b presents the fitting results of ln(Ct/C0) versus VLI time by using BBHP, BiEuO3, BiTmDySbO7, or N-T. As a result, an excellent correlation coefficient R2 was calculated to be 0.9966, 0.9981, 0.9947, or 0.9954 by using BBHP, BiEuO3, BiTmDySbO7, or N-T.
In order to assess the complete mineralization efficiency of the MNZ, the Reduction in TOC concentration was monitored. Figure 9c presents the curves of TOCt/TOC0 versus VLI time for the photocatalytic degradation of the MNZ by using BBHP, BiTmDySbO7, BiEuO3, or N-T. TOCt referred to the TOC concentration in the system at a specific VLI time; correspondingly, TOC0 meant the initial TOC concentration. BBHP exhibited a superior mineralization efficiency of the TOC concentration during the degradation process of the MNZ, and possessed the strongest ability for obtaining the highest photocatalytic degradation rate of the MNZ compared with BiTmDySbO7, BiEuO3, or N-T. Figure 9d shows the fitting results of ln(TOC0/TOCt) versus VLI time during the degradation process of MNZ by using BBHP, BiEuO3, BiTmDySbO7, or N-T. Above TOC data, which were gained during the degradation process of the MNZ by using BBHP, BiEuO3, BiTmDySbO7, or N-T, also fitted a first-order kinetic model, which was evidenced by the linear plots of ln(TOC0/TOCt) versus VLI time and the high correlation coefficient value of R2, which was achieved according to Figure 9d.
Figure 9e evaluates the photodegradation rate of the MNZ and the kinetic constants after VLI of 130 min by using BBHP, BiEuO3, BiTmDySbO7, or N-T. It could be found from Figure 9e that the degradation rate of the MNZ by using BBHP was significantly higher compared with that by using BiEuO3, BiTmDySbO7, or N-T. The removal rate of the MNZ was 99.56%, 87.22%, 79.22%, or 37.59% by using BBHP, BiEuO3, BiTmDySbO7, or N-T. The removal rate of the MNZ by using BBHP was 1.14 times that by using BiEuO3, 1.26 times that by using BiTmDySbO7, and 2.65 times that by using N-T. Four photocatalysts exhibit different removal rates for degrading the MNZ, with the descending order as follows: BBHP > BiEuO3 > BiTmDySbO7 > N-T. In addition, the reaction rate of degrading the MNZ was 3.61 × 10−9 mol·L−1·s−1, 3.12 × 10−9 mol·L−1·s−1, 2.77 × 10−9 mol·L−1·s−1, or 1.54 × 10−9 mol·L−1·s−1 by using BBHP, BiEuO3, BiTmDySbO7, or N-T. The irradiation photon flux was 4.76 × 10−6 Einstein L−1·s−1 after VLI was measured by a radiometer. In accordance with Equation (3), the calculated photonic efficiency (PHEY) value for degrading the MNZ was 0.0758%, 0.0655%, 0.0581% or 0.0323% by using BBHP, BiEuO3, BiTmDySbO7, or N-T. The first-order kinetic constant kC was calculated to be 0.0345 min−1, 0.0131 min−1, 0.0098 min−1, or 0.0031 min−1 during the degradation process of the MNZ by using BBHP, BiEuO3, BiTmDySbO7, or N-T. Obviously, the value of the kinetic constant kC, which was acquired during the degradation process of the MNZ by using BBHP, was higher compared with that by using BiEuO3, BiTmDySbO7, or N-T.
The PHEY was calculated by using Equation (3):
ϕ = R / I 0
The Φ was the PHEY (%), simultaneously, the R was the degradation velocity of the MNZ (mol·L−1·s−1), and the I0 was the irradiation photon flux (Einstein L−1 s −1).
Similarly, Figure 9f presents the mineralization efficiency of the TOC concentration and the kinetic constants after VLI of 130 min by using BBHP, BiEuO3, BiTmDySbO7, or N-T. The removal rate of the TOC concentration was 98.11%, 84.02%, 75.93% or 34.27% during the degradation process of the MNZ by using BBHP, BiEuO3, BiTmDySbO7, or N-T. The mineralization rate of the TOC concentration during the degradation process of MNZ by using BBHP was 1.17 times that by using BiEuO3, 1.29 times that by using BiTmDySbO7, and 2.86 times that by using N-T; thereby, after VLI of 130 min, BBHP demonstrated superior mineralization efficiency of the TOC concentration compared with BiEuO3, BiTmDySbO7, or N-T. These results unequivocally demonstrated that BBHP played a crucial role in degrading the MNZ and complete mineralization of the TOC concentration. When the BiTmDySbO7/BiEuO3 heterojunction, BiEuO3, BiTmDySbO7, or N-T was utilized as the catalyst for degrading the MNZ, the kinetic constant kTOC was 0.0257 min−1, 0.0118 min−1, 0.0093 min−1, or 0.0027 min−1. Therefore, the value of the kinetic constant kTOC during the degradation process of the MNZ by using the BBHP was the highest compared with that by using BiEuO3, BiTmDySbO7, or N-T.
In order to assess the long-term viability of the synthesized BBHP, a cyclic performance test was conducted. Figure 10a displays the effect of the MNZ concentration on VLI time during five cyclic experiments by using the BBHP. As shown in Figure 10a, the ratio variation of Ct/C0, which was derived from the contribution of the remaining MNZ concentration at facultative illumination time and the initial concentration, was monitored over multiple cyclic experiments. Figure 10b presents the fitting results of ln(C0/Ct) versus VLI time during the cyclic degradation process of the MNZ by using BBHP. The outcomes indicated that the above variation curves of ln(C0/Ct) versus VLI time conformed to the first-order kinetic model during the cyclic degradation process of the MNZ by using BBHP. Figure 10c displays the change curves of the TOC concentration during five cyclic experiments for degrading MNZ by using the BBHP. Figure 10d shows the fitting results of ln(TOC0/TOCt) versus VLI time during the cyclic degradation process of the MNZ by using BBHP. The kinetic analysis results of the cyclic degradation process of the MNZ by using BBHP are shown in Figure 10d. The kinetic analysis results confirmed that the changes in the kinetic constants, which were derived from the contribution of ln(TOC0/TOCt) and VLI time, continued to obey the first-order kinetic model across all cyclic experiments. Figure 10e displays the removal efficiencies of MNZ and the kinetic constants for degrading MNZ during five cyclic experiments by using the BBHP. After five cycles of photocatalytic degradation of the MNZ by using BBHP, the removal rate of the MNZ decreased by only 4.85%. Figure 10f presents the mineralization efficiency of the TOC concentration and the kinetic constants derived from the contribution of ln(TOC0/TOCt) and VLI time during five cyclic experiments by using the BBHP. After five cycles of photocatalytic degradation of the MNZ by using BBHP, the removal rate of the TOC concentration decreased by merely 4.72%. Figure 10e,f provided a clear summary of the performance for BBHP over five consecutive cyclic experiments. After five repeated photodegradation processes of MNZ, BBHP was able to maintain a degradation efficiency of 94.71% and a mineralization efficiency of 93.39%. Figure S2 shows the XRD spectrum of the BBHP after the quintic cyclic degradation process of the MNZ by using the BBHP under the condition of VLI. Therefore, the XRD pattern, which was shown in Figure S2, confirmed that the crystal structure of the BBHP remained unchanged after the quintic cyclic degradation process of the MNZ. Figure S3 presents the ultraviolet and visible diffuse reflection spectrum of the BBHP after the quintic cyclic degradation process of the MNZ by using the BBHP under the condition of VLI. The intrinsic transition edge of the absorption spectrum, which belonged to the BBHP, almost did not exhibit a change and remained within the visible light region. The above result provided strong evidence for the high durability and robust reusability of BBHP, confirming that BBHP possessed excellent stability for the sustainable removal of the MNZ from wastewater. Meanwhile, after VLI of 130 min, the removal rate of the MNZ by using BBHP was 1.14 times that by using the BiEuO3, 1.26 times that by using the BiTmDySbO7, and 2.65 times that by using the N-T. The above experimental results demonstrated that the BBHP possessed outstanding Z-scheme performance and photocatalytic activity.
For the purpose of identifying the key reactive species involved in the photocatalytic degradation of the MNZ by using BBHP, a series of radical scavenging experiments was conducted. BQ, IPA, or EDTA was utilized as a scavenger for capturing superoxide anions (•O2), hydroxyl radicals(•OH), or holes [87]. Figure 11a shows the change curves that derive from the effect of the MNZ concentration on the VLI time with the addition of the radical scavenger, such as IPA, BQ, or EDTA. Figure 11b presents the removal rate of the MNZ during the degradation process of the MNZ by using BBHP with IPA, BQ, or EDTA as scavengers under the condition of VLI. Due to the removal of the •OH by using the IPA, the removal rate for degrading the MNZ by using the BBHP had decreased by 47.90% under the condition of VLI in the free radical scavenging experiment. Additionally, the removal rate for degrading the MNZ by using the BBHP had decreased by 32.78% under the condition of VLI because the BQ captured the •O2 in the pharmacy wastewater reaction solution system. Eventually, the removal efficiency for degrading the MNZ by using the BBHP was reduced by 11.32% under the condition of VLI due to the scavenging of h+ by using the EDTA. The above results, which were derived from the free radical scavenging experiment, indicated that •OH, •O2, and h+ are involved in the degradation process of the MNZ by using the BBHP under the condition of VLI. The contribution descending order of the active species was as follows: •OH > •O2 > h+. The •OH played a major role compared with •O2 or h+ in the pharmacy wastewater reaction solution system. The OH underwent oxidation reaction with h+ for generating the •OH. However, the capture of the h+ by using the EDTA did not significantly reduce the removal rate of the MNZ. In the free radical scavenging experiment, most of the h+ was consumed for the oxidation of the OH to generate •OH, thereby, the consumed h+ which was utilized for generating the •OH was not captured by EDTA, as a result, only a small amount of the free h+ which was not used to oxidize OH for generating the •OH was captured by EDTA. Therefore, the removal rate for degrading the MNZ by using the BBHP decreased by only 11.31% after capturing the h+ with EDTA under the condition of VLI.
Figure 12 shows the electron paramagnetic resonance (EPR) spectrum of DMPO·O2 or DMPO·OH by using BBHP during the degradation process of the MNZ by using BBHP under the condition of VLI. It could be found from Figure 12 that the unique signal of •OH was detected in the EPR spectrum, which possessed a four-peak intensity ratio of 1:2:2:1, which indicated the generation of •OH during the degradation process of the MNZ by using BBHP under the condition of VLI [87,88]. Moreover, the signal peak of •O2 was also detected from the EPR spectrum, which had four peak intensity ratios of 1:1:1:1, which implied the formation of •O2 during the degradation process of the MNZ. Figure S5 shows the EPR spectrum of DMPO•O2 or DMPO•OH during the photocatalytic degradation process of the MNZ by using the BiEuO3 photocatalyst under the condition of VLI. According to the experimental results, which were gained from Figure S5, the detected •OH was observed and generated during the photocatalytic degradation process of the MNZ by using the BiEuO3 photocatalyst under the condition of VLI, in contrast, the detected •O2 was not generated during the photocatalytic degradation process of the MNZ by using the BiEuO3 photocatalyst under the condition of VLI. Additionally, Figure S6 displays the EPR spectrum of DMPO•O2 or DMPO•OH during the photocatalytic degradation process of the MNZ by using the BiTmDySbO7 photocatalyst under the condition of VLI. According to the experimental results, which were obtained from Figure S6, the detected •O2 was observed and generated during the photocatalytic degradation process of the MNZ by using the BiTmDySbO7 photocatalyst under the condition of VLI, by contrast, the detected •OH was not generated during the photocatalytic degradation process of the MNZ by using the BiTmDySbO7 photocatalyst under the condition of VLI. The EPR experimental results, which were obtained according to Figure 12, confirmed that •OH and •O2 played major roles during the photocatalytic degradation process of the MNZ by using BBHP, and provided crucial insight into the photocatalytic degradation mechanism.

3.3. Research on the Photocatalytic Mechanism of BBHP

In order to gain a deeper understanding of the degradation mechanism of the BBHP, the VB edge potential of BiEuO3 or BiTmDySbO7 was determined by using ultraviolet photoelectron spectroscopy (UPS). Figure 13a shows the UPS spectrum of BiTmDySbO7. Figure 13b displays the UPS spectrum of BiEuO3. As shown in Figure 13a, the starting value of the binding energy, which originated from BiTmDySbO7, was 1.048 eV; concurrently, the ending value of the binding energy, which derived from BiTmDySbO7, was 19.955 eV. At the same time, as shown in Figure 13b, the starting value of the binding energy, which originated from BiEuO3, was 0.721 eV; Meanwhile, the ending value of the binding energy, which derived from BiEuO3, was 18.711 eV. Given the excitation energy of 21.2 eV [89,90,91], the ionization potential of VB for BiTmDySbO7 was 2.293 eV; simultaneously, the ionization potential of VB for BiEuO3 was 3.215 eV. Combined with the band gap energy values, which were calculated according to the previous ultraviolet and visible diffuse reflection spectra that belonged to BiTmDySbO7 or BiEuO3, we could determine that the electrochemical potential of the CB for the BiTmDySbO7 photocatalyst or the BiEuO3 photocatalyst was −0.491 eV or 1.282 eV [91].
Figure 14 displays the schematic diagram of the photodegradation mechanism of the MNZ under the condition of VLI by using BBHP. The possible charge transfer pathways within the BBHP could be found according to Figure 14. Based on the determination value of the UPS spectrum, which was derived from BiTmDySbO7 or BiEuO3, the ionization potential of VB and the electrochemical potential of CB for BiTmDySbO7 or BiEuO3 were achieved; thus, it could be inferred from the above results that the heterojunction structure of BBHP might conform to either a Z-scheme or a Type II mechanism [92,93,94,95,96]. The Z-scheme heterojunction structure for BBHP was identified as the most plausible pathway. Based on the EPR experiments and the free radical capture experiments, it was determined that the primary active species which were responsible for degrading the MNZ by using the BBHP were •OH and •O2. In order to utilize the photoinduced electrons for reducing O2 into •O2, the CB potential of the photocatalyst had to be more negative than the standard reduction potential of −0.33 eV (vs. NHE). Similarly, for the sake of utilizing the photoinduced holes for oxidizing OH into •OH, the VB potential of the photocatalyst should be more positive than the standard oxidizing potential of 2.38 eV (vs. NHE). When BBHP absorbed visible light energy, the electrons were excited from the VB of BiEuO3 or BiTmDySbO7 to the CB of BiEuO3 or BiTmDySbO7. In the Z-scheme heterojunction structure for BBHP, the photogenerated electrons, which are located in the CB of BiEuO3, combine with holes, which are located in the VB of BiTmDySbO7. This specific recombination pathway effectively preserved the photogenerated electrons, which possessed the strongest reducing power, and the holes, which possessed the strongest oxidizing power, which played a crucial role in improving the photocatalytic activity of BBHP. This left the electrons in the CB of BiTmDySbO7, and left the holes in the VB of BiEuO3 for driving the redox reactions. The CB potential of BiTmDySbO7 was −0.491 eV, which was more negative than the standard potential of O2/•O2 (−0.33 eV vs. NHE) [97]. The above potential was sufficient to reduce oxygen for forming •O2. Concurrently, the VB potential of BiEuO3 was 3.215 eV, which was more positive than the standard potential of OH/•OH (2.38 eV vs. NHE), allowing the holes that existed in the VB of BiEuO3 to oxidize hydroxide ions for generating highly reactive •OH [98,99]. The VB potential of the BiTmDySbO7 was 2.293 eV, which was insufficient for generating the •OH through the photocatalytic oxidative reaction with the photoinduced holes. The CB potential of the BiEuO3 was 1.282 eV, which could not reduce O2 for forming •O2 by using the photoinduced electrons. Consequently, when the BBHP absorbed visible light, the photoinduced electrons migrated from the CB position of the BiEuO3 to the VB position of the BiTmDySbO7 and recombined with the photoinduced holes, which existed in the VB position of the BiTmDySbO7. The generation of these key active species, such as •O2 and •OH, was consistent with our experimental findings, which were derived from the free radical scavenging experimental results and the EPR experimental results. Alternatively, a Type II mechanism for BBHP would suggest that the photogenerated electrons are transferred from the higher-energy CB of BiTmDySbO7 to the lower-energy CB of BiEuO3. However, this pathway would not produce the highly reactive radical species, which were observed in our experiments; as a result, the Z-scheme for BBHP became the more probable mechanism.

3.4. Degradation Pathway of the MNZ

In the light of the LC-MS results, three possible degradation pathways for the MNZ (m/z = 171) were proposed. The detailed path diagram is shown in Figure 15 under the condition of VLI by using BBHP. Figure S7 indicates the mass spectra and the retention time (rt) of the possible intermediate products for the metronidazole during the photocatalytic degradation process of the metronidazole by using the BBHP under the condition of VLI. The degradation of the MNZ was initiated by the attack of the reactive species, which was produced by using BBHP; thereby, the attack of the reactive species could lead to transformations such as hydroxyethyl cleavage, denitration, and oxidation of the MNZ molecule [100,101,102,103,104]. The first degradation pathway of the MNZ was initiated by hydroxyethyl cleavage, which formed an intermediate product with an m/z = 127. This intermediate product then underwent ring opening to generate oxalic acid (m/z = 91). The second degradation pathway of the MNZ involved the denitration of MNZ, resulting in the formation of 1H-imidazole-2,5-diol (m/z = 101). This intermediate product subsequently underwent hydroxyethyl cleavage; as a result, the oxalic acid was directly formed. The third degradation pathway of the MNZ began with the oxidation of MNZ to form (2-methyl-5-nitroimidazol-1-yl) acetic acid (m/z = 185). This product then followed one of two routes: either denitration played a crucial role in producing 2-(5-hydroxy-2-methyl-1H-imidazol-1-yl) acetic acid (m/z = 157), or ring opening played an important role in producing acetylglycine (m/z = 118). Ultimately, all intermediate products, including oxalic acid, 2-(5-hydroxy-2-methyl-1H-imidazol-1-yl) acetic acid, and acetylglycine, were further degraded into the final mineralized products: carbon dioxide (CO2), water (H2O), and nitrate (NO3) [105,106,107,108,109].

4. Conclusions

In this study, a novel Z-scheme BiTmDySbO7/BiEuO3 heterojunction photocatalyst was successfully synthesized via a hydrothermal method. Comprehensive characterization by using XRD, FTIR, Raman spectroscopy, XPS, and various microscopy techniques such as TEM, HRTEM, and EDS confirmed the successful synthesis of the BiTmDySbO7/BiEuO3 heterojunction and the existence of a Z-scheme structure, indicating that the BBHP possessed high photocatalytic activity potential. The superior charge carrier dynamics of the BBHP were verified by the photoelectrochemical analysis results and the optical analysis results, including the PL spectra results, the TRPL spectra results, the ultraviolet and visible diffuse reflection spectra results, the photocurrent intensity variation results, and the EIS results, obviously, above experimental results provided strong evidence for the high photocatalytic activity of the BBHP. Under VLI, the BBHP demonstrated exceptional photocatalytic activity for the degradation of the metronidazole compared with the BiTmDySbO7 or the BiEuO3. After VLI of 130 min, the degradation rate of the MNZ by using the BBHP was found to be 1.14 times that by using the BiEuO3, 1.26 times that by using the BiTmDySbO7, and 2.65 times that by using the N-T. Similarly, the mineralization rate for removing the TOC concentration during the photocatalytic degradation process of the MNZ by using the BBHP was 1.17 times that by using the BiEuO3, 1.29 times that by using the BiTmDySbO7, and 2.86 times that by using the N-T. Both the degradation process of the MNZ and the mineralization process for removing the TOC concentration by using the BBHP, the BiTmDySbO7, or the BiEuO3 conformed to a first-order kinetic model. Furthermore, the cyclic experiments by using the BBHP confirmed that the BBHP possessed strong stability and reusability. Studies that included the radical scavenging experiments and the EPR test experiments clarified that the •OH and the •O2 were the main active species that provided a dedication for the photocatalytic degradation of the MNZ. Eventually, in accordance with the LC-MS results and the existing research results, three potential degradation pathways for the MNZ were proposed. In conclusion, the BBHP not only offered a highly effective solution for purifying the wastewater which was contaminated by MNZ, but also represented a major devotion that was related to the progress of the Z-scheme photocatalyst for environmental applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su172210024/s1, Table S1: Structural parameters of BiTmDySbO7; Table S2: Structural parameters of BiEuO3; Table S3: XPS peak position values of various elements deriving from the BiTmDySbO7 and the BBHP; Table S4: XPS peak position values of various elements deriving from the BiEuO3 and the BBHP; Table S5: The time constant components and average fluorescence lifetime of BiEuO3, BiTmDySbO7 and BBHP; Section S1: Preparation of Nitrogen-Doped TiO2; Section S2: Feature Description; Figure S1: The ultraviolet and visible diffuse reflection spectrum of the N-doped TiO2; Figure S2: The XRD spectrum of the BBHP after quintic cyclic degradation process of the metronidazole by using the BBHP under the condition of visible light irradiation; Figure S3: The ultraviolet and visible diffuse reflection spectrum of the BBHP after quintic cyclic degradation process of the metronidazole by using the BBHP under the condition of visible light irradiation; Figure S4: The relationship curve between the irradiation light wavelength and absorbance for the photocatalytic degradation of the metronidazole by using the BBHP under the different condition of the visible light irradiation time; Figure S5: The EPR spectrum of DMPO•O2 or DMPO•OH during the photocatalytic degradation process of the metronidazole by using the BiEuO3 photocatalyst under the condition of visible light irradiation; Figure S6: The EPR spectrum of DMPO•O2 or DMPO•OH during the photocatalytic degradation process of the metronidazole by using the BiTmDySbO7 photocatalyst under the condition of visible light irradiation; Figure S7: Mass spectra and retention time (rt) of the possible intermediate products for the metronidazole during the photocatalytic degradation process of the metronidazole by using the BBHP under the condition of visible light irradiation.

Author Contributions

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

Funding

Funding from the Steeple-Crowned Talent Development Fund of Department of Human Resource and Social Security of Jilin Province of China (GrantNo. JiCaiSheZhi [2024] 0451).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Guo, Y.H.; Peng, B.; Liao, J.G.; Cao, W.C.; Liu, Y.J.; Nie, X.D.; Li, Z.W.; Rui, O.Y. Recent advances in the role of dissolved organic matter during antibiotics photodegradation in the aquatic environment. Sci. Total Environ. 2024, 916, 170101. [Google Scholar] [CrossRef] [PubMed]
  2. Chamkal, N.; Lhlou, I.; Bandadi, L.; Ounine, K. Hospital Antibiotics Usage: Environmental Hazard and Promotion of Antibiotic Resistant Bacteria. Ann. Ig. Med. Prev. Comunita 2022, 34, 266–278. [Google Scholar]
  3. Chen, J.; Xiang, Q.; Wu, J.Y.; Huang, X.B.; Wang, C.; Wei, D.Q.; Lv, Y. Different Effects of Antibiotics on Klebsiella pneumoniae and Escherichia coli Resistance Induced by Antibiotics: A Retrospective Study from China. Microb. Drug Resist. 2022, 28, 660–669. [Google Scholar] [CrossRef] [PubMed]
  4. Le, T.H.; Truong, T.; Tran, L.-T.; Nguyen, D.-H.; Pham, T.P.T.; Ng, C. Antibiotic resistance in the aquatic environments: The need for an interdisciplinary approach. Int. J. Environ. Sci. Technol. 2023, 20, 3395–3408. [Google Scholar] [CrossRef]
  5. Sampaio, C.P.P.; Biondo-Simoes, M.D.P.; Trindade, C.T.; Olandowski, M.; Matias, J.E.F. Metronidazole concentration in the bloodstream following its topical application, at different concentration levels, on experimental skin wounds during healing by secondary intention. Acta Cir. Bras. 2019, 34, e20190010000004. [Google Scholar] [CrossRef]
  6. Acikan, I.; Sayeste, E.; Bozoglan, A.; Artas, G.; Isayev, A.; Kirtay, M.; Ozercan, I.H.; Yaman, F.; Dundar, S.; Icen, V. Evaluation of the Effects of Topical Application of Chlorhexidine, Ozone, and Metronidazole on Palatal Wound Healing: A Histopathological Study. J. Craniofac. Surg. 2022, 33, 1929–1933. [Google Scholar] [CrossRef]
  7. Rafiq, M.; Ullah, R.; Ahmed, A.; Usman, M.; Yu, B.; Shen, Y.Q.; Cong, H.L. Highly efficient adsorption assisted photocatalytic decontamination of metronidazole by Co3O4 doped graphitic carbon nitrides. Ferroelectrics 2024, 618, 1715–1730. [Google Scholar] [CrossRef]
  8. Vasseghian, Y.; Dragoi, E.N.; Almomani, F.; Le, V.T. Graphene-based materials for metronidazole degradation: A comprehensive review. Chemosphere 2022, 286, 131727. [Google Scholar] [CrossRef]
  9. Banu, T.; Rashid, M.H.; Tofail, S.A.M.; Ul Haq, E.; Gulshan, F. Photocatalytic degradation of metronidazole (MNZ) antibiotic by CuO nanoparticles for environmental protection from pharmaceutical pollution. Surf. Interface Anal. 2023, 55, 430–436. [Google Scholar] [CrossRef]
  10. Saha, M.; Sarkar, A. Review on Multiple Facets of Drug Resistance: A Rising Challenge in the 21st Century. J. Xenobiot. 2021, 11, 197–214. [Google Scholar] [CrossRef] [PubMed]
  11. Thai, V.; Dang, V.D.; Thuy, N.T.; Pandit, B.; Vo, T.K.Q.; Khedulkar, A.P. Fluoroquinolones: Fate, effects on the environment and selected removal methods. J. Clean. Prod. 2023, 418, 137762. [Google Scholar] [CrossRef]
  12. Bueno, I.; He, H.; Kinsley, A.C.; Ziemann, S.J.; Degn, L.R.; Nault, A.J.; Beaudoin, A.L.; Singer, R.S.; Wammer, K.H.; Arnold, W.A. Biodegradation, photolysis, and sorption of antibiotics in aquatic environments: A scoping review. Sci. Total Environ. 2023, 897, 165301. [Google Scholar] [CrossRef]
  13. Zhu, T.T.; Su, Z.X.; Lai, W.X.; Zhang, Y.B.; Liu, Y.W. Insights into the fate and removal of antibiotics and antibiotic resistance genes using biological wastewater treatment technology. Sci. Total Environ. 2021, 776, 145906. [Google Scholar] [CrossRef]
  14. Jin, Q.; Liu, W.; Dong, Y.B.; Lu, Y.R.; Yang, C.; Lin, H. Single atom catalysts for degradation of antibiotics from aqueous environments by advanced oxidation processes: A review. J. Clean. Prod. 2023, 423, 138688. [Google Scholar] [CrossRef]
  15. Zheng, N.; Shi, J.T.; Nie, L.J.; Xue, K.K.; Gao, Y.H.; Shi, J.H. Photo-Fenton degradation of oxytetracycline by g-C3N4/CQDs/FeOCl with in-situ hydrogen peroxide production: Degradation pathway and toxicity analysis of intermediate products. J. Water Process. Eng. 2025, 69, 106615. [Google Scholar] [CrossRef]
  16. Nasiri, A.; Golestani, N.; Rajabi, S.; Hashemi, M. Facile and green synthesis of recyclable, environmentally friendly, chemically stable, and cost-effective magnetic nanohybrid adsorbent for tetracycline adsorption. Heliyon 2024, 10, e24179. [Google Scholar] [CrossRef]
  17. Li, J.H.; Zhao, W.; Guo, Y.; Wei, Z.B.; Han, M.S.; He, H.; Yang, S.G.; Sun, C. Facile synthesis and high activity of novel BiVO4/FeVO4 heteroj unction photocatalyst for degradation of metronidazole. Appl. Surf. Sci. 2015, 351, 270–279. [Google Scholar] [CrossRef]
  18. Derikvandi, H.; Nezamzadeh-Ejhieh, A. Designing of experiments for evaluating the interactions of influencing factors on the photocatalytic activity of NiS and SnS2: Focus on coupling, supporting and nanoparticles. J. Colloid Interface Sci. 2017, 490, 628–641. [Google Scholar] [CrossRef]
  19. Derikvandi, H.; Nezamzadeh-Ejhieh, A. Synergistic effect of p-n heterojunction, supporting and zeolite nanoparticles in enhanced photocatalytic activity of NiO and SnO2. J. Colloid Interface Sci. 2017, 490, 314–327. [Google Scholar] [CrossRef] [PubMed]
  20. Boxi, S.S.; Paria, S. Visible light induced enhanced photocatalytic degradation of organic pollutants in aqueous media using Ag doped hollow TiO2 nanospheres. RSC Adv. 2015, 5, 37657–37668. [Google Scholar] [CrossRef]
  21. Hu, X.; Hu, X.J.; Peng, Q.Q.; Zhou, L.; Tan, X.F.; Jiang, L.H.; Tang, C.F.; Wang, H.; Liu, S.H.; Wang, Y.Q.; et al. Mechanisms underlying the photocatalytic degradation pathway of ciprofloxacin with heterogeneous TiO2. Chem. Eng. J. 2020, 380, 122366. [Google Scholar] [CrossRef]
  22. Guo, J.F.; Li, S.M.; Duan, L.; Guo, P.; Li, X.J.; Cui, Q.Q.; Wang, H.; Jiang, Q. Preparation of Si doped molecularly imprinted TiO2 photocatalyst and its degradation to antibiotic wastewater. Integr. Ferroelectr. 2016, 168, 170–182. [Google Scholar] [CrossRef]
  23. Zhou, Y.W.; Cai, F.; Li, W.; Qiu, J.; Xu, L.; Liu, Q.Y.; Zhang, Y.L. Photocatalytic degradation of a typical macrolide antibiotic roxithromycin using polypropylene fibre sheet supported N-TiO2/graphene oxide composite materials. Environ. Technol. 2023, 44, 3354–3366. [Google Scholar]
  24. Rameel, M.I.; Wali, M.; Al-Humaidi, J.Y.; Liaqat, F.; Khan, M.A. Enhanced photocatalytic degradation of levofloxacin over heterostructured C3N4/Nb2O5 system under visible light. Heliyon 2023, 9, e20479. [Google Scholar] [CrossRef]
  25. Wen, X.J.; Niu, C.G.; Zhang, L.; Liang, C.; Zeng, G.M. A novel Ag2O/CeO2 heterojunction photocatalysts for photocatalytic degradation of enrofloxacin: Possible degradation pathways, mineralization activity and an in depth mechanism insight. Appl. Catal. B-Environ. Energy 2018, 221, 701–714. [Google Scholar] [CrossRef]
  26. Ahmed, Y.; Dutta, K.R.; Akhtar, P.; Hossen, M.A.; Alam, M.J.; Alharbi, O.A.; Almohamadi, H.; Mohammad, A.W. Emerging strategies in the sustainable removal of antibiotics using semiconductor-based photocatalysts. Beilstein J. Nanotechnol. 2025, 16, 264–285. [Google Scholar] [CrossRef]
  27. Khan, J.A.; Ahamad, S.; Ansari, M.A.H.; Tauqeer, M.; Park, C.H.; Park, J.P.; Choi, C.H.; Mohammad, A. State-of-the-art in ZnS-based nanoarchitects for visible-light photocatalytic degradation of antibiotics and organic dyes. J. Water Process. Eng. 2024, 67, 106151. [Google Scholar] [CrossRef]
  28. Gong, Y.Y.; Wu, Y.J.; Xu, Y.; Li, L.; Li, C.; Liu, X.J.; Niu, L.Y. All-solid-state Z-scheme CdTe/TiO2 heterostructure photocatalysts with enhanced visible-light photocatalytic degradation of antibiotic waste water. Chem. Eng. J. 2018, 350, 257–267. [Google Scholar] [CrossRef]
  29. Fan, G.D.; Zhou, J.F.; Ruan, F.Y.; Li, Y.; Tian, H.Y.; Fan, D.; Chen, Q.Q.; Li, N. The Z-scheme photocatalyst S-BiOBr/Bi2Sn2O7 with 3D/0D interfacial structure for the efficient degradation of organic pollutants. Sep. Purif. Technol. 2023, 309, 123099. [Google Scholar] [CrossRef]
  30. Lv, D.D.; Zhang, D.F.; Sun, Q.Z.; Wu, J.D.; Zhang, L.; Pu, X.P.; Ma, H.Y.; Dou, J.M. A novel method for the synthesis of BiOCl/Bi2Sn2O7 heterojunction photocatalysts with enhanced visible light photocatalytic properties. Nanotechnology 2016, 27, 385602. [Google Scholar] [CrossRef] [PubMed]
  31. Khodabakhsh, M.; Yilmaz, B.; Firoozi, S.; Haghshenas, D.F.; Unal, U. Enhanced Photocatalytic Properties of Restacked Unilamellar [SrTa2O7]2− Nanosheets of Aurivillius Phase Layered Perovskites. ACS Omega 2023, 8, 10607–10617. [Google Scholar] [CrossRef]
  32. Zhang, W.J.; Ma, Z.; Du, L.; Li, H. Role of PEG4000 in sol-gel synthesis of Sm2Ti2O7 photocatalyst for enhanced activity. J. Alloys Compd. 2017, 704, 26–31. [Google Scholar] [CrossRef]
  33. Naceur, B.; Abdelkader, E.; Nadjia, L.; Sellami, M.; Noureddine, B. Synthesis and characterization of Bi1.56Sb1.48Co0.96O7 pyrochlore sun-light-responsive photocatalyst. Mater. Res. Bull. 2016, 74, 491–501. [Google Scholar] [CrossRef]
  34. Murcia-López, S.; Hidalgo, M.C.; Navío, J.A. Synthesis characterization and photocatalytic activity of Bi-doped TiO2 photocatalysts under simulated solar irradiation. Appl. Catal. A-Gen. 2011, 404, 59–67. [Google Scholar] [CrossRef]
  35. Mezyen, M.; El Fidha, G.; Bitri, N.; Harrathi, F.; Ly, I.; Llobet, E. Visible light activated SnO2: Dy thin films for the photocatalytic degradation of methylene blue. RSC Adv. 2023, 13, 31151–31166. [Google Scholar] [CrossRef] [PubMed]
  36. Rao, Z.P.; Shi, G.S.; Wang, Z.; Mahmood, A.; Xie, X.F.; Sun, J. Photocatalytic degradation of gaseous VOCs over Tm3+-TiO2: Revealing the activity enhancement mechanism and different reaction paths. Chem. Eng. J. 2020, 395, 125078. [Google Scholar] [CrossRef]
  37. Gandelman, H.; da Silva, A.L.; Ramos, B.; Gouvêa, D. Interface excess on Sb-doped TiO2 photocatalysts and its influence on photocatalytic activity. Ceram. Int. 2021, 47, 619–625. [Google Scholar] [CrossRef]
  38. Ran, Y.; Zhong, J.; Li, J.; Li, M.; Tian, C. Substantially boosted photocatalytic detoxification activity of TiO2 benefited from Eu doping. Environ. Technol. 2023, 44, 1313–1321. [Google Scholar] [CrossRef]
  39. Alzard, R.H.; Siddig, L.A.; Abdelhamid, A.S.; Ramachandran, T.; Alzamly, A. Structural analysis and photocatalytic activities of bismuth-lanthanide oxide perovskites. J. Solid State Chem. 2024, 329, 124359. [Google Scholar] [CrossRef]
  40. Zhi, G.W.; Hao, L.Y.; Chen, W.J.; Sheng, Q.Y.; Liu, L.; Wang, W.; Yao, H.Y. Z-scheme CuSbS2/ZnO Heterojunction for Enhanced Photocatalytic Degradation of RhB. Chem. Sel. 2022, 7, e202202457. [Google Scholar] [CrossRef]
  41. Tang, Z.L.; Deng, X.; Chen, X.T.; Jiang, C.M.; Cai, L.; Guo, T.L. Ag2S/TiO2 Z-scheme heterojunction under magnetic field: Enhanced photocatalytic degradation of tetracycline. J. Alloys Compd. 2025, 1010, 177752. [Google Scholar] [CrossRef]
  42. Zhou, S.; Wang, Y.K.; Zhao, G.Q.; Li, C.F.; Liu, L.K.; Jiao, F.P. Enhanced visible light photocatalytic degradation of rhodamine B by Z-scheme CuWO4/g-C3N4 heterojunction. J. Mater. Sci.-Mater. Electron. 2021, 32, 2731–2743. [Google Scholar] [CrossRef]
  43. Liu, J.H.; Zhang, B.B.; Huang, Z.Y.; Wang, W.Y.; Xi, X.G.; Dong, P.Y. MOF-Derived In2O3 Microrod-Decorated MgIn2S4 Nanosheets: Z-Scheme Heterojunction for Efficient Photocatalytic Degradation of Tetracycline. Langmuir 2023, 39, 17458–17470. [Google Scholar] [CrossRef]
  44. Pan, B.; Chen, W.; Zhou, L.X.; Lai, X.J.; Qin, J.N.; Wang, C.Y. CoFe2O4/carbon nitride Z-scheme heterojunction photocatalytic PMS activation for efficient tetracycline degradation: Accelerated electron transfer. Process Saf. Environ. Protect. 2024, 191, 2522–2532. [Google Scholar] [CrossRef]
  45. Wang, C.; Liu, H.Y.; Wang, G.F.; Huang, W.Y.; Wei, Z.W.; Fang, H.Y.; Shen, F. Visible light driven S-scheme heterojunction Zn3In2S6/Bi2MoO6 for efficient degradation of metronidazole. J. Alloys Compd. 2022, 917, 165507. [Google Scholar] [CrossRef]
  46. Wang, X.Y.; Zhang, R.Y.; Luo, J.; Zhang, H.F.; Yan, Q.Z.; Zhao, J.W.; Liang, Z.; Shi, W.L.; Lu, C.Y. Constructing an S-scheme ZnSnO3-CuBi2O4 heterojunction modulation to boost visible-light-induced photothermal-photocatalytic degradation of tetracycline. J. Environ. Chem. Eng. 2025, 13, 115612. [Google Scholar] [CrossRef]
  47. Saha, S.; Chanda, S.; Dutta, A.; Kumar, U.; Ranjan, R.; Sinha, T.P. Dielectric relaxation and anti-ferromagnetic coupling of BiEu03 and BiGdO3. J. Magn. Magn. Mater. 2014, 360, 80–86. [Google Scholar] [CrossRef]
  48. Li, S.G.; Chen, F.; An, Q.; Tang, R.F.; Huang, H.W. Triggering Overall Crystal Polarization by Octahedron Elongation Distortion for Highly Boosted and Selective Aerobic Photo-Oxidation. Adv. Funct. Mater. 2024, 34, 2409035. [Google Scholar] [CrossRef]
  49. Peng, Z.Y.; Jiang, Y.H.; Wang, X.Q.; Zhang, R.X.; Xu, H.Q.; Xiao, Y.; Jing, X.; Zhang, J.M.; Liu, Y.; Ni, L. Novel CdIn2S4 nano-octahedra/TiO2 hollow hybrid heterostructure: In-situ synthesis, synergistic effect and enhanced dual-functional photocatalytic activities. Ceram. Int. 2019, 45, 15942–15953. [Google Scholar] [CrossRef]
  50. Guo, B.; Zhao, C.C.; Liu, X.X.; Yu, Z.; Zhou, L.J.; Yuan, H.M.; Zhao, Z. Effects of Surface Hydrothermal Carbon Layer on the Photocatalytic Activity of Magnetic NiFe2O4 Octahedron. Chem. J. Chin. Univ.-Chin. 2022, 43, 20220472. [Google Scholar]
  51. Nguyen, V.Q.; Mady, A.H.; Mahadadalkar, M.A.; Baynosa, M.L.; Kumar, D.R.; Rabie, A.M.; Lee, J.; Kim, W.K.; Shim, J.J. Highly active Z-scheme heterojunction photocatalyst of anatase TiO2 octahedra covered with C-MoS2 nanosheets for efficient degradation of organic pollutants under solar light. J. Colloid Interface Sci. 2022, 606, 337–352. [Google Scholar] [CrossRef]
  52. Ma, D.C.; Tang, J.W.; He, G.W.; Xue, Y.; Pan, S.; Liu, F.J.; Zhao, J.Z. Enhancing photocatalytic degradation of rhodamine B with visible-light-driven HCl-assisted Bi2O3 photocatalysts: Activity, mechanism, and pathways. Mater. Sci. Semicond. Process 2024, 181, 108672. [Google Scholar] [CrossRef]
  53. Bouziani, A.; Yahya, M.; Bianchi, C.L.; Falletta, E.; Celik, G. Ternary Polyaniline@Bi2O3-BiOCl Nanocomposites as Innovative Highly Active Photocatalysts for the Removal of the Dye under Solar Light Irradiation. Nanomaterials 2023, 13, 713. [Google Scholar] [CrossRef] [PubMed]
  54. Punitha, T.; Balamurugan, P.; Kumar, G.S.; Sasikruba, S. Ferromagnetic nanostructures of chromium-doped dysprosium iron oxide (DyCrxFe1−xO3) with visible light photocatalysis capabilities. Results Surf. Interfaces 2025, 18, 100452. [Google Scholar] [CrossRef]
  55. Zargar, R.A.; Imran, M.; Arora, M.; Nagal, V.; Mearaj, T.; Manthrammel, M.A.; Shkir, M.; Hafiz, A.K. Low-cost synthesis of lanthanides (Eu3+ and Sm3+)-intercalated TiO2 nanostructures: A detailed study on structural, optical and photocatalytic applications. J. Mater. Sci.-Mater. Electron. 2022, 33, 26931–26942. [Google Scholar] [CrossRef]
  56. Karnchana, N.; Phuruangrat, A.; Thongtem, S.; Thongtem, T. Combustion synthesis and characterization of visible-light-driven Tm-doped ZnO nanoparticles used for photodegradation of methylene blue. Dig. J. Nanomater. Biostruct. 2021, 16, 1243–1252. [Google Scholar] [CrossRef]
  57. Makhloufi, R.; Hachani, S.E.; Fettah, A.; Messai, B. Wet Chemical Synthesis of Sb4O5Cl2 Used as an Effective Photocatalyst for Methylene Blue and Crystal Violet Degradation under Visible Light Irradiation. Ann. Chim.-Sci. Mater. 2022, 46, 69–74. [Google Scholar] [CrossRef]
  58. Thirumal, V.; Yuvakkumar, R.; Kumar, P.S.; Keerthana, S.P.; Ravi, G.; Thambidurai, M.; Dang, C.; Velauthapillai, D. Facile hydrothermal synthesis of MXene@antimony nanoneedle composites for toxic pollutants removal. Environ. Res. 2022, 210, 112904. [Google Scholar] [CrossRef]
  59. Rauf, N.; Ilyas, S.; Heryanto, H.; Rahmat, R.; Fahri, A.N.; Rahmi, M.H.; Tahir, D. The Correlation between Structural and Optical Properties of Zinc Hydroxide Nanoparticle in Supports Photocatalytic Performance. Opt. Mater. 2021, 112, 110780. [Google Scholar] [CrossRef]
  60. Jalali, S.; Ardjmand, M.; Ramavandi, B.; Nosratinia, F. Elimination of amoxicillin using zeolite Y-sea salt as a good catalyst for activation of hydrogen peroxide: Investigating degradation pathway and the effect of wastewater chemistry. J. Environ. Manag. 2022, 302, 114045. [Google Scholar] [CrossRef]
  61. Asaithambi, S.; Murugan, R.; Sakthivel, P.; Karuppaiah, M.; Rajendran, S.; Ravi, G. Influence of Ni Doping in SnO2 Nanoparticles with Enhanced Visible Light Photocatalytic Activity for Degradation of Methylene Blue Dye. J. Nanosci. Nanotechnol. 2019, 19, 4438–4446. [Google Scholar] [CrossRef]
  62. Cadis, A.I.; Rus, F.S.; Goncalves, J.N.; Ivanovici, M. Preparing a Ca-Bi-O System by the Precipitation Method and Studying Its Intermediate Structural Properties for Applications in Water Treatment. Inorganics 2023, 11, 79. [Google Scholar] [CrossRef]
  63. Regmi, C.; Kshetri, Y.K.; Ray, S.K.; Pandey, R.P.; Lee, S.W. Utilization of visible to NIR light energy by Yb+3, Er+3 and Tm+3 doped BiVO4 for the photocatalytic degradation of methylene blue. Appl. Surf. Sci. 2017, 392, 61–70. [Google Scholar] [CrossRef]
  64. Slimani, Y.; Khan, A.; Nawaz, M.; Hossain, M.K.; Thakur, A. Efficient rhodamine B dye photocatalytic degradation in aqueous media using novel ZnO nanomaterials co-doped with Ce and Dy. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2024, 321, 124725. [Google Scholar] [CrossRef] [PubMed]
  65. Nozari, M.; Dehaghi, S.M. Structural properties of TiO2/Sb nanoparticles with different molar ratios and investigation the photocatalytic degradation of AZO dyes under UV irradiation in different condition. Fresenius Environ. Bull. 2016, 25, 3962–3970. [Google Scholar]
  66. Singh, A.; Dutta, D.P.; Roy, M.; Tyagi, A.K.; Fulekar, M.H. Sonochemical synthesis, characterization, and photocatalytic properties of Bi2−xSbxWO6 nanorods. J. Mater. Sci. 2014, 49, 2085–2097. [Google Scholar] [CrossRef]
  67. Patil, M.S.; Kitchamsetti, N.; Mulani, S.R.; Rondiya, S.R.; Deshpande, N.G.; Patil, R.A.; Cross, R.W.; Dzade, N.Y.; Sharma, K.K.; Patil, P.S.; et al. Photocatalytic behavior of Ba(Sb/Ta)2O6 perovskite for reduction of organic pollutants: Experimental and DFT correlation. J. Taiwan Inst. Chem. Eng. 2021, 122, 201–209. [Google Scholar] [CrossRef]
  68. Shandilya, P.; Mittal, D.; Soni, M.; Raizada, P.; Lim, J.H.; Jeong, D.Y.; Dewedi, R.P.; Saini, A.K.; Singh, P. Islanding of EuVO4 on high-dispersed fluorine doped few layered graphene sheets for efficient photocatalytic mineralization of phenolic compounds and bacterial disinfection. J. Taiwan Inst. Chem. Eng. 2018, 93, 528–542. [Google Scholar] [CrossRef]
  69. Cao, S.B.; Wu, L.; Li, J.; Li, Y.; Da, K.; Chen, W.T.; Xue, R.T.; Yang, J.; Fan, X.M. Unraveling the generation mechanism of singlet oxygen in Bi2WxMo1−xO6 solid solution and roles in photocatalytic degradation of gaseous toluene under visible-light irradiation. J. Environ. Chem. Eng. 2024, 12, 112321. [Google Scholar] [CrossRef]
  70. Liang, M.J.; Deng, N.; Xiang, X.Y.; Mei, Y.; Yang, Z.Y.; Yang, Y.; Yang, S.J. Bi/BiVO4&Bi4V2O11 Composite Catalysts: Preparation and Photocatalytic Performance. Chin. J. Inorg. Chem. 2019, 35, 263–270. [Google Scholar]
  71. Wang, Q.; Wu, H.; Gao, Q.Y.; Lin, D.G.; Fan, Y.J.; Duan, R.; Cong, Y.Q.; Zhang, Y. Fabrication of visible-light-active Bi/BiOI-Bi2O3 composite with enhanced photocatalytic activity. J. Colloid Interface Sci. 2019, 548, 255–264. [Google Scholar] [CrossRef]
  72. Heryanto, H.; Mutmainna, I.; Rahmi, M.H.; Ola, A.T.T.; Tang, N.F.R.; Mohamed, M.A.; Tahir, D. Favourable peak diffraction shift moments as a function of Mg doping on ZnO matrix as a promising catalyst for methylene blue waste. Mater. Chem. Phys. 2024, 313, 128772. [Google Scholar] [CrossRef]
  73. Wang, C.; Liu, H.Y.; Wang, G.F.; Ye, F.; Tao, S.K.; Li, X.; Qiu, P.; Huang, Z.H. Ni (II) doping induced lattice distortion in Zn3In2S6/BiOBr-OVs for boosting photocatalytic removal of antibiotics and Cr (VI) performance. Sep. Purif. Technol. 2023, 324, 124457. [Google Scholar] [CrossRef]
  74. Li, X.Y.; Li, W.J.; Gu, S.N.; Liu, X.T.; Li, H.D.; Ren, C.J.; Ma, X.H.; Zhou, H.L. Efficient ytterbium-doped Bi2WO6 photocatalysts: Synthesis, the formation of oxygen vacancies and boosted superoxide yield for enhanced visible-light photocatalytic activity. J. Alloys Compd. 2021, 851, 156935. [Google Scholar] [CrossRef]
  75. 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]
  76. Zou, Z.G.; Ye, J.H.; Arakawa, H. Substitution effects of In3+ by Fe3+ on photocatalytic and structural properties of Bi2InNbO7 photocatalysts. J. Mol. Catal. A-Chem. 2001, 168, 289–297. [Google Scholar] [CrossRef]
  77. Wu, S.Q.; Hu, H.Y.; Lin, Y.; Zhang, J.L.; Hu, Y.H. Visible light photocatalytic degradation of tetracycline over TiO2. Chem. Eng. J. 2020, 382, 122842. [Google Scholar] [CrossRef]
  78. Nain, P.; Pawar, M.; Rani, S.; Kumar, S.; Khan, M.A.M. (Ho, Yb) co-doping effects on structural, optical, photocatalytic and thermal properties of TiO2 NPs prepared by hydrothermal approach. Phys. B 2025, 706, 417132. [Google Scholar] [CrossRef]
  79. Sukhadeve, G.K.; Janbandhu, S.Y.; Kumar, R.; Lataye, D.H.; Ramteke, D.D.; Gedam, R.S. Visible light assisted photocatalytic degradation of Indigo Carmine dye and NO2 removal by Fe doped TiO2 nanoparticles. Ceram. Int. 2022, 48, 29121–29135. [Google Scholar] [CrossRef]
  80. Bansal, N.; Kumar, A. Photocatalytic degradation of organophosphorus pesticide Chlorpyrifos using CdS/NiS nanocomposite. Ionics 2024, 30, 4901–4916. [Google Scholar] [CrossRef]
  81. Geldasa, F.T.; Kebede, M.A.; Shura, M.W.; Andoshe, D.M.; Tegegne, N.A.; Hone, F.G. Facile synthesis of different metals doped α-PbO nanoparticles for photocatalytic degradation of Methylene Blue dye. Phys. Scr. 2023, 98, 065701. [Google Scholar] [CrossRef]
  82. Beura, R.; Pachaiappan, R.; Thangadurai, P. A detailed study on Sn4+ doped ZnO for enhanced photocatalytic degradation. Appl. Surf. Sci. 2018, 433, 887–898. [Google Scholar] [CrossRef]
  83. Agulo, I.J.; Rosario, P.; Yague, K.; Balod, M.J.; Empizo, M.J.F.; Agulto, V.C.; Shimizu, T.; Madarang, M.A.; Ngaloy, R.; Sarukura, N. Photoluminescence and enhanced photocatalytic activity of mechanically activated graphite-zinc oxide composites. Mater. Res. Express 2023, 10, 065601. [Google Scholar] [CrossRef]
  84. Dang, H.F.; Li, B.Q.; Li, C.P.; Zang, Y.H.; Xu, P.R.; Zhao, X.C.; Fan, H.B.; Qiu, Y.F. One-dimensional Au/SiC heterojunction nanocomposites with enhanced photocatalytic and photoelectrochemical performances: Kinetics and mechanism insights. Electrochim. Acta 2018, 267, 24–33. [Google Scholar] [CrossRef]
  85. Zhu, X.D.; Song, H.J.; Feng, W.; Wen, G.L.; Li, H.Y.; Zhou, J. Photocurrent response and photocatalytic activity of Nd-doped TiO2 thin films prepared by sol-gel method. J. Adv. Oxid. Technol. 2017, 20, 20160190. [Google Scholar] [CrossRef]
  86. Al Bacha, S.; Saitzek, S.; Kabbour, H.; McCabe, E.E. Iron Oxychalcogenides and Their Photocurrent Responses. Inorg. Chem. 2024, 63, 3292–3302. [Google Scholar] [CrossRef] [PubMed]
  87. Wang, Y.X.; Li, X.; Liu, S.N.; Liu, Y.; Kong, T.; Zhang, H.Y.; Duan, X.G.; Chen, C.M.; Wang, S.B. Roles of Catalyst Structure and Gas Surface Reaction in the Generation of Hydroxyl Radicals for Photocatalytic Oxidation. ACS Catal. 2022, 12, 2770–2780. [Google Scholar] [CrossRef]
  88. Zhuang, Y.; Luan, J.F. Improved photocatalytic property of peony-like InOOH for degrading norfloxacin. Chem. Eng. J. 2020, 382, 122770. [Google Scholar] [CrossRef]
  89. Wei, F.; Xing, H.D.; Xiu, Z.Y.; Li, J.D.; Xing, D.F.; Han, X.J. Z-scheme TiO2-Au@CN heterojunction for simultaneous water purification of disinfection and organic pollutant removal by simulated solar light. Mater. Res. Bull. 2023, 168, 100829. [Google Scholar] [CrossRef]
  90. Yuan, Y.; Guo, R.T.; Hong, L.F.; Ji, X.Y.; Lin, Z.D.; Li, Z.S.; Pan, W.G. A review of metal oxide-based Z-scheme heterojunction photocatalysts: Actualities and developments. Mater. Today Energy 2021, 21, 100829. [Google Scholar] [CrossRef]
  91. 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, 118661. [Google Scholar] [CrossRef]
  92. Shabbir, A.; Sardar, S.; Mumtaz, A. Mechanistic investigations of emerging type-II, Z-scheme and S-scheme heterojunctions for photocatalytic applications—A review. J. Alloys Compd. 2024, 1003, 175683. [Google Scholar] [CrossRef]
  93. Yin, N.; Chen, H.Y.; Yuan, X.Z.; Zhang, Y.; Zhang, M.J.; Guo, J.Y.; Zhang, Y.Y.; Qiao, L.; Liu, M.S.; Song, K.X. Highly efficient photocatalytic degradation of norfloxacin via Bi2Sn2O7/PDIH Z-scheme heterojunction: Influence and mechanism. J. Hazard. Mater. 2022, 436, 129317. [Google Scholar] [CrossRef]
  94. Jia, T.; Wu, J.; Xiao, Y.X.; Liu, Q.Z.; Wu, Q.; Qi, Y.F.; Qi, X.M. Self-grown oxygen vacancies-rich CeO2/BiOBr Z-scheme heterojunction decorated with rGO as charge transfer channel for enhanced photocatalytic oxidation of elemental mercury. J. Colloid Interf. Sci. 2021, 587, 402–416. [Google Scholar] [CrossRef]
  95. Lu, C.Y.; Guo, F.; Yan, Q.Z.; Zhang, Z.J.; Li, D.; Wang, L.P.; Zhou, Y.H. Hydrothermal synthesis of type II ZnIn2S4/BiPO4 heterojunction photocatalyst with dandelion-like microflower structure for enhanced photocatalytic degradation of tetracycline under simulated solar light. J. Alloys Compd. 2019, 811, 151976. [Google Scholar] [CrossRef]
  96. Nkudede, E.; Xing, Q.Z.; Pang, Z.Y.; Liu, N.H.; Yan, X.W.; Sulemana, H.; Yusuf, B.A.; Di, J.; Yin, S.; Xia, J.X. A novel Type-II Bi-based oxyhalide: Bi4O5I2/BiOBr microspheres for improved photocatalytic performance. Mater. Sci. Semicond. Process. 2024, 182, 108660. [Google Scholar] [CrossRef]
  97. Xue, M.; Zhao, J.Z.; Chen, J.J.; Ouyang, W.; Wang, Y.Q.; Yin, S.; Li, H.M.; Xia, J.X. Improving charge separation Bi19S27Cl3/BiOCl Z-scheme heterojunction for efficient visible-light photocatalytic degradation of 2-mercaptobenzothiazole. Colloid Surf. A-Physicochem. Eng. Asp. 2025, 710, 136136. [Google Scholar] [CrossRef]
  98. Feng, C.; Hu, M.; Alharbi, J.; Rueping, M.; Zhang, H. Bridging Scales in Solar-Driven Water Splitting: Pathways to System Integration. Adv. Mater. 2025, e06690. [Google Scholar] [CrossRef] [PubMed]
  99. Xiang, H.C.; Tuo, B.Y.; Song, X. Photocatalytic degradation of methyl orange by Bi20TiO32–montmorillonite composite. Micro Nano Lett. 2020, 15, 561–565. [Google Scholar] [CrossRef]
  100. Khan, S.S.; Subhiksha, V.; Harini, G.; Syed, A.; Janani, B.; Steffy, J.P.; Elgorban, A.M.; Abid, I.; Wong, L.S. Integrating N doped KNbO3 tailored with Sr2V2O7 for escalated performance towards photocatalytic degradation of metronidazole: Insights into mechanism, degradation pathway and its toxicity evaluation. Surf. Interfaces 2024, 51, 104610. [Google Scholar] [CrossRef]
  101. Sre, V.V.; Janani, B.; Syed, A.; Elgorban, A.M.; Abid, I.; Wong, L.S.; Khan, S.S. Unveiling the enhanced photocatalytic degradation of metronidazole over green-synthesized Ag@ZnCdS QDs Schottky heterojunction: Pathway, toxicity evaluation and mechanistic insights. J. Water Process. Eng. 2024, 62, 105325. [Google Scholar] [CrossRef]
  102. Li, Z.; Meng, A.; Li, W.; Xiong, G.; Ye, M.; Meng, Y.; Li, Z. Co-Catalyst-Free Al6Si2O13/Cd8.05Zn1.95S10 Nanocomposites for Visible-Light-Driven Stable H2 Evolution and DDVP Degradation. Catalysts 2025, 15, 564. [Google Scholar] [CrossRef]
  103. Kaiba, A.; Alansi, A.M.; Oubelkacem, A.; Chabri, I.; Hameed, S.T.; Afzal, N.; Rafique, M.; Qahtan, T.F. Sunlight-Driven Synthesis of TiO2/(MA)2SnCl4 Nanocomposite Films for Enhanced Photocatalytic Degradation of Organic Pollutants. Catalysts 2025, 15, 214. [Google Scholar] [CrossRef]
  104. Saira; Abd El-Fattah, W.; Shahid, M.; Ashraf, S.; Sandhu, Z.A.; Guesmi, A.; Ben Hamadi, N.; Farhan, M.; Raza, M.A. Solar-Activated Titanium-Based Cu4O3/ZrO2/TiO2 Ternary Nano-Heterojunction for Rapid Photocatalytic Degradation of the Textile Dye Everzol Yellow 3RS. Catalysts 2025, 15, 751. [Google Scholar] [CrossRef]
  105. Si, P.F.; Feng, Y.J.; Wang, X.; Yue, X.P.; Li, D.F.; Zhang, S.; Liang, B.; Gao, Y.J.; Zhou, A.J. Enhanced degradation of Metronidazole by the coupling of photocatalytic and microbial fuel cell: Mechanism and electrochemistry characteristic. J. Environ. Chem. Eng. 2023, 11, 109707. [Google Scholar] [CrossRef]
  106. Zhao, J.; He, Q.; Yao, B.H.; Zhang, Q.; Zhang, T.; Yang, M.K.; Liu, Y. Preparation and Photocatalytic Property of Er3NbO7/LSC-100 Chelating Resin Composites. Integr. Ferroelectr. 2015, 159, 73–86. [Google Scholar] [CrossRef]
  107. Wu, Y.; Xia, F.F.; Li, C.Z.; Jayswal, M.R.; Prakash, O.; Muddassir, M. Two new Ni(II) and Co(II)-based coordination polymers based on mixed linkers as photocatalysts for antibiotics decomposition. Inorg. Chim. Acta 2023, 558, 121723. [Google Scholar] [CrossRef]
  108. Tran, M.L.; Fu, C.C.; Juang, R.S. Effects of water matrix components on degradation efficiency and pathways of antibiotic metronidazole by UV/TiO2 photocatalysis. J. Mol. Liq. 2019, 276, 32–38. [Google Scholar] [CrossRef]
  109. Abbasi-Asl, H.; Sabzehmeidani, M.M.; Ghaedi, M. Efficient degradation of metronidazole antibiotic by TiO2/Ag3PO4/g–C3N4 ternary composite photocatalyst in a continuous flow-loop photoreactor. J. Environ. Chem. Eng. 2021, 9, 105963. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of BBHP, BiTmDySbO7, and BiEuO3.
Figure 1. XRD patterns of BBHP, BiTmDySbO7, and BiEuO3.
Sustainability 17 10024 g001
Figure 2. XRD patterns and Rietveld refinement results of (a) BiTmDySbO7 and (b) BiEuO3. Atomic structure of (c) BiTmDySbO7 and (d) BiEuO3.
Figure 2. XRD patterns and Rietveld refinement results of (a) BiTmDySbO7 and (b) BiEuO3. Atomic structure of (c) BiTmDySbO7 and (d) BiEuO3.
Sustainability 17 10024 g002
Figure 3. FTIR spectra of BBHP, BiTmDySbO7, and BiEuO3.
Figure 3. FTIR spectra of BBHP, BiTmDySbO7, and BiEuO3.
Sustainability 17 10024 g003
Figure 4. The green curve represents the Raman spectrum of the BBHP. The blue curve represents the Raman spectrum of the BiTmDySbO7. The orange curve represents the Raman spectrum of the BiEuO3.
Figure 4. The green curve represents the Raman spectrum of the BBHP. The blue curve represents the Raman spectrum of the BiTmDySbO7. The orange curve represents the Raman spectrum of the BiEuO3.
Sustainability 17 10024 g004
Figure 5. (a) TEM; (b) HRTEM; (c) EDS layered; (d) EDS composition content distribution image; (e) EDS elemental mapping images of BBHP.
Figure 5. (a) TEM; (b) HRTEM; (c) EDS layered; (d) EDS composition content distribution image; (e) EDS elemental mapping images of BBHP.
Sustainability 17 10024 g005
Figure 6. XPS spectra of synthesized BBHP, BiTmDySbO7, and BiEuO3: (a) Full scan spectra; (bf) High-resolution spectra of Bi 4f, Tm 4d, Dy 4d, Eu 3d, Sb 3d, and O 1s.
Figure 6. XPS spectra of synthesized BBHP, BiTmDySbO7, and BiEuO3: (a) Full scan spectra; (bf) High-resolution spectra of Bi 4f, Tm 4d, Dy 4d, Eu 3d, Sb 3d, and O 1s.
Sustainability 17 10024 g006aSustainability 17 10024 g006b
Figure 7. (a) The ultraviolet and visible diffuse reflection spectra of BBHP, BiTmDySbO7, and BiEuO3; (b) Corresponding plots of (αhν)1/2 and for BBHP, BiTmDySbO7, and BiEuO3; The dashed lines are the tangent to the curves. (c) the photoluminescence spectra of BBHP, BiTmDySbO7, and BiEuO3; (d) TRPL spectrum of BBHP; (e) TRPL spectrum of BiEuO3; (f) TRPL spectrum of BiTmDySbO7.
Figure 7. (a) The ultraviolet and visible diffuse reflection spectra of BBHP, BiTmDySbO7, and BiEuO3; (b) Corresponding plots of (αhν)1/2 and for BBHP, BiTmDySbO7, and BiEuO3; The dashed lines are the tangent to the curves. (c) the photoluminescence spectra of BBHP, BiTmDySbO7, and BiEuO3; (d) TRPL spectrum of BBHP; (e) TRPL spectrum of BiEuO3; (f) TRPL spectrum of BiTmDySbO7.
Sustainability 17 10024 g007
Figure 8. (a) The effect of the photocurrent density on VLI time with BBHP, BiEuO3, or BiTmDySbO7 as photoelectrode; (b) The EIS plots of BBHP, BiTmDySbO7, and BiEuO3.
Figure 8. (a) The effect of the photocurrent density on VLI time with BBHP, BiEuO3, or BiTmDySbO7 as photoelectrode; (b) The EIS plots of BBHP, BiTmDySbO7, and BiEuO3.
Sustainability 17 10024 g008
Figure 9. (a) The curves of Ct/C0 versus VLI time for the photocatalytic degradation of the MNZ by using BBHP, BiTmDySbO7, BiEuO3 or N-T; (b) The fitting results of ln(Ct/C0) versus VLI time by using BBHP, BiEuO3, BiTmDySbO7 or N-T; (c) The curves of TOCt/TOC0 versus VLI time for the photocatalytic degradation of the MNZ by using BBHP, BiTmDySbO7, BiEuO3 or N-T; (d) The fitting results of ln(TOC0/TOCt) versus VLI time during the degradation process of MNZ by using BBHP, BiEuO3, BiTmDySbO7 or N-T; (e) The photodegradation rate of the MNZ and the kinetic constants after VLI of 130 min by using BBHP, BiEuO3, BiTmDySbO7 or N-T; (f) The mineralization efficiency of the TOC concentration and the kinetic constants after VLI of 130 min by using BBHP, BiEuO3, BiTmDySbO7 or N-T.
Figure 9. (a) The curves of Ct/C0 versus VLI time for the photocatalytic degradation of the MNZ by using BBHP, BiTmDySbO7, BiEuO3 or N-T; (b) The fitting results of ln(Ct/C0) versus VLI time by using BBHP, BiEuO3, BiTmDySbO7 or N-T; (c) The curves of TOCt/TOC0 versus VLI time for the photocatalytic degradation of the MNZ by using BBHP, BiTmDySbO7, BiEuO3 or N-T; (d) The fitting results of ln(TOC0/TOCt) versus VLI time during the degradation process of MNZ by using BBHP, BiEuO3, BiTmDySbO7 or N-T; (e) The photodegradation rate of the MNZ and the kinetic constants after VLI of 130 min by using BBHP, BiEuO3, BiTmDySbO7 or N-T; (f) The mineralization efficiency of the TOC concentration and the kinetic constants after VLI of 130 min by using BBHP, BiEuO3, BiTmDySbO7 or N-T.
Sustainability 17 10024 g009
Figure 10. (a) The effect of the MNZ concentration on VLI time during five cyclic experiments by using the BBHP; (b) The fitting results of ln(C0/Ct) versus VLI time during the cyclic degradation process of the MNZ by using BBHP; (c) The change curves of the TOC concentration during five cyclic experiments for degrading MNZ by using the BBHP; (d) The fitting results of ln(TOC0/TOCt) versus VLI time during the cyclic degradation process of the MNZ by using BBHP; (e) The removal efficiencies of MNZ and the kinetic constants for degrading MNZ during five cyclic experiments by using the BBHP; (f) The mineralization efficiency of the TOC concentration and the kinetic constants which derived from the contribution of ln(TOC0/TOCt) and VLI time during five cyclic experiments by using the BBHP.
Figure 10. (a) The effect of the MNZ concentration on VLI time during five cyclic experiments by using the BBHP; (b) The fitting results of ln(C0/Ct) versus VLI time during the cyclic degradation process of the MNZ by using BBHP; (c) The change curves of the TOC concentration during five cyclic experiments for degrading MNZ by using the BBHP; (d) The fitting results of ln(TOC0/TOCt) versus VLI time during the cyclic degradation process of the MNZ by using BBHP; (e) The removal efficiencies of MNZ and the kinetic constants for degrading MNZ during five cyclic experiments by using the BBHP; (f) The mineralization efficiency of the TOC concentration and the kinetic constants which derived from the contribution of ln(TOC0/TOCt) and VLI time during five cyclic experiments by using the BBHP.
Sustainability 17 10024 g010
Figure 11. (a) The change curves that derive from the effect of the MNZ concentration on the VLI time with the addition of the radical scavenger such as IPA, BQ, or EDTA; The green area represents the dark treatment stage of the free radical scavenging experiment; The white area represents the experimental stage of free radical scavenging under the condition VLI. (b) The removal rate of the MNZ during the degradation process of the MNZ by using BBHP with IPA, BQ, or EDTA as scavenger under the condition of VLI.
Figure 11. (a) The change curves that derive from the effect of the MNZ concentration on the VLI time with the addition of the radical scavenger such as IPA, BQ, or EDTA; The green area represents the dark treatment stage of the free radical scavenging experiment; The white area represents the experimental stage of free radical scavenging under the condition VLI. (b) The removal rate of the MNZ during the degradation process of the MNZ by using BBHP with IPA, BQ, or EDTA as scavenger under the condition of VLI.
Sustainability 17 10024 g011
Figure 12. The EPR spectra of DMPO·O2 and DMPO·OH during the degradation process of the MNZ by using BBHP under the condition of VLI.
Figure 12. The EPR spectra of DMPO·O2 and DMPO·OH during the degradation process of the MNZ by using BBHP under the condition of VLI.
Sustainability 17 10024 g012
Figure 13. (a) The ultraviolet photoelectron spectroscopy spectrum of BiTmDySbO7; (b) The ultraviolet photoelectron spectroscopy spectrum of BiEuO3.
Figure 13. (a) The ultraviolet photoelectron spectroscopy spectrum of BiTmDySbO7; (b) The ultraviolet photoelectron spectroscopy spectrum of BiEuO3.
Sustainability 17 10024 g013
Figure 14. The schematic diagram of the photodegradation mechanism of the MNZ under the condition of VLI by using BBHP.
Figure 14. The schematic diagram of the photodegradation mechanism of the MNZ under the condition of VLI by using BBHP.
Sustainability 17 10024 g014
Figure 15. The feasible photodegradation pathways of MNZ under the condition of VLI by using BBHP.
Figure 15. The feasible photodegradation pathways of MNZ under the condition of VLI by using BBHP.
Sustainability 17 10024 g015
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Luan, J.; Li, Z.; Yao, Y.; Wang, J.; Hao, L. Metronidazole Degradation via Visible Light-Driven Z-Scheme BiTmDySbO7/BiEuO3 Heterojunction Photocatalyst. Sustainability 2025, 17, 10024. https://doi.org/10.3390/su172210024

AMA Style

Luan J, Li Z, Yao Y, Wang J, Hao L. Metronidazole Degradation via Visible Light-Driven Z-Scheme BiTmDySbO7/BiEuO3 Heterojunction Photocatalyst. Sustainability. 2025; 17(22):10024. https://doi.org/10.3390/su172210024

Chicago/Turabian Style

Luan, Jingfei, Zhe Li, Ye Yao, Jian Wang, and Liang Hao. 2025. "Metronidazole Degradation via Visible Light-Driven Z-Scheme BiTmDySbO7/BiEuO3 Heterojunction Photocatalyst" Sustainability 17, no. 22: 10024. https://doi.org/10.3390/su172210024

APA Style

Luan, J., Li, Z., Yao, Y., Wang, J., & Hao, L. (2025). Metronidazole Degradation via Visible Light-Driven Z-Scheme BiTmDySbO7/BiEuO3 Heterojunction Photocatalyst. Sustainability, 17(22), 10024. https://doi.org/10.3390/su172210024

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