Next Article in Journal
Pretreatment and Nanoparticles as Catalysts for Biogas Production Reactions in Pepper Waste and Pig Manure
Next Article in Special Issue
Special Issue on “Advanced Catalytic Material for Water Treatment”
Previous Article in Journal
Long-Term Hydrogen Production from a Methanol–Water Solution Catalyzed by an Iridium Complex
Previous Article in Special Issue
Preparation of Mn-Co-MCM-41 Molecular Sieve with Thermosensitive Template and Its Degradation Performance for Rhodamine B
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Dual Z-Scheme CuBi2O4/Bi2Sn2O7/Sn3O4 Photocatalysts with Enhanced Photocatalytic Performance for the Degradation of Tetracycline under Visible Light Irradiation

Jiangsu Key Laboratory of Atmospheric Environment Monitoring and Pollution Control, School of Environmental Science and Engineering, Collaborative Innovation Center of Atmospheric Environment and Equipment Technology, Jiangsu Engineering Technology Research Center of Environmental Cleaning Materials, Nanjing University of Information Science and Technology, Nanjing 210044, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(7), 1028; https://doi.org/10.3390/catal13071028
Submission received: 11 May 2023 / Revised: 8 June 2023 / Accepted: 16 June 2023 / Published: 21 June 2023
(This article belongs to the Special Issue Advanced Catalytic Material for Water Treatment)

Abstract

:
The dual Z-scheme heterojunction CuBi2O4/Bi2Sn2O7/Sn3O4(CBS) was successfully constructed through in situ growth methods, and its photocatalytic performance was evaluated via degradation of tetracycline hydrochloride under visible light. Out of all samples, CBS-2 exhibited the highest photocatalytic activity, with an apparent rate constant of 2.34, 20.16, and 44.17 times that of Bi2Sn2O7, CuBi2O4, and Sn3O4, respectively. Even after four cycles, the photocatalytic efficiency remained above 85%. The improvement can be attributed to the construction of the Z-scheme heterojunction, which effectively promotes the separation and migration of photogenerated carriers. The possible photocatalytic degradation mechanism of dual Z-scheme heterojunction CBS was deduced based on the theory of free radical capture and energy band.

1. Introduction

In recent years, antibiotics have been widely used in many fields, but at the same time, they are difficult for the human body to completely metabolize, and a large quantity of them enter the water environment through the digestive system, causing increasing water pollution. Furthermore, they are reabsorbed by the human body through the water environment, which is harmful to human health [1,2,3,4]. Therefore, antibiotic pollution has become a major problem in water pollution control. Among them, tetracycline, as a class of common broad-spectrum antibiotics, is widely used in animal husbandry and animal medicine, but due to years of overuse, its residues exist in the water environment for a long time and are difficult to degrade [5]. Therefore, safe and effective removal methods have received extensive attention.
Commonly used methods for the removal of antibiotics are adsorption [6,7] and biodegradation [8,9], but their shortcomings are also obvious: the enrichment transfer process producing secondary pollution, the low removal efficiency, the high price, the more toxic by-products, and the large energy consumption. Photocatalysis can be used as a safe, effective, and environmentally friendly method to degrade tetracycline [10,11,12,13]. Mukhtar et al. prepared dual Z-scheme core–shell PANI-CeO2-Fe2O3-NiO heterojunctions to remove MB and MO dyes, with removal rates of 99% and 98%. Increased photocatalytic activity can be attributed to dual Z-scheme formation and the core–shell morphology of PANl, which effectively promote the separation of photogenerated carriers [14]. Munawar et al. prepared a dual Z-scheme Sm2O3-WO3-La2O3 nanocomposite to remove MB and MO dyes, with removal rates of 99% and 96% [15]. However, its photocatalytic stability, absorption spectrum range, and degradation capability limit its further application. Therefore, the development of stable, efficient, and visible light photocatalysts is one of the most important objectives in the field of photocatalysis.
Bi2Sn2O7(BSO) is a Bi-based pyrochlore structure photocatalyst with the advantages of a moderate band gap (2.7 eV), adjustable structure, low cost, easy doping modification, and non-toxicity [16,17,18,19,20]. However, the high recombination rate of photogenerated electrons and holes and poor quantum efficiency limit the improvement of its photocatalytic performance [21,22,23]. Among many methods for improving photocatalytic performance, the construction of heterojunction structures is one of the most effective methods [24,25,26]. The Z-scheme heterojunction couples two semiconductors with matching band gaps, based on maintaining the light absorption range, by which the separation efficiency and redox capacity of carriers are improved, therefore improving the photocatalytic performance [19,23,27]. Considering the relatively positive CB potential of BSO, it is necessary to select photocatalysts with better matching energy band structures to form a Z-scheme structure.
CuBi2O4(CBO), a novel Bi-based photocatalyst with a relatively negative CB potential and narrow band gap (1.7 eV), has attracted much attention due to its wide light absorption range and high light absorption efficiency. At the same time, it also has the disadvantage of a high carrier recombination rate [28,29,30]. Hence, CuBi2O4 can be selected as suitable semiconductors to construct a heterojunction structure with a matched energy band structure.
Sn3O4(SNO) is also a promising candidate for the construction of Z-scheme heterojunctions due to the ideal band gap (2.5 eV) and relatively negative CB potential [31,32,33]. Therefore, we can speculate that the construction of a dual Z-scheme heterojunction structure CBS can promote carrier separation while retaining redox capability, which is a feasible method to improve photocatalytic performance.
In this paper, we successfully synthesized Z-scheme CBS photocatalysts using the hydrothermal method. Then, the photocatalytic performance of the CBS composite was tested via degradation of tetracycline under visible light irradiation. Heterojunction materials can effectively broaden the spectral response range of the photocatalyst, promote carrier separation, inhibit recombination, and enhance photocatalytic activity. Finally, the possible mechanism of photocatalytic degradation of tetracycline was proposed based on our analysis.

2. Results and Discussion

2.1. Characterization

The phase of the BSO, CBO, SNO, and CBS-2 composites was characterized via XRD. Figure 1 displays the XRD patterns of pure BSO, CBO, SNO, and CBS-2. For pure BSO, the distinct diffraction peaks at 28.8°, 33.4°, 48.0°, 56.9°, 59.7°, and 77.6° are in good agreement with the (222), (400), (440), (622), (444), and (662) crystal planes of BSO (JCPDS NO. 87-0284). For pure CBO, the principal peaks at 14.7°, 20.9°, 28.0°, 30.7°, 33.2°, 37.4°, 43.2°, 45.1°, 46.7°, 47.8°, 53.0°, 55.6°, 57.9°, 60.7°, 64.3°, 66.0°, 73.0°, 74.4°, and 78.2° are consistent with the (110), (200), (211), (220), (310), (202), (222), (330), (411), (420), (213), (332), (422), (521), (512), (413), (541), (602), and (622) crystal planes of CBO (JCPDS NO. 72-0493). For pure SNO, the main peaks at 27.0°, 32.3°, 37.0°, 50.0°, and 51.7° are well-matched with the (111), (−121), (130), (−301), and (−1−32) crystal planes of Sn3O4 (JCPDS NO. 16-0737). There are no extra peaks from the SnO2 or SnO phase appearing in the XRD pattern of SNO, implying that the as-synthesized sample is pure SNO. The sharp feature peaks of the pure sample represent high crystallinity, and no extra impurity peaks indicate its high purity. Because of the low content and overlapping characteristic peaks, the main diffraction peaks of the composite samples are similar to BSO; however, they are consistent with the (111) crystal plane of SNO at 27° and the (202) crystal plane of CBO at 37.5°, which proves the coexistence of BSO, CBO, and SNO phases in the CBS composites. In summary, the double Z-scheme CBS-2 was successfully obtained.
TEM and HRTEM indicate the morphology and crystalline structure of the as-prepared photocatalyst. As shown in Figure 2a–c, pure BSO is composed of many microspheres, CBO presented rod-like morphology, and SNO exhibited the flower structure. Figure S1 shows the TEM image of composite CBS-2, in which CBO and SNO can be identified; parts of BSO microspheres were successfully distributed on the surface of SNO and CBO due to the in situ generation method. In Figure 2d, we can display the 0.31 nm crystal lattice corresponding to the (222) crystal plane of BSO, the 0.24 nm crystal lattice corresponding to the (211) crystal plane of CBO, and the 0.32 nm crystal lattice corresponding to the (111) crystal plane of SNO, respectively. BSO, CBO, and SNO are observed to be closely combined. Therefore, the TEM and HRTEM images prove the existence of CBO and SNO in the composite material and they are in close connection with the basal BSO material and jointly construct the heterojunction structure.
The micro-strain variation is linked with the crystallite size changes, and δ (δ = 1/D2) has a direct impact on lattice imperfection. We can calculate the average crystalline size from the XRD results using the Debye–Scherrer formula [34]:
D = K λ/β cosθ
The average crystalline size of BSO, CBO, SNO, and CBS-2 is about 17.5 nm, 11.6 nm, 10.3 nm, and 17.2 nm, respectively, which also supported TEM results.
XPS testing was used to further confirm the elemental composition and valence state of the prepared materials. The XPS survey spectra of BSO and CBS-2 are reproduced in Figure S2. The BSO contains Bi, Sn, and O elements, while the composite samples contain Bi, Sn, O, and Cu elements, which suggest the coexistence of CBO in CBS-2. In Figure 3a, the Bi 4f high-resolution spectrum, the peaks near 164.7 eV and 159.4 eV correspond to the orbital binding energies of Bi 4f5/2 and Bi 4f7/2, respectively, which proves that in BSO and composites, Bi exists in the form of Bi3+ [35,36]. In Figure 3b, the Sn 3d high-resolution spectrum of BSO, the orbital binding energies of 3d3/2 and 3d5/2 correspond to the peaks near 486.9 eV and 495.6 eV, respectively, which proves that it exists in the form of Sn4+ [37]. In the Sn 3d high-resolution spectrum of the complex, 3d3/2 and 3d5/2 can be deconvoluted into four peaks, indicating that there is more than one valence state of Sn in the sample. The peaks near 486.8 eV and 495.5 eV correspond to Sn4+, and the peak at 486.3 eV and 494.8 eV corresponds to Sn2+ [38]. In Figure 3c, the Cu 2p high-resolution spectrum, the binding energies of Cu 2p3/2 and Cu 2p1/2 peaks are 934.7 eV and 954.8 eV, representing Cu2+ [39]. In Figure 3d, the O 1s high-resolution spectrum, the binding energy of 529.8 eV corresponds to the Bi-O bond, 531.7 eV corresponds to the Sn-O bond, and 532.2 eV corresponds to the surface adsorbed oxygen [40].
We measured the average particle size using dynamic light scattering. The average particle size of BSO, CBO, and SNO is 1142 nm, 463.1 nm, and 287.6 nm, CBO/BSO is 522.1 nm, and CBS-1, CBS-2, and CBS-3 is 207.3 nm, 208.6 nm, and 211.5 nm, respectively, which is similar to the TEM images. Although the average crystalline size of CBS-2 is similar to that of BSO, its average particle size is smaller, and the particles are more dispersed, which means that it has a larger specific surface area and more active sites.
The light absorption range of a photocatalyst is closely related to its photocatalytic capacity. The as-prepared samples were characterized via UV–Vis diffuse reflectance spectroscopy to investigate their optical absorption characteristics. According to Figure 4a, the absorption edge range of BSO is about 450 nm–500 nm, CBO is about 750 nm–800 nm, and that of SNO is about 500–550 nm. The composite photocatalyst CBS-2 is slightly red-shifted on the base of pure BSO, which confirms the visible-light response ability of CBS-2. Their band gaps can be obtained from the following formula [41]:
Ahν = A(hν − Eg)n/2
where α, h, ν, and Eg represent proportionality constant, Planck constant, and optical frequency, and n is related to semiconductor type. According to the literature, BSO and SNO are indirect semiconductors, n = 4, and CBO is a direct semiconductor, n = 1 [35,39,42]. Therefore, it can be roughly deduced that the Eg of BSO is 2.76 eV, the Eg of CBO is 1.55 eV, and the Eg of SNO is 2.39 eV. The calculation results are shown in Figure 4b,c, which is consistent with the previous literature.
In addition, the band structure of photocatalysts is also one of the important factors affecting the photocatalytic ability of photocatalysts. We can implement the Mott–Schottky (M–S) curves to measure its flat-band potential, then estimate the position of the energy band of the photocatalytic materials. As illustrated in Figure 4d–f, the slopes of BSO and SNO tangents are positive, indicating that both BSO and SNO are n-type semiconductors. The slope of the CBO tangent is negative, indicating that CBO is a p-type semiconductor. The tangents of BSO, CBO, and SNO are extended, and the flat band potential (Efb) of BSO, CBO, and SNO is about 0.3 eV, 0.63 eV, and −0.85 eV, respectively (Efb|Ag/AgCl). We can obtain standard hydrogen electrode values by using the following equation [43]:
Efb|NHE = Efb|Ag/AgCl + 0.197
The minimum value of the n-type conduction band (CB) semiconductor is about 0.1 eV higher than the flat-band potential, and the maximum value of the p-type semiconductor valence band (VB) is about 0.1 eV lower than the flat-band potential. Therefore, the ECB values of BSO and SNO are calculated as 0.40 eV and −0.75 eV, and the EVB values of CBO are calculated as 0.93 eV. Then, we calculate their corresponding potential according to the following formula:
EVB = ECB + Eg
Therefore, it can be concluded that the EVB values of BSO and SNO are calculated as 3.16 eV and 1.64 eV, and the EVB values are calculated as −0.92 eV [30,40,44].

2.2. Photocatalytic Activity

We tested the photocatalytic performance of the photocatalyst via the degradation efficiency of the model pollutant, tetracycline solution, under visible light. Results are as shown in Figure 5a,b: pure samples reveal poor photocatalytic activities, the photocatalytic activities of BSO, CBO, and SNO are only 56.2%, 10.3%, and 5%, and the CWO/BSO heterojunction material showed higher photocatalytic activity than pure samples, but lower than the CBS Z-scheme heterojunction under the same conditions. The best photocatalytic activity of CBS-2 is 89.7% under the same conditions. This may be because, due to the increase in CBO and SNO content, more heterojunction structures were generated, thus improving the photocatalytic activity. However, with the further increase in CBO and SNO content, the photocatalytic activity of the composite decreases, which may be because the decrease in active reaction sites reduces the optical adsorption capacity of the composite catalyst [45,46].
The comparison of the tetracycline degradation activity of CBS-2 and other photocatalytic materials is shown in Table 1.
Figure 5c,d shows the apparent kinetic constants of different catalysts degrading tetracycline solution, which can be calculated according to the following equation [52]:
ln(Ct/C0) = kt
C0 represents the concentration of the tetracycline solution before illumination starts after dark treatment, and Ct represents a concentration of the tetracycline solution at t min. As illustrated in Figure 5c,d, the apparent rate constant values of BSO, CBO, and SNO are 0.1674, 0.0195, and 0.0089 min−1, that of CBO/BSO is 0.2572 min−1, and those of CBS-1, CBS-2, and CBS-3 are 0.3381, 0.3931, and 0.2852 min−1. The composite CBS-2 possessed the highest k value, about 2.34, 20.16, and 44.17 times those of pure BSO, CBO, and SNO.
The R2 value of BSO, CBO, and SNO are 0.8455, 0.8397, and 0.994, that of CBO/BSO is 0.967, and those of CBS-1, CBS-2, and CBS-3 are 0.9389, 0.9475, and 0.9785. The R2 value of most photocatalysts is greater than 0.9, except for BSO and CBO.
In addition to photocatalytic activity, the stability and reusability of photocatalysts are also key considerations restricting their applications [53]. XRD tests were carried out before and after the photocatalytic cycling experiment, respectively. Figure 5e shows that the crystal structure of the photocatalyst has no obvious change after four cycles of photocatalysis. In Figure 5f, the degradation efficiency remained stable after four cycles, the degradation rates of the four repeated experiments were 88.9%, 87.2%, 86.1%, and 85.6%, and the photocatalytic activity of CBS-2 was only slightly decreased by about 4.3% after four photocatalytic cycling experiments. The decrease in activity may be due to the adsorption of TC by the active site of the catalyst. In short, the composite material can be reused.

2.3. Photocatalytic Mechanism

To investigate the transfer and separation efficiency of photogenerated carriers, photocurrent cycle experiments were carried out on samples. A high photocurrent intensity reflects high separation efficiency [54]. As shown in Figure 6a, the photocurrent intensity of CBS-2 is higher than that of the pure samples, which indicates that the CBS-2 composite material has the strongest carrier transfer and separation efficiency.
The relatively weak PL intensity represents a high separation efficiency [55]. In Figure 6b, the high PL emission peaks of BSO, CBO, and SNO imply a high photogenerated carrier recombination rate, while CBS-2 shows a lower emission peak. Constructing a double Z-scheme heterostructure between BSO, CBO, and SNO, the carrier separation efficiency is higher than that of the pure sample, and the recombination of electron–hole pairs can be suppressed effectively so that the composite material exhibits enhanced photocatalytic activity.
To confirm the difference in the charge transfer separation efficiency and impedance between different samples, electrochemical impedance spectroscopy (EIS) tests were performed on BSO, CBO, SNO, and the composite material. The smaller radius of the circle in the diagram means lower resistance, and faster electron migration rate [56]. Figure 6c shows the relative radius of the pure and composite photocatalyst, sorted by size as CBO > SNO > BSO > CBS-2.
According to characterization results, due to the successful construction of the heterojunction, the average particle size was reduced, the optical absorption range of CBS was widened, the carrier separation efficiency was improved, and the photogenerated carrier recombination rate was reduced; therefore, the photocatalytic activity was significantly improved.
To inspect the main active radicals in the photocatalytic process of the CBS-2 composite, we performed a radical trapping experiment. The experimental results are shown in Figure 6d: the addition of IPA has almost no effect on the degradation efficiency, but the degradation efficiency is significantly inhibited by the addition of BQ, and the addition of AO also has a certain inhibitory effect on the degradation efficiency. The degradation efficiency dropped slightly from 88.9% to 32.8% and 68.8% when BQ and AO were added. These results indicate that •O2 and h+ act synergistically in the photocatalytic process of CBS-2, of which •O2 is more important.
To elucidate the possible mechanism, ESR characterization and trapping experiments were used to detect the main active radicals. From Figure 6e, •O2 characteristic signals of pure BSO cannot be obtained under visible light or in the dark, while that of CBS-2 can, which illustrates the presence of •O2 and its participation in the photocatalytic process. As shown in Figure 6f, both BSO and CBS-2 can receive •OH characteristic signals after irradiation. Trapping experiments also illustrate that h+ and •O2 play critical roles in the degradation of TC.
Figure 7 illustrates the possible mechanism of the photodegradation of TC. The double Z-scheme photocatalytic is constructed with BSO, CBO, and SNO. Under visible light, they are all motivated to generate photoproduced electron–hole pairs. According to the classic type II heterojunction transfer mechanism, the CB of BSO (0.4 eV) is more positive, and the electrons on the CB of CBO and SNO will transfer to the CB of BSO. Meanwhile, the h+ shift from BSO to CBO and SNO occurs because of the more positive VB of BSO. The CB of BSO is more positive than O2/•O2 (−0.33 eV), and the VB of CBO and SNO is more negative than H2O/•OH (2.27 eV), so it is impossible to generate •O2 thermodynamically. However, that inference contrasts with the results of trapping experiments and ESR analysis. It can be inferred that the photogenerated carrier migration follows a Z-scheme mechanism. When BSO, CBO, and SNO are motivated to generate holes and electrons, the electrons on the CB of BSO recombine with the holes on the VB of CBO and SNO, which not only retains the e of CBO and SNO with stronger reduction and the h+ of BSO with oxidation ability in the CBS-2 system but also maintains the separation of charge carriers. Some holes on the VB of BSO can directly degrade the TC solution, and others can react with H2O to generate •OH to further degrade the TC solution. In addition, the electrons on the CB of CBO and SBO can convert the adsorbed oxygen into •O2, thereby further oxidatively degrading the TC solution. The construction of the double Z-scheme mechanism improves the efficiency of carrier separation and migration while retaining the strong redox ability of the catalyst, thereby enhancing the photocatalytic performance of the composite.

3. Experimental

3.1. Preparation of Photocatalysts

3.1.1. Synthesis of BSO

Based on previous reports, Bi2Sn2O7 was synthesized using a simple hydrothermal method [40]. Generally, SnCl4·5H2O (2.80 g) was dissolved in 80 mL H2O, Bi(NO3)3·5H2O (3.88 g) was added slowly to it under rapid stirring, and 80 mL 2 M NaOH solution was gradually dropped into it afterward. Then, the suspension was transferred to a 200 mL stainless steel hydrothermal kettle and kept at 180 °C for 24 h, filtered and cleaned with a large amount of water and absolute ethanol after the reaction to separate the powder from the liquid, and finally dried at 60 °C to collect the yellow powder. The material is recorded as BSO.

3.1.2. Synthesis of CBO/BSO

Firstly, Bi(NO3)3·5H2O (2.42 g), Cu(NO3)2·3H2O (0.60 g) and NaOH (0.87 g) were distracted into 80 mL water and stirred for 3 h. Then, a certain amount of BSO was added and stirred for 1 h. Finally, the mixed solution was transferred to a 100 mL stainless steel hydrothermal kettle and kept at 180 °C for 24 h, filtered and cleaned with a large amount of water and absolute ethanol after the reaction to separate the powder from the liquid, and finally dried at 60 °C for use. The mass ratio between CBO and BSO in the prepared samples was 1%, which was denoted as CBO/BSO. Pure CBO samples were prepared through the same protocol without the addition of BSO [30].

3.1.3. Synthesis of CBS

We fully dissolved Na3C6H5O7·2H2O (2.94 g) and SnCl2·2H2O (0.90 g) in 10 mL of water, gradually dropped 10 mL of 0.2 M NaOH into it under continuous stirring, added a certain amount of CBO/BSO after stirring for half an hour, and stirred for another hour. The suspension was put into a 50 mL hydrothermal reactor, and reacted at 180 °C for 12 h; after cooling, the product was filtered and washed with a large amount of water and absolute ethanol, and finally dried at 60 °C for later use. The sample with a mass of CBO accounting for 1% of the total mass and a mass of SNO accounting for 1% of the total mass was labeled as CBS-1, the sample with a the mass of CBO accounting for 1% of the total mass and a mass of SNO accounting for 2% of the total mass was labeled as CBS-2, and the sample with a mass of CBO accounting for 2% of the total mass and a mass of SNO accounting for 5% of the total mass was labeled as CBS-3. Pure SNO samples were prepared through the same protocol without the addition of CBO/BSO [44].

3.2. Characterization Methods

The composition and crystallinity characteristics of the as-prepared photocatalysts were measured with an X-ray powder diffractometer (Shimaduz-6100, Kyoto, Japan) at a scan rate of 10°/min from 10° to 80°. Transmission electron microscope (TEM) images and high-resolution transmission electron microscope (HRTEM) images were used to observe the size and microstructure of the photocatalytic materials. X-ray photoelectron spectroscopy (XPS) was carried out on a Thermo (Waltham, USA) ESCALAB 250Xi spectrometer, which was devoted to analyzing the elemental composition and its valence state in the photocatalysts. The particle size of the characterized photocatalysts was analyzed with a Malvern Zetasizer Nano ZS90. Taking BaSO4 as the reflection standard to test the absorption edge of the sample, the UV-Vis diffuse reflectance spectra (UV-DRS) were performed on a Shimadzu UV-3600 spectrophotometer in the range 200 nm to 800 nm. Electrochemical tests were obtained with a CHI-660D electrochemical workstation. We used 0.1 M sodium sulfate solution as an electrolyte solution in photoelectric tests. The sample electrode played the role of the working electrode. In the beginning, 10 mg of photocatalysts and microscale Nafion were uniformly dispersed in a minute quantity of ethanol via ultrasonic treatment. Then, the mixture was spread evenly on one square centimeter area of FTO glass. Finally, it was dried at 120 °C for 2 h. Photoluminescence (PL) spectra were characterized using a fluorescence spectrometer (Cary Eclipse, Agilent, Santa Clara, CA, USA) at an excitation wavelength of 520 nm. The electron paramagnetic resonance spectra of the as-prepared photocatalysts were characterized with the JES FA200 EPR Spectrometer. Radicals were measured via electron spin resonance (ESR, EOL JES-FA200 EER) using 5,5-dimethyl-1-pyrroline oxynitride (DMPO) as a spin trap.

3.3. Photocatalytic Evaluation

The photocatalytic performance of the samples was evaluated by calculating the degradation rate of tetracycline under 300 W xenon lamp irradiation with a filter (λ > 400 nm). Firstly, an aqueous solution containing tetracycline (100 mL, 5 ppm) and 50 mg photocatalyst was stirred uniformly in the dark for half an hour to reach adsorption–desorption equilibrium. Then, it was placed in a 250 mL cylindrical quartz reactor cooled with cold water circulating outside the reactor. The stirrer was used to ensure that the suspension stayed homogeneous during the photocatalytic reaction, removing the supernatant from the reactor at intervals. Finally, after sampling and centrifugation, the supernatant was filtered through a 0.22 m filter membrane and then quantitatively analyzed via high-performance liquid chromatography. The mobile phase was methanol, acetonitrile, and 0.01 M oxalic acid in a 1:2:7 volume ratio.

4. Conclusions

In summary, we synthesized a double Z-scheme CBS-2 photocatalyst using the in situ generation method and found that it exhibited stronger photocatalytic performance than the constituent parts. The best sample, CBS-2 composite, achieved 89.7% degradation efficiency of TC solution under visible light irradiation. The value of k was about 2.34, 20.16, and 44.17 times those of BSO, CBO, and SNO. The active species •O2 and h+ all participated in the removal of TC, in which •O2 plays a major role. Considering the results of the capture experiment and ESR characterization, the electron transfer pattern follows the Z-scheme heterojunction. The Z-scheme heterojunction suppressed the recombination of photogenerated carriers and retained the strong redox ability of the catalyst. Furthermore, it possessed good stability after cycle experiments. Hence, the dual Z-scheme CBS-2 photocatalyst has a huge application prospect for removing wastewater under visible light. In a word, we advance the study of Z-heterojunction photocatalysis, then provide a new method to improve the photocatalytic activity of BSO, and finally find a photocatalytic material that can degrade tetracycline efficiently.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13071028/s1, Figure S1: TEM image of CBS-2; Figure S2: XPS survey spectra of BSO and CBS-2.

Author Contributions

Conceptualization, J.X. and M.C.; methodology, Y.Z. (Yanlin Zhu); validation, Z.L., X.T., H.G. and Y.Z. (Yaxin Zhao); formal analysis, Y.Z. (Yanlin Zhu); investigation, Y.Z. (Yanlin Zhu); data curation, Y.Z. (Yanlin Zhu); writing—original draft preparation, Y.Z. (Yanlin Zhu); writing—review and editing, Y.Z. (Yanlin Zhu); supervision, J.X.; project administration, J.X. All authors have read and agreed to the published version of the manuscript.

Funding

The Natural Science Foundation of China (51978342), the funding of the Jiangsu Provincial Graduate Research and Innovation Program (SJKY19-0976), and PAPD.

Data Availability Statement

Not applicable.

Acknowledgments

We are grateful for grants from the Natural Science Foundation of China (51978342), the funding of the Jiangsu Provincial Graduate Research and Innovation Program (SJKY19-0976), and PAPD.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. He, Y.; Gao, M.; Zhou, Y.; Zhou, Y. Efficient photocatalytic remediation of typical antibiotics in water via Mn3O4 decorated carbon nitride nanotube. Chemosphere 2023, 311, 136925. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, B.; Li, Z.; Ma, H.; Zhang, J.; Jiao, L.; Hao, H.; Liu, E.; Xu, L.; Wang, C.; Zhou, B. Dynamic construction of self-assembled supramolecular H12SubPcB-OPhCOOH/Ag3PO4 S-scheme arrays for visible photocatalytic oxidation of antibiotics. Appl. Catal. B Environ. 2022, 318, 121882. [Google Scholar] [CrossRef]
  3. Yu, J.; Zhang, P.; Zhang, Y.; Sun, K.; Shi, X.; Li, L. The preparation of conjugated microporous polymer composite materials with montmorillonite template and its improvement in photocatalytic degradation for multiple antibiotics. Appl. Clay Sci. 2023, 231, 106752. [Google Scholar] [CrossRef]
  4. Chin, J.Y.; Ahmad, A.L.; Low, S.C. Evolution of photocatalytic membrane for antibiotics degradation: Perspectives and insights for sustainable environmental remediation. J. Water Process Eng. 2023, 51, 103342. [Google Scholar] [CrossRef]
  5. Hu, X.; Zhang, Y.; Chen, Z.; Gao, Y.; Teppen, B.; Boyd, S.A.; Zhang, W.; Tiedje, J.M.; Li, H. Tetracycline accumulation in biofilms enhances the selection pressure on Escherichia coli for expression of antibiotic resistance. Sci. Total Environ. 2023, 857, 159441. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Li, Y.; Xu, W.; Cui, M.; Wang, M.; Chen, B.; Sun, Y.; Chen, K.; Li, L.; Du, Q. Filtration and adsorption of tetracycline in aqueous solution by copper alginate-carbon nanotubes membrane which has the muscle-skeleton structure. Chem. Eng. Res. Des. 2022, 183, 424–438. [Google Scholar] [CrossRef]
  7. Varadharajan, V.; Senthilkumar, D.S.; Senthilkumar, K.; Sundramurthy, V.P.; Manikandan, R.; Senthilarasan, H.; Ganesan, H.; Kesavamoorthy, I.; Ramasamy, A. Process modeling and toxicological evaluation of adsorption of tetracycline onto the magnetized cotton dust biochar. J. Water Process Eng. 2022, 49, 103046. [Google Scholar] [CrossRef]
  8. Han, K.; Liu, Y.; Hu, J.; Jia, J.; Sun, S. Effect of live and inactivated Chlamydomonas reinhardtii on the removal of tetracycline in aquatic environments. Chemosphere 2022, 309, 136666. [Google Scholar] [CrossRef] [PubMed]
  9. Hu, P.; Shao, J.; Qian, G.; Adeleye, A.S.; Hao, T. Removal of tetracycline by aerobic granular sludge from marine aquaculture wastewater: A molecular dynamics investigation. Bioresour. Technol. 2022, 355, 127286. [Google Scholar] [CrossRef]
  10. Das, S.; Ahn, Y.-H. Synthesis and application of CdS nanorods for LED-based photocatalytic degradation of tetracycline antibiotic. Chemosphere 2022, 291, 132870. [Google Scholar] [CrossRef] [PubMed]
  11. Wang, Y.; Wu, X.; Liu, J.; Zhai, Z.; Yang, Z.; Xia, J.; Deng, S.; Qu, X.; Zhang, H.; Wu, D. Mo-modified band structure and enhanced photocatalytic properties of tin oxide quantum dots for visible-light driven degradation of antibiotic contaminants. J. Environ. Chem. Eng. 2022, 10, 107091. [Google Scholar] [CrossRef]
  12. Pei, Z.; Guo, H.; Zhu, L.; Li, C.; Fu, Z.; Xu, J. Photocatalytic degradation of various antibiotics under visible light irradiation by CdS-doped SiO2@ BiOX (X= Br, Cl) prepared by mixed solvothermal method. Mater. Sci. Eng. B 2023, 287, 116134. [Google Scholar] [CrossRef]
  13. Pan, J.; Wang, L.; Shi, Y.; Li, L.; Xu, Z.; Sun, H.; Guo, F.; Shi, W. Construction of nanodiamonds/UiO-66-NH2 heterojunction for boosted visible-light photocatalytic degradation of antibiotics. Sep. Purif. Technol. 2022, 284, 120270. [Google Scholar] [CrossRef]
  14. Mukhtar, F.; Munawar, T.; Nadeem, M.S.; ur Rehman, M.N.; Khan, S.A.; Koc, M.; Batool, S.; Hasan, M.; Iqbal, F. Dual Z-scheme core-shell PANI-CeO2-Fe2O3-NiO heterostructured nanocomposite for dyes remediation under sunlight and bacterial disinfection. Environ. Res. 2022, 215, 114140. [Google Scholar] [CrossRef] [PubMed]
  15. Munawar, T.; Nadeem, M.S.; Mukhtar, F.; Manzoor, S.; Ashiq, M.N.; Batool, S.; Hasan, M.; Iqbal, F. Multifunctional dual Z-scheme heterostructured Sm2O3-WO3-La2O3 nanocomposite: Enhanced electrochemical, photocatalytic, and antibacterial properties. Adv. Powder Technol. 2023, 34, 104061. [Google Scholar] [CrossRef]
  16. Yin, N.; Chen, H.; Yuan, X.; Zhang, Y.; Zhang, M.; Guo, J.; Zhang, Y.; Qiao, L.; Liu, M.; Song, K. Highly efficient photocatalytic degradation of norfloxacin via Bi2Sn2O7/PDIH Z-scheme heterojunction: Influence and mechanism. J. Hazard. Mater. 2022, 436, 129317. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, Y.; Di, J.; Zhu, X.; Ji, M.; Chen, C.; Liu, Y.; Li, L.; Wei, T.; Li, H.; Xia, J. Chemical Bonding Interface in Bi2Sn2O7/BiOBr S-Scheme Heterojunction Triggering Efficient N2 Photofixation. Appl. Catal. B Environ. 2022, 323, 122148. [Google Scholar] [CrossRef]
  18. Zhou, J.; Fan, G.; Ruan, F.; Li, Y.; Fan, D.; Chen, Q. Construction of Bi2Sn2O7/Ag/Ag3PO4 heterojunction and its photocatalytic degradation properties. J. Taiwan Inst. Chem. Eng. 2022, 138, 104443. [Google Scholar] [CrossRef]
  19. Zhang, D.; Yang, Z.; Hao, J.; Zhang, T.; Sun, Q.; Wang, Y. Boosted charge transfer in dual Z-scheme BiVO4@ ZnIn2S4/Bi2Sn2O7 heterojunctions: Towards superior photocatalytic properties for organic pollutant degradation. Chemosphere 2021, 276, 130226. [Google Scholar] [CrossRef]
  20. Wu, C.; Shen, Q.; Liu, J.; Jiang, L.; Sheng, J.; Li, Y.; Yang, H. Regulation of charge transfer in ZnO/Bi2Sn2O7 heterojunction for enhanced photocatalytic performance towards antibiotic degradation. J. Photochem. Photobiol. A Chem. 2022, 433, 114142. [Google Scholar] [CrossRef]
  21. Wu, C.; Shen, Q.; Zheng, S.; Zhang, X.; Sheng, J.; Yang, H. Fabrication of Bi2Sn2O7@ MIL-100 (Fe) composite photocatalyst with enhanced superoxide-radical-dominated photocatalytic activity for ciprofloxacin degradation. J. Mol. Struct. 2022, 1258, 132657. [Google Scholar] [CrossRef]
  22. Zhang, S.; Rong, Y.; Wei, J.; Li, Z.; Liang, T.; Yu, Z.; Zhu, H.; Wang, S.; Hou, Y. Flower-like microspheres Z-scheme Bi2Sn2O7/NiAl-LDH heterojunction for boosting photocatalytic CO2 reduction under visible light. J. Colloid Interface Sci. 2023, 629, 604–615. [Google Scholar] [CrossRef]
  23. Zhu, P.; Lin, J.; Liu, M.; Duan, M.; Luo, D.; Wu, X.; Zhang, S. Nd2Sn2O7/Bi2Sn2O7/Ag3PO4 double Z-type heterojunction for antibiotic photodegradation under visible light irradiation: Mechanism, optimization and pathways. Sep. Purif. Technol. 2022, 300, 121897. [Google Scholar] [CrossRef]
  24. Lv, D.; Zhang, D.; Pu, X.; Li, Y.; Yin, J.; Ma, H.; Dou, J. Facile synthesis of BiOBr/Bi2Sn2O7 heterojunction photocatalysts with improved photocatalytic activities. J. Am. Ceram. Soc. 2016, 99, 3973–3979. [Google Scholar] [CrossRef]
  25. Wu, X.-F.; Sun, Y.; Li, H.; Wang, Y.-J.; Zhang, C.-X.; Zhang, J.-R.; Su, J.-Z.; Wang, Y.-W.; Zhang, Y.; Wang, C. In-situ synthesis of novel pn heterojunction of Ag2CrO4-Bi2Sn2O7 hybrids for visible-light-driven photocatalysis. J. Alloys Compd. 2018, 740, 1197–1203. [Google Scholar] [CrossRef]
  26. Zhu, Z.; Xia, H.; Li, H.; Han, S. Facile Construction of Bi2Sn2O7/g-C3N4 Heterojunction with Enhanced Photocatalytic Degradation of Norfloxacin. Inorganics 2022, 10, 131. [Google Scholar] [CrossRef]
  27. Zhang, D.; Wu, M.; Hao, J.; Zheng, S.; Yang, Y.; Yao, T.; Wang, Y. Construction of Z-scheme heterojunction by coupling Bi2Sn2O7 and BiOBr with abundant oxygen vacancies: Enhanced photodegradation performance and mechanism insight. J. Colloid Interface Sci. 2022, 612, 550–561. [Google Scholar] [CrossRef]
  28. Zhang, X.; Wang, X.; Chai, J.; Xue, S.; Wang, R.; Jiang, L.; Wang, J.; Zhang, Z.; Dionysiou, D.D. Construction of novel symmetric double Z-scheme BiFeO3/CuBi2O4/BaTiO3 photocatalyst with enhanced solar-light-driven photocatalytic performance for degradation of norfloxacin. Appl. Catal. B Environ. 2020, 272, 119017. [Google Scholar] [CrossRef]
  29. Chen, J.; Yao, L.; Chen, X.; Li, N.; Yu, C.; Zhang, Y.; Lai, Y. Hydrothermal synthesis of dendritic CuBi2O4 and its photocatalytic performance towards tetracycline degradation under different light conditions. Mater. Sci. Semicond. Process. 2022, 142, 106503. [Google Scholar] [CrossRef]
  30. Shi, W.; Guo, F.; Yuan, S. In situ synthesis of Z-scheme Ag3PO4/CuBi2O4 photocatalysts and enhanced photocatalytic performance for the degradation of tetracycline under visible light irradiation. Appl. Catal. B Environ. 2017, 209, 720–728. [Google Scholar] [CrossRef]
  31. Jiang, X.; Wang, M.; Luo, B.; Yang, Z.; Li, W.; Zhang, D.; Pu, X.; Cai, P. Magnetically recoverable flower-like Sn3O4/SnFe2O4 as a type-II heterojunction photocatalyst for efficient degradation of ciprofloxacin. J. Alloys Compd. 2022, 926, 166878. [Google Scholar] [CrossRef]
  32. Li, S.; Liu, Z.; Qu, Z.; Piao, C.; Liu, J.; Xu, D.; Li, X.; Wang, J.; Song, Y. An all-solid-state Z-scheme NaNbO3-Au-Sn3O4 photocatalyst for effective degradation of carbofuran under sunlight irradiation. J. Photochem. Photobiol. A Chem. 2020, 389, 112246. [Google Scholar] [CrossRef]
  33. Li, C.; Yu, S.; Dong, H.; Liu, C.; Wu, H.; Che, H.; Chen, G. Z-scheme mesoporous photocatalyst constructed by modification of Sn3O4 nanoclusters on g-C3N4 nanosheets with improved photocatalytic performance and mechanism insight. Appl. Catal. B Environ. 2018, 238, 284–293. [Google Scholar] [CrossRef]
  34. Mukhtar, F.; Munawar, T.; Nadeem, M.S.; Khan, S.A.; Koc, M.; Batool, S.; Hasan, M.; Iqbal, F. Enhanced sunlight-absorption of Fe2O3 covered by PANI for the photodegradation of organic pollutants and antimicrobial inactivation. Adv. Powder Technol. 2022, 33, 103708. [Google Scholar] [CrossRef]
  35. Sun, X.; Li, J.; Wang, Y.; Zhang, Y.; Yuan, H. The g-C3N4/Bi2Sn2O7@ PAN nanofibes: Enhanced photocatalytic activity in H2 evolution by the formation of heterojunction and in-situ growth. Appl. Surf. Sci. 2023, 608, 155228. [Google Scholar] [CrossRef]
  36. Wang, J.-C.; Wang, B.; Shi, W.; Qiao, X.; Yang, X.; Zhang, L.; Zhang, W.; Li, R.; Hou, Y. Natural-sunlight-driven synchronous degradation of 4-nitrphenol and rhodamine B over S-scheme heterojunction of α-Fe2O3 nanoparticles decorated CuBi2O4 rods. J. Environ. Chem. Eng. 2022, 10, 108565. [Google Scholar] [CrossRef]
  37. Li, S.; Wang, C.; Liu, Y.; Cai, M.; Wang, Y.; Zhang, H.; Guo, Y.; Zhao, W.; Wang, Z.; Chen, X. Photocatalytic degradation of tetracycline antibiotic by a novel Bi2Sn2O7/Bi2MoO6 S-scheme heterojunction: Performance, mechanism insight and toxicity assessment. Chem. Eng. J. 2022, 429, 132519. [Google Scholar] [CrossRef]
  38. Jin, C.-Z.; Yang, X.-A.; Zhai, X.-M.; Wang, S.-B.; Zhang, W.-B. ZnO/Sn3O4 amorphous-crystalline heterojunctions for Cr (VI) visible photocatalysis: Simple synthesis with excellent performance. Appl. Surf. Sci. 2023, 608, 155263. [Google Scholar] [CrossRef]
  39. Sabri, M.; Habibi-Yangjeh, A.; Ghosh, S. Novel ZnO/CuBi2O4 heterostructures for persulfate-assisted photocatalytic degradation of dye contaminants under visible light. J. Photochem. Photobiol. A Chem. 2020, 391, 112397. [Google Scholar] [CrossRef]
  40. Shen, J.-C.; Zeng, H.-Y.; Chen, C.-R.; Xu, S. Novel plasmonic pn heterojunction Ag-Ag2CO3/Bi2Sn2O7 photocatalyst for Cr (VI) reduction. J. Taiwan Inst. Chem. Eng. 2021, 120, 106–115. [Google Scholar] [CrossRef]
  41. Tang, Q.-Y.; Yang, M.-J.; Yang, S.-Y.; Xu, Y.-H. Enhanced photocatalytic degradation of glyphosate over 2D CoS/BiOBr heterojunctions under visible light irradiation. J. Hazard. Mater. 2021, 407, 124798. [Google Scholar] [CrossRef] [PubMed]
  42. Liu, Y.-S.; Yamaguchi, A.; Yang, Y.; El Aisnada, A.N.; Uchida, S.; Abe, H.; Ueda, S.; Yamaguchi, K.; Tanabe, T.; Miyauchi, M. Synthesis and Characterization of the Orthorhombic Sn3O4 Polymorph. Angew. Chem. Int. Ed. 2023, 62, e202300640. [Google Scholar] [CrossRef]
  43. Zhong, D.; Liu, W.; Tan, P.; Zhu, A.; Liu, Y.; Xiong, X.; Pan, J. Insights into the synergy effect of anisotropic {001} and {230} facets of BaTiO3 nanocubes sensitized with CdSe quantum dots for photocatalytic water reduction. Appl. Catal. B Environ. 2018, 227, 1–12. [Google Scholar] [CrossRef]
  44. Zhu, Y.; Cui, Y.; Xiao, B.; Ou-yang, J.; Li, H.; Chen, Z. Z-scheme 2D/2D g-C3N4/Sn3O4 heterojunction for enhanced visible-light photocatalytic H2 evolution and degradation of ciprofloxacin. Mater. Sci. Semicond. Process. 2021, 129, 105767. [Google Scholar] [CrossRef]
  45. Feng, S.; Chen, T.; Liu, Z.; Shi, J.; Yue, X.; Li, Y. Z-scheme CdS/CQDs/g-C3N4 composites with visible-near-infrared light response for efficient photocatalytic organic pollutant degradation. Sci. Total Environ. 2020, 704, 135404. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, B.; Lv, M.; Jiang, W.; Gao, B.; Li, Y.; Zhou, S.; Wang, D.; Liu, C.; Che, G. Self-photoreduced Ag 0-doped Ag (i)–organic frameworks with efficient visible-light-driven photocatalytic performance. CrystEngComm 2021, 23, 7496–7501. [Google Scholar] [CrossRef]
  47. Yang, C.; Feng, S.; Ma, C.; Zhou, Y.; Dai, X.; Ye, Z.; Wang, Y. Bi2Sn2O7/UiO-66-NH2 heterojunction photocatalyst simultaneously adsorbed and photodegraded tetracycline. J. Environ. Chem. Eng. 2023, 11, 109664. [Google Scholar] [CrossRef]
  48. Fan, G.; Zhou, J.; Ruan, F.; Li, Y.; Tian, H.; Fan, D.; Chen, 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]
  49. Heidari, S.; Haghighi, M.; Shabani, M. Ultrasound assisted dispersion of Bi2Sn2O7-C3N4 nanophotocatalyst over various amount of zeolite Y for enhanced solar-light photocatalytic degradation of tetracycline in aqueous solution. Ultrason. Sonochem. 2018, 43, 61–72. [Google Scholar] [CrossRef]
  50. Huang, S.; Zhang, J.; Qin, Y.; Song, F.; Du, C.; Su, Y. Direct Z-scheme SnO2/Bi2Sn2O7 photocatalyst for antibiotics removal: Insight on the enhanced photocatalytic performance and promoted charge separation mechanism. J. Photochem. Photobiol. A Chem. 2021, 404, 112947. [Google Scholar] [CrossRef]
  51. Heidari, S.; Haghighi, M.; Shabani, M. Sono-photodeposition of Ag over sono-fabricated mesoporous Bi2Sn2O7-two dimensional carbon nitride: Type-II plasmonic nano-heterojunction with simulated sunlight-driven elimination of drug. Chem. Eng. J. 2020, 389, 123418. [Google Scholar] [CrossRef]
  52. Wu, S.; Yu, X.; Zhang, J.; Zhang, Y.; Zhu, Y.; Zhu, M. Construction of BiOCl/CuBi2O4 S-scheme heterojunction with oxygen vacancy for enhanced photocatalytic diclofenac degradation and nitric oxide removal. Chem. Eng. J. 2021, 411, 128555. [Google Scholar] [CrossRef]
  53. Zhang, Y.; Sun, A.; Xiong, M.; Macharia, D.K.; Liu, J.; Chen, Z.; Li, M.; Zhang, L. TiO2/BiOI pn junction-decorated carbon fibers as weavable photocatalyst with UV–vis photoresponsive for efficiently degrading various pollutants. Chem. Eng. J. 2021, 415, 129019. [Google Scholar] [CrossRef]
  54. Habibi-Yangjeh, A.; Asadzadeh-Khaneghah, S.; Ghosh, S. Anchoring Bi4O5I2 and AgI nanoparticles over g-C3N4 nanosheets: Impressive visible-light-induced photocatalysts in elimination of hazardous contaminates by a cascade mechanism. Adv. Powder Technol. 2020, 31, 2618–2628. [Google Scholar] [CrossRef]
  55. Shafiq, I.; Hussain, M.; Shehzad, N.; Maafa, I.M.; Akhter, P.; Shafique, S.; Razzaq, A.; Yang, W.; Tahir, M.; Russo, N. The effect of crystal facets and induced porosity on the performance of monoclinic BiVO4 for the enhanced visible-light driven photocatalytic abatement of methylene blue. J. Environ. Chem. Eng. 2019, 7, 103265. [Google Scholar] [CrossRef]
  56. Feng, Z.; Zeng, L.; Zhang, Q.; Ge, S.; Zhao, X.; Lin, H.; He, Y. In situ preparation of g-C3N4/Bi4O5I2 complex and its elevated photoactivity in Methyl Orange degradation under visible light. J. Environ. Sci. 2020, 87, 149–162. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of BSO, CBO, SNO, and CBS-2 composite samples.
Figure 1. XRD patterns of BSO, CBO, SNO, and CBS-2 composite samples.
Catalysts 13 01028 g001
Figure 2. TEM images of (a) BSO, (b) CBO, (c) SNO, (d) HRTEM image of CBS-2.
Figure 2. TEM images of (a) BSO, (b) CBO, (c) SNO, (d) HRTEM image of CBS-2.
Catalysts 13 01028 g002
Figure 3. XPS spectra of (a) Bi 4f, (b) Sn 3d, (c) Cu 2p, (d) O 1s.
Figure 3. XPS spectra of (a) Bi 4f, (b) Sn 3d, (c) Cu 2p, (d) O 1s.
Catalysts 13 01028 g003
Figure 4. (a) UV-Vis diffuse reflectance absorption spectra of BSO, CBO, SNO, and CBS-2, (b) plots of (αhν)1/2 versus hν of BSO and SNO, (c) plots of (αhν)2 versus hν of CBO, Motte-Schottky (M-S) plots of (d) BSO, (e) CBO and (f) SNO.
Figure 4. (a) UV-Vis diffuse reflectance absorption spectra of BSO, CBO, SNO, and CBS-2, (b) plots of (αhν)1/2 versus hν of BSO and SNO, (c) plots of (αhν)2 versus hν of CBO, Motte-Schottky (M-S) plots of (d) BSO, (e) CBO and (f) SNO.
Catalysts 13 01028 g004
Figure 5. (a) Dynamic degradation curves of TC by different samples under visible light illumination, (b) dynamic degradation curves of TC by CBS with different composite proportions, (c) the apparent kinetic constants corresponding to BSO, CBO, SNO, CBO/BSO, and CBS-2 TC degradation, (d) the apparent kinetic constants corresponding to CBS with different composite proportions TC degradation, (e) XRD patterns comparison before and after photocatalytic cycling, (f) photocatalytic cycling tests for TC degradation.
Figure 5. (a) Dynamic degradation curves of TC by different samples under visible light illumination, (b) dynamic degradation curves of TC by CBS with different composite proportions, (c) the apparent kinetic constants corresponding to BSO, CBO, SNO, CBO/BSO, and CBS-2 TC degradation, (d) the apparent kinetic constants corresponding to CBS with different composite proportions TC degradation, (e) XRD patterns comparison before and after photocatalytic cycling, (f) photocatalytic cycling tests for TC degradation.
Catalysts 13 01028 g005
Figure 6. (a) Transient photocurrent response of BSO, CBO, SNO, and CBS-2 composite samples, (b) photoluminescence spectra of BSO, CBO, SNO, and CBS-2 composite samples, (c) electrochemical impedance spectroscopy of BSO, CBO, SNO, and CBS-2, (d) effects of different trapping agents on TC degradation of CBS-2, (e) the ESR signals of •O2 of pure BSO and CBS-2 photocatalysts, (f) the ESR signals of •OH of pure BSO and CBS-2 photocatalysts.
Figure 6. (a) Transient photocurrent response of BSO, CBO, SNO, and CBS-2 composite samples, (b) photoluminescence spectra of BSO, CBO, SNO, and CBS-2 composite samples, (c) electrochemical impedance spectroscopy of BSO, CBO, SNO, and CBS-2, (d) effects of different trapping agents on TC degradation of CBS-2, (e) the ESR signals of •O2 of pure BSO and CBS-2 photocatalysts, (f) the ESR signals of •OH of pure BSO and CBS-2 photocatalysts.
Catalysts 13 01028 g006
Figure 7. Possible mechanism of TC degradation by double Z-scheme CBS-2 composite photocatalyst under visible light.
Figure 7. Possible mechanism of TC degradation by double Z-scheme CBS-2 composite photocatalyst under visible light.
Catalysts 13 01028 g007
Table 1. Comparison of the photocatalytic performance of different photocatalysts.
Table 1. Comparison of the photocatalytic performance of different photocatalysts.
PhotocatalystsQualityTC ContentLight Source EfficiencyRef
Bi2Sn2O7/UiO-66-NH210 mg50 mL
25 mg/L
400 W Xe lamp
λ > 400 nm
86%
(120 min)
[47]
Bi2Sn2O7/Bi2MoO635 mg100 mL
20 mg/L
300 W Xe lamp
λ > 400 nm
99.4%
(100 min)
[37]
BiOBr/Bi2Sn2O720 mg20 mL
20 mg/L
500 W Xe lamp
λ > 400 nm
51.3%
(80 min)
[48]
C3N4/Bi2Sn2O7200 mg100 mL
20 mg/L
300 W Tungsten
Halogen lamp
80.4%
(90 min)
[49]
SnO2/Bi2Sn2O725 mg50 mL
20 mg/L
300 W Xe lamp
λ > 400 nm
88.4%
(180 min)
[50]
Ag/Bi2Sn2O7-C3N4200 mg200 mL
20 mg/L
400 W Tungsten
Halogen lamp
89.1%
(90 min)
[51]
BiOBr/Bi2Sn2O710 mg30 mL
10 mg/L
300 W Xe lamp
λ > 400 nm
50.4%
(60 min)
[27]
Ag@ZnO/Bi2Sn2O750 mg100 mL
10 mg/L
300 W Xe lamp
λ > 400 nm
74%
(60 min)
[20]
Nd2Sn2O7/Bi2Sn2O7/Ag3PO420 mg50 mL
15 mg/L
45 W lamp
λ > 400 nm
97.1%
(120 min)
[23]
CuBi2O4/Bi2Sn2O7/Sn3O450 mg100 mL
5 mg/L
300 W Xe lamp
λ > 400 nm
89.7%
(100 min)
This work
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

Xu, J.; Zhu, Y.; Liu, Z.; Teng, X.; Gao, H.; Zhao, Y.; Chen, M. Synthesis of Dual Z-Scheme CuBi2O4/Bi2Sn2O7/Sn3O4 Photocatalysts with Enhanced Photocatalytic Performance for the Degradation of Tetracycline under Visible Light Irradiation. Catalysts 2023, 13, 1028. https://doi.org/10.3390/catal13071028

AMA Style

Xu J, Zhu Y, Liu Z, Teng X, Gao H, Zhao Y, Chen M. Synthesis of Dual Z-Scheme CuBi2O4/Bi2Sn2O7/Sn3O4 Photocatalysts with Enhanced Photocatalytic Performance for the Degradation of Tetracycline under Visible Light Irradiation. Catalysts. 2023; 13(7):1028. https://doi.org/10.3390/catal13071028

Chicago/Turabian Style

Xu, Jingjing, Yanlin Zhu, Zeyu Liu, Xueyu Teng, Haiqing Gao, Yaxin Zhao, and Mindong Chen. 2023. "Synthesis of Dual Z-Scheme CuBi2O4/Bi2Sn2O7/Sn3O4 Photocatalysts with Enhanced Photocatalytic Performance for the Degradation of Tetracycline under Visible Light Irradiation" Catalysts 13, no. 7: 1028. https://doi.org/10.3390/catal13071028

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