Next Article in Journal
Carrier Recombination in Nitride-Based Light-Emitting Devices: Multiphonon Processes, Excited Defects, and Disordered Heterointerfaces
Previous Article in Journal
Transient Pressure Behavior of CBM Wells during the Injection Fall-Off Test Considering the Quadratic Pressure Gradient
Previous Article in Special Issue
Bromine Ion-Intercalated Layered Bi2WO6 as an Efficient Catalyst for Advanced Oxidation Processes in Tetracycline Pollutant Degradation Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Tight-Connection g-C3N4/BiOBr (001) S-Scheme Heterojunction Photocatalyst for Boosting Photocatalytic Degradation of Organic Pollutants

1
Key Laboratory of Superlight Materials & Surface Technology of Ministry of Education, Harbin Engineering University, Harbin 150001, China
2
Qilu Institute of Technology, College of Biological and Chemical Engineering, Jinan 250200, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2024, 14(13), 1071; https://doi.org/10.3390/nano14131071
Submission received: 19 May 2024 / Revised: 19 June 2024 / Accepted: 20 June 2024 / Published: 22 June 2024

Abstract

:
The efficient separation of photogenerated charge carriers and strong oxidizing properties can improve photocatalytic performance. Here, we combine the construction of a tightly connected S-scheme heterojunction with the exposure of an active crystal plane to prepare g-C3N4/BiOBr for the degradation of high-concentration organic pollutants. This strategy effectively improves the separation efficiency of photogenerated carriers and the number of active sites. Notably, the synthesized g-C3N4/BiOBr displays excellent photocatalytic degradation activity towards various organic pollutants, including methylene blue (MB, 90.8%), congo red (CR, 99.2%), and tetracycline (TC, 89%). Furthermore, the photocatalytic degradation performance of g-C3N4/BiOBr for MB maintains 80% efficiency under natural water quality (tap water, lake water, river water), and a wide pH range (pH = 4–10). Its excellent photocatalytic activity is attributed to the tight connection between g-C3N4 and BiOBr in the S-scheme heterojunction interface, as well as the exposure of highly active (001) crystal planes. These improve the efficiency of the separation of photogenerated carriers, and maintain their strong oxidation capability. This work presents a simple approach to improving the separation of electrons and holes by tightly combining two components within a heterojunction.

1. Introduction

In recent years, the rapid development of industrialization has unavoidably discharged a large amount of wastewater containing organic pollutants, such as dyes, antibiotics, and pesticides, causing health hazards [1,2,3]. Especially high concentrations of pollutants can lead to more serious health hazards. Notably, high concentrations of pollutants can lead to more serious health pitfalls. According to reports in the literature, the wastewater from the production of antibiotics still contains high concentrations of antibiotic residues and has been identified as an important source of antibiotic contamination [4]. Among them, the main source of high-concentration TC is concentrated wastewater discharged by manufacturers [5]. The wastewater produced from concentrated antibiotic production mainly comes from high-concentration organic wastewater that is separated, extracted, and refined, such as waste mother liquor, waste acid water, etc., which is difficult-to-treat wastewater. Dye production wastewater is diluted before discharge, but the dye concentration in the treated wastewater remains at 100–500 mg L−1 [6]. High-concentration organic pollutants are generally difficult to degrade, and traditional physical, chemical, or biological treatment methods often struggle to achieve ideal treatment results. For example, advanced oxidation technologies (such as the Fenton method and ozone oxidation method) can achieve the effective degradation of organic pollutants, but their operating costs are high and difficult to promote and apply. Photocatalytic technology using semiconductors for degrading organic pollutants in wastewater is considered as a promising strategy [7,8,9].
Graphite carbon nitride (g-C3N4) is a typical two-dimensional non-metallic catalyst, presenting suitable energy band positions, a relatively narrow band gap, a unique electronic structure, and stable physical and chemical properties [10]. The positions of the band gaps in g-C3N4 and BiOBr appropriately match (VBg-C3N4 ≈ 1.59 V, CBg-C3N4 ≈ −0.94 V), allowing them to construct a S-scheme heterojunction. Due to having a more negative conduction band (CB) than E0 (O2/O2) (−0.33 V), the photogenerated electrons of g-C3N4 possess the reducing capacity to generate superoxide radicals (·O2), which participate in the degradation process of organic pollutants.
Nowadays, BiOBr has attracted much attention for photocatalytic degradation due to its suitable band gap (2.7–2.9 eV) and unique structures [11,12]. The tetragonal structure of BiOBr consists of alternating the stacking of [Bi2O2]2+ layers and double Br layers, with strong intra-layer covalent bonds and weak inter-layer force [13,14,15,16]. This layer structure provides the space for polarization to improve the separation of photogenerated carriers [17,18,19]. For instance, Zuo et al. synthesized BiOBr/Mn-Ti3C2Tx through in situ ion modification, and the synergistic effect of surface modification and built-in electric field construction effectively improved the carrier separation efficiency and electron utilization efficiency [20]. Zhen et al. optimized the band structure of BiOBr by doping N while introducing abundant oxygen vacancies, resulting in a degradation efficiency of 99.7% for sulfamethoxazole (SIZ) [21]. Yang et al. discovered that S-doping BiOBr enhanced the polarization between the [Bi2O2S]2+ and [Br2]2− layers, resulting in the efficient separation of photogenerated carriers [15].
According to the photocatalytic mechanism, the photocatalytic performance of BiOBr can be promoted by improving the carrier separation efficiency and enhancing surface activity. Currently, the effective methods to improve the photocatalytic efficiency include noble metal modification (Pt, Au, Ag, etc.) [22,23,24,25], crystal plane engineering (exposure of active crystal surface) [26,27,28,29], and heterojunction construction [30,31,32,33]. Among them, the S-scheme heterojunction separates photogenerated carriers and maintains a strong redox ability effectively [34,35].
In 2019, Yu et al. first proposed a stepped scheme charge transfer method based on the traditional Z-scheme charge transfer process to describe the enhanced activity of WO3/g-C3N4 systems [36]. Commonly, a S-scheme heterojunction catalyst combined an oxide semiconductor (OP) with a reduction photocatalyst (RP) [37]. After contract, a built-in electric field (IEF) is formed, which can be attributed to the difference in Fermi energy levels (Ef) [38,39]. The IEF from RP to OP effectively promotes the migration of charge carriers [40]. The reserved holes in the positive valence band (VB) allow BiOBr to be an oxidative photocatalyst in the S-scheme heterojunction.
Furthermore, the modulation of the crystal surface generally is considered to optimize the surface atomic arrangement and surface chemistry. Thus, the crystal surface is a critical factor to enhance the photocatalytic activity of a semiconductor. Sun et al. reported that the exposure of the (001) crystal plane in BiOBr increased the active sites, thereby improving photocatalytic efficiency [41].
Here, we combine the exposure of the active crystal plane with S-heterojunctions to enhance the photocatalytic performance of BiOBr. The exposure of the (001) crystal plane provides more active sites. The tightly connected S-scheme heterojunction between BiOBr and g-C3N4 was observed to provide a channel for electron transfer. The S-scheme heterojunction helps to separate the carriers and retains the strong redox ability of e and h+ to degrade organic pollutants. This study provides a strategy for constructing tight structures in S-scheme heterojunction.

2. Experimental

2.1. Catalyst Preparation

2.1.1. Preparation of Bulk g-C3N4

The bulk g-C3N4 was prepared through pyrolysis of melamine in air. Specifically, melamine (AR, 5 g) was grinded thoroughly in a mortar, and heated at 500 °C for 4 h with a heating rate of 3 °C min−1. The bulk g-C3N4 was obtained after grinding.

2.1.2. Preparation of x-g-C3N4/BiOBr

The x-g-C3N4/BiOBr was synthesized by a hydrothermal method. The KBr (AR, 0.3570 g) was added in the solution of Bi(NO3)3·6H2O (AR, 70 mL, 20.79 g L−1), and stirred for 2 h to obtain a white suspension. According to the weight ratios of the obtained BiOBr/g-C3N4 (10:4, 10:6, 10:8), the g-C3N4 (0.3657 g, 0.5488 g, 0.7317 g) was dispersed in the above white suspension with stirring for 30 min. Then, this suspension was transferred into a kettle (100 mL) and heated at 160 °C for 12 h. The obtained precipitate was washed three times using deionized water and ethanol. Eventually, the x-g-C3N4/BiOBr was collected by drying at 60 °C for 6 h and full grinding. The collected samples were labeled as 4-g-C3N4/BiOBr, 6-g-C3N4/BiOBr (g-C3N4/BiOBr), and 8-g-C3N4/BiOBr.
In addition, the physical mixed samples were for comparison. The physical mixed sample was obtained by grinding the above g-C3N4 (0.5488 g) and BiOBr (0.9147 g) for 10 min.
The preparation of pure BiOBr follows the same steps as above without the addition of g-C3N4.

2.2. Characterization

The crystal structure and phase composition of the prepared samples are characterized by X-ray diffraction (XRD, TTR-III) with Cu-Kα radiation under 40 kV and 150 mV. Fourier transform infrared (FT-IR) spectroscopy is used to characterize the vibrations of functional groups on the catalyst surface. The microscopic morphology of the samples is observed and analyzed by scanning electron microscopy (SEM, SU70-HSD) and transmission electron microscopy (TEM, JEM2010). The thermogravimetric analysis (TGA) curve is measured by a differential thermal analyzer (NETZSCH STA 449 F3), which reflects the relationship between the quality of the reaction material and temperature or time. X-ray photoelectron spectroscopy is utilized to identify the chemical valence states of different elements. The UV-Vis absorption spectra are measured by UV-vis spectrophotometer (UV-2450). Electrochemical experiments are investigated using a three-electrode system (on CHI 660E electrochemical workstation). The photoluminescence (PL) spectra are measured at the excitation wavelength of an Edinburgh FLS 980 instrument.

2.3. Photocatalytic Degradation Experiments

The photocatalytic activity of the photocatalysts was evaluated by the visible-light-induced photodegradation reaction of an organic pollutant (MB, CR, TC) in aqueous solution. A 300 W Xe lamp (λ > 420 nm) was used as the experimental light source for the whole photodegradation reaction. Briefly, 0.05 g of catalyst was put into 100 mL of pollutant (MB (20 mg L−1), CR (20 mg L−1), and TC (50 mg L−1)) solution. Prior to the light exposure, the suspension was stirred continuously for 30 min under dark conditions to reach the absorption–desorption equilibrium. During the light exposure, a certain amount of the contaminant solution was taken from the beaker every 10 min, the upper clear solution was taken after centrifugation, and the absorbance of the filtrate was analyzed by UV-vis spectrophotometer. Calculation of degradation efficiency and kinetic analysis of photocatalyst were done using the following Formulas (1) and (2):
E t = 1 C t / C 0
ln C t C 0 = k t
where Et represents the degradation efficiency, Ct is the concentration of the pollutant, C0 represents the initial concentration of the pollutant, and k represents the rate constant of the photocatalytic reaction.

2.4. Electrochemical Measurements

The slurry of the working electrode was mixed with obtained catalyst (5 mg), isopropanol (IPA, 1 mL), and Nafion (50 μL). The working electrode was prepared by applying it on indium-tin oxide conductive film glass (ITO) and then drying it under the vacuum. The electrolyte solution, counter electrode, and reference electrode were sodium sulfate (0.5 mol L−1, 100 mL), platinum, and Ag/AgCl, respectively.

3. Results and Discussion

3.1. Structure and Morphology

The X-ray diffraction (XRD) patterns of BiOBr, g-C3N4, and x-g-C3N4/BiOBr (x = 4, 6, 8) are indexed to BiOBr (PDF # 09-0393) and g-C3N4 (Figure 1a and Figure S2a). The diffraction peaks at 27.6° from the (002) plane of g-C3N4 were detected in the XRD pattern of g-C3N4/BiOBr, suggesting successful preparation of the g-C3N4/BiOBr [35]. The results of the Fourier transform infrared spectroscopy of the g-C3N4/BiOBr are shown in Figure 1b. The peaks at 515 cm−1, 807 cm−1, and 1032–1772 cm−1 are assigned to the stretching vibration of the Bi-O, 3-s-triazine ring structure and the C-N bond, respectively, indicating that g-C3N4 couples with BiOBr to form a heterojunction [42,43,44]. In addition, the wider absorption peak at 3000–3500 cm−1 corresponds to the stretching vibration of -OH in the water absorbed by the sample surface [14].
The morphologies of the BiOBr, g-C3N4, and g-C3N4/BiOBr are characterized using scanning electron microscopy (SEM) and transmission electron microscopy (TEM). The BiOBr and g-C3N4 possess petal-like and block structures, respectively (Figure 1c,d). It can be observed that the BiOBr embedded into the surface of the block of g-C3N4, showing a tightly connected structure (Figure 1e,f). Figure 2a–e are TEM images of BiOBr, g-C3N4, and g-C3N4/BiOBr. Notably, the distinct boundary is observed in g-C3N4/BiOBr (Figure 2d), implying tight contact. This tight contact structure enables the charge separation and enhances the photocatalytic degradation performance. The lattice stripe (0.28 nm) corresponding to the (110) crystal plane for BiOBr is observed (Figure S4a). Additionally, the amorphous structure of g-C3N4 is evident as the lattice stripes are not detected in g-C3N4 (Figure S4b), which is consistent with the result of the XRD. As seen in the inset of Figure 2f, the selective area electron diffraction (SAED) confirms that the BiOBr platelets are monocrystalline. The base plane is classified as the (001) plane due to the direction of the electron beam being perpendicular to its base plane [45]. Additionally, the presence of Bi, Br, O, C, and N is demonstrated by the TEM mapping (Figure 2g–l).
The X-ray photoelectron spectroscopy (XPS, Figure 3a) survey spectra reveal the existence of Bi, Br, O, C, and N elements in the g-C3N4/BiOBr. The positions of the Bi 4f (Figure 3b), Br 3d (Figure 3c), and O 1s (Figure 3d) peaks in g-C3N4/BiOBr shift slightly to higher binding energy compared those in BiOBr. Conversely, the binding energy of N 1s (Figure 3f) moves towards a lower binding energy position than that of the g-C3N4. This shift in binding energy indicates a strong interaction at the g-C3N4/BiOBr interface instead of a simple physical mixing, which is consistent with the TEM images.

3.2. Band Structure of Catalysts

The UV-visible diffuse reflectance spectra are shown in Figure 4a. The BiOBr and g-C3N4 display absorption edges of 445 nm and 468 nm, respectively. Obviously, the light absorption edge of the g-C3N4/BiOBr exhibits a red shifting to 450 nm, indicating that the optical absorption range is improved by the introduction of g-C3N4.
The energy band structures of g-C3N4/BiOBr are determined by a Tauc plot (Figure 4b) and the Formulas (S1) and (S2) [46,47]. It is calculated that the CB position of the BiOBr and g-C3N4 is 0.33 V and −0.94 V, and the VB position is 3.01 V and 1.59 V, respectively. According to the valence band XPS spectra (Figure 4c,d), the distance from the VB to the Ef is 2.03 eV (BiOBr) and 1.65 eV (g-C3N4). Therefore, the Ef of BiOBr and g-C3N4 is 0.98 V and −0.06 V, respectively. Thus, the CB (−0.94 V) and Ef (−0.06 V) of g-C3N4 are more negative than BiOBr (CB: 0.33 V, Ef: 0.98 V). As depicted in Figure 4e, when g-C3N4 contacts BiOBr, the difference in Ef drives the electrons migrating from BiOBr to g-C3N4 until their Fermi levels equalize. This electronic transfer promotes the separation of carriers and enables the high redox capability of photogenerated e-h+.

3.3. Photocatalytic Degradation

Photocatalytic degradation experiments are carried out under visible light irradiation. The adsorption efficiencies of BiOBr, g-C3N4, and g-C3N4/BiOBr are as slight as 5.2%, 6.6%, and 8.9%, respectively (Figure 5a). After 90 min of visible light irradiation, the photocatalytic efficiency in degrading MB is 79.6% (BiOBr), 70.5% (g-C3N4), and 90.8% (g-C3N4/BiOBr), respectively. The first-order kinetic analysis demonstrates that the photocatalytic kinetic constant of g-C3N4/BiOBr (26.89 min−1 × 10−3) is 1.56 and 1.98 times faster than that of BiOBr (17.22 min−1 × 10−3) and g-C3N4 (13.59 min−1 × 10−3), respectively (Figure 5b). This result indicates the successful construction of S-scheme heterojunctions and the important role of the tightly connected IEF at the interface in photocatalytic degradation.
It is well-known that the degradation of pollutants in a real water environment is not only challenging but also has practical value. The degradation efficiency of g-C3N4/BiOBr toward CR (20 mg L−1), TC (50 mg L−1), and MB (20 mg L−1) is 99.2%, 89%, and 90.8%, respectively, meaning an excellent practical application potential of g-C3N4/BiOBr (Figure 5c). Compared with other materials, g-C3N4/BiOBr maintains relatively good photocatalytic degradation performance (Table S3). In addition, we degraded TC at different concentrations (2 mg L−1,20 mg L−1, and 50 mg L−1, Figure S7) and found that g-C3N4/BiOBr still maintained excellent degradation efficiency for higher and lower concentrations of pollutants.
The degradation performance is commonly affected by catalyst dosage, pH value, and water quality during the process of photocatalytic degradation. As shown in Figure S8a, the photocatalytic degradation efficiency of MB increases from 80.3% (0.25 g L−1) to 92.2% (1.0 g L−1) with a higher catalyst dosage. The effect of the pH value (pH = 4–10) on degradation is tested (Figure 5d). It can be observed that the degradation efficiency remains at 80% over a wide pH range (pH = 4–10).
Photocatalytic degradation experiments in natural water quality are conducted to further investigate the potential application of the g-C3N4/BiOBr. The relevant parameters of natural water bodies have been listed in Table S5. Before degrading the MB solution, the impurities are filtered out to avoid interfering with subsequent measurements. The photocatalytic degradation efficiencies of MB by g-C3N4/BiOBr still remained 95.9% (tap water), 97.8% (lake water), and 99.5% (river water). Thus, the change in water quality has a minimal impact on the degradation of MB (Figure 5e). These findings indicate the excellent practical application potential of g-C3N4/BiOBr.
To evaluate the reusability of the g-C3N4/BiOBr, five consecutive photocatalytic experiments are conducted by adding the same amount of recovered g-C3N4/BiOBr into fresh MB solution after each reaction cycle. Figure 5f exposes that the photocatalytic activity of the g-C3N4/BiOBr remains 80% even after five photo-degradation cycles, certifying the good cycling stability. Simultaneously, XRD patterns (Figure S13a) and SEM images (Figure S13b) of the catalyst after cycling show no significant change, indicating the good cycling stability of g-C3N4/BiOBr.

3.4. Photoelectric Properties

Photoluminescence (PL) and electrochemical impedance spectroscopy (EIS) are utilized to analyze the recombination of photogenerated carriers. The g-C3N4/BiOBr presents the lowest PL intensity (Figure 6a) compared to that of g-C3N4 and BiOBr, implying the efficient separation of electron and hole. In the EIS analysis (Figure 6b), the semicircle radius of g-C3N4/BiOBr is smaller than that of BiOBr and g-C3N4, indicating a lower charge transfer resistance of g-C3N4/BiOBr. This result demonstrates that the tightly connected S-scheme heterojunction structure is beneficial to the migration of photogenerated carriers. Subsequently, the photocurrent intensity of g-C3N4/BiOBr is higher than that of BiOBr and g-C3N4, demonstrating that more photogenerated electrons and holes are generated (Figure 6c). These results exhibit that the S-scheme heterojunction of g-C3N4/BiOBr significantly promotes the separation of photogenerated carriers.

3.5. Photocatalytic Degradation Mechanism

To verify the types of active free radicals, quenching experiments are tested (Figure 6d). Isopropanol (IPA), p-benzoquinone (BQ), and formic acid (FA) were chosen to quench ·OH, ·O2−, and h+, respectively [48,49,50,51,52]. The degradation efficiency is decreased by 8.3%, 24.7%, and 31.7% after introducing ·OH, ·O2, and h+ scavenger, respectively. Therefore, in the g-C3N4/BiOBr degradation system, h+ and ·O2 are the main active species, and the ·OH plays weakly degrades MB.
According to the results of the above characterizations and experiments, we believe that the high activity of photocatalysts comes from three aspects. Firstly, the exposure of (001) crystal planes of BiOBr provides more active sites for photocatalytic reactions. Secondly, the tightly connected S-scheme heterojunction structure provides a fast channel for photogenerated electrons (Figure 2d and Figure 6b). Finally, the construction of S-scheme heterojunctions effectively accelerates the separation of photogenerated carriers and keeps the redox ability. Therefore, the photocatalytic efficiency is significantly improved under the synergistic effect of high-activity (001) crystal planes exposure, a tightly connected structure, and S-type heterojunction.
Based on the above results, the degradation mechanism can be described as shown in Figure 7. An IEF from g-C3N4 to BiOBr is formed. This IEF induces the photogenerated e in the CB of BiOBr to combine with the h+ in the VB of g-C3N4, significantly promoting carrier separation under the irradiation of visible light. Thus, the retained h+ in the VB of BiOBr (3.01 V) can react with H2O/OH to form ·OH or directly react with pollutants, while the retained e in the CB of g-C3N4 can react with O2 dissolved in the solution to form ·O2−. Due to the tightly connected S-scheme heterojunction, effective charge transfer facilitates the generation of ·OH, ·O2−, and h+ to achieve the effective degradation of organic pollutants. The photocatalytic process is as follows (Formulas (3)–(6)):
g - C 3 N 4 / BiOBr   h v e + h +
OH + h+ → ·OH
O2 + e → ·O2
Org. (MB, CR, TC) + ROS (h+, ·OH, ·O2) → CO2 + H2O

4. Conclusions

In summary, we have prepared efficient and closely contacted g-C3N4/BiOBr (001) S-scheme heterojunction photocatalysts using a simple hydrothermal method. The synergy of highly active (001) crystal plane exposure, a tightly connected structure, and the construction of a S-scheme heterojunctions structure significantly enhances the photocatalytic performance of g-C3N4/BiOBr. The exposure of a highly active (001) crystal plane provides more active sites for photocatalytic reactions. The formation of close-contact structures improves the efficiency of electron transfer. And the construction of S-scheme heterojunctions accelerates the separation of charge carriers. Additionally, g-C3N4/BiOBr exhibits an excellent degradation performance for different pollutants (MB: 90.8%, CR: 99.2%, and TC: 89%), wide pH conditions (pH = 4–10), and different water qualities. This work provides a feasible strategy for preparing novel stepped photocatalysts for water treatment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano14131071/s1. Table S1. ICP of x-g-C3N4/BiOBr (x = 4, 6, 8); Figure S1. TG curve of (a) 4-g-C3N4/BiOBr, (b) 6-g-C3N4/BiOBr (g-C3N4/BiOBr), and (c) 8-g-C3N4/BiOBr; Table S2. Composition of x-g-C3N4/BiOBr (x = 4, 6, 8) according TG; Table S3. The degradation efficiency of high concentration TC by different catalysts; Table S4. The water intake location; Table S5. Natural water matrix parameters; Formulas; Figure S2. (a) XRD patterns and (b) FT-IR spectra of g-C3N4, BiOBr, x-g-C3N4/BiOBr (x = 4, 6, 8); Figure S3. SEM images of (a) 4-g-C3N4/BiOBr and (b) 8-g-C3N4/BiOBr; Figure S4. HRTEM images of (a) BiOBr and (b) g-C3N4; Figure S5. Photocatalytic degradation of MB by 4-g-C3N4/BiOBr, 6-g-C3N4/BiOBr, and 8-g-C3N4/BiOBr; Figure S6. The pseudo-first-order kinetic rate plots (a), and the corresponding rate constants in the photocatalytic degradation system (b) of g-C3N4, BiOBr, 6-g-C3N4/BiOBr, and physical mixing; Figure S7. The degradation efficiency of TC at different concentrations (2 mgL−1,20 mgL−1, and 50 mgL−1); Figure S8. (a) Effect of catalyst dosage (0.25–1.00 g L−1), (b) the pseudo-first-order kinetic rate plots, and (c) the corresponding rate constants in the photocatalytic degradation system of different 6-g-C3N4/BiOBr dosage; Figure S9. (a) The pseudo-first-order kinetic rate plots, and (b) the corresponding rate constants in the photocatalytic degradation system of 6-g-C3N4/BiOBr under different initial pH (4–10); Figure S10. The pseudo-first-order kinetic rate plots (a), and the corresponding rate constants in the photocatalytic degradation system (b) of 6-g-C3N4/BiOBr r at actual water quality (tap, lake, and river water); Figure S11. Cycle performance with photocatalytic degradation of MB by 6-g-C3N4/BiOBr; Figure S12. Photocatalytic degradation of MB by 6-g-C3N4/BiOBr in the presence of different scavengers: IPA for quenching ·OH, BQ for quenching ·O2, and FA for quenching h+; Figure S13. (a) XRD and (b) SEM of 6-g-C3N4/BiOBr catalyst before and after cycling. References [53,54,55,56,57,58,59,60] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, X.Z., W.L. and J.F.; methodology, X.Z. and W.L.; formal analysis, X.Z., W.L. and J.F.; investigation, L.H.; writing—original draft preparation, X.Z. and W.L.; writing—review and editing, X.Z., W.L., J.F. and M.G.; supervision, J.F. and M.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cao, N.; Zhao, X.; Gao, M.; Li, Z.; Ding, X.; Li, C.; Liu, K.; Du, X.; Li, W.; Feng, J.; et al. Superior selective adsorption of MgO with abundant oxygen vacancies to removal and recycle reactive dyes. Sep. Purif. Technol. 2021, 275, 119236. [Google Scholar] [CrossRef]
  2. Gao, M.; Li, W.; Su, X.; Li, Z.; Ding, X.; Du, X.; Ren, Y.; Zhang, H.; Feng, J.; Wei, T. A regenerable Cu2O/BiOBr S-scheme heterojunction photocatalysts for efficient photocatalytic degradation of mixed organic pollutants. Sep. Purif. Technol. 2023, 313, 123447. [Google Scholar] [CrossRef]
  3. Qing, Y.; Huang, E.; Cao, L.; Gao, H.; Zhao, S.; Li, Y.; Li, W. Facile synthesis of PA photocatalytic membrane based on C-PANI@BiOBr heterostructures with enhanced photocatalytic removal of 17β-estradiol. Chem. Eng. J. 2024, 490, 151617. [Google Scholar] [CrossRef]
  4. Yi, Q.; Gao, Y.; Zhang, H.; Zhang, H.; Zhang, Y.; Yang, M. Establishment of a pretreatment method for tetracycline production wastewater using enhanced hydrolysis. Chem. Eng. J. 2016, 300, 139–145. [Google Scholar] [CrossRef]
  5. Li, X.; Lu, Z.; Wu, B.; Xie, H.; Liu, G. Antibiotics and antibiotic resistance genes removal in biological aerated filter. Bioresour. Technol. 2024, 395, 130392. [Google Scholar] [CrossRef]
  6. Chen, D.; Hu, X.; Yip, A.C.; Lam FL, Y. Bentonite/cerium-modified LDH composite catalyst for catalytic ozonation of high-concentration indigo carmine dye solution. Appl. Catal. A Gen. 2024, 671, 119560. [Google Scholar] [CrossRef]
  7. Parida, V.K.; Srivastava, S.K.; Chowdhury, S.; Gupta, A.K. Facile synthesis of 2D/0D Bi2O3/MnO2 Z-scheme heterojunction for enhanced visible light-assisted photocatalytic degradation of acetaminophen. Chem. Eng. J. 2023, 472, 144969. [Google Scholar] [CrossRef]
  8. Wang, B.; Zhang, X.; Zhang, R.; Li, Z.; Tian, B.; Ma, H.; Zheng, Z.; Zhou, B.; Ji, M.; Shi, C.; et al. Supramolecularly engineered S-scheme SubPc-Br/MoS2 photocatalyst nanosheets for enhanced photocatalytic degradation of antibiotics. Chem. Eng. J. 2023, 477, 147193. [Google Scholar] [CrossRef]
  9. Tsao, C.-W.; Fang, M.-J.; Hsu, Y.-J. Modulation of interfacial charge dynamics of semiconductor heterostructures for advanced photocatalytic applications. Coord. Chem. Rev. 2021, 438, 213876. [Google Scholar] [CrossRef]
  10. Gao, M.; Li, Z.; Su, X.; Zhang, X.; Chang, J.; Geng, D.; Lu, Y.; Zhang, H.; Wei, T.; Feng, J. 2D/2D MgO/g-C3N4 S-scheme heterogeneous tight with Mg–N bonds for efficient photo-Fenton degradation: Enhancing both oxygen vacancy and charge migration. Chemosphere 2023, 343, 140285. [Google Scholar] [CrossRef]
  11. Jin, Y.; Zhang, J.; Wang, S.; Wang, X.; Fang, M.; Zuo, Q.; Kong, M.; Liu, Z.; Tan, X. Stress modulation on photodegradation of tetracycline by Sn-doped BiOBr. J. Environ. Chem. Eng. 2022, 10, 107675. [Google Scholar] [CrossRef]
  12. Zhang, K.; Zhang, Y.; Zhang, D.; Liu, C.; Zhou, X.; Yang, H.; Qu, J.; He, D. Efficient photocatalytic water disinfection by a novel BP/BiOBr S-scheme heterojunction photocatalyst. Chem. Eng. J. 2023, 468, 143581. [Google Scholar] [CrossRef]
  13. Guo, J.; Liao, X.; Lee, M.-H.; Hyett, G.; Huang, C.-C.; Hewak, D.W.; Mailis, S.; Zhou, W.; Jiang, Z. Experimental and DFT insights of the Zn-doping effects on the visible-light photocatalytic water splitting and dye decomposition over Zn-doped BiOBr photocatalysts. Appl. Catal. B Environ. 2019, 243, 502–512. [Google Scholar] [CrossRef]
  14. Hao, J.; Zhang, Y.; Zhang, L.; Shen, J.; Meng, L.; Wang, X. Restructuring surface frustrated Lewis acid-base pairs of BiOBr through isomorphous Sn doping for boosting photocatalytic CO2 reduction. Chem. Eng. J. 2023, 464, 142536. [Google Scholar] [CrossRef]
  15. Jin, Y.; Li, F.; Li, T.; Xing, X.; Fan, W.; Zhang, L.; Hu, C. Enhanced internal electric field in S-doped BiOBr for intercalation, adsorption and degradation of ciprofloxacin by photoinitiation. Appl. Catal. B Environ. 2022, 302, 120824. [Google Scholar] [CrossRef]
  16. Liu, Y.; Hu, Z.; Yu, J.C. Photocatalytic degradation of ibuprofen on S-doped BiOBr. Chemosphere 2021, 278, 130376. [Google Scholar] [CrossRef] [PubMed]
  17. Meng, L.Y.; Qu, Y.; Jing, L.Q. Recent advances in BiOBr-based photocatalysts for environmental remediation. Chin. Chem. Lett. 2021, 32, 3265–3276. [Google Scholar] [CrossRef]
  18. Wei, X.-X.; Cui, H.; Guo, S.; Zhao, L.; Li, W. Hybrid BiOBr–TiO2 nanocomposites with high visible light photocatalytic activity for water treatment. J. Hazard. Mater. 2013, 263, 650–658. [Google Scholar] [CrossRef] [PubMed]
  19. Zhong, S.; Wang, Y.; Chen, Y.; Jiang, X.; Lin, M.; Lin, C.; Lin, T.; Gao, M.; Zhao, C.; Wu, X. Improved piezo-photocatalysis for aquatic multi-pollutant removal via BiOBr/BaTiO3 heterojunction construction. Chem. Eng. J. 2024, 488, 151002. [Google Scholar] [CrossRef]
  20. Zuo, S.; Huang, S.; Wang, Y.; Wan, J.; Yan, Z.; Ma, Y.; Wang, S. Surface modify BiOBr/Mn-Ti3C2Tx to enhance the carrier separation efficiency and weak electron utilization for photo-assisted peroxymonosulfate activation to degrade emerging contaminants. Chem. Eng. J. 2023, 475, 146230. [Google Scholar] [CrossRef]
  21. Li, Z.; Wen, C.; Li, D.; Fang, Z.; Lin, Z.; Liu, D.; Wang, Y.; Zhang, X.; Chen, P.; Lv, W.; et al. Insights into nitrogen-doped BiOBr with oxygen vacancy and carbon quantum dots photocatalysts for the degradation of sulfonamide antibiotics: Actions to promote exciton dissociation and carrier migration. Chem. Eng. J. 2024, 492, 152449. [Google Scholar] [CrossRef]
  22. Ren, G.; Shi, M.; Li, Z.; Zhang, Z.; Meng, X. Electronic metal-support interaction via defective-induced platinum modified BiOBr for photocatalytic N2 fixation. Appl. Catal. B Environ. 2023, 327, 122462. [Google Scholar] [CrossRef]
  23. Hu, J.; Chen, Y.; Zhou, Y.; Zeng, L.; Huang, Y.; Lan, S.; Zhu, M. Piezo-enhanced charge carrier separation over plasmonic Au-BiOBr for piezo-photocatalytic carbamazepine removal. Appl. Catal. B Environ. 2022, 311, 121369. [Google Scholar] [CrossRef]
  24. Tian, J.; Zhong, K.; Zhu, X.; Yang, J.; Mo, Z.; Liu, J.; Dai, J.; She, Y.; Song, Y.; Li, H.; et al. Highly exposed active sites of Au nanoclusters for photocatalytic CO2 reduction. Chem. Eng. J. 2023, 451, 138392. [Google Scholar] [CrossRef]
  25. Liu, G.; Wang, L.; Wang, B.; Zhu, X.; Yang, J.; Liu, P.; Zhu, W.; Chen, Z.; Xia, J. Synchronous activation of Ag nanoparticles and BiOBr for boosting solar-driven CO2 reduction. Chin. Chem. Lett. 2023, 34, 107962. [Google Scholar] [CrossRef]
  26. Allagui, L.; Chouchene, B.; Gries, T.; Medjahdi, G.; Girot, E.; Framboisier, X.; Amara, A.B.H.; Balan, L.; Schneider, R. Core/shell rGO/BiOBr particles with visible photocatalytic activity towards water pollutants. Appl. Surf. Sci. 2019, 490, 580–591. [Google Scholar] [CrossRef]
  27. Jiang, H.; Wang, Q.; Chen, P.; Zheng, H.; Shi, J.; Shu, H.; Liu, Y. Photocatalytic degradation of tetracycline by using a regenerable (Bi)BiOBr/rGO composite. J. Clean. Prod. 2022, 339, 130771. [Google Scholar] [CrossRef]
  28. Ji, M.; Zhang, Z.; Xia, J.; Di, J.; Liu, Y.; Chen, R.; Yin, S.; Zhang, S.; Li, H. Enhanced photocatalytic performance of carbon quantum dots/BiOBr composite and mechanism investigation. Chin. Chem. Lett. 2018, 29, 805–810. [Google Scholar] [CrossRef]
  29. Wu, M.; Wang, H.; Zhang, B.; Wang, J.; Fan, M.; Dong, L.; Li, B.; Chen, Z.; Chen, G. Dual electron reservoirs: Synergistic photocatalytic degradation of antibiotics by oxygen vacancy-rich BiOBr nanoflowers and magnetic carbon nanotubes. Surf. Interfaces 2024, 45, 103947. [Google Scholar] [CrossRef]
  30. Liu, T.; Hu, K.; Li, Y.; Wang, Y.; Han, D.; Wang, Z.; Gu, F. The Z-Scheme MIL-88B(Fe)/BiOBr Heterojunction Promotes Fe(III)/Fe(II) Cycling and Photocatalytic-Fenton-Like Synergistically Enhances the Degradation of Ciprofloxacin. Small 2024, 2309541. [Google Scholar] [CrossRef]
  31. Low, B.Q.L.; Jiang, W.; Yang, J.; Zhang, M.; Wu, X.; Zhu, H.; Zhu, H.; Heng, J.Z.X.; Tang, K.Y.; Wu, W.-Y.; et al. 2D/2D Heterojunction of BiOBr/BiOI Nanosheets for In Situ H2O2 Production and Activation toward Efficient Photocatalytic Wastewater Treatment. Small Methods 2023, 8, 2301368. [Google Scholar] [CrossRef]
  32. Tang, D.; Chen, X.; Yan, J.; Xiong, Z.; Lou, X.; Ye, C.; Chen, J.; Qiu, T. Facile one-pot synthesis of a BiOBr/Bi2WO6 heterojunction with enhanced visible-light photocatalytic activity for tetracycline degradation. Chin. J. Chem. Eng. 2023, 53, 222–231. [Google Scholar] [CrossRef]
  33. Ma, J.; Xu, L.; Yin, Z.; Li, Z.; Dong, X.; Song, Z.; Chen, D.; Hu, R.; Wang, Q.; Han, J.; et al. “One stone four birds” design atom co-sharing BiOBr/Bi2S3 S-scheme heterojunction photothermal synergistic enhanced full-spectrum photocatalytic activity. Appl. Catal. B Environ. 2024, 344, 123601. [Google Scholar] [CrossRef]
  34. Hu, H.; Zhang, X.; Zhang, K.; Ma, Y.; Wang, H.; Li, H.; Huang, H.; Sun, X.; Ma, T. Construction of a 2D/2D Crystalline Porous Materials Based S-Scheme Heterojunction for Efficient Photocatalytic H2 Production. Adv. Energy Mater. 2024, 14, 2303638. [Google Scholar] [CrossRef]
  35. Zu, D.; Ying, Y.; Wei, Q.; Xiong, P.; Ahmed, M.S.; Lin, Z.; Li, M.M.-J.; Li, M.; Xu, Z.; Chen, G.; et al. Oxygen Vacancies Trigger Rapid Charge Transport Channels at the Engineered Interface of S-scheme Heterojunction for Boosting Photocatalytic Performance. Angew. Chem. Int. Ed. 2024, e202405756. [Google Scholar] [CrossRef]
  36. Zhang, B.; Hu, X.; Liu, E.; Fan, J. Novel S-scheme 2D/2D BiOBr/g-C3N4 heterojunctions with enhanced photocatalytic activity. Chin. J. Catal. 2021, 42, 1519–1529. [Google Scholar] [CrossRef]
  37. Xu, Q.; Wageh, S.; Al-Ghamdi, A.A.; Li, X. Design principle of S-scheme heterojunction photocatalyst. J. Mater. Sci. Technol. 2022, 124, 171–173. [Google Scholar] [CrossRef]
  38. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-Scheme Heterojunction Photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  39. Gu, X.; Lin, S.; Qi, K.; Yan, Y.; Li, R.; Popkov, V.; Almjasheva, O. Application of tungsten oxide and its composites in photocatalysis. Sep. Purif. Technol. 2024, 345, 127299. [Google Scholar] [CrossRef]
  40. Bhat, S.S.M.; Jang, H.W. Recent Advances in Bismuth-Based Nanomaterials for Photoelectrochemical Water Splitting. ChemSusChem 2017, 10, 3001–3018. [Google Scholar] [CrossRef]
  41. Sun, J.; Li, X.; Zhao, Q.; Liu, B. Ultrathin nanoflake-assembled hierarchical BiOBr microflower with highly exposed {001} facets for efficient photocatalytic degradation of gaseous ortho-dichlorobenzene. Appl. Catal. B Environ. 2021, 281, 119478. [Google Scholar] [CrossRef]
  42. Dou, X.C.; Zhang, C.L.; Shi, H.F. The simultaneous promotion of Cr (VI) photoreduction and tetracycline removal over 3D/2D Cu2O/BiOBr S-scheme nanostructures. Sep. Purif. Technol. 2022, 282, 120023. [Google Scholar] [CrossRef]
  43. Lv, J.; Dai, K.; Zhang, J.; Liu, Q.; Liang, C.; Zhu, G. Facile constructing novel 2D porous g-C3N4/BiOBr hybrid with enhanced visible-light-driven photocatalytic activity. Sep. Purif. Technol. 2017, 178, 6–17. [Google Scholar] [CrossRef]
  44. Ni, J.; Wang, W.; Liu, D.; Zhu, Q.; Jia, J.; Tian, J.; Li, Z.; Wang, X.; Xing, Z. Oxygen vacancy-mediated sandwich-structural TiO2−x/ultrathin g-C3N4/TiO2−x direct Z-scheme heterojunction visible-light-driven photocatalyst for efficient removal of high toxic tetracycline antibiotics. J. Hazard. Mater. 2021, 408, 124432. [Google Scholar] [CrossRef]
  45. Luo, Z.; Ye, X.; Zhang, S.; Xue, S.; Yang, C.; Hou, Y.; Xing, W.; Yu, R.; Sun, J.; Yu, Z.; et al. Unveiling the charge transfer dynamics steered by built-in electric fields in BiOBr photocatalysts. Nat. Commun. 2022, 13, 2230. [Google Scholar] [CrossRef]
  46. Gao, M.; Feng, J.; He, F.; Zeng, W.; Wang, X.; Ren, Y.; Wei, T. Carbon microspheres work as an electron bridge for degrading high concentration MB in CoFe2O4@carbon microsphere/g-C3N4 with a hierarchical sandwich-structure. Appl. Surf. Sci. 2020, 507, 145167. [Google Scholar] [CrossRef]
  47. Li, H.; Deng, F.; Zheng, Y.; Hua, L.; Qu, C.; Luo, X. Visible-light-driven Z-scheme rGO/Bi2S3–BiOBr heterojunctions with tunable exposed BiOBr (102) facets for efficient synchronous photocatalytic degradation of 2-nitrophenol and Cr(vi) reduction. Environ. Sci. Nano 2019, 6, 3670–3683. [Google Scholar] [CrossRef]
  48. Huang, C.; Wen, Y.; Ma, J.; Dong, D.; Shen, Y.; Liu, S.; Ma, H.; Zhang, Y. Unraveling fundamental active units in carbon nitride for photocatalytic oxidation reactions. Nat. Commun. 2021, 12, 320. [Google Scholar] [CrossRef]
  49. Zhao, Y.; An, H.; Feng, J.; Ren, Y.; Ma, J. Impact of Crystal Types of AgFeO2 Nanoparticles on the Peroxymonosulfate Activation in the Water. Environ. Sci. Technol. 2019, 53, 4500–4510. [Google Scholar] [CrossRef]
  50. Sá, J.; Agüera, C.A.; Gross, S.; Anderson, J.A. Photocatalytic nitrate reduction over metal modified TiO2. Appl. Catal. B Environ. 2009, 85, 192–200. [Google Scholar] [CrossRef]
  51. Su, X.; Cao, N.; Feng, J.; Li, W.; Ding, X.; Li, Z.; Gao, M.; Hu, L.; Zhang, H.; Ren, Y. The enhanced photocatalytic performance of the amorphous carbon/MgO nanofibers: Insight into the role of the oxygen vacancies and 1D morphology. Appl. Surf. Sci. 2023, 616, 156470. [Google Scholar] [CrossRef]
  52. Yang, W.; Sun, K.; Wan, J.; Ma, Y.-A.; Liu, J.; Zhu, B.; Liu, L.; Fu, F. Boosting holes generation and O2 activation by bifunctional NiCoP modified Bi4O5Br2 for efficient photocatalytic aerobic oxidation. Appl. Catal. B Environ. 2023, 320, 121978. [Google Scholar] [CrossRef]
  53. Lu, W.; Xu, L.; Shen, X.; Meng, L.; Pan, Y.; Zhang, Y.; Han, J.; Mei, X.; Qiao, W.; Gan, L. Highly efficient activation of sulfite by p-type S-doped g-C3N4 under visible light for emerging contaminants degradation. Chem. Eng. J. 2023, 472, 144708. [Google Scholar] [CrossRef]
  54. Wang, S.; Wu, J.; Lu, X.; Xu, W.; Gong, Q.; Ding, J.; Dan, B.; Xie, P. Removal of acetaminophen in the Fe2+/persulfate system: Kinetic model and degradation pathways. Chem. Eng. J. 2019, 358, 1091–1100. [Google Scholar] [CrossRef]
  55. Luo, C.; Wang, S.; Wu, D.; Cheng, X.; Ren, H. UV/Nitrate photocatalysis for degradation of Methylene blue in wastewater: Kinetics, transformation products, and toxicity assessment. Environ. Technol. Innov. 2022, 25, 102198. [Google Scholar] [CrossRef]
  56. Yuan, X.; Yang, J.; Yao, Y.; Shen, H.; Meng, Y.; Xie, B.; Ni, Z.; Xia, S. Preparation, characterization and photodegradation mechanism of 0D/2D Cu2O/BiOCl S-scheme heterojunction for efficient photodegradation of tetracycline. Sep. Purif. Technol. 2022, 291, 120965. [Google Scholar]
  57. Yang, Q.; Tan, G.; Yin, L.; Liu, W.; Zhang, B.; Feng, S.; Bi, Y.; Liu, T.; Wang, Z.; Ren, H.; et al. Full-spectrum broad-spectrum degradation of antibiotics by BiVO4@BiOCl crystal plane S-type and Z-type heterojunctions. Chem. Eng. J. 2023, 467, 143405. [Google Scholar] [CrossRef]
  58. Shi, W.; Hao, C.; Fu, Y.; Guo, F.; Tang, Y.; Yan, X. Enhancement of synergistic effect photocatalytic/persulfate activation for degradation of antibiotics by the combination of photo-induced electrons and carbon dots. Chem. Eng. J. 2022, 433, 133741. [Google Scholar] [CrossRef]
  59. Lian, Z.; Wu, T.; Zhang, X.; Cai, S.; Xiong, Y.; Yang, R. Synergistic degradation of tetracycline from Mo2C/MoOx films mediated peroxymonosulfate activation and visible-light triggered photocatalysis. Chem. Eng. J. 2023, 469, 143774. [Google Scholar] [CrossRef]
  60. Hang, Y.; Zhou, J.; Chen, X.; Wang, L.; Cai, W. Coupling of heterogeneous advanced oxidation processes and photocatalysis in efficient degradation of tetracycline hydrochloride by Fe-based MOFs: Synergistic effect and degradation pathway. Chem. Eng. J. 2019, 369, 745–757. [Google Scholar]
Figure 1. (a) XRD patterns; (b) FT-IR spectra; and (cf) SEM images of BiOBr, g-C3N4, and g-C3N4/BiOBr.
Figure 1. (a) XRD patterns; (b) FT-IR spectra; and (cf) SEM images of BiOBr, g-C3N4, and g-C3N4/BiOBr.
Nanomaterials 14 01071 g001
Figure 2. TEM images of (a) BiOBr, (b) g-C3N4, and (c) g-C3N4/BiOBr; (df) HRTEM images of g-C3N4/BiOBr; (gl) TEM mapping: Bi, Br, O, C, and N.
Figure 2. TEM images of (a) BiOBr, (b) g-C3N4, and (c) g-C3N4/BiOBr; (df) HRTEM images of g-C3N4/BiOBr; (gl) TEM mapping: Bi, Br, O, C, and N.
Nanomaterials 14 01071 g002
Figure 3. (a) XPS spectra for the survey, (b) Bi 4f, (c) Br 3d, (d) O 1s, (e) C 1s, and (f) N 1s XPS spectra of BiOBr, g-C3N4, and g-C3N4/BiOBr.
Figure 3. (a) XPS spectra for the survey, (b) Bi 4f, (c) Br 3d, (d) O 1s, (e) C 1s, and (f) N 1s XPS spectra of BiOBr, g-C3N4, and g-C3N4/BiOBr.
Nanomaterials 14 01071 g003
Figure 4. (a) UV-visible diffuse-reflectance spectra; (b) the Tauc plot of BiOBr and g-C3N4; (c,d) the valance band XPS spectra of BiOBr and g-C3N4; and (e) the band structure of BiOBr and g-C3N4.
Figure 4. (a) UV-visible diffuse-reflectance spectra; (b) the Tauc plot of BiOBr and g-C3N4; (c,d) the valance band XPS spectra of BiOBr and g-C3N4; and (e) the band structure of BiOBr and g-C3N4.
Nanomaterials 14 01071 g004
Figure 5. (a) Photocatalytic degradation efficiency of g-C3N4, BiOBr, physical mixed samples, and g-C3N4/BiOBr; (b) first-order kinetic constants of MB degradation; (c) photocatalytic degradation of CR, TC, and MB; (d) initial pH (4–10); (e) photocatalytic degradation of MB in natural water quality; (f) the cyclic catalytic degradation of MB in the presence of g-C3N4/BiOBr.
Figure 5. (a) Photocatalytic degradation efficiency of g-C3N4, BiOBr, physical mixed samples, and g-C3N4/BiOBr; (b) first-order kinetic constants of MB degradation; (c) photocatalytic degradation of CR, TC, and MB; (d) initial pH (4–10); (e) photocatalytic degradation of MB in natural water quality; (f) the cyclic catalytic degradation of MB in the presence of g-C3N4/BiOBr.
Nanomaterials 14 01071 g005
Figure 6. (a) PL spectra; (b) EIS spectra; (c) transient photocurrent responses; (d) quenching experiments.
Figure 6. (a) PL spectra; (b) EIS spectra; (c) transient photocurrent responses; (d) quenching experiments.
Nanomaterials 14 01071 g006
Figure 7. Photocatalytic mechanism of MB by the g-C3N4/BiOBr.
Figure 7. Photocatalytic mechanism of MB by the g-C3N4/BiOBr.
Nanomaterials 14 01071 g007
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

Zhang, X.; Li, W.; Hu, L.; Gao, M.; Feng, J. A Tight-Connection g-C3N4/BiOBr (001) S-Scheme Heterojunction Photocatalyst for Boosting Photocatalytic Degradation of Organic Pollutants. Nanomaterials 2024, 14, 1071. https://doi.org/10.3390/nano14131071

AMA Style

Zhang X, Li W, Hu L, Gao M, Feng J. A Tight-Connection g-C3N4/BiOBr (001) S-Scheme Heterojunction Photocatalyst for Boosting Photocatalytic Degradation of Organic Pollutants. Nanomaterials. 2024; 14(13):1071. https://doi.org/10.3390/nano14131071

Chicago/Turabian Style

Zhang, Xinyi, Weixia Li, Liangqing Hu, Mingming Gao, and Jing Feng. 2024. "A Tight-Connection g-C3N4/BiOBr (001) S-Scheme Heterojunction Photocatalyst for Boosting Photocatalytic Degradation of Organic Pollutants" Nanomaterials 14, no. 13: 1071. https://doi.org/10.3390/nano14131071

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