Next Article in Journal
The Influence of Zn Substitution on Physical Properties of CoFe2O4 Nanoparticles
Previous Article in Journal
Surface Defects Improved SERS Activity of Nanoporous Gold Prepared by Electrochemical Dealloying
Previous Article in Special Issue
Self-Sustained Three-Dimensional Macroporous TiO2-Graphene Photocatalyst for Sunlight Decolorization of Methyl Orange
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Photo-Assisted Fenton Degradation of Antibiotics over Iron-Doped Bi-Rich Bismuth Oxybromide Photocatalyst

School of Chemistry and Chemical Engineering, Qingdao University, Qingdao 266071, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(1), 188; https://doi.org/10.3390/nano13010188
Submission received: 19 November 2022 / Revised: 25 December 2022 / Accepted: 27 December 2022 / Published: 31 December 2022
(This article belongs to the Special Issue Hybrid Nanomaterials Applied to Photocatalysis)

Abstract

:
Herein, combining photocatalysis and Fenton oxidation, a photo-assisted Fenton system was conducted using Fe-doped Bi4O5Br2 as a highly efficient photocatalyst to realize the complete degradation of Tetracycline antibiotics under visible light. It has been observed that the optimized photocatalyst 5%Fe-doped Bi4O5Br2 exhibits a degradation efficiency of 100% for Tetracycline with H2O2 after 3 h visible-light irradiation, while a degradation percentage of 59.8% over the same photocatalyst and 46.6% over pure Bi4O5Br2 were obtained without the addition of H2O2 (non-Fenton process). It is unambiguous that a boost photo-assisted Fenton system for the degradation of Tetracycline has been established. Based on structural analysis, it demonstrated that the Fe atoms in place of the Bi sites may result in the distortion of the local structure, which induced the occurrence of the spontaneous polarization and thus enhanced the built-in electric field. The charge separation efficiency is enhanced, and the recombination of electrons and holes is inhabited so that more charges are generated to reach the surface of the photocatalyst and therefore improve the photocatalytic degradation efficiency. Moreover, more Fe (II) sites formed on the 5%Fe-Bi4O5Br2 photocatalyst and facilitated the activation of H2O2 to form oxidative species, which greatly enhanced the degradation efficiency of Tetracycline.

1. Introduction

Antibiotics are quite effective in resisting microorganisms and treating biological diseases [1,2]. Tetracycline (TC), the quintessential antibiotic, is commonly employed, possessing the broadest antibacterial spectrum as well as a tough chemical structure [3]. However, the massive misuse of TC must be accompanied by incomplete metabolism and accidental spills due to its intrinsic toxicity and constancy. Tiny TC can seriously poison the aquatic environment and conversely threaten life on Earth [4,5]. Therefore, removing tetracycline and other antibiotics has become one of the important research topics in the environmental field.
The photocatalytic technique as an advanced oxidation processes (AOPs) stands out among various technologies to purify antibiotic sewage, which is apparently a “green” process driven by sustainable solar energy [6]. It is presently significant to harvest visible light (at least 50% of the solar spectrum) for efficient photocatalytic elimination of TC. As a novel photocatalyst, bismuth oxybromide (BiOBr) exhibits potential applications on energy generation, environmental remediation, bacterial disinfection and so on, particularly in wastewater treatment [7,8]. Pristine BiOBr possesses a bandgap of ~2.8 eV determined by a sandwich-like molecular structure with [Bi2O2]2+ and a double layer of Br; however, the weak visible-light harvest and slow carrier separation efficiency hindered its wide applications [9]. To alleviate the above symptom, a “Bi-rich” strategy is employed to optimize the unique molecular structure, and a series of Bi-rich bismuth oxybromide (Bi4O5Br2, Bi3O4Br, Bi24O31Br10, Bi5O7Br, et al.) are employed [10,11,12,13,14]. The suitable bandgap (~2.5 eV) and the great charge transfer property of Bi4O5Br2 demonstrate more photogenerated carriers excited by visible light, which is used to modulate the valence band (VB, contributed by O 2p and Br 4p) and conduction band (CB, contributed by Bi 6p) potentials through enhancing the local structure [15]. Li et al. claimed that more superoxide radicals (O2) may be generated in the negative CB potential of Bi4O5Br2 compared to that of BiOBr, which is beneficial for the degradation of ciprofloxacin [16]. Thus, tuning the band structure and optimizing the carrier dynamics are of great importance to expedite a photocatalytic reaction.
It is worthy to mention that the band structure engineering of Bi4O5Br2 is also carried out via deliberately doping the foreign heteroatoms (anions, cations), which influences the potential difference due to the spontaneous polarization and internal electric field [17,18,19]. Therefore, the photogenerated carriers are effectively separated and transferred driven by the built-in electric field [19]. Wang et al. found that iodine, also in the halogen family, was doped in Bi4O5Br2, resulting in a negative shift of VB potential, and thus improved the separation of the carriers [20]. Thus, more O2 and h+ generated from I0.7-Bi4O5Br2 attacked the parabens with a higher efficiency. Zhang et al. studied the photocatalytic NO removement of Mn-doped Bi4O5Br2 and observed a superior photocatalytic oxidation activity owing to the excellent O2 capture ability and the strong oxidative hydroxyl radical (OH) generated from the photoelectrons [21]. However, unfortunately it is still a critical task for an entire degradation of TC contained in wastewater.
Recently, a “green” process that combines the advanced photocatalytic technique with the intensive Fenton oxidation process (heterogeneous photo-assisted Fenton oxidation) was reported [22,23,24], by which an entire degradation of TC can be achieved efficiently [25,26]. In this photo-Fenton process, more photogenerated charges shifted onto the photocatalyst surface and played a decisive role on the degradation efficiency of organic pollutants because more charges may activate the iron ions by reacting with H2O2 to produce more O2 and OH [27,28,29]. The GO-FePO4 was first employed as a heterogeneous photo−Fenton catalyst to the degradation of rhodamine B by Yu et al., which reported the H2O2 could be easily activated under solar energy excitation [30]. Peng et al. introduced Fe into g-C3N4 and created a Fe-g-C3N4/Bi2WO6 photocatalyst, which is able to activate H2O2 to facilitate the degradation of TC [31]. Zhou et al. claimed the doping of Fe into Rectorite not only showed the quick decomposition performance of H2O2 but also enhanced the charge transfer on the material [32]. Cheng et al. prepared a LaFeO3/BiOI heterojunction photocatalyst, combining the superiority of charge separation in the heterojunction and advanced Fenton activity of LaFeO3 [33]. Thus, the photogenerated electrons are transferred to Fe (II)/Fe (III) and activated H2O2, which makes the degradation of organic pollutants [34,35,36].
In this manuscript, Fe-doped Bi4O5Br2 nanosheets were synthesized to enhance the degradation of TC under visible-light irradiation. The successful doping of Fe ions resulted in the variation of the local structure distortion, which enhanced the spontaneous polarization and thus improved the built-in electric field. This improvement boosted the photons harvest and photocarriers separation. Moreover, the activation of H2O2 by the doped Fe2+/Fe3+ formed O2 and OH. The dependence of photo-assisted Fenton oxidative antibiotics on the structure of the photocatalysts is investigated in detail.

2. Materials and Methods

2.1. Materials

All the chemical reagents used in the experiments, including bismuth nitrate pentahydrate, iron nitrate nonahydrate, potassium bromide, ethylene glycol, sodium hydroxide and ethanol absolute are analytically pure and purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China) without any further purification.

2.2. Preparation of the Photocatalysts

The one-step solvothermal method was used for the preparation of all the photocatalysts in this study. The steps are illustrated as Scheme 1: first, bismuth nitrate pentahydrate and iron nitrate nonahydrate were dissolved in ethylene glycol (25 mL) at various mass ratios (Bi:Fe = 99:1, 97:3, 95:5 and 93:7) and labeled as solution A, and 2 mmol potassium bromide was dissolved into 10 mL ethylene glycol, labeled as solution B. Then solution B was slowly added into solution A by stirring vigorously for 30 min until they were homogeneously mixed. The pH of the mixed solution was regulated to 10.5 using a 2.0 mol/L sodium hydroxide solution. Finally, the mixture was charged into a 100 mL Teflon-lined autoclave and kept at 160 °C for 12 h. The product was washed with deionized water and anhydrous ethanol 3 times, respectively, and then dried at 70 °C. The resulting products were labeled as Bi4O5Br2, 1%Fe-Bi4O5Br2, 3%Fe-Bi4O5Br2, 5%Fe-Bi4O5Br2 and 7%Fe-Bi4O5Br2, respectively.

2.3. Characterizations

Raman vibrational spectra were achieved on a DXR2 spectrometer (Thermo, Waltham MA, USA), which was excited by a 532 nm laser. The crystal structures of the samples were performed on an Ultima IV (XRD, Rigaku, Tokyo, Japan) with Cu Kα irradiation (λ = 0.15418 nm), and the range of diffraction angles was controlled at 5–80°. The morphology and element mapping images were taken on a Regulus 8100 (SEM, JOEL, Tokyo, Japan). The local images were taken on a 2100F (TEM, HRTEM, JOEL, Tokyo, Japan). The XPS spectra were conducted on Escalab Xi+ (Thermo) with Al-Kα (hν = 1486.6 eV), which were calibrated based on C1s of adventitious carbon at 284.8 eV. The diffuse reflectance UV-Vis spectra were obtained on a UV-2700 (Shimadzu, Tokyo, Japan) using BaSO4 as background. The nitrogen adsorption and desorption isotherm and pore size distribution were measured and analyzed by Autosorb-iQ-MP-C (Quantachrome, Norcross, GA, USA). The PL spectra were conducted on a FLS 980 (Ediburgh, livingston, Scotland, UK) using a Xenon lamp with an excitation wavelength of 375 nm at room temperature (293 K). The in-situ ESR signals were collected on an EMX plus (Bruker, Berlin, Germany) under visible-light irradiation (λ ≥ 420 nm). The photoelectrochemical measurements were carried out on PARSTAT 4000A (Princeton, NJ, USA). The ICP–MS test was measured on an Agilent 730ES (Palo Alto, CA, USA).

2.4. Degradation Experiments of Tetracycline Hydrochloride

The photocatalytic activity of the catalyst was evaluated by the degradation of Tetracycline hydrochloride (TC) under visible-light irradiation. Specifically, 30 mg of photocatalyst was dispersed into a 20 mg/L 50 mL TC solution and the suspension was stirred at a certain speed. Firstly, a 4 mL reaction mixture was taken every 30 min under dark conditions to ensure that the degradation reached the adsorption–desorption equilibrium. The light source used in this experiment was a 300 W Xe lamp with a 420 nm cutoff filter. Following this step, 4 mL samples were taken every 30 min after turning on the light. The obtained reaction mixture was centrifuged, and then the concentration of the tetracycline was analyzed by measuring the absorbance at 356 nm on a UV–Vis spectrophotometer. The photo-Fenton process was operated similarly to the above photocatalytic degradation of TC, except that a concentration of H2O2 was 0.28 mol/L in the reaction system. The experiments under darkness are the same as the above steps without light.

3. Results and Discussion

3.1. The Structure of the Synthesized Samples

X-ray diffraction (XRD) patterns were carried out to determine the crystal structure of the samples. Figure 1a clearly shows five characteristic peaks located at 24.2°, 29.2°, 31.8°, 42.9° and 45.6° (JCPDS No. 037-0699), which can be indexed to (112), (11-3), (020), (105) and (422) lattice planes of pure Bi4O5Br2, respectively. The Fe-doped samples showed peaks at similar locations without new diffraction peaks observed. The XRD patterns demonstrated that pure Bi4O5Br2 and Fe-doped Bi4O5Br2 structures were formed. It should be noted that the peak corresponding to the (11-3) lattice plane shifts gradually to a higher angle for Fe-doped Bi4O5Br2 samples. This observation can be explained by the fact that a smaller radius of the Fe atom replaces the Bi atom, resulting in the distortion of the local structure of Bi4O5Br2. It also indicated that the Fe atoms have been doped into the Bi4O5Br2 framework in the form of substituting Bi atoms on the basis of Goldschmidt’s rule [37].
The molecular structure variation and the lattice distortion induced by the Fe atom doping can also be confirmed by the Raman spectra. As illustrated in Figure 1b, the characteristic peaks at 96.6 cm−1 and 156.3 cm−1 could be assigned to A1g and E1g of the internal Bi−Br stretching mode of Bi4O5Br2, respectively [38]. The peak intensity represents the content of the vibratory group, and the half-peak width is related to the number of layers in the same space [39]. It is clear that variations in the half peak width and shift in the vibrational frequency of the peak can be observed with the introduction of Fe atoms into the framework of Bi4O5Br2. Moreover, the intensity of peaks assigned to A1g and E1g modes gradually reduced with the increasing amount of Fe dopant, and even these peaks almost disappeared in the 7%Fe-Bi4O5Br2 sample. This result indicates that the number of Bi-Br groups decreased due to the substitution of the Fe atom doping [40]. In contrast, the widening in the half-peak of Fe-doped Bi4O5Br2 indicates that the Bi-Br layers in the same space reduced in number and thus the lattice distortion is formed, which resulted from the doping of Fe atoms with a smaller radius.
The specific surface area (SBET) and pore distribution of the samples were investigated by low temperature N2 physical adsorption. As shown in Figure 1c, it indicated that all samples showed a type IV N2 adsorption-desorption isotherm with an H3 hysteresis loop. This result shows that there might be a large number of active vacancies on the surface of the sample, which further confirmed that the doping of surface Fe atoms resulted in the increase in vacancies. The SBET of pure Bi4O5Br2 is 24.6 m2/g, while all the Fe-doped Bi4O5Br2 increased in value. Among them, the SBET of 7%Fe-Bi4O5Br2 is the largest one attaining to 75.0 m2/g. Similarly, the 5%Fe-Bi4O5Br2 increases 2.63-fold in the SBET than that of pure Bi4O5Br2, which is 64.7 m2/g. The increases in the SBET due to the doping of Fe atoms could offer a large number of reactive sites to participate in photodegradation reactions.
The pore size distribution of samples is mainly in the range of 2–50 nm (mesoporous) as shown in Figure 1d. Notably, the sample of 5%Fe-Bi4O5Br2 shows many macropores in the range of 80–140 nm. The formation of these mesopores is beneficial for the adsorption of pollutants in wastewater, and then effectively decomposed the pollutants [41]. As displayed in Table 1, it was found that the pore volume of the sample increased due to iron doping. Comparatively, the pore volume of 5%Fe-Bi4O5Br2 is 0.431 m2/g compared to that of pure Bi4O5Br2 (0.328 m2/g).
The FESEM, TEM and HRTEM are conducted to show the morphologies and the microstructure of the samples. It is obvious that both Bi4O5Br2 (Figure 2a and Figure 3a) and 5%Fe-Bi4O5Br2 (Figure 2b and Figure 3b) show nanosheet-interlaced morphologies; the thickness of 5%Fe-Bi4O5Br2 is close to 8.3 nm, which is much thinner than that of Bi4O5Br2 (about 12.5 nm). This observation indicates that the doping of Fe atoms reduced the thickness of the nanosheet, which thus reduced the transfer distance of photogenerated carriers to the surface and therefore facilitates more carriers to participate in the reaction. The lattice fringes of the samples can be clearly observed in the HRTEM of Figure 3c,d, and the lattice spacing of both Bi4O5Br2 and 5%Fe-Bi4O5Br2 are 0.28 nm, which is consistent with the index of the (11-3) lattice plane. This result indicates that Fe-doped Bi4O5Br2 with high crystallinity is successfully synthesized. The elemental mappings (Figure 2c) show that the O, Fe, Bi and Br elements are uniformly dispersed in 5%Fe-Bi4O5Br2, which indicates that the Fe element is highly distributed in the sample.
The elemental composition and states on the surface of samples are determined by X-ray photoelectron spectroscopy (XPS). As shown in Figure 4a, the obvious signals of Bi, Br and O are collected in the survey spectrum of Bi4O5Br2, and on the basis, an extra signal of Fe is captured on 5%Fe-Bi4O5Br2. In detail, the Bi 4f5/2 and Bi 4f7/2 peaks at 164.5 eV and 159.2 eV, representing the presence of Bi3+, are detected in the Bi 4f refined spectra for Bi4O5Br2 (Figure 4b). In comparison, these peaks shift ~0.15 eV to low binding energy for 5%Fe-Bi4O5Br2, confirming that the density of the electron cloud around the Bi atom increased. This result can be explained by the fact that the electronegativity of the Fe element (1.83) is lower than that of the Bi element (1.9) [42,43]. Therefore, the electron from the doping of Fe is easily injected into the Bi sites. As shown in Figure 4c, the Br 3d signal in Bi4O5Br2 can be deconvoluted into Br 3d5/2 and Br 3d3/2 at 68.6 eV and 69.6 eV, belonging to the Br, respectively. Similarly, a low binding energy shift of about 0.15 eV occurred in the Br 3d spectrum of 5%Fe-Bi4O5Br2. Therefore, the electron density of the Br atom plate is stronger, which is affected by the doping of Fe possessing more valence electrons compared to the Bi atom [44]. Furthermore, in the high-resolution O 1s spectra (Figure 4d), two peaks at 529.8 eV and 531.2 eV are observed in Bi4O5Br2 and 5%Fe-Bi4O5Br2, which are ascribed to the intrinsic lattice oxygen and absorbed oxygen [9]. In addition, Bi4O5Br2 possessed an additional peak at 533.1 eV assigned to the absorbed water [45]. Deeply, the signal of Fe 2p can be fitted into four peaks for 5%Fe-Bi4O5Br2 corresponding to Fe 2p1/2 and Fe 2p3/2, each of which contains two different valence states of Fe (Figure 4e). Among them, the two peaks at 725.8 eV and 712.5 eV are assigned to the Fe3+. Meanwhile, the other two peaks at 722.6 eV and 710.0 eV are assigned to the Fe2+ [45,46]. In terms of contributing to the Fe 2p, the content of Fe2+ accounted for 0.587%, which is larger than Fe3+ (0.413%), indicating the surface content of Fe2+ is higher. Therefore, the abundant redox Fe2+/Fe3+ on the surface are facilitated to excite H2O2 to boost the degradation performance of the TC.
The photophysical properties of the samples are measured by UV–vis diffuse reflectance spectroscopy. As shown in Figure 5a, the absorbance edge of pure Bi4O5Br2 is around 500 nm, indicating that this sample is a visible-light responsive narrow bandgap photocatalyst. It is obvious that the red-shift in the absorbance edge with the introduction of the Fe atom can be observed, demonstrating that the absorbance spectrum in the range of visible light is broadened with the increasing amount of dopant. In addition, the band gap of Fe-Bi4O5Br2 becomes narrow compared to that of pure Bi4O5Br2, reflected in the Tauc plot (Figure 5b). The bandgaps of Bi4O5Br2, 1%Fe-Bi4O5Br2, 3%Fe-Bi4O5Br2, 5%Fe-Bi4O5Br2 and 7%Fe-Bi4O5Br2 were calculated to be about 2.26, 1.87, 1.91, 1.71 and 1.49 eV, respectively. The doping of Fe atoms may result in the formation of an intermediate state energy level, thus narrowing the bandgap of the sample [47,48]. The flat-band potentials of the samples were measured by the electrochemical method reflected in the Mott-Schottky plots (Figure 5c). Both samples showed curves of n-type semiconductors in which the slope of the fitted linear graph is positive. The flat-band potential of pure Bi4O5Br2 and 5%Fe-Bi4O5Br2 are −0.42 eV and −0.20 eV (vs. Ag/AgCl), respectively. The conduction band (CB) position is at a negative 0.1~0.2 eV compared to the flat-band potential for N-type semiconductors [49]. Thus by Equation (1), the CB of pure Bi4O5Br2 and 5%Fe-Bi4O5Br2 are −0.43 and −0.21 eV, respectively (vs. NHE). The valance band (VB) position is calculated via Equation (2), and finally, the distinct band structure is plotted in Figure 5d. In detail, the 5%Fe-Bi4O5Br2 gives a more positive conduction band compared to the single-electron reduction in O2 (E [O2/O2] = −0.33 eV vs. NHE). Moreover, based on the bandgap of 1.71 eV, the VB potential of 5%Fe-Bi4O5Br2 is 1.50 eV (vs. NHE), which is more negative than that of Bi4O5Br2 (1.83 eV vs. NHE).
E NHE = E Ag / AgCl + 0.0591 × pH + E Ag / AgCl   0 ; E Ag / AgCl 0 = 0.1976 V   vs .   NHE ( 25   ° C )
E V B = E g + E C B
Figure 6a shows the transient photocurrent density response of Bi4O5Br2 and Fe-doped Bi4O5Br2 samples. It can be observed that the series of Fe-doped samples produce intensive current responses compared to Bi4O5Br2 under visible-light irradiation. The photocurrent intensity of the 5%Fe-Bi4O5Br2 photocatalyst exhibited much stronger than that of pure Bi4O5Br2. The enhanced photocurrent confirmed that the effective separation of the photogenerated electron−hole pairs is driven by a boosted internal electric field generated by the doping of Fe. Furthermore, the transfer resistance of carriers within the samples could be evaluated from the electrochemical impedance spectroscopy (EIS). Generally, the shorter semi-cycle arc radius in the Nyquist plot means a lower resistance to the charge transfer. Therefore, as shown in Figure 6b, the semicircular radius in the Nyquist plots of the Fe-Bi4O5Br2 series samples decreases with the introduction of Fe, which confirmed that the transfer resistances of carriers in the Fe-Bi4O5Br2 series samples are less than that in Bi4O5Br2. The semi−cycle arc radius in the Nyquist plot of 5%Fe-Bi4O5Br2 is the shortest one, indicative of the quickest carrier transferring rate of this sample being generated. In summary, the doping of Fe atoms into the framework of Bi4O5Br2 resulted in the occurrence of spontaneous polarization and therefore enhancement in the internal electric field, which effectively induced the separation and rapid migration of the photogenerated electron–hole pairs [50].
The recombination kinetics of photogenerated carriers can be comprehensively simulated by the measurement of fluorescence spectra. In the steady-state photoluminescence (PL) spectra (Figure 6c), 5%Fe-Bi4O5Br2 responds with a weaker fluorescence intensity compared to Bi4O5Br2 upon excitation at a 375 nm exciting light source. The result indicated that the recombination of photogenerated carriers was inhibited in 5%Fe-Bi4O5Br2, which is consistent with the result observed in the photocurrent. In addition, the average fluorescence lifetime (τ) of photogenerated carriers is evaluated in the time−resolved photoluminescence spectra (TRPL). As shown in Figure 6d, the τ value of 5%Fe-Bi4O5Br2 is 3.112 ns, which is effectively extended than that of Bi4O5Br2 (2.909 ns), suggesting that facilitated carriers with longer residence time are generated in the sample. These results further demonstrated that the photogenerated carriers can be effectively separated and transferred driven by the built−in electron field enhanced by 5%Fe-Bi4O5Br2.

3.2. Degradation Performance of TC

The photocatalytic activities of the Bi4O5Br2 and Fe-Bi4O5Br2 samples were investigated by the degradation of TC under visible-light irradiation. As shown in Figure 7a, it is clear that almost no loss of TC in the absence of the photocatalyst in the control experiment can be observed, which confirmed that the degradation of TC is entirely derived from the action of the photocatalysts. Moreover, a dark treatment for 1 h is enough for the TC to reach an adsorption–desorption equilibrium on the surface of all the samples. The 5%Fe-Bi4O5Br2 sample shows the strongest adsorption for TC, almost three times as much as Bi4O5Br2, which is attributed to a larger specific surface area exposing more adsorption sites. Significantly, 5%Fe-Bi4O5Br2 possesses the best photocatalytic degradation activity for TC under visible-light irradiation for 3 h, reaching 60%. However, a further increase in the amount of the dopant into Bi4O5Br2 induces an adverse effect on the photocatalytic performance. It can be observed that the photocatalytic activity of 7%Fe-Bi4O5Br2 for the removal of TC decreases to only 28.9%, even worse than that of pure Bi4O5Br2 (46.6%). Therefore, the appropriate ratio of Fe doping can induce the spontaneous polarization and enhance the built-in electric field, thus achieving effective separation of the photogenerated carrier. Moreover, the photocatalytic degradation of TC is evaluated in the first-order reaction by Equation (3). Further, as shown in Figure 7b, the reaction rate constant (k) of 5%Fe-Bi4O5Br2 is fitted to 0.006 min−1, which increases nearly 1.5-fold than that of Bi4O5Br2 (0.004 min−1).
In   C t = In C 0 k 1 t
Furthermore, the photo−Fenton system is employed to improve the degradation efficiency of TC, and the dependence of the activity on the structure of the photocatalyst in the presence of H2O2 is studied. As shown in Figure 7c, without the photocatalyst while only the H2O2 system is added, 29.4% of TC degradation activity is achieved under visible-light irradiation. It is interesting that when the Fe-Bi4O5Br2 series of samples are supplied, the degradation performance of TC in the presence of H2O2 under visible-light irradiation is promoted significantly and is much higher than that of Bi4O5Br2 with the H2O2 system (57.9%). The highest photocatalytic activity over the optimized sample, 5%Fe-Bi4O5Br2, can reach the entire removal of TC in 3 h under visible-light irradiation. Similarly, the degradation reaction rate constant of TC in the photo−Fenton system is also estimated according to the first-order reaction. As shown in Figure 7d, the highest k is 0.016 min−1 for the 5%Fe-Bi4O5Br2-Fenton system, which is almost 3.2 times greater compared to the Bi4O5Br2-H2O2 system (0.005 min−1). This result indicates that the optimized recombination kinetics of the photogenerated carrier over 5%Fe-Bi4O5Br2 facilitates to activate H2O2 sufficiently to generate strong oxidation OH and expedites the degradation reaction kinetic.
As shown in Figure 7e, the Fenton process for the degradation of TC is evaluated under darkness. Specifically, almost 17% of TC is removed only with H2O2, and the Bi4O5Br2-H2O2 shows the degradation ratio of TC is about 27%, which is comparable to that of only with H2O2 regardless of the adsorption capacity of the photocatalyst. Apparently, the degradation performance is enhanced in the 5%Fe-Bi4O5Br2-H2O2 Fenton system (42%). It demonstrates that the 5%Fe-Bi4O5Br2 could activate H2O2 via the exposed Fe ions under darkness, whereas Bi4O5Br2 could not.
From the above results, the 5%Fe-Bi4O5Br2 photocatalyst showed excellent activity with light and H2O2. Finally, in order to study the photocatalytic stability of the catalyst, a cyclic experiment was carried out, as shown in Figure 7f. After five cycles for the degradation of TC, the 5%Fe-Bi4O5Br2 still retains its excellent performance (85%), with only a 15% loss of degradation ratio compared to the fresh reaction. In addition, it is found that without obvious variations of the crystal structure of the samples after five cycles can be observed. Moreover, the concentration of Fe ions in the solution was negligible (about 0.022 mg/L) in the used 5%Fe-Bi4O5Br2 after the cycle experiments in Figure 7h.
In order to explore the active species over Fe-Bi4O5Br2 in the photocatalytic process, an in-situ ESR technique is employed. The radical capture reagent TEMPO (2,2,6,6-tetramethyl-1-piperidinyloxy) is utilized in the photocatalytic system to remove the h+. As shown in Figure 8a,b, the triple peak in the ESR signal shows the characteristic species of TEMPO with h+. The decreases in the peak intensity represent that the generated h+ is captured by TEMPO. It is obvious that the decrease in peak intensity of the 5%Fe-Bi4O5Br2 is more prominent than that of the Bi4O5Br2 after 10 min visible-light irradiation, indicating that the 5%Fe-Bi4O5Br2 may be able to generate holes more efficiently during the photocatalytic degradation. The generated OH is captured by the 5,5-dimethyl-1-pyrroline (DMPO) in the 0.28 M H2O2 aqueous solution. As shown in Figure 8c,d, four characteristic peaks with an intensity of 1:2:2:1 are collected, demonstrating that the Bi4O5Br2 and the 5%Fe-Bi4O5Br2 can activate H2O2 efficiently under visible-light irradiation [51]. Additionally, the more intensive signal of DMPO confirmed that more photogenerated carriers in the 5%Fe-Bi4O5Br2 could be used to activate H2O2 to produce a greater number of OH.
As shown in Figure 8e, the characteristic signal of DMPO-O2 was not observed on the 5%Fe-Bi4O5Br2 regardless of the irradiation or not. This result indicated that the O2 species could not be generated through the single−electron reduction in oxygen under visible-light irradiation; for that the CB potential of Fe-Bi4O5Br2 is −0.21 eV vs. NHE (E [O2/O2] = −0.33 eV vs. NHE). Interestingly, the ESR signal of O2 can be observed with 10 min irradiation in the presence of H2O2 in Figure 8f. It is therefore demonstrated that the O2 species could be generated by the activation of H2O2 through excited 5%Fe-Bi4O5Br2. So, both the OH and O2 could be generated by visible-light irradiation of the Fe-Bi4O5Br2 photocatalyst to realize the efficient degradation of tetracycline.
In summary, combined with the above discussion of free radicals, the mechanism and diagram can be presumed and shown in Equations (4)–(10) and Figure 9. Under the irradiation of visible light, the carriers are separated and shifted efficiently to the photocatalyst surface (Equation (4)). Due to the enhancement in the spontaneous polarization effect, a great number of photogenerated charges reach the surface of 5%Fe-Bi4O5Br2. More holes are transferred to the surface of the 5%Fe-Bi4O5Br2 photocatalyst from the in−situ ESR spectrum. Photogenerated holes may oxidize TC (Equation (5)), while the Fe3+ species is reduced to Fe2+ by the photogenerated electrons (Equation (6)). Moreover, more Fe(II) sites were synthetized on the 5%Fe-Bi4O5Br2 from the XPS spectrum. Due to the addition of hydrogen peroxide, the Fenton effect in the reaction, that is, Fe2+ reacts with H2O2 to form Fe3+ and produce OH (Equation (7)), and generated OH can degrade TC (Equation (9)). Finally, Fe3+ combined with hydrogen peroxide will produce O2 (Equation (8)) [52], which can further degrade TC pollutants (Equation (10)).
Photocatalyst + hv h + + e
h + + TC Degradation   products H 2 O + CO 2
e + Fe 3 + Fe 2 +
Fe 2 + + H 2 O 2 Fe 3 + + OH + OH
Fe 3 + + H 2 O 2 Fe 2 + + O 2 + 2 H +
OH + + TC Degradation   products H 2 O + CO 2
O 2 + TC Degradation   products H 2 O + CO 2

4. Conclusions

In conclusion, the photo-Fenton degradation system was constructed via Fe-doped Bi4O5Br2 nanosheet. The obtained 5%Fe-Bi4O5Br2-photo-Fenton system realized excellent degradation performance (almost 100%) for the removal of Tetracycline under visible-light irradiation. The apparent reaction rate constant of 5%Fe-Bi4O5Br2 reaches 0.016 min−1, which is almost 3.2 times faster than that of Bi4O5Br2 (0.005 min−1). The complete removal of Tetracycline and the enhanced reaction rate are primarily attributed to the multiple effects of Fe doping in the 5%Fe-Bi4O5Br2-photo-Fenton system. On one hand, the doping of Fe induces the spontaneous polarization, which enhances the built-in electric field to modulate the photogenerated carrier separation kinetics of 5%Fe-Bi4O5Br2. On the other hand, the Fe(II) in the molecular framework becomes the activation center of H2O2 and the abundant photogenerated carriers accelerate the transition between Fe(II) and Fe(III), thus efficiently activating H2O2 to generate O2 and more OH. The holes also play a decisive role to oxidize Tetracycline under visible-light irradiation. These active species are confirmed by in-situ ESR. Therefore, the 5%Fe-Bi4O5Br2-photo-Fenton system is a potential candidate for “green” photocatalytic removal of antibiotics.

Author Contributions

Conceptualization, Y.Z.; methodology, F.Z.; formal analysis, F.Z., Z.L. and X.Y.; investigation, F.Z.; composing the original draft, F.Z.; writing—review and editing, Y.Z. and Y.P.; supervision, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The financial support from “National Natural Science Foundation of China (22005164)”, “Shandong Excellent Young Scientist Research Award Fund (No. BS2015CL002)”, “Basic Research Project of Qingdao Source Innovation Program Fund (17-1-1-82-jch)” and “Shanghai Xuntian Technology Co., Ltd.” are greatly appreciated.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Catteau, L.; Zhu, L.; Van Bambeke, F.; Quetin-Leclercq, J. Natural and hemi−synthetic pentacyclic triterpenes as antimicrobials and resistance modifying agents against Staphylococcus aureus: A review. Phytochem. Rev. 2018, 17, 1129–1163. [Google Scholar] [CrossRef]
  2. Luo, Y.; Xu, L.; Rysz, M.; Wang, Y.; Zhang, H.; Alvarez, P.J.J. Occurrence and Transport of Tetracycline, Sulfonamide, Quinolone, and Macrolide Antibiotics in the Haihe River Basin, China. Environ. Sci. Technol. 2011, 45, 1827–1833. [Google Scholar] [CrossRef]
  3. Chen, Q.; Guo, X.; Hua, G.; Li, G.; Feng, R.; Liu, X. Migration and degradation of swine farm tetracyclines at the river catchment scale: Can the multi-pond system mitigate pollution risk to receiving rivers? Environ. Pollut. 2017, 220, 1301–1310. [Google Scholar] [CrossRef] [PubMed]
  4. Kumar, M.; Jaiswal, S.; Sodhi, K.K.; Shree, P.; Singh, D.K.; Agrawal, P.K.; Shukla, P. Antibiotics bioremediation: Perspectives on its ecotoxicity and resistance. Environ. Int. 2019, 124, 448–461. [Google Scholar] [CrossRef] [PubMed]
  5. Sarmah, A.K.; Meyer, M.T.; Boxall, A.B.A. A global perspective on the use, sales, exposure pathways, occurrence, fate and effects of veterinary antibiotics (VAs) in the environment. Chemosphere 2006, 65, 725–759. [Google Scholar] [CrossRef]
  6. Xia, L.; Sun, Z.; Wu, Y.; Yu, X.-F.; Cheng, J.; Zhang, K.; Sarina, S.; Zhu, H.-Y.; Weerathunga, H.; Zhang, L.; et al. Leveraging doping and defect engineering to modulate exciton dissociation in graphitic carbon nitride for photocatalytic elimination of marine oil spill. Chem. Eng. J. 2022, 439, 135668. [Google Scholar] [CrossRef]
  7. Lyu, J.; Hu, Z.; Li, Z.; Ge, M. Removal of tetracycline by BiOBr microspheres with oxygen vacancies: Combination of adsorption and photocatalysis. J. Phys. Chem. Solids 2019, 129, 61–70. [Google Scholar] [CrossRef]
  8. 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]
  9. Wu, Z.; Shen, J.; Ma, N.; Li, Z.; Wu, M.; Xu, D.; Zhang, S.; Feng, W.; Zhu, Y. Bi4O5Br2 nanosheets with vertical aligned facets for efficient visible-light-driven photodegradation of BPA. Appl. Catal. B Environ. 2021, 286, 119937. [Google Scholar] [CrossRef]
  10. Mao, D.; Ding, S.; Meng, L.; Dai, Y.; Sun, C.; Yang, S.; He, H. One-pot microemulsion−mediated synthesis of Bi-rich Bi4O5Br2 with controllable morphologies and excellent visible-light photocatalytic removal of pollutants. Appl. Catal. B Environ. 2017, 207, 153–165. [Google Scholar] [CrossRef]
  11. Wang, S.; Hai, X.; Ding, X.; Chang, K.; Xiang, Y.; Meng, X.; Yang, Z.; Chen, H.; Ye, J. Light-Switchable Oxygen Vacancies in Ultrafine Bi5O7Br Nanotubes for Boosting Solar−Driven Nitrogen Fixation in Pure Water. Adv. Mater. 2017, 29, 1701774. [Google Scholar] [CrossRef] [PubMed]
  12. Xiao, X.; Zheng, C.; Lu, M.; Zhang, L.; Liu, F.; Zuo, X.; Nan, J. Deficient Bi24O31Br10 as a highly efficient photocatalyst for selective oxidation of benzyl alcohol into benzaldehyde under blue LED irradiation. Appl. Catal. B Environ. 2018, 228, 142–151. [Google Scholar] [CrossRef]
  13. Zhang, W.; Peng, Y.; Yang, Y.; Zhang, L.; Bian, Z.; Wang, H. Bismuth−rich strategy intensifies the molecular oxygen activation and internal electrical field for the photocatalytic degradation of tetracycline hydrochloride. Chem. Eng. J. 2022, 430, 132963. [Google Scholar] [CrossRef]
  14. Bai, Y.; Yang, P.; Wang, L.; Yang, B.; Xie, H.; Zhou, Y.; Ye, L. Ultrathin Bi4O5Br2 nanosheets for selective photocatalytic CO2 conversion into CO. Chem. Eng. J. 2019, 360, 473–482. [Google Scholar] [CrossRef]
  15. Jin, X.; Lv, C.; Zhou, X.; Xie, H.; Sun, S.; Liu, Y.; Meng, Q.; Chen, G. A bismuth rich hollow Bi4O5Br2 photocatalyst enables dramatic CO2 reduction activity. Nano Energy 2019, 64, 103955. [Google Scholar] [CrossRef]
  16. Di, J.; Xia, J.; Ji, M.; Yin, S.; Li, H.; Xu, H.; Zhang, Q.; Li, H. Controllable synthesis of Bi4O5Br2 ultrathin nanosheets for photocatalytic removal of ciprofloxacin and mechanism insight. J. Mater. Chem. A 2015, 3, 15108–15118. [Google Scholar] [CrossRef]
  17. 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]
  18. Shao, L.; Liu, Y.; Wang, L.; Xia, X.; Shen, X. Electronic structure tailoring of BiOBr (0 1 0) nanosheets by cobalt doping for enhanced visible-light photocatalytic activity. Appl. Surf. Sci. 2020, 502, 143895. [Google Scholar] [CrossRef]
  19. Li, J.; Cai, L.; Shang, J.; Yu, Y.; Zhang, L. Giant Enhancement of Internal Electric Field Boosting Bulk Charge Separation for Photocatalysis. Adv. Mater. 2016, 28, 4059–4064. [Google Scholar] [CrossRef]
  20. Xiao, X.; Lu, M.; Nan, J.; Zuo, X.; Zhang, W.; Liu, S.; Wang, S. Rapid microwave synthesis of I-doped Bi4O5Br2 with significantly enhanced visible-light photocatalysis for degradation of multiple parabens. Appl. Catal. B Environ. 2017, 218, 398–408. [Google Scholar] [CrossRef]
  21. Zhang, W.-B.; Xiao, X.; Wu, Q.-F.; Fan, Q.; Chen, S.; Yang, W.-X.; Zhang, F.-C. Facile synthesis of novel Mn-doped Bi4O5Br2 for enhanced photocatalytic NO removal activity. J. Alloy Compd. 2020, 826, 154204. [Google Scholar] [CrossRef]
  22. Noorjahan, M.; Durga Kumari, V.; Subrahmanyam, M.; Panda, L. Immobilized Fe(III)−HY: An efficient and stable photo-Fenton catalyst. Appl. Catal. B Environ. 2005, 57, 291–298. [Google Scholar] [CrossRef]
  23. Zhao, Y.; Hu, J.; Jin, W. Transformation of Oxidation Products and Reduction of Estrogenic Activity of 17β-Estradiol by a Heterogeneous Photo−Fenton Reaction. Environ. Sci. Technol. 2008, 42, 5277–5284. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, H.-L.; Liang, W.-Z.; Zhang, Q.; Jiang, W.-F. Solar−light−assisted Fenton oxidation of 2,4-dinitrophenol (DNP) using Al2O3−supported Fe(III)-5-sulfosalicylic acid (ssal) complex as catalyst. Chem. Eng. J. 2010, 164, 115–120. [Google Scholar] [CrossRef]
  25. Zhang, X.; Ren, B.; Li, X.; Liu, B.; Wang, S.; Yu, P.; Xu, Y.; Jiang, G. High-efficiency removal of tetracycline by carbon-bridge-doped g-C3N4/Fe3O4 magnetic heterogeneous catalyst through photo-Fenton process. J. Hazard. Mater. 2021, 418, 126333. [Google Scholar] [CrossRef] [PubMed]
  26. Qin, L.; Wang, Z.; Fu, Y.; Lai, C.; Liu, X.; Li, B.; Liu, S.; Yi, H.; Li, L.; Zhang, M.; et al. Gold nanoparticles−modified MnFe2O4 with synergistic catalysis for photo−Fenton degradation of tetracycline under neutral pH. J. Hazard. Mater. 2021, 414, 125448. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, T.; You, H.; Chen, Q. Heterogeneous photo-Fenton degradation of polyacrylamide in aqueous solution over Fe(III)-SiO2 catalyst. J. Hazard. Mater. 2009, 162, 860–865. [Google Scholar] [CrossRef]
  28. Chen, Q.; Wu, P.; Li, Y.; Zhu, N.; Dang, Z. Heterogeneous photo−Fenton photodegradation of reactive brilliant orange X-GN over iron-pillared montmorillonite under visible irradiation. J. Hazard. Mater. 2009, 168, 901–908. [Google Scholar] [CrossRef]
  29. Chen, Q.; Ji, F.; Liu, T.; Yan, P.; Guan, W.; Xu, X. Synergistic effect of bifunctional Co-TiO2 catalyst on degradation of Rhodamine B: Fenton-photo hybrid process. Chem. Eng. J. 2013, 229, 57–65. [Google Scholar] [CrossRef]
  30. Guo, S.; Zhang, G.; Yu, J.C. Enhanced photo−Fenton degradation of rhodamine B using graphene oxide-amorphous FePO4 as effective and stable heterogeneous catalyst. J. Colloid Interface Sci. 2015, 448, 460–466. [Google Scholar] [CrossRef]
  31. Liu, C.; Dai, H.; Tan, C.; Pan, Q.; Hu, F.; Peng, X. Photo−Fenton degradation of tetracycline over Z-scheme Fe-g-C3N4/Bi2WO6 heterojunctions: Mechanism insight, degradation pathways and DFT calculation. Appl. Catal. B Environ. 2022, 310, 121326. [Google Scholar] [CrossRef]
  32. Guo, S.; Yang, W.; You, L.; Li, J.; Chen, J.; Zhou, K. Simultaneous reduction of Cr(VI) and degradation of tetracycline hydrochloride by a novel iron-modified rectorite composite through heterogeneous photo−Fenton processes. Chem. Eng. J. 2020, 393, 124758. [Google Scholar] [CrossRef]
  33. Cao, Z.; Zhao, Y.; Zhou, Z.; Wang, Q.; Mei, Q.; Cheng, H. Efficiency LaFeO3 and BiOI heterojunction for the enhanced photo−Fenton degradation of tetracycline hydrochloride. Appl. Surf. Sci. 2022, 590, 153081. [Google Scholar] [CrossRef]
  34. Cao, Z.; Jia, Y.; Wang, Q.; Cheng, H. High-efficiency photo−Fenton Fe/g-C3N4/kaolinite catalyst for tetracycline hydrochloride degradation. Appl. Clay Sci. 2021, 212, 106213. [Google Scholar] [CrossRef]
  35. Ma, J.; Wang, K.; Wang, C.; Chen, X.; Zhu, W.; Zhu, G.; Yao, W.; Zhu, Y. Photocatalysis-self-Fenton system with high-fluent degradation and high mineralization ability. Appl. Catal. B Environ. 2020, 276, 119150. [Google Scholar] [CrossRef]
  36. Dong, Y.; Han, Z.; Liu, C.; Du, F. Preparation and photocatalytic performance of Fe (III)-amidoximated PAN fiber complex for oxidative degradation of azo dye under visible light irradiation. Sci. Total Environ. 2010, 408, 2245–2253. [Google Scholar] [CrossRef]
  37. Liu, Y.; Zhu, G.; Gao, J.; Hojamberdiev, M.; Zhu, R.; Wei, X.; Guo, Q.; Liu, P. Enhanced photocatalytic activity of Bi4Ti3O12 nanosheets by Fe3+-doping and the addition of Au nanoparticles: Photodegradation of Phenol and bisphenol A. Appl. Catal. B Environ. 2017, 200, 72–82. [Google Scholar] [CrossRef]
  38. Cai, L.; Zhang, G.; Zhang, Y.; Wei, Y. Mediation of band structure for BiOBrxI1−x hierarchical microspheres of multiple defects with enhanced visible-light photocatalytic activity. CrystEngComm 2018, 20, 3647–3656. [Google Scholar] [CrossRef]
  39. Zhang, D.; Li, J.; Wang, Q.; Wu, Q. High {001} facets dominated BiOBr lamellas: Facile hydrolysis preparation and selective visible-light photocatalytic activity. J. Mater. Chem. A 2013, 1, 8622–8629. [Google Scholar] [CrossRef]
  40. Zhao, Y.; Zhao, Y.; Shi, R.; Wang, B.; Waterhouse, G.I.N.; Wu, L.-Z.; Tung, C.-H.; Zhang, T. Tuning Oxygen Vacancies in Ultrathin TiO2 Nanosheets to Boost Photocatalytic Nitrogen Fixation up to 700 nm. Adv. Mater. 2019, 31, 1806482. [Google Scholar] [CrossRef]
  41. Li, J.-F.; Zhong, C.-Y.; Huang, J.-R.; Chen, Y.; Wang, Z.; Liu, Z.-Q. Carbon dots decorated three−dimensionally ordered macroporous bismuth−doped titanium dioxide with efficient charge separation for high performance photocatalysis. J. Colloid Interface Sci. 2019, 553, 758–767. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, H.; Yang, M.; Liu, J.; Lu, G.; Feng, X. Insight into the effects of electronegativity on the H2 catalytic activation for CO2 hydrogenation: Four transition metal cases from a DFT study. Catal. Sci. Technol. 2020, 10, 5641–5647. [Google Scholar] [CrossRef]
  43. Zhang, X.; Li, L.; Wu, Z.; Zhu, H.; Xie, Y.; Schaefer, H.F. Heteroatom (N, P, As, Sb, Bi) Effects on the Resonance-Stabilized 2-, 3-, and 4-Picolyl Radicals. Inorg. Chem. 2021, 60, 5860–5867. [Google Scholar] [CrossRef] [PubMed]
  44. Li, X.; Pi, Y.; Wu, L.; Xia, Q.; Wu, J.; Li, Z.; Xiao, J. Facilitation of the visible light−induced Fenton-like excitation of H2O2 via heterojunction of g-C3N4/NH2-Iron terephthalate metal−organic framework for MB degradation. Appl. Catal. B Environ. 2017, 202, 653–663. [Google Scholar] [CrossRef]
  45. Li, X.; Tang, Y.; Jiang, Y.; Mu, M.; Yin, X. Fe-MIL tuned and bound with Bi4O5Br2 for boosting photocatalytic reduction of CO2 to CH4 under simulated sunlight. Catal. Sci. Technol. 2021, 11, 2864–2872. [Google Scholar] [CrossRef]
  46. Zhu, G.; Hojamberdiev, M.; Zhang, W.; Taj Ud Din, S.; Joong Kim, Y.; Lee, J.; Yang, W. Enhanced photocatalytic activity of Fe-doped Bi4O5Br2 nanosheets decorated with Au nanoparticles for pollutants removal. Appl. Surf. Sci. 2020, 526, 146760. [Google Scholar] [CrossRef]
  47. Wu, Y.; Ji, H.; Liu, Q.; Sun, Z.; Li, P.; Ding, P.; Guo, M.; Yi, X.; Xu, W.; Wang, C.-C.; et al. Visible light photocatalytic degradation of sulfanilamide enhanced by Mo doping of BiOBr nanoflowers. J. Hazard. Mater. 2022, 424, 127563. [Google Scholar] [CrossRef]
  48. Geng, Q.; Xie, H.; Cui, W.; Sheng, J.; Tong, X.; Sun, Y.; Li, J.; Wang, Z.; Dong, F. Optimizing the Electronic Structure of BiOBr Nanosheets via Combined Ba Doping and Oxygen Vacancies for Promoted Photocatalysis. J. Phys. Chem. C 2021, 125, 8597–8605. [Google Scholar] [CrossRef]
  49. Idris, A.M.; Liu, T.; Hussain Shah, J.; Shahid Malik, A.; Zhao, D.; Han, H.; Li, C. Correction to “Sr2NiWO6 Double Perovskite Oxide as a Novel Visible-light −Responsive Water Oxidation Photocatalyst”. ACS Appl. Mater. Interfaces 2020, 12, 56659–56660. [Google Scholar] [CrossRef]
  50. Yu, Z.; Yang, K.; Yu, C.; Lu, K.; Huang, W.; Xu, L.; Zou, L.; Wang, S.; Chen, Z.; Hu, J.; et al. Steering Unit Cell Dipole and Internal Electric Field by Highly Dispersed Er atoms Embedded into NiO for Efficient CO2 Photoreduction. Adv. Funct. Mater. 2022, 32, 2111999. [Google Scholar] [CrossRef]
  51. Yang, X.; Yang, X.; Peng, Y.; Li, Z.; Yu, J.; Zhang, Y. Regulating the Built−In Electric Field of BiOBr by a Piezoelectric Mineral Tourmaline and the Enhanced Photocatalytic Property. Ind. Eng. Chem. Res. 2022, 61, 1704–1714. [Google Scholar] [CrossRef]
  52. Huang, X.; Xiao, J.; Yi, Q.; Li, D.; Liu, C.; Liu, Y. Construction of core-shell Fe3O4@GO-CoPc photo-Fenton catalyst for superior removal of tetracycline: The role of GO in promotion of H2O2 to •OH conversion. J. Environ. Manag. 2022, 308, 114613. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Schematic illustration of the Fe-Bi4O5Br2 preparation processes.
Scheme 1. Schematic illustration of the Fe-Bi4O5Br2 preparation processes.
Nanomaterials 13 00188 sch001
Figure 1. (a) XRD patterns, (b) Raman spectra, (c) N2—adsorption desorption isotherm and (d) pore size distribution of samples.
Figure 1. (a) XRD patterns, (b) Raman spectra, (c) N2—adsorption desorption isotherm and (d) pore size distribution of samples.
Nanomaterials 13 00188 g001
Figure 2. FESEM (a,b) of Bi4O5Br2 and 5%Fe-Bi4O5Br2. (c) Element mapping image of 5%Fe-Bi4O5Br2.
Figure 2. FESEM (a,b) of Bi4O5Br2 and 5%Fe-Bi4O5Br2. (c) Element mapping image of 5%Fe-Bi4O5Br2.
Nanomaterials 13 00188 g002
Figure 3. TEM (a,b) and HRTEM (c,d) of Bi4O5Br2 and 5%Fe-Bi4O5Br2.
Figure 3. TEM (a,b) and HRTEM (c,d) of Bi4O5Br2 and 5%Fe-Bi4O5Br2.
Nanomaterials 13 00188 g003
Figure 4. Binding energy of 5%Fe-Bi4O5Br2 and Bi4O5Br2. (a) High-resolution XPS, (b) Bi 4f, (c) Br 3d, (d) O 1s spectrum of 5%Fe-Bi4O5Br2 and Bi4O5Br2 and (e) Fe2p spectrum of 5%Fe-Bi4O5Br2.
Figure 4. Binding energy of 5%Fe-Bi4O5Br2 and Bi4O5Br2. (a) High-resolution XPS, (b) Bi 4f, (c) Br 3d, (d) O 1s spectrum of 5%Fe-Bi4O5Br2 and Bi4O5Br2 and (e) Fe2p spectrum of 5%Fe-Bi4O5Br2.
Nanomaterials 13 00188 g004
Figure 5. (a) UV–vis–NIR DRS spectra, (b) Tauc plots, (c) Mott-Schottky plots of Bi4O5Br2 and 5%Fe-Bi4O5Br2 and (d) Schematic illustration of band structure over Bi4O5Br2 and 5%Fe-Bi4O5Br2.
Figure 5. (a) UV–vis–NIR DRS spectra, (b) Tauc plots, (c) Mott-Schottky plots of Bi4O5Br2 and 5%Fe-Bi4O5Br2 and (d) Schematic illustration of band structure over Bi4O5Br2 and 5%Fe-Bi4O5Br2.
Nanomaterials 13 00188 g005
Figure 6. (a) Transient photocurrent response of samples, (b) EIS Nyquist plots, (c) PL spectra of Bi4O5Br2 and 5%Fe-Bi4O5Br2 and (d) TRPL spectra of Bi4O5Br2 and 5%Fe-Bi4O5Br2.
Figure 6. (a) Transient photocurrent response of samples, (b) EIS Nyquist plots, (c) PL spectra of Bi4O5Br2 and 5%Fe-Bi4O5Br2 and (d) TRPL spectra of Bi4O5Br2 and 5%Fe-Bi4O5Br2.
Nanomaterials 13 00188 g006
Figure 7. (a) Photocatalytic activity for the degradation of TC for samples, (b) The reaction kinetics of samples for the degradation of TC, (c) Photocatalytic activity for the degradation of TC for samples with H2O2, (d) The reaction kinetics of samples for the degradation of TC with H2O2, (e) The degradation of TC over samples under darkness, (f) Cyclic experiments of 5%Fe-Bi4O5Br2 for the degradation of TC, (g) XRD patterns of 5%Fe-Bi4O5Br2 before and after five cycles of reaction and (h) The Fe content of the solution before and after the photo−Fenton reaction by ICP-MS.
Figure 7. (a) Photocatalytic activity for the degradation of TC for samples, (b) The reaction kinetics of samples for the degradation of TC, (c) Photocatalytic activity for the degradation of TC for samples with H2O2, (d) The reaction kinetics of samples for the degradation of TC with H2O2, (e) The degradation of TC over samples under darkness, (f) Cyclic experiments of 5%Fe-Bi4O5Br2 for the degradation of TC, (g) XRD patterns of 5%Fe-Bi4O5Br2 before and after five cycles of reaction and (h) The Fe content of the solution before and after the photo−Fenton reaction by ICP-MS.
Nanomaterials 13 00188 g007
Figure 8. The in-situ ESR spectra of Bi4O5Br2 and 5%Fe-Bi4O5Br2. (a,b) h+ spectra of Bi4O5Br2 and 5%Fe-Bi4O5Br2 using TEMPO in water, (c,d) OH spectra of Bi4O5Br2 and 5%Fe-Bi4O5Br2 by adding DMPO and H2O2 in water and (e,f) O2 spectra of 5%Fe-Bi4O5Br2 by adding DMPO and H2O2 in methanol and without.
Figure 8. The in-situ ESR spectra of Bi4O5Br2 and 5%Fe-Bi4O5Br2. (a,b) h+ spectra of Bi4O5Br2 and 5%Fe-Bi4O5Br2 using TEMPO in water, (c,d) OH spectra of Bi4O5Br2 and 5%Fe-Bi4O5Br2 by adding DMPO and H2O2 in water and (e,f) O2 spectra of 5%Fe-Bi4O5Br2 by adding DMPO and H2O2 in methanol and without.
Nanomaterials 13 00188 g008
Figure 9. Mechanism of photo-Fenton degradation of TC over the 5%Fe-Bi4O5Br2 photocatalyst.
Figure 9. Mechanism of photo-Fenton degradation of TC over the 5%Fe-Bi4O5Br2 photocatalyst.
Nanomaterials 13 00188 g009
Table 1. Pore volume and mean pore size of samples.
Table 1. Pore volume and mean pore size of samples.
SamplePore Volume/cm3 g−1Mean Pore Size/nm
Bi4O5Br20.32837.7
1%Fe-Bi4O5Br20.36727.8
3%Fe-Bi4O5Br20.48127.8
5%Fe-Bi4O5Br20.43123.5
7%Fe-Bi4O5Br20.47323.5
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, F.; Peng, Y.; Yang, X.; Li, Z.; Zhang, Y. Enhanced Photo-Assisted Fenton Degradation of Antibiotics over Iron-Doped Bi-Rich Bismuth Oxybromide Photocatalyst. Nanomaterials 2023, 13, 188. https://doi.org/10.3390/nano13010188

AMA Style

Zhang F, Peng Y, Yang X, Li Z, Zhang Y. Enhanced Photo-Assisted Fenton Degradation of Antibiotics over Iron-Doped Bi-Rich Bismuth Oxybromide Photocatalyst. Nanomaterials. 2023; 13(1):188. https://doi.org/10.3390/nano13010188

Chicago/Turabian Style

Zhang, Fengjiao, Yanhua Peng, Xiaolong Yang, Zhuo Li, and Yan Zhang. 2023. "Enhanced Photo-Assisted Fenton Degradation of Antibiotics over Iron-Doped Bi-Rich Bismuth Oxybromide Photocatalyst" Nanomaterials 13, no. 1: 188. https://doi.org/10.3390/nano13010188

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