Metal–organic frameworks (MOFs) have garnered significant interest in the field of photocatalysis. In this study, Z-scheme heterojunction BM-x composites consisting of bismuth bromide oxide (BiOBr) and iron-based metal–organic backbone (MIL-100(Fe)) were successfully synthesized using ethylene glycol as a solvent. The composites were characterized using various techniques. BM-x exhibit abundant functional groups, large specific surface areas, and narrow band gap energy, thus provide numerous active sites for catalytic reactions and respond well to visible light. Notably, BM-7 displays remarkable catalytic activity in a visible light-activated permonosulfate (PMS) system and achieves a degradation rate of 99.02% over 100 mg/L gold orange II (AO7) within 60 min. The effects of BM-7 and PMS addition, initial AO7 concentration, initial pH, inorganic anions, and humic acid on the degradation system were investigated. The proposed mechanism of the Z-scheme heterojunction in the BM-7 photocatalyst demonstrates effective photoelectron transfer from the BiOBr conduction band to the MIL-100(Fe) valence band, resulting in excellent catalytic activity. Radical burst experiments identified 1O2, h+, and ·O2 as the main active substances. BM-7 has high stability and reusability, with a degradation rate reduction of only 14.48% after three recycles. These findings provide valuable insights into using persulfate combined with visible light.

  • A Z-scheme photocatalyst BiOBr/MIL-100(Fe) was prepared using a solvothermal method.

  • BiOBr/MIL-100(Fe) enhanced the photocatalytic-persulfate activity toward AO7.

  • The effects of water environmental factors on the degradation efficiency were investigated.

  • 1O2, h+, and ·O2 radicals were the vital radicals of the photocatalytic-persulfate system.

Water pollution has become a significant global environmental issue in recent years, leading to a scarcity of usable water resources due to water quality deterioration and water resource mismanagement (Qutub et al. 2022). The discharge of various organic pollutants, including synthetic dyes, pesticides, and antibiotics, into aquatic environments on a daily basis has resulted in continuous damage to ecosystems (Gholami et al. 2023; Tiwari et al. 2023; Xu et al. 2023). To address this problem, sulfate radical-based advanced oxidation processes (SR-AOPs) have emerged as a rapidly developing solution. Compared with the more oxidized hydroxyl radical (OH·), has the following advantages: (1) (standard redox potential (2.5 ∼ 3.1 V) is higher than OH (1.8–2.7 V)); (2) (longer half-life (30–40 μs), while OH· (half-life less than 1 μs); and (3) (adapted to the wider pH range (2–9)), while OH· (the optimal pH range is 2–4) (Gao et al. 2021; Qiu et al. 2022; Xu et al. 2022). SR-AOPs can effectively break down organic pollutants into less toxic substances (e.g., CO2 and H2O) or smaller molecular structures, and thus are effective treatment approaches for organic pollutants.

Metal–organic frameworks (MOFs) are structured materials composed of organic ligands and inorganic metal ions or clusters. These components self-assemble through coordination bonds to form ordered porous mesh structures (Wan et al. 2020; Su et al. 2023). MOFs exhibit several desirable characteristics for the application of SR-AOPs (Wang et al. 2023a, b). They possess tunable porosity, open topology, large specific surface areas, and abundant active sites. MOFs also have semiconductor-like photoinduced properties, as the internal ligand-metal charge transfer showcases high charge separation capability. MIL-100(Fe), a typical MOF, can activate permonosulfate and degrade pollutants under visible light irradiation, and thus has garnered significant interest in the field of photocatalysis. Li et al. (2022a, b) prepared MIL-100(Fe), and the photoelectrons generated by MIL-100(Fe) under visible light irradiation will drive Fe(III) on the surface to Fe(II) in situ, which effectively activated permonosulfate to degrade the pollutant. The iron-based metal–organic backbone (MIL-100(Fe)/PDI) heterojunction constructed by You et al. (2023) effectively improved the separation efficiency of photogenerated electron–hole pairs under visible light and persulfate conditions, resulting in almost 100% degradation of bisphenol A (BPA) (Lv et al. 2020). This enhanced separation led to nearly complete pollutant degradation, as the heterojunction facilitated efficient redox reactions on the photocatalyst surface and thus accelerated the production of active substances for pollutant decomposition.

Bismuth bromide oxide (BiOBr), a semiconductor material with an indirect band gap, is composed of (Bi2O2)2+ and Br layers in a distinct layered structure. This material has been extensively studied owing to its favorable properties, such as chemical stability, non-toxicity, corrosion resistance, and excellent photocatalytic capability (Zhao et al. 2022; Ighnih et al. 2023). Despite these advantages, the practical application of BiOBr is hindered by its relatively large band gap (∼2.7 eV) and the rapid recombination of electron–hole pairs generated upon light absorption. To overcome these limitations, researchers have proposed the development of heterojunctions within BiOBr. The type II BiOBr/MoS2 heterojunctions were successfully synthesized by Zhang et al. (2021) and they greatly enhanced the transfer of charge carriers and effectively activated PMS molecules (Xu et al. 2017). Additionally, these heterojunctions improved the oxidation ability of the generated holes and enabled the efficient degradation of various organic dyes and the reduction of Cr(VI).

Inspired by the previous studies, we focus on preparing heterojunction composite photocatalysts using the solvothermal method. Specifically, MIL-100(Fe) and BiOBr were combined to form the composite photocatalysts. The main objective was to examine the degradation ability of the BiOBr/MIL-100(Fe) + PMS + Vis system toward the azodye dye gold orange II (AO7). Various factors such as photocatalyst dosage, initial AO7 concentration, permonosulfate (PMS) dosage, initial pH, common inorganic anions, and HA were systematically investigated to understand their impacts on AO7 degradation. Furthermore, an AO7 degradation mechanism by the BiOBr/MIL-100(Fe) + PMS + Vis system was proposed.

Chemicals

Ferric nitrate hydrate (Fe(NO3)3·9H2O, ≥98.5%), sodium chloride (NaCl, ≥99.5%), potassium iodide (KI, ≥99.5%), and potassium bromide (KBr, ≥ 99.5%) were obtained from Sinopharm Chemical Reagents (Shanghai, China). 1,3,5-Benzenetricarboxylic acid (C6H3(CO2H)3, ≥98.0%), potassium persulfate (KHSO5, ≥99.5%), L-histidine (C6H9N3O2, ≥99.5%), and p-benzoquinone (C6H4O2, ≥99.0%) were purchased from Shanghai Maclin company. Orange II (C16H11N2NaO4S, AR), sodium nitrate (NaNO3, ≥99.5%), anhydrous ethanol (C2H6O, ≥99.7%), ethylene glycol (C2H6O2, ≥99.5%), methanol (CH3OH, ≥99.9%), sodium bicarbonate (NaHCO3, ≥99.5%), sodium carbonate (Na2CO3, ≥99.5%), tert-butanol (C4H10O, ≥99.5%), and humic acid (HA) (C9H9NO6, ≥90.0%) were obtained from Xilong Scientific Co. All the above chemicals were an analytical reagent unless otherwise specified and used without further purification.

Preparation of catalytic materials

Preparation of MIL-100(Fe)

MIL-100(Fe) is prepared by a simple one-step hydrothermal method. First, 2.020 g of Fe(NO3)3·9H2O (5.0 mmol) and 1.050 g of 1,3,5-benzenetricarboxylic acid (H3BTC) (5.0 mmol) are dispersed in a beaker containing 50 mL of deionized water. Then, the mixed solution is stirred for 60 min at room temperature to ensure uniform mixing. Subsequently, the mixed solution is transferred to a 100 mL high-pressure reaction vessel lined with polytetrafluoroethylene and subjected to a hydrothermal reaction at 160 °C in a muffle furnace for 12 h. After cooling to room temperature, it is washed three times with deionized water and anhydrous ethanol. Finally, it is dried at 60 °C for 12 h to obtain a pale yellow powder, which is MIL-100(Fe).

Preparation of BiOBr/MIL-100(Fe) composites

BM-x composite materials are prepared using a solvothermal method. First, 0.485 g of Bi(NO3)3·5H2O (1.0 mmol) is dissolved in 20 mL of EG solution, referred to as solution A. Then, 0.166 g of KBr (1.0 mmol) and a certain amount of MIL-100(Fe) are dispersed in 20 mL of EG solution, referred to as solution B. The two solutions are stirred evenly. Subsequently, solution A is slowly added dropwise into solution B to form a mixed solution. The mixed solution is transferred to a 100 mL high-pressure reaction vessel lined with polytetrafluoroethylene and reacted at 160 °C for 12 h. After naturally cooling to room temperature, it is washed three times with deionized water and anhydrous ethanol to remove residual chemicals. Finally, the composite material is dried at 60 °C for 12 h to obtain a BiOBr/MIL-100(Fe) composite photocatalytic material. Pure BiOBr is prepared without adding MIL-100(Fe). According to different mass ratios of MIL-100(Fe) to BiOBr [m(MIL-100(Fe)):m(BiOBr) = 0.3, 0.5, 0.7, 0.9], they are named as BM-3, BM-5, BM-7, and BM-9.

Catalyst characterization

The surface morphology of the photocatalysts was analyzed using a scanning electron microscope (SEM, Model X130W/TMP, The Netherlands); the surface functional groups were identified by Fourier transform infrared spectroscopy (FTIR, Thermo Scientific Nicolet iS20, USA); the surface area of the photocatalysts was determined using a fully automated specific surface and porosity analyzer; BET (Micromeritics ASAP 2460, USA) was used to determine the specific surface area, the total pore volume, and the average pore diameter of the photocatalysts; the crystal structure of the photocatalysts was analyzed by an X-ray polycrystalline diffraction analyzer (XRD, SMART APEX II, Germany); X-ray photoelectron spectroscopy (XPS, Scientific K-Alpha, USA) was used to analyze the elemental composition and valence states; and a UV–Vis spectrometer (UV–Vis DRS, UV-3600 model, Japan) was used to study the optical properties.

Photocatalytic degradation experiment

The photocatalytic degradation performance of the catalyst was evaluated by degrading the AO7 solution using a 350 W xenon lamp (λ ≥ 420 nm) as the simulated light source. First, a certain mass of photocatalyst was added to 100 mL of 100 mg/L AO7 solution and stirred for 30 min under dark reaction conditions to achieve the adsorption–desorption equilibrium. Subsequently, a certain mass of PMS was added at the same time under light conditions, 2 mL was modified to a dye wastewater sample that was added at regular intervals, and finally, the absorbance was measured by a UV–Vis spectrophotometer.

Microstructure characterization

X-ray diffraction

The crystal structures and crystallinity of the catalysts were analyzed using XRD (Figure 1(a)). The water-synthesized MIL-100(Fe) exhibits consistent results with previous reports. The diffraction peaks at 2θ = 6.3, 10.2, 11.0, and 20.0° correspond to the (333), (660), (428), and (4814) crystal planes, respectively (Jiang et al. 2021). Other peaks at 10.9, 21.9, 25.2, 31.8, 32.3, 39.4, 46.3, 50.8, 53.5, 57.3, 67.6, 71.2, and 76.7° stand for the (001), (002), (011), (012), (110), (112), (020), (014), (211), (212), (220), (124), and (032) crystal planes of the tetragonal BiOBr phase, respectively (JCPDS:73-2061) (Senasu et al. 2021; Yang et al. 2024). The BM-x composites mainly exhibit the characteristic peaks of BiOBr. However, with the increase in MIL-100(Fe) content, the diffraction peaks belonging to BiOBr are gradually weakened, and the intensity of the (012) crystal plane significantly decreases. It is inferred when the two materials are combined, MIL-100(Fe) preferentially couples with the (012) crystal plane of BiOBr.
Figure 1

(a) XRD and (b) FTIR images of the prepared photocatalysts.

Figure 1

(a) XRD and (b) FTIR images of the prepared photocatalysts.

Close modal

Fourier transform infrared spectroscopy

The surface functional groups of the photocatalysts were investigated using FTIR within 400–4,000 cm−1 (Figure 1(b)). The peak at 504 cm−1 is attributed to the stretching vibration of Bi–O in BiOBr (Song et al. 2014). The prominent peak at 711 cm−1 is due to the stretching of the benzene ring (Chen et al. 2021). The characteristic peaks at 1,620, 1,574, 1,446, and 1,376 cm−1 are associated with the asymmetric/symmetric vibrations of the carboxyl group. The broad absorption band at 3,000–3,500 cm−1 is caused by the O–H vibration of the adsorbed H2O molecules or coordinated hydroxyl groups (Ahmad et al. 2020). As reported, the abundant surface functional groups of nanomaterials play a key role in the adsorption and degradation of pollutants (Zhang et al. 2023). In the BM-x composites, the corresponding characteristic peaks are enhanced with the increase of MIL-100(Fe) proportion, and no other impurity peaks are observed. FTIR confirms the formation of the BM-x composites.

Scanning electron microscopy

SEM was used to analyze the morphology of the catalysts (Figure 2). BiOBr exhibits a rod-shaped stacked spherical structure (Figure 2(a)), while MIL-100(Fe) appears as distinct blocks with a smooth surface and particle size ranging from 0.5 to 1 μm (Figure 2(f)). With the increase of m(MIL-100(Fe)):m(BiOBr), the structure of BM-x transits to a spherical shape, and the grain size gradually decreases. BM-3 maintains the basic structure of BiOBr with partially loose rod-like structures (Figure 2(b)). The rod-like structure of BM-5 becomes thinner and gradually smoother (Figure 2(c)). When the mass ratio exceeds 0.7:1, significant morphological changes occur, and nanoparticle-like structures aggregate with each other (Figure 2(d) and 2(e)). The grain size decreases, and the formation of nanoscale structures enlarges the specific surface area, exposing a large number of active sites. The energy dispersive spectrometry (EDS) spectrum of the BM-7 composite shows that elements Br, Bi, and Fe are uniformly distributed within the scanning area (Figure 2(g)), which indicates the successful combination of MIL-100(Fe) and BiOBr.
Figure 2

SEM images of (a) BiOBr, (b) BM-3, (c) BM-5, (d) BM-7, (e) BM-9, (f) MIL-100(Fe), and (g) the EDS elemental mappings of the BM-7 composite.

Figure 2

SEM images of (a) BiOBr, (b) BM-3, (c) BM-5, (d) BM-7, (e) BM-9, (f) MIL-100(Fe), and (g) the EDS elemental mappings of the BM-7 composite.

Close modal

Nitrogen adsorption–desorption isotherms

The specific surface areas, pore volumes, and pore size distributions of BiOBr, MIL-100(Fe), and BM-7 composite were analyzed using low-temperature N2 adsorption–desorption curves (Figure 3). MIL-100(Fe) exhibits the distribution characteristics of Type I isotherm, indicating mesoporous adsorption (Lu et al. 2023), and BiOBr displays the distribution characteristics of Type IV isotherm. In comparison, BM-7 shows a comprehensive isotherm between Type I and Type IV, with an H3 hysteresis loop, indicating the presence of both micropores and mesopores. Table 1 shows the specific surface areas, pore volumes, and pore sizes of BiOBr, MIL-100(Fe), and BM-7. Compared with MIL-100(Fe), the specific surface area of BM-7 decreased to 301.56 m2 g−1, which was because nanospheres were constituted after the rod-like stacked spherical structure of BiOBr combined with block-like MIL-100(Fe). Moreover, the pore volume and pore size increased to a certain extent, providing more active sites and improving its photocatalytic performance (Wang et al. 2023a, b).
Table 1

N2 adsorption–desorption isotherms test results of different catalysts

CatalyzerSurface area (m2g−1)Pore volume (cm3g−1)Pore size (nm)
BiOBr 6.13 0.033571 21.92 
MIL-100(Fe) 501.46 0.323693 2.58 
BM-7 301.56 0.346934 4.60 
CatalyzerSurface area (m2g−1)Pore volume (cm3g−1)Pore size (nm)
BiOBr 6.13 0.033571 21.92 
MIL-100(Fe) 501.46 0.323693 2.58 
BM-7 301.56 0.346934 4.60 
Figure 3

(a, b) N2 adsorption–desorption isotherms of BiOBr, MIL-100(Fe), and BM-7.

Figure 3

(a, b) N2 adsorption–desorption isotherms of BiOBr, MIL-100(Fe), and BM-7.

Close modal

Ultraviolet–visible diffuse reflectance spectroscopy

The optical properties of different catalysts were studied by UV–Vis DRS. The pure MIL-100(Fe), BiOBr, and BM-7 have absorption edges at 560, 440, and 620 nm, respectively (Figure 3(a)). The visible light absorption edge of BM-7 significantly red shifts, indicating an enhanced range of visible light absorption. The band gap width (Eg) of the photocatalytic materials was calculated using the Kubelka–Munk formula (Equation (1)):
(1)
where α, , and A represent the absorption coefficient, photon energy, and absorbance, respectively. The value of n depends on the transition mode of the semiconductor. MIL-100(Fe) is a direct transition semiconductor with n = 1 (Wang et al. 2022), while BiOBr is an indirect transition semiconductor with n = 4 (Ning et al. 2022). The calculated results are shown in Figure 4(b). The band gap widths of MIL-100(Fe) and BiOBr are 2.88 and 2.66 eV, respectively, which are similar to other reported results. However, the band gap width of BM-7 narrows to 2.60 eV, which indicates a more excellent visible light response and an enhanced light absorption capability, thereby accelerating the generation rate of photogenerated electron–hole pairs.
Figure 4

UV–Vis diffuse reflectance spectrum (a) and energy gap (b).

Figure 4

UV–Vis diffuse reflectance spectrum (a) and energy gap (b).

Close modal

X-ray photoelectron spectroscopy

The surface elemental composition and chemical valence states of the photocatalysts MIL-100(Fe), BiOBr, and BM-7 were studied using XPS (Figure 5). The spectral measurements in Figure 5(a) reveal the presence of C, O, Fe, Bi, and Br in BM-7, which indicate the successful composite formation of MIL-100(Fe) and BiOBr. The O 1s spectrum of MIL-100(Fe) exhibits peaks at 531.86 and 533.63 eV (Figure 5(b)), which are related to Fe–O clusters and oxygen elements in H3BTC (Liu et al. 2021). The O 1s peaks of BiOBr at 529.68 and 531.08 eV correspond to Bi–O bonds and surface hydroxyl groups (Shi et al. 2022). The O 1s peaks of BiOBr and MIL-100(Fe) shift to higher and lower binding energy, respectively, which collectively form the BM-7 composite. The binding energies at 159.08 and 164.38 eV in Figure 5(c) correspond to Bi 4f7/2 and Bi 4f5/2, respectively (Liu et al. 2024). Compared with BiOBr, the binding energy of Bi 4f peak in BM-7 increases. The Fe 2p3/2 and Fe 2p1/2 peaks of MIL-100(Fe) at 711.61, 714.26, 724.91, 727.02, 717.63, and 730.85 eV (You et al. 2023) shift to 710.86, 713.48, 724.18, and 726.90 eV in BM-7 (Figure 5(d)). Additionally, the Fe 2p peak in BM-7 slightly shifts toward lower binding energy. The high-resolution Br 3d spectra of BiOBr and BM-7 in Figure 5(e) show two distinct peaks at 68.08 and 69.09 eV, which are attributed to Br 3d5/2 and Br 3d3/2, respectively (Yan et al. 2023). A shift in the center and a change in the intensity of the Br 3d peak in BM-7 compared with BiOBr indicate possible interactions between MIL-100(Fe) and BiOBr during the synthesis process, altering the chemical environment of Br. These results demonstrate that the peak shifts and intensity changes observed in the materials are evidence of electron density variation due to the formation of heterojunctions, which can enhance the photocatalytic activity of BM-7 (Huang et al. 2023a, b).
Figure 5

XPS spectra of BM-7: full spectrum (a), O 1s (b), Bi 4f (c), Fe 2p (d), and Br 3d (e) energy regions.

Figure 5

XPS spectra of BM-7: full spectrum (a), O 1s (b), Bi 4f (c), Fe 2p (d), and Br 3d (e) energy regions.

Close modal

AO7 degradation performance of catalysts

AO7 was used as the target dye to study the visible light-activated degradation performance of MIL-100(Fe), BiOBr, and their composite catalysts at different ratios (BM-3, BM-5, BM-7, and BM-9) for PMS (Figure 6). With an initial concentration of AO7 at 100 mg/L, PMS concentration at 0.5 mmol/L, catalyst dosage at 0.2 g/L, light intensity at 350 W, and pH unchanged (pH 6.47), the adsorption–desorption equilibrium dark reaction of the catalysts was studied first to explore the effect on adsorption. The first 30 min were dedicated to adsorption experiments, which showed the adsorption performance of BiOBr was limited and MIL-100(Fe) reached an AO7 adsorption capacity of 25.43%. In the composite materials, the adsorption capacity gradually increased from 16.15 to 25.15% with the rise in the MIL-100(Fe) ratio. The reason for this increase is that the large specific surface area of MIL-100(Fe) effectively enhances its adsorption capacity. At the same time, the AO7 degradation effect of BiOBr alone under photocatalytic activation of PMS was unsatisfactory. Under the same conditions, MIL-100(Fe) alone had a better AO7 degradation effect and thus possessed higher photocatalytic activity. This is because the photogenerated electrons on the surface of MIL-100(Fe) can reduce Fe(III) to Fe(II) in situ, and the Fe(III)/Fe(II) redox cycle can effectively activate persulfate. The composite materials all exhibited high degradation ability, and as the composite ratio of MIL-100(Fe) continued to increase in the BiOBr/MIL-100(Fe) composite, the degradation rates of BH-3, BH-5, BH-7, and BH-9 at 60 min were 93.68, 98.61, 99.02, and 99.01%, respectively. This result is related to the red shift in visible light absorption and the formation of heterojunctions in the composite materials. Among them, BH-7 showed the best degradation efficiency and reached a degradation rate of 96.34% at 40 min. To further analyze the degradation reaction kinetics of the photocatalytic materials on AO7, a pseudo-first-order kinetic model was used to simulate AO7 degradation (Figure 6(b)). The apparent rate constants (kobs) of BiOBr, MIL-100(Fe), BM-3, BM-5, BM-7, and BM-9 were calculated to be 0.00196, 0.0472, 0.0385, 0.06616, 0.08664, and 0.07536 min−1, respectively. With the increase in the composite ratio, the apparent rate constants increased and then decreased. Among them, BM-7 had the fastest reaction rate. It is speculated that the number of heterojunctions formed in combination also increases with a rise in the mass ratio of MIL-100(Fe) to BiOBr. However, excessive MIL-100(Fe) will encapsulate the formed heterojunctions and thereby reduce the number of exposed active sites, lowering the photocatalytic activity of the material. Therefore, the optimal composite ratio for the materials is BM-7. Table 2 presents a comparison of degradation performance.
Table 2

Comparison of degradation performance

CatalystCatalyst amount (g/L)PollutantOxidizerTemperature (°C)Illumination time (min)Efficiency (%)Ref.
MIL-53(Fe) 0.6 0.05 mmol/L AO7 2 mmol/L PMS 25 90 82 Gao et al. (2017)  
BiOBr/PBCD-B-D 0.2 mmol/L AO7 1 mmol/L PMS 30 60 98.7 Du et al. (2021)  
Co-MIL-101(Fe) 0.3 0.1 mmol/L AO7 8 mmol/L PDS 25 150 98  
Cu-MIL-101(Fe)      92  
Fe3O4@MIL-101 1.0 25 mg/L AO7 25 mmol/L PDS 25 60 100  
AgBr/LaFeO3 0.6 30 mg/L AO7 0.5 mmol/L PMS 30 60 99.41  
BiOBr/MIL-100(Fe) 0.2 100 mg/L AO7 0.5 mmol/L PMS 25 60 99.02 This text 
CatalystCatalyst amount (g/L)PollutantOxidizerTemperature (°C)Illumination time (min)Efficiency (%)Ref.
MIL-53(Fe) 0.6 0.05 mmol/L AO7 2 mmol/L PMS 25 90 82 Gao et al. (2017)  
BiOBr/PBCD-B-D 0.2 mmol/L AO7 1 mmol/L PMS 30 60 98.7 Du et al. (2021)  
Co-MIL-101(Fe) 0.3 0.1 mmol/L AO7 8 mmol/L PDS 25 150 98  
Cu-MIL-101(Fe)      92  
Fe3O4@MIL-101 1.0 25 mg/L AO7 25 mmol/L PDS 25 60 100  
AgBr/LaFeO3 0.6 30 mg/L AO7 0.5 mmol/L PMS 30 60 99.41  
BiOBr/MIL-100(Fe) 0.2 100 mg/L AO7 0.5 mmol/L PMS 25 60 99.02 This text 
Figure 6

Degradation of AO7 by different photocatalysts under visible light irradiation (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Figure 6

Degradation of AO7 by different photocatalysts under visible light irradiation (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Close modal

Effect of different systems on AO7 removal

To study the synergistic effect of the BM-7 photocatalyst and visible light co-catalyzed persulfate on AO7 degradation, a control experiment was designed to compare the degradation effects of seven systems: BM-7 + PMS + Vis, BM-7 + PMS, BM-7 + Vis, PMS + Vis, PMS, BM-7, and Vis. The results are shown in Figure 7. AO7 was highly stable and minimally degraded under visible light conditions. The degradation abilities of PMS alone and the PMS + Vis system were both very limited. Similarly, in the BM-7 system and the BM-7 + Vis system, the AO7 degradation was primarily due to the adsorption capacity of BM-7. However, when BM-7 activates PMS (BM-7 + PMS), it generates AO7 degrading radicals, with a degradation rate of only 44.94%. Introducing visible light (BM-7 + PMS + Vis) significantly increases the AO7 degradation rate to 99.02% within 60 min. These results indicate that BM-7, PMS, and Vis have a synergistic effect that significantly improves degradation efficiency.
Figure 7

Effectiveness of different systems in removing AO7 (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Figure 7

Effectiveness of different systems in removing AO7 (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Close modal

Effect of reaction conditions on AO7 degradation

Effect of catalyst dosage

The impact of BM-7 dosage on the catalytic degradation system was examined, as shown in Figure 8. Results revealed that the efficiency of photocatalytic AO7 degradation was improved as the BM-7 dosage gradually increased. Specifically, the AO7 degradation rate was exceptional when the dosage was raised from 0.1 to 0.4 g/L. This increment in dosage led to a rise in kobs from 0.03961 to 0.14516 min−1 and an increase in adsorption amount. The reason for these results is the amount of catalyst is increased, the number of adsorption sites and active sites also increases, and the system is able to capture more visible light photons and generate more photogenerated electrons and holes. Consequently, the activation of PMS is facilitated to generate more active substances for AO7 degradation. Hence, the AO7 degradation rate continues to accelerate due to the combined effects of adsorption and degradation. However, when the dosage increases to 0.5 g/L, the kobs only reach 0.14623 min−1, indicating the enhancement is not significant. This result can be attributed to the limited PMS concentration, which restricts the production of additional free radicals.
Figure 8

Effect of BM-7 dosage on AO7 degradation rate (a) and corresponding kobs. (b) Experimental conditions: PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Figure 8

Effect of BM-7 dosage on AO7 degradation rate (a) and corresponding kobs. (b) Experimental conditions: PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Close modal

Effect of PMS concentration

Figure 9 demonstrates the influence of the initial PMS concentration on the catalytic system. Evidently, the degradation rate only reached 96.80% after 60 min at a PMS concentration of 0.25 mmol/L. As the PMS concentration rose to 0.75 mmol/L, the AO7 degradation rate reached 97.96% within 40 min, and the apparent rate constant kobs increased from 0.0499 to 0.0948 min−1. This increase can be attributed to the higher concentration of active substances in the reaction system. However, when the PMS concentration further rose to 1.0 mmol/L, the kobs decreased to 0.0664 min−1. Notably, increasing the PMS concentration alone cannot sustainably accelerate AO7 degradation. Previous research suggests that excessive PMS dose can trigger a self-burst reaction of generated (Equations (2) and (3)), which can interfere with AO7 degradation (Xu et al. 2013).
(2)
(3)
Figure 9

Effect of PMS dosage on the degradation rate of AO7 (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, AO7 = 100 mg/L, and T = 25 °C.

Figure 9

Effect of PMS dosage on the degradation rate of AO7 (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, AO7 = 100 mg/L, and T = 25 °C.

Close modal

Effect of initial pollutant concentration

The impact of initial AO7 concentration (50, 75, 100, 125, and 150 mg/L) on the efficiency of the degradation system was investigated. As shown in Figure 10, the kobs at the five initial AO7 concentrations are 0.16992, 0.09933, 0.8664, 0.07653, and 0.03935 min−1, respectively. The degradation rates gradually decreased as the pollutant concentration increased. Nevertheless, more than 90% degradation was achieved within 60 min at all initial mass concentrations, which demonstrates an excellent degradation effect. These findings indicate the BM-7 + PMS + Vis system can maintain a favorable degradation rate at a specific initial mass concentration of AO7.
Figure 10

Effect of different AO7 concentrations on the catalytic system (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, and T = 25 °C.

Figure 10

Effect of different AO7 concentrations on the catalytic system (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, and T = 25 °C.

Close modal

Effect of pH

The protonation or deprotonation of pollutants or catalysts, which in turn affects the adsorption and performance of a catalyst, is influenced by the concentrations of hydrogen and hydroxide ions in the solution. The initial pH of the solution was adjusted to 3.0, 5.0, 7.0, 9.0, and 11.0 (the AO7 solution at initial pH 6.47) using 0.1 mol/L sodium hydroxide (NaOH) and 0.1 mol/L concentrated sulfuric acid (H2SO4). The AO7 degradation efficacy of the BM-7 + PMS + Vis system was then investigated at these different initial pHs (Figure 11). Results revealed that the AO7 degradation rate reached about 99% within pH 3–9. As the pH increased, the degradation rate decreased gradually. Specifically, the highest degradation rate was observed at pH 3, with a kobs of 0.10342 min−1. Under this condition, BM-7 exhibited stronger adsorption performance possibly due to the acidic environment, which led to increased adsorption of H+ on the surface of BM-7 and thus resulted in a positively charged catalyst surface and enhanced adsorption of AO7. Additionally, the positive charge on the catalyst surface facilitated the movement of photoelectrons (e) toward the Fe3+ center, thereby inhibiting the formation of photoinducted electron–hole pairs. Moreover, oxygen molecules adsorbed to the material surface reacted with the e moving toward the catalyst surface to generate superoxide radicals O2 (Xue et al. 2018; Pi et al. 2019), which accelerated AO7 degradation. However, at pH 11, the degradation rate and kobs decreased significantly to only 26.01% and 0.0374 min−1, respectively. These results were attributed that some of the iron ions precipitated by the catalyst formed a precipitate, which affected the mesoporous structure of the catalyst surface and reduced its adsorption capacity (Guan et al. 2011). Furthermore, under highly alkaline conditions, self-burst reactions of and ·OH (Equations (4) and (5)) further inhibited AO7 degradation (Wang et al. 2021).
(4)
(5)
Figure 11

Effect of different initial pH on the degradation rate of AO7 (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Figure 11

Effect of different initial pH on the degradation rate of AO7 (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Close modal

Effects of different inorganic anions

Inorganic anions are widely distributed in natural waters and can interact with free radicals produced in the degradation system through electron transfer. This interaction can impact the efficiency of pollutant degradation (Quan et al. 2023). In the present study, the degradation system was supplemented with 10 mmol/L Cl, , , , or to investigate their effects on AO7 degradation. The results in Figure 12 show the addition of Cl, , or minimally impacted AO7 degradation compared with the blank experiment. However, the degradation was significantly inhibited after the addition of or , as the degradation rates were only 45.52 and 18.96%, respectively. The reason for this inhibition is that captures and in the system, generating less oxidizable HCO3· and (Equations (6) and (7)) (Jia et al. 2022; Chen et al. 2023). These less reactive species consume the free radicals in the system and decelerate the reaction rate. Similarly, the addition of to the degradation system led to the hydrolysis of , which affected the degradation efficiency. Additionally, the alkaline conditions caused by and the effects of pH discussed above further inhibited AO7 degradation. The combined effect of these factors further inhibited AO7 degradation by compared with
(6)
(7)
Figure 12

Effect of inorganic anions on the degradation rate of AO7 (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Figure 12

Effect of inorganic anions on the degradation rate of AO7 (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Close modal

Effect of humic acid

HA is a significant constituent of natural organic matter and is commonly found in surface water. Its complex molecular structure makes it have a complex impact on the oxidation process. Experiments were conducted to understand the effects of different mass concentrations of HA (5, 10, 20, and 50 mg/L) on AO7 degradation. Results showed that AO7 degradation was slightly accelerated when low concentration HA was added (Figure 13). Reportedly, HA as a photosensitizer becomes excited under visible light and transfers electrons to the conduction band, thus initiating a series of electron transfer processes to degrade organic pollutants (Fang et al. 2013). However, as the HA concentration increased to 50 mg/L, the AO7 degradation was somewhat inhibited, resulting in a final degradation rate of 91.24%. The reason for this inhibition is that the phenol, hydroxyl, and carboxyl functional groups on HA deactivate the active sites on the catalyst after adsorption (Zhao et al. 2021; Li et al. 2022a, b). Additionally, HA can capture and ·OH, further inhibiting AO7 degradation. However, this inhibition is only minor, indicating that non-radical substances play a crucial role in the degradation process.
Figure 13

Effect of HA on the degradation rate of AO7 (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Figure 13

Effect of HA on the degradation rate of AO7 (a) and corresponding kobs. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Close modal

Possible mechanisms

Radical burst experiment chemistry

Free radical burst experiments were conducted to investigate the active substances involved in the AO7 degradation by the BM-7 + PMS + Vis system. The results are presented in Figure 14. Previous studies show that methanol (MeOH) captures and ·OH at similar reaction rates, falling within 1.2–2.8 × 109 and 1.6–7.7 × 108M−1S−1, respectively. Moreover, tertiary butyl alcohol captures ·OH at a faster rate (3.8–7.6 × 108M−1S−1) compared with (4.0–9.1 × 108M−1S−1) (Song et al. 2017; Krishnan et al. 2022). Based on this information, we used methanol to simultaneously capture and ·OH, and employed tert-butanol to capture ·OH. The addition of methanol and tert-butanol reduced the AO7 degradation rate from 99.02 to 69.13% and to 94.77%, respectively. These findings indicate that both and ·OH are involved in the degradation process, with having a stronger effect. Additionally, potassium iodide and p-benzoquinone are commonly used to trap holes (h+) and superoxide radicals () to assess their contributions in the degradation system. The inclusion of potassium iodide and p-benzoquinone decreased the AO7 degradation rate by 52.47 and 41.01%, respectively, suggesting that h+ and play significant roles in the degradation process. Finally, L-histidine was introduced to capture 1O2 and evaluate the role of non-radicals in the degradation system. The addition of L-histidine significantly inhibited the degradation, which indicates the production and major role of 1O2 in the degradation process.
Figure 14

Effect of different bursting agents on the catalytic system. Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Figure 14

Effect of different bursting agents on the catalytic system. Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Close modal

Mechanisms for degradation

To confirm the type of heterojunction formed by the composite of BiOBr and MIL-100(Fe), the values of the conduction and valence bands of BiOBr and MIL-100(Fe) were calculated using Mulliken's theoretical formulas for electronegativity (Equations (8) and (9)):
(8)
(9)

The geometric average electronegativity of the constituent atoms in the semiconductor, denoted as χ, plays a key role in the energy calculations of the system. χ is 5.10 eV for MIL-100(Fe) (Cui et al. 2024) and 6.45 eV for BiOBr (He et al. 2021). Additionally, the energy of free electrons (Ee) is around 4.5 eV vs. NHE. For MIL-100(Fe), ECB is −0.84 eV and EVB is 2.04 eV. For BiOBr, ECB is 0.62 eV and EVB is 3.28 eV. Based on experimental results and energy band calculations, it can be concluded that the composite formation of a Z-scheme heterojunction between BiOBr and MIL-100(Fe) not only effectively separates the photogenerated electron–hole pairs, but also exhibits a stronger redox capacity (Hu et al. 2023). The standard electrode potential is −0.33 eV vs. NHE for and E(·OH/OH) is 2.40 eV vs. NHE for ·OH. The photogenerated electrons on the valence band of BiOBr cannot combine with O2 to form , and the holes on the conduction band of MIL-100(Fe) face difficulty in oxidizing H2O to ·OH. However, and ·OH are generated in the composite BM-7 system, indicating that the photogenerated electrons and holes are concentrated in the conduction band of MIL-100(Fe) and the valence band of BiOBr, respectively. These species further react with O2 and H2O to produce and ·OH, which are highly effective in degrading pollutants.

In light of the above experiments, we propose a possible photocatalytic activation mechanism for the AO7 degradation by BM-7 (Figure 15). Under visible light conditions, both BiOBr and MIL-100(Fe) are excited to form photogenerated electrons and holes. The photoelectrons in the conduction band of BiOBr can transfer to the valence band of MIL-100(Fe) and combine with the holes in the valence band through Z-scheme heterojunction migration. This migration is facilitated by the staggered positions in the energy band structures of the two photocatalytic materials. The transfer of photoelectrons from the conduction band of BiOBr to the valence band of MIL-100(Fe), along with their combination with holes in the valence band, effectively enhances the separation efficiency of photogenerated electrons and holes through Z-scheme heterojunction migration. The photoelectrons remaining in the conduction band of MIL-100(Fe) play a crucial role in three aspects.
  • (1) The photoelectrons reduce the dissolved oxygen in water to , which further reacts with H+ to produce 1O2 (Equations (10) and (11)) (Silva et al. 2023).

  • (2) The photoelectrons transfer to BM-7, reducing Fe(III) to Fe(II). As an active center, Fe(II) effectively activates PMS to produce and ·OH (Equations (12) and (13)), establishing the Fe(III)/Fe(II) cycle and accelerating the generation of free radicals (Chen et al. 2024).

  • (3) The photoelectrons are directly captured by PMS to produce (Equations (14) and (15)) (Chen et al. 2024).

Figure 15

Photocatalytic activation mechanism for the AO7 degradation by BM-7.

Figure 15

Photocatalytic activation mechanism for the AO7 degradation by BM-7.

Close modal
Furthermore, the accumulative holes in the valence band of BiOBr can directly participate in AO7 degradation or react with water to generate ·OH for AO7 degradation (Equation (16)). In conclusion, the combined effect of BM-7, PMS, and visible light promotes the generation of radicals and non-radicals in the system, which contribute to the efficient degradation of pollutants.
(10)
(11)
(12)
(13)
(14)
(15)
(16)

Reusable performance and stability

Reusability and stability are crucial properties to consider in the practical applications of photocatalysts. In this study, the BM-7 + PMS + Vis system was developed to repeatedly degrade AO7 (Figure 16(a)). The degradation curves of different cycles showed that the system consistently achieved a degradation rate above 85.82% after three cycles, which indicates its good reusable performance. XRD conducted on BM-7 after recycling (Figure 16(b)) revealed no significant changes in the characteristic diffraction peaks of the catalyst, indicating its structural stability and a good crystal structure. These findings confirm the excellent reusability and structural stability of the BM-7 composite, which makes it a promising photocatalyst for pollutant degradation.
Figure 16

Cyclic test for the degradation rate of AO7 by the BM-7 + PMS + Vis system (a) and XRD plots before and after reaction. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Figure 16

Cyclic test for the degradation rate of AO7 by the BM-7 + PMS + Vis system (a) and XRD plots before and after reaction. (b) Experimental conditions: catalysts dosage = 0.20 g/L, PMS = 0.5 mmol/L, AO7 = 100 mg/L, and T = 25 °C.

Close modal

A visible-light-responsive photocatalyst BiOBr/MIL-100(Fe) was successfully prepared using a simple solvothermal method. The crystal structure, morphology, and photoelectrochemical properties of the samples were characterized using various physical and chemical characterization methods, which confirmed the successful synthesis of the nanocomposite photocatalysts. Under the optimal composite ratio (BM-7), the activated persulfate system demonstrated superior degradation performance under visible light irradiation compared with BiOBr and MIL-100(Fe). This achievement can be attributed to the construction of a Z-scheme heterojunction in BM-7 that enabled it to exhibit stronger visible light response and electron transfer ability. The BM-7 + PMS + Vis system achieved a remarkable degradation efficiency of 99.02% over AO7 (100 mg/L) within 60 min under optimal conditions. However, the degradation efficiency of the system decreased with increasing pH. The presence of and significantly inhibited the AO7 degradation, but the system exhibited strong adaptability to HA. Free radical burst experiments identified 1O2, h+, and as the main active substances in the BM-7 + PMS + Vis system, and the proposed mechanism for AO7 degradation involved the Z-scheme heterojunction constructed by BM-7. Furthermore, the system can equally and efficiently degrade other organic dyes. After three cycle tests, the AO7 degradation rate remained at 85.82%, indicating its potential for reuse. Therefore, the BM-7 + PMS + Vis system holds promise for advanced oxidation applications in the degradation of organic dyes.

This study was supported by the National Natural Science Foundation of China (No. 51808268).

X.L. conceptualized the study, wrote, reviewed, and edited the article. X.C. arranged the resources, wrote the review, and edited the article. J.L. investigated the work and rendered support in data curation. J.T. investigated the work and rendered support in data curation. R.W. investigated the work and edited the article.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Ahmad
S. Z. N.
,
Salleh
W. N. W.
,
Ismail
A. F.
,
Yusof
N.
,
Yusop
M. Z. M.
&
Aziz
F.
(
2020
)
Adsorptive removal of heavy metal ions using graphene-based nanomaterials: toxicity, roles of functional groups and mechanisms
,
Chemosphere
,
248
,
126008
.
https://doi.org/10.1016/j.chemosphere.2020.126008
.
Chen
L.
,
Wang
X.
,
Rao
Z. P.
,
Tang
Z. X.
,
Wang
Y.
,
Shi
G. S.
,
Lu
G. H.
,
Xie
X. F.
,
Chen
D. L.
&
Sun
J.
(
2021
)
In-situ synthesis of Z-scheme MIL-100(Fe)/α-Fe2O3 heterojunction for enhanced adsorption and visible-light photocatalytic oxidation of O-xylene
,
Chemical Engineering Journal
,
416
,
129112
.
https://doi.org/10.1016/j.cej.2021.129112
.
Chen
L.
,
Zhang
J. K.
,
Li
Q. X.
,
Zhang
Y. Y.
,
Huang
J. J.
,
Shen
J. Q.
,
Gu
M. Y.
&
Yang
J. X.
(
2023
)
Removal of p-nitrophenols by BC@nZVI activated persulfate: a study of key factors and mechanisms
,
Journal of Environmental Chemical Engineering
,
11
,
111483
.
http://dx.doi.org/10.1016/j.jece.2023.111483
.
Cui
S. C.
,
Cong
Y.
,
Zhao
W. S.
,
Guo
R.
,
Wang
X. H.
,
Lv
B. H.
,
Liu
H. B.
,
Liu
Y.
&
Zhang
Q.
(
2024
)
A novel multifunctional magnetically recyclable BiOBr/ZnFe2O4-GO S-scheme ternary heterojunction: photothermal synergistic catalysis under Vis/NIR light and NIR-driven photothermal detection of tetracycline
,
Journal of Colloid and Interface Science
,
654
,
356
370
.
https://doi.org/10.1016/j.jcis.2023.10.051
.
Du
H.
,
Wang
X. Y.
,
Zhang
Y. L.
,
Xu
G. Z.
,
Tu
Y. Z.
,
Jia
X.
,
Wu
D. S.
,
Xie
X.
&
Liu
Y.
(
2021
)
Insights into the photocatalytic activation of peroxymonosulfate by visible light over BiOBr-cyclodextrin polymer complexes for efficient degradation of dye pollutants in water
,
Environmental Research
,
207
,
112160
.
https://doi.org/10.1016/j.envres.2021.112160
.
Fang
G. D.
,
Gao
J.
&
Dionysiou
D. D.
(
2013
)
Activation of persulfate by quinones: free radical reactions and implication for the degradation of PCBs
,
Environmental Science & Technology
,
47
(
9
),
4605
4611
.
https://doi.org/10.1021/es400262n
.
Gao
Y.
,
Li
S.
,
Li
Y.
,
Yao
L.
&
Zhaung
H.
(
2017
)
Accelerated photocatalytic degradation of organic pollutant over metal-organic framework MIL-53(Fe) under visible LED light mediated by persulfate
,
Applied Catalysis B: Environmental
,
174
,
202165
.
https://doi.org/10.1016/j.apcatb.2016.09.005
.
Gao
Y.
,
Zhou
Y.
,
Pang
S. Y.
,
Jiang
J.
,
Shen
Y. M.
,
Song
Y.
,
Duan
J. B.
&
Guo
Q.
(
2021
)
Enhanced peroxymonosulfate activation via complexed Mn(II): a novel non-radical oxidation mechanism involving manganese intermediates
,
Water Research
,
193
,
116856
.
https://doi.org/10.1016/j.watres.2021.116856
.
Gholami
P.
,
Heidari
A.
,
Khataee
A.
&
Ritala
M.
(
2023
)
Oxygen and nitrogen plasma modifications of ZnCuCo LDH-graphene nanocomposites for photocatalytic hydrogen production and antibiotic degradation
,
Separation and Purification Technology
,
325
,
124706
.
https://doi.org/10.1016/j.seppur.2023.124706
.
Guan
Y. H.
,
Ma
J.
,
Li
X. C.
,
Fang
J. Y.
&
Chen
L. W.
(
2011
)
Influence of pH on the formation of sulfate and hydroxyl radicals in the UV/peroxymonosulfate system
,
Environmental Science & Technology
,
45
(
21
),
9308
9314
.
https://doi.org/10.1021/es2017363
.
He
Y. Y.
,
Wang
D. B.
,
Li
X. P.
,
Fu
Q. Z.
,
Yin
L. M.
,
Yang
Q.
&
Chen
H.
(
2021
)
Photocatalytic degradation of tetracycline by metal-organic frameworks modified with Bi2WO6 nanosheet under direct sunlight
,
Chemosphere
,
284
,
131386
.
https://doi.org/10.1016/j.chemosphere.2021.131386
.
Hu
H. L.
,
Lin
H.
,
Chen
X. C.
,
Pan
Y. P.
,
Li
X. W.
,
Zhuang
Z. L.
,
Chen
H. D.
,
Wang
X.
,
Luo
M.
,
Zheng
K. Z.
,
Zhang
L. G.
&
Chen
F. M.
(
2023
)
Mn-Prussian blue analogue@MXene composites for efficient photocatalytic peroxydisulfate activated degradation of tetracycline hydrochloride and photoelectrochemical desalination
,
Chemical Engineering Journal
,
476
,
146682
.
https://doi.org/10.1016/j.cej.2023.146682
.
Huang
X. D.
,
Liu
C. H.
,
Zhang
Z. Y.
,
Vasanthakumar
V.
,
Ai
H. Y.
,
Xu
L.
,
Fu
M. L.
&
Yuan
B. L.
(
2023a
)
Stable Cu-Co/C carbon-based composites for efficiency catalytic degradation of orange II by PMS: effect factors, application potential analysis, and mechanism
,
Journal of Industrial and Engineering Chemistry
,
126
,
307
316
.
http://dx.doi.org/10.1016/j.jiec.2023.06.021
.
Huang
S. Y.
,
Yuan
H. Q.
,
Dong
W.
,
Wur
X.
&
Wang
G. H.
(
2023b
)
Visible-light-driven removal of AO7 by peroxymonosulfate-assisted Z-scheme AgBr/LaFeO3 heterojunction vis accelerated electronic transfer
,
Materials Science in Semiconductor Processing
,
160
,
107414
.
https://doi.org/10.1016/j.mssp.2023.107414
.
Ighnih
H.
,
Haounati
R.
,
Malekshah
R. E.
,
Ouachtak
H.
,
Toubi
Y.
,
Alakhras
F.
,
Jada
A.
&
Addi
A. A.
(
2023
)
Sunlight driven photocatalytic degradation of RhB dye using composite of bismuth oxy-bromide kaolinite BiOBr@Kaol: experimental and molecular dynamic simulation studies
,
Journal of Photochemistry and Photobiology A: Chemistry
,
445
,
115071
.
https://doi.org/10.1016/j.jphotochem.2023.115071
.
Jia
J. L.
,
Liu
D. M.
,
Wang
Q.
,
Li
H. R.
,
Ni
J. X.
,
Cui
F. Y.
&
Tian
J. Y.
(
2022
)
Comparative study on bisphenols oxidation via TiO2 photocatalytic activation of peroxymonosulfate: effectiveness, mechanism and pathways
,
Journal of Hazardous Materials
,
424
,
127434
.
http://dx.doi.org/10.1016/j.jhazmat.2021.127434
.
Jiang
L. S.
,
Xie
Y.
,
He
F.
,
Ling
Y.
,
Zhao
J. S.
,
Ye
H.
,
Li
S. Q.
,
Wang
J. L.
&
Hou
Y.
(
2021
)
Facile synthesis of GO as middle carrier modified flower-like BiOBr and C3N4 nanosheets for simultaneous treatment of chromium(VI) and tetracycline
,
Chinese Chemical Letters
,
32
(
07
),
2187
2191
.
https://doi.org/10.1016/j.cclet.2020.12.010
.
Krishnan
S. A. G.
,
Sasikumar
B.
,
Arthanareeswaran
G.
,
László
Z.
,
Santos
E. N.
,
Veréb
G.
&
Kertész
S.
(
2022
)
Surface-initiated polymerization of PVDF membrane using amine and bismuth tungstate (BWO) modified MIL-100(Fe) nanofillers for pesticide photodegradation
,
Chemosphere
,
304
,
135286
.
https://doi.org/10.1016/j.chemosphere.2022.135286
.
Li
D.
,
Ali
J.
,
Shahzad
A.
,
Gendy
E. A.
,
Nie
H.
,
Jiang
W.
,
Xiao
H. L.
,
Chen
Z. Q.
&
Wang
S. L.
(
2022a
)
Persulfate coupled with Cu2+/LDH-MoS4: a novel process for the efficient atrazine abatement, mechanism and degradation pathway
,
Chemical Engineering Journal
,
436
,
134933
.
https://doi.org/10.1016/j.cej.2022.134933
.
Li
X.
,
Chen
X. G.
,
Wang
B.
&
Yu
G.
(
2022b
)
Boosting Fe(II) generation in MOFs under visible-light irradiation for accumulated micropollutants decomposition
,
Journal of Environmental Chemical Engineering
,
10
,
108833
.
https://doi.org/10.1016/j.jece.2022.108833
.
Liu
Y.
,
Hu
Z. F.
&
Yu
J. C.
(
2021
)
Photocatalytic degradation of ibuprofen on S-doped BiOBr
,
Chemosphere
,
278
,
130376
.
https://doi.org/10.1016/j.chemosphere.2021.130376
.
Liu
Y. K.
,
Zhang
X. X.
,
Li
X.
&
Zhou
Z. W.
(
2024
)
Visible light driven S-scheme GQDs/BiOBr heterojunction with enhanced photocatalytic degradation of rhodamine B and mechanism insight
,
Journal of Water Process Engineering
,
57
,
104588
.
https://doi.org/10.1016/j.jwpe.2023.104588
.
Lu
M. L.
,
Xiao
X. Y.
,
Xiao
Y.
,
Li
J. J.
&
Wang
F.
(
2023
)
In situ fabrication of type-II BiOIO3/Bi-MOF heterostructures with enhanced photocatalytic activity
,
Journal of Photochemistry and Photobiology A: Chemistry
,
444
,
114957
.
https://doi.org/10.1016/j.jphotochem.2023.114957
.
Lv
X. C.
,
Yan
D. Y.
,
Lam
F. L.
,
Ng
Y. H.
,
Yin
S. M.
&
An
A. K.
(
2020
)
Solvothermal synthesis of copper-doped BiOBr microflowers with enhanced adsorption and visible-light driven photocatalytic degradation of norfloxacin
,
Chemical Engineering Journal
,
401
,
126012
.
https://doi.org/10.1016/j.cej.2020.126012
.
Ning
R.
,
Pang
H.
,
Yan
Z.
,
Lu
Z.
,
Wang
Q.
,
Wu
Z.
,
Dai
W.
,
Liu
L.
,
Li
Z.
,
Fan
G.
&
Fu
X.
(
2022
)
An innovative S-scheme AgCl/MIL-100(Fe) heterojunction for visible-light-driven degradation of sulfamethazine and mechanism insight
,
Journal of Hazardous Materials
,
435
,
129061
.
https://doi.org/10.1016/j.jhazmat.2022.129061
.
Pi
Z. J.
,
Li
X. M.
,
Wang
D. B.
,
Xu
Q. X.
,
Tao
Z. T.
,
Huang
X. D.
,
Yao
F. B.
,
Wu
Y.
,
He
L.
&
Yang
Q.
(
2019
)
Persulfate activation by oxidation biochar supported magnetite particles for tetracycline removal: performance and degradation pathway
,
Journal of Cleaner Production
,
235
,
1103
1115
.
https://doi.org/10.1016/j.jclepro.2019.07.037
.
Qiu
Z.
,
Xiao
X.
,
Yu
W. T.
,
Zhu
X. Y.
,
Chu
C. H.
&
Chen
B. L.
(
2022
)
Selective separation catalysis membrane for highly efficient water and soil decontamination via a persulfate-based advanced oxidation process
,
Environmental Science & Technology
,
56
,
3234
3244
.
https://doi.org/10.1021/acs.est.1c06721
.
Quan
X.
,
Zhang
J.
,
Yin
L. L.
,
Zuo
W.
&
Tian
Y.
(
2023
)
Fe3O4 decorated with α-cyclodextrin for enhanced peroxymonosulfate (PMS)-activated degradation of PPCPs
,
Separation and Purification Technology
,
317
,
123904
.
http://dx.doi.org/10.1016/j.seppur.2023.123904
.
Qutub
N.
,
Singh
P. i.
,
Sabir
S.
,
Sagadevan
S.
&
Oh
W.
(
2022
)
Enhanced photocatalytic degradation of acid blue dye using CdS/TiO2 nanocomposite
,
Scientific Reports
,
12
,
5759
.
https://doi.org/10.1038/s41598-022-09479-0
.
Senasu
T.
,
Nijpanich
S.
,
Juabrum
S.
,
Chanlek
N.
&
Nanan
S.
(
2021
)
Cds/BiOBr heterojunction photocatalyst with high performance for solar-light-driven degradation of ciprofloxacin and norfloxacin antibiotics
,
Applied Surface Science
,
567
,
150850
.
https://doi.org/10.1016/j.apsusc.2021.150850
.
Shi
Y. X.
,
Wang
L. Y.
,
Dong
S. H.
,
Miao
X.
,
Zhang
M. X.
,
Sun
K.
,
Zhang
Y. F.
,
Cao
Z. Z.
&
Sun
J. M.
(
2022
)
Wool-ball-like BiOBr@ZnFe-MOF composites for degradation organic pollutant under visible-light: synthesis, performance, characterization and mechanism
,
Optical Materials
,
131
,
112580
.
https://doi.org/10.1016/j.optmat.2022.112580
.
Silva
R. R. M.
,
Ruotolo
L. A. M.
&
Nogueira
F. G. E.
(
2023
)
Peroxymonosulfate activation by magnetic NiFe2O4/g-C3N4 for tetracycline hydrochloride degradation under visible light
,
Chemical Engineering Journal
,
476
,
146621
.
http://dx.doi.org/10.1016/j.cej.2023.146621
.
Song
G. Q.
,
Wang
Z. Q.
,
Wang
L.
,
Huang
M. J.
&
Yan
F. X.
(
2014
)
Preparation of MOF(Fe) and its catalytic activity for oxygen reduction reaction in an alkaline electrolyte
,
Chinese Journal of Catalysis
,
35
(
02
),
185
195
.
https://doi.org/10.1016/S1872-2067(12)60729-3
.
Song
H. R.
,
Yan
L. X.
,
Ma
J.
,
Jiang
J.
,
Cai
G. Q.
,
Zhang
W. J.
,
Zhang
Z. X.
,
Zhang
J. M.
&
Yang
T.
(
2017
)
Nonradical oxidation from electrochemical activation of peroxydisulfate at Ti/Pt anode: efficiency, mechanism and influencing factors
,
Water Research
,
116
,
182
193
.
http://dx.doi.org/10.1016/j.watres.2017.03.035
.
Tiwari
S.
,
Ryu
W. H.
,
Kim
K. J.
&
Park
E. S.
(
2023
)
Development of (Fe-Co-Ni)-Si-B metallic glass catalyst for promoting degradation of acid orange II azo-dye
,
Journal of Alloys and Compounds
,
961
,
171027
.
https://doi.org/10.1016/j.jallcom.2023.171027
.
Wan
Y. J.
,
Wan
J. Q.
,
Ma
Y. W.
,
Wang
Y.
&
Luo
T.
(
2020
)
Sustainable synthesis of modulated Fe-MOFs with enhanced catalyst performance for persulfate to degrade organic pollutants
,
Science of the Total Environment
,
701
,
134806
.
https://doi.org/10.1016/j.scitotenv.2019.134806
.
Wang
Y. P.
,
Yan
P. W.
,
Dou
X. M.
,
Liu
C.
,
Zhang
Y. T.
,
Song
Z. L.
,
Chen
Z. L.
,
Xu
B. B.
&
Qi
F.
(
2021
)
Degradation of benzophenone-4 by peroxymonosulfate activated with microwave synthesized well-distributed CuBi2O4 microspheres: theoretical calculation of degradation mechanism
,
Applied Catalysis B: Environmental
,
290
,
120048
.
https://doi.org/10.1016/j.apcatb.2021.120048
.
Wang
Y. H.
,
Ding
L. Z.
,
Liu
C.
,
Lu
Y.
,
Wu
Q. Y.
,
Wang
C.
&
Hu
Q.
(
2022
)
0D/2D/2D ZnFe2O4/Bi2O2CO3/BiOBr double Z-scheme heterojunctions for the removal of tetracycline antibiotics by permonosulfate activation: photocatalytic and non-photocatalytic mechanisms, radical and non-radical pathways
,
Separation and Purification Technology
,
283
,
120164
.
https://doi.org/10.1016/j.seppur.2021.120164
.
Wang
D. K.
,
Suo
M. J.
,
Lai
S. Q.
,
Deng
L. Q.
,
Liu
J. Y.
,
Yang
J.
,
Chen
S. Q.
,
Wu
M. F.
&
Zou
J. P.
(
2023a
)
Photoinduced acceleration of Fe3+/Fe2+ cycle in heterogeneous FeNi-MOFs to boost peroxodisulfate activation for organic pollutant degradation
,
Applied Catalysis B: Environmental
,
321
,
122054
.
https://doi.org/10.1016/j.apcatb.2022.122054
.
Wang
Q.
,
Zhang
K. J.
,
Zheng
S. Z.
,
Hu
X.
,
Wang
L. Y.
,
Du
H.
,
Hao
D.
&
Yang
G. X.
(
2023b
)
An innovative AgI/MIL-100(Fe) Z-scheme heterojunction for simultaneously enhanced photoreduction of Cr(VI) and antibacterial activity
,
Applied Surface Science
,
616
,
156528
.
https://doi.org/10.1016/j.apsusc.2023.156528
.
Xu
L.
,
Zang
H. M.
,
Zhang
Q.
,
Chen
Y.
,
Wei
Y. G.
,
Yan
J. H.
&
Zhao
Y. H.
(
2013
)
Photocatalytic degradation of atrazine by H3PW12O40/Ag-TiO2: kinetics, mechanism and degradation pathways
,
Chemical Engineering Journal
,
232
,
174
182
.
http://dx.doi.org/10.1016/j.cej.2013.07.095
.
Xu
B.
,
Chen
Z. M.
,
Han
B.
&
Li
C. C.
(
2017
)
Glycol assisted synthesis of MIL-100(Fe) nanospheres for photocatalytic oxidation of benzene to phenol
,
Catalysis Communications
,
98
,
112
115
.
http://dx.doi.org/10.1016/j.catcom.2017.04.041
.
Xu
L.
,
Cao
S. Y.
,
Bai
X.
,
Jin
X.
,
Shi
X.
,
Han
J.
,
Gao
Y. H.
&
Jin
P. K.
(
2022
)
Efficient Fe(III) reduction and persulfate activation induced by ligand-to-metal charge transfer under visible light enhances degradation of organics
,
Chemical Engineering Journal
,
446
,
137052
.
https://doi.org/10.1016/j.cej.2022.137052
.
Xu
P.
,
Wei
R.
,
Wang
P.
,
Li
X.
,
Yang
C. Y.
,
Shen
T. Y.
,
Zheng
T.
&
Zhang
G. S.
(
2023
)
CuFe2O4/diatomite actuates peroxymonosulfate activation process: mechanism for active species transformation and pesticide degradation
,
Water Research
,
235
,
119843
.
https://doi.org/10.1016/j.watres.2023.119843
.
Yan
X. R.
,
Zhang
T. T.
,
Meng
S. C.
,
Zhou
P. Y.
,
Xu
Y. G.
,
Wang
Y.
&
Xie
M.
(
2023
)
BaFe12O19/BiOBr S-scheme heterojunction magnetic nanosheets for high-efficiency photocatalytic degradation of 2-mercaptobenzothiazole
,
Separation and Purification Technology
,
326
,
124746
.
http://dx.doi.org/10.1016/j.seppur.2023.124746
.
Yang
Z. L.
,
Peng
D. Y.
,
Zeng
H. Y.
,
Xiong
J.
,
Li
S.
,
Ji
W. Y.
&
Wu
B.
(
2024
)
Enhanced photocatalytic performance of heterostructure BiOBr/PPy for Cr(VI) reduction and dye degradation
,
Colloids and Surfaces A: Physicochemical and Engineering Aspects
,
680
,
132647
.
https://doi.org/10.1016/j.colsurfa.2023.132647
.
You
Z. W.
,
Wang
Z.
,
Tian
K. X.
&
Yao
X. Y.
(
2023
)
Activation of persulfate by MIL-100(Fe) coated with perylene diimide heterojunction photocatalyst for efficient removal of bisphenol A in visible light
,
Journal of Environmental Chemical Engineering
,
11
,
110159
.
https://doi.org/10.1016/j.jece.2023.110159
.
Zhang
B. F.
,
Zhang
M. T.
,
Zhang
L.
,
Bingham
P. A.
,
Tanaka
M.
,
Li
W.
&
Kubuki
S.
(
2021
)
BiOBr/MoS2 catalyst as heterogenous peroxymonosulfate activator toward organic pollutant removal: energy band alignment and mechanism insight
,
Journal of Colloid and Interface Science
,
594
,
635
649
.
http://dx.doi.org/10.1016/j.jcis.2021.03.066
.
Zhang
X. J.
,
Kang
X. Q.
,
Mi
J. G.
,
Jin
J. S.
&
Meng
H.
(
2023
)
In-situ confined growth of defective MIL-100(Fe) in macroporous polyacrylate spherical substrate at room temperature for high-efficient toluene removal
,
Separation and Purification Technology
,
324
,
124623
.
https://doi.org/10.1016/j.seppur.2023.124623
.
Zhao
C.
,
Wang
J. S.
,
Chen
X.
,
Wang
Z. H.
,
Ji
H. D.
,
Chen
L.
,
Liu
W.
&
Wang
C. C.
(
2021
)
Bifunctional Bi12O17Cl2/MIL-100(Fe) composites toward photocatalytic Cr(VI) sequestration and activation of persulfate for bisphenol A degradation
,
Science of the Total Environment
,
752
,
141901
.
https://doi.org/10.1016/j.scitotenv.2020.141901
.
Zhao
Y. Y.
,
Li
Z. Y.
,
Wei
J.
,
Li
X. L.
,
Shi
H. X.
,
Cao
B. Y.
&
Fan
J.
(
2022
)
Efficient photodegradation of cefixime catalyzed by a direct Z-scheme CQDs-BiOBr/CN composite: performance, toxicity evaluation and photocatalytic mechanism
,
Chemosphere
,
292
,
133430
.
https://doi.org/10.1016/j.chemosphere.2021.133430
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).