Catalysts | Free Full-Text | Self-Assembly of Porous Hierarchical BiOBr Sub-Microspheres for Efficient Aerobic Photooxidation of Benzyl Alcohol under Simulated Sunlight Irradiation
Next Article in Journal
Effect of Re Addition on the Water–Gas Shift Activity of Ni Catalyst Supported by Mixed Oxide Materials for H2 Production
Next Article in Special Issue
Facile Synthesis of Ni-Doped WO3-x Nanosheets with Enhanced Visible-Light-Responsive Photocatalytic Performance for Lignin Depolymerization into Value-Added Biochemicals
Previous Article in Journal
Catalytic Hydrogenation of Anthracene on Binary (Bimetallic) Composite Catalysts
Previous Article in Special Issue
The Effect of Different g-C3N4 Precursor Nature on Its Structural Control and Photocatalytic Degradation Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Self-Assembly of Porous Hierarchical BiOBr Sub-Microspheres for Efficient Aerobic Photooxidation of Benzyl Alcohol under Simulated Sunlight Irradiation

1
School of Chemistry and Chemical Engineering, Hubei Polytechnic University, Huangshi 435003, China
2
State Key Laboratory of New Textile Materials & Advanced Processing Technologies, Wuhan Textile University, Wuhan 430200, China
3
School of Chemistry and Environmental Engineering, Wuhan Institute of Technology, Wuhan 430205, China
4
Henan Institute of Advanced Technology, Zhengzhou University, Zhengzhou 450002, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(6), 958; https://doi.org/10.3390/catal13060958
Submission received: 7 May 2023 / Revised: 21 May 2023 / Accepted: 30 May 2023 / Published: 1 June 2023
(This article belongs to the Special Issue Nanocatalysts for the Degradation of Refractory Pollutants)

Abstract

:
Semiconductor photocatalytic performances can be modulated through morphology modification. Herein porous hierarchical BiOBr microspheres (BiOBr-MS) of ~3 μm was firstly self-assembled without the assistance of a template via a facile solvothermal synthesis in triethylene glycol (TEG) at 150 °C for 3 h. KBrO3 was exploited as a bromine source, which slowly provided bromide ions upon reduction in TEG and controlled the growth and self-assembly of primary BiOBr nanoplates. The addition of PVP during solvothermal synthesis of BiOBr-MS reduced the particle size by about three-fold to generate BiOBr sub-microspheres (BiOBr-sMS) of <1 μm. BiOBr-sMS exhibited significantly higher photocatalytic activity than BiOBr-MS for aerobic photooxidation of benzyl alcohol (BzOH) to benzaldehyde (BzH) under simulated sunlight irradiation (conversions of BzOH (50 mM) over BiOBr-sMS and BiOBr-MS were, respectively, 51.3% and 29.6% with 100% selectivity to BzH after Xenon illumination for 2 h at 25 °C). The photogenerated holes and ·O2 were found to be main reactive species for the BzOH oxidation over BiOBr spheres by scavenging tests and spin-trapping EPR spectra. The higher photocatalytic activity of BiOBr-sMS was attributed to its more open hierarchical structure, efficient charge separation, more negative conduction-band position and the generation of larger amounts of ·O2.

Graphical Abstract

1. Introduction

Benzaldehyde (BzH) is an important starting material for the production of some chemical intermediates, dyes, flavors and pharmaceuticals (e.g., benzylamine, benzoin, mandelic acid, triphenylmethane dyes, cinnamaldehyde and chloramphenicol) [1]. It is industrially produced by the hydrolysis of benzal chloride or partial oxidation of toluene [1,2]. The former generates large quantities of HCl as a by-product, while the latter produces BzH only as a by-product in the production of less valuable benzoic acid.
Sunlight-driven photocatalytic aerobic oxidation of benzyl alcohol (BzOH) is a potential green route to synthesize BzH because O2 is a green oxidant and natural sunlight is clean, safe, widely available and inexhaustible. This technology relies on the design and synthesis of sunlight-responsive semiconductor photocatalysts with suitable conduction and valance-band positions to selectively oxidize BzOH to BzH [3]. In order to utilize solar energy, it is important for potential photocatalysts to harness visible light since UV light only accounts for less than 5% in the whole solar spectrum [4].
Various strategies, including heterojunction construction, doping, noble metal deposition and photosensitization, have been adopted to develop efficient solar photocatalysts for aerobic photooxidation of BzOH. However, most of these methods required complex architecture engineering with multiple components (e.g., WO2.72/rGO nanocomposites [5], CdS@MoS2 [6], hybrid-SiW11O39 [7], fluorinated mesoporous WO3 [8] and Ni-OTiO2 [9]). Therefore, it would be more attractive if a single semiconductor with a narrow-band gap can effectively catalyze the sunlight-driven aerobic photooxidation of BzOH.
Among many visible light-responsive semiconductors (e.g., Pt-BiFeO3 [10], Bi2WO6/ZnO composite [11]), bismuth oxybromide (BiOBr) has attracted considerable attention owing to its unique layered structure, appropriate-band gap (~2.7 eV) and good chemical stability [12,13]. The layered structure of BiOBr consists of [Bi2O2]2+ layers interleaved with double layers of Br along the c-axis (Scheme 1), and is beneficial for the separation of photogenerated charge carriers owing to the internal electric field normal to each layer. However, the photocatalytic performance of pure BiOBr is still limited due to high recombination rate of photogenerated electron-hole pairs and low quantum efficiency. After visible light (>420 nm) irradiation for 4 h at 25 °C, only 10.9% of BzOH was converted to BzH in the reaction system comprising 20 mg BiOBr and 8.0 mM BzOH in 2.5 mL benzotrifluoride (BTF) saturated with O2 [14]. Even after visible light (>420 nm) irradiation for 8 h at 50 °C, the conversion of BzOH to BzH was still low (21.6%) in the reaction system comprising 50 mg BiOBr and 0.5 mM BzOH in 10 mL CH3CN in the presence of O2 [3].
Many efforts have been made to enhance sunlight-driven photocatalytic efficiency of BiOBr, including heterojunction construction (e.g., BiOBr/Bi2WO6 [14] and CuO-Bi-BiOBr [15]), surface regulation (e.g., BiOBr m [16]), metal loading (e.g., Ag-Fe-Cu/BiOBr [17], Eu-doped BiOBr [18]), defect engineering (e.g., BiOBr-OV [19]) and synergistic effects (e.g., Zn-VBOB [20]). It was also reported that morphology control of BiOBr (e.g., ultrathin nanosheets and three-dimensional (3D) hierarchical structures) could promote its photocatalytic performances [13,21,22,23,24]. Compared with low-dimensional counterparts, three-dimensional hierarchical BiOBr microspheres built from nanosheets or nanoplates not only alleviate nanoparticle aggregation and facilitate the separation and recovery of catalysts, but also exhibit enhanced visible-light photocatalytic performances (e.g., in photocatalytic coupling of amines to imines [22] and degradation of rhodamine B, phenol or tetracycline [21,23,24]) due to synergistic effects of enhanced visible-light utilization via multiple reflections and scattering, efficient separation of photo-generated electrons and holes and large surface area. In previous reports, hierarchical BiOBr microspheres of 1.0–6.0 μm in diameter were prepared in a two-component solvent (ethylene glycol and isopropanol [21] or water and ethanol [22]) with the assistance of a soft template (e.g., cetyltrimethylammonium bromide [21,23], reactive ionic liquids [24] and poly(vinylpyrrolidone) (PVP) [22]) for the formation of self-assembled microspheres.
Herein hierarchical BiOBr microspheres of 2.57 ± 0.49 μm (designated BiOBr-MS) were directly self-assembled without the assistance of a template via a facile solvothermal synthesis in triethylene glycol (TEG) at 150 °C for 3 h. The addition of PVP during solvothermal synthesis of BiOBr-MS significantly reduces the particle size to generate BiOBr sub-microspheres of 0.87 ± 0.25 μm (designated BiOBr-sMS). Both BiOBr-MS and BiOBr-sMS were evaluated as photocatalysts for aerobic photooxidation of BzOH under simulated sunlight irradiation. BiOBr-sMS had significantly improved photocatalytic performances (BzOH conversion and BzH selectivity and yield were 76.4%, 85.4% and 65.2%, respectively, upon Xenon irradiation for 3 h at 25 °C), which was attributed to its efficient charge separation, more negative conduction-band position and generation of larger amounts of ·O2 species. The h+ and ·O2 species were found to be main reactive species for the BzOH oxidation over BiOBr spheres by scavenging tests and spin-trapping EPR spectra. The present work demonstrated that the photocatalytic performance of the BiOBr semiconductor could be effectively enhanced by morphology modulation. It was achieved by using KBrO3 as a bromine source rather than organic templates, which slowly provides bromide ions during solvothermal treatment in TEG.

2. Results

2.1. Characterization of BiOBr-(s)MS

The XRD patterns of BiOBr-sMS and BiOBr-MS (Figure 1) are consistent with the standard pattern of tetragonal BiOBr (PDF No. 78-0348) without any impurity peak, indicating that pure BiOBr particles with good crystallinity were obtained whether PVP was added to the initial reaction mixture or not. In comparison with the standard XRD pattern, the relative intensity of the (110) diffraction peak was obviously enhanced, as observed in the literature [21,23]. This may be related to the hierarchical spherical morphology of BiOBr-(s)MS formed by self-assembly of nanoplates building units (SEM images shown below). During the XRD measurement, most nanoplates in hierarchical spheres would stand upright rather than lie flat on the XRD holder, which facilitates the exposure of more (110) facets to the incident X-ray beam and leads to the intensified (110) diffraction peak.
In SEM images (Figure 2), BiOBr-MS particles are mostly microspheres of 2.57 ± 0.49 μm in diameter. Each microsphere possesses porous flower-like hierarchical structure formed by self-assembly of nanoplates of 30.2 ± 3.5 nm in thickness (Figure 2c). It is worthy to note that the BiOBr-MS microspheres are directly self-assembled without the assistance of a template. In our method, TEG is a reducing solvent with a high-boiling point (~285 °C) [25,26,27,28]. During solvothermal treatment in TEG at 150 °C, KBrO3 will be gradually reduced to provide bromide ions for controlled growth and self-assembly of BiOBr-MS microspheres.
The addition of PVP during solvothermal synthesis of BiOBr-MS reduces the particle size by about three-fold to generate BiOBr-sMS sub-microspheres of 0.87 ± 0.25 μm in diameter, as revealed by its SEM and TEM images (Figure 3). In addition, the constituent nanoplates (24.8 ± 2.8 nm thick) in BiOBr-sMS microspheres are also thinner than those in BiOBr-MS. PVP has been generally exploited to make thin BiOX (X = Cl, Br or I) nanosheets irrespective of different halogen ions [13,29,30]. Therefore, PVP mainly interacts with Bi3+ ions in the [Bi2O2]2− layers of BiOBr-sMS via its carbonyl or amine functional groups of PVP molecules, and restricts the growth along the [001] direction to generate thinner BiOX nanoplates. Meanwhile, the polyvinyl chains of PVP adsorbed on the BiOBr-sMS nanoplates also contribute to their self-assembly process via van der Waals forces [22], inducing the formation of BiOBr-sMS with smaller particle sizes.
The porous structures of BiOBr-(s)MS were also supported by N2 adsorption–desorption isotherms (Figure 4). Both BiOBr-sMS and BiOBr-MS showed Type IV isotherm with H4 hysteresis loop between P/P0 = 0.4 and 1.0 according to IUPAC recommendation [31]. Type IV isotherms are typical of mesoporous adsorbents. The Type H4 loop is often associated with narrow slit-like pore. The SBET and Vp values are 18.8 m2/g and 0.068 cm3/g for BiOBr-sMS, and 36.9 m2/g and 0.118 cm3/g for BiOBr-MS. The larger SBET of BiOBr-MS is ascribed to its rich micropore, as manifested by higher adsorption at lower relative pressure in its N2 adsorption isotherm (Figure 4). The average pore diameter dp (nm) was estimated by dp = 4000VP/SBET [32,33,34,35] to be 14.5 nm for BiOBr-sMS and 12.8 nm for BiOBr-MS, which indicates that BiOBr-sMS has a more open hierarchical structure than BiOBr-MS.

2.2. Photocatalytic Performances of BiOBr-(s)MS

The photocatalytic performances of BiOBr-sMS and BiOBr-MS were evaluated by simulated sunlight-driven aerobic photooxidation of BzOH to BzH (Figure 5). After illumination with Xenon lamp for 2 h, the BzOH conversions (CBzOH) over BiOBr-sMS and BiOBr-MS are 51.3% and 29.6%, respectively, with nearly 100% selectivity to BzH (SBzH). The high selectivity of BzH may be related to selective adsorption of BzOH via –CH2OH group on BiOBr, which is beneficial for its selective oxidation by photo-generated surface reactive species (e.g., photo-holes and ·O2). However, as the photooxidation of BzOH on BiOBr-sMS proceeded, SBzH gradually decreased (e.g., 67.0% at 4 h in Figure 6) due to further oxidation of BzH to benzoic acid. The photocatalytic activity of BiOBr-sMS is significantly higher than that of BiOBr-MS under Xenon lamp illumination. Since both BiOBr-sMS and BiOBr-MS were synthesized via the same synthetic procedure, except for the addition of PVP during the solvothermal synthesis of BiOBr-sMS, BiOBr-MS was post-treated with PVP as follows: 0.30 g of BiOBr-MS was stirred in 30 mL of DI water containing 0.40 g of PVP at room temperature for 1 h, centrifuged and then dried in a 60 °C vacuum oven overnight. The photocatalytic activity of the resulting control sample (denoted PVP/BiOBr-MS) was evaluated to clarify the role of PVP on the enhanced photocatalytic performance of BiOBr-sMS. As shown in Figure 5, PVP/BiOBr-MS has the lowest activity among the three samples (2 h BzOH conversion in the order of 11.6%, 29.6% and 51.3%). It indicated that the post-modification of BiOBr-MS by PVP deteriorated the photocatalytic activity of BiOBr-MS, possibly by blocking some active surface sites. Since physically adsorbed PVP could not improve the activity of BiOBr-sMS, it implied the importance of PVP in modulating the growth and self-assembly of BiOBr-sMS during its synthesis, which resulted in smaller, three-dimensional, hierarchical porous sub-microspheres made up of thinner nanoplates.
The photooxidation products of BzOH over BiOBr-sMS at different illumination times were analyzed to obtain more information about its kinetic process. As shown in Figure 6, CBzOH increased with the illumination time. Upon illumination by Xe lamp for 4 h, 96.5% of 50 mM BzOH was converted, which is among the highest conversion achieved on single-component BiOBr photocatalysts for visible light-driven photooxidation of BzOH with a high-initial concentration (50 mM) (Table 1). However, SBzH gradually decreased with illumination (SBzH: 100% at 2 h and 67.0% at 4 h). The decrease in SBzH after illumination for 4 h may be due to the high concentration of BzH accumulated in the reaction system, which was further oxidized to benzoic acid. As a result, the yield of BzH (YBzH) was almost constant after illumination for 3 h (64.7–65.2%). It is interesting to note that the CBzOH–time plot in Figure 6 was excellently fitted with a linear equation, CBzOH = 0.249t (R2 = 0.999). It indicated that the photocatalytic oxidation of BzOH over BiOBr-sMS conformed to the zero-order reaction model with the rate constant of k0 = 12.5 mmol L−1 h−1. Similarly, the CBzOH–time plots over BiOBr-MS and PVP/BiOBr-MS could also be fitted into the zero-order reaction model with the corresponding k0 value of 6.7 and 3.0 mmol L−1 h−1 (Figure S1). The largest k0 value of BiOBr-sMS is consistent with its highest activity.

2.3. Mechanistic Studies of Enhanced Photocatalytic Activity of BiOBr-sMS

It is generally accepted that the photocatalytic activity is affected by various factors, such as specific surface area, band-gap energy, band position and separation efficiency of photo-induced electrons and holes. The BET surface areas of BiOBr-sMS and BiOBr-MS are 18.8 and 36.9 m2/g, respectively. Therefore, the surface area may not be the key factor contributing to the enhanced photocatalytic activity of BiOBr-sMS. However, BiOBr-sMS has larger average pore diameter than BiOBr-MS (14.5 vs. 12.8 nm), which should benefit the mass transport and diffusion in BiOBr-sMS.
UV-vis absorption spectra of BiOBr-(s)MS are shown in Figure 7a. As compared to BiOBr-MS, the absorption onset of BiOBr-sMS shifted to a shorter wavelength by about 17 nm (398 vs. 415 nm) with lower absorbance. Thus BiOBr-sMS utilizes less visible and ultra-violet light of the total solar energy, which cannot explain its improved photocatalytic activity either. The band-gap energy, Eg, was estimated by Tauc plots according to αhν = A(hν−Eg)n/2, where α, ν and A are the absorption coefficient, light frequency and a constant, respectively, and n is a constant that depends on the characteristics of the optical transition in a semiconductor (n = 1 for a direct transition and n = 4 for an indirect transition). For BiOBr, as an indirect gap semiconductor, the value of n is 4 [13,19,37]. By extrapolating the αhν—(hν)1/2 plots to the x axis (Figure 8 inset), the Eg values of BiOBr-MS and BiOBr-sMS are estimated to be 2.75 eV and 2.85 eV, respectively. The larger band-gap energy of BiOBr-sMS is ascribed to its smaller particle size.
Theoretical calculation reveals that the conduction and valance bands of BiOBr consist of Bi 6p, O 2p and Br 4p orbital [37,38]. The O 2p and Br 4p states dominate the valance-band (VB) maximum, while Bi 6p states contribute the most to the conduction-band (CB) minimum [38]. Both BiOBr-MS and BiOBr-sMS have similar compositions except the involvement of PVP during the solvothermal synthesis of BiOBr-sMS. Since PVP mainly interacts with the Bi3+ ions via its carbonyl or amine functional groups, it is reasonable to assume that PVP mainly influences the CB position of BiOBr-SMS while both BiOBr-MS and BiOBr-sMS have similar VB positions. As the band-gap energy of BiOBr-sMS is larger than that of BiOBr-MS by 0.10 eV, the CB minimum of BiOBr-sMS is expected to shift upwards by 0.10 eV. As compared to BiOBr-MS, the more negative CB position of BiOBr-sMS is expected to facilitate the reduction of adsorbed molecular oxygen to generate more superoxide radical anions (·O2), which was collaborated by the following scavenging tests and spin-trapped EPR spectra. The generation of more·O2 contributes to the enhanced photocatalytic activity of BiOBr-sMS.
Photoluminescence (PL) spectroscopy is generally used to evaluate the recombination rate of photo-excited charge carriers. The stronger PL intensity suggests rapid recombination of photoelectrons and holes. As displayed in Figure 7b, the PL spectrum of BiOBr-sMS had lower intensity than that of BiOBr-MS due to the reduced charge recombination in the former. It implied more efficient separation of photo-induced electrons and holes in BiOBr-sMS, which also contributes to its enhanced photocatalytic performance.
A series of active species trapping experiments were conducted to further investigate the photooxidation mechanism of BzOH. When triethanolamine (TEA) was added to the BzOH system to trap the holes (h+) [3], the BzOH conversion decreased significantly (e.g., from 51.0% to 14.2% over BiOBr-sMS in Figure 8), revealing that the photo-generated holes are the major oxidative species for the selective oxidation of BzOH into BzH. The addition of p-benzoquinone to trap superoxide radical anions (·O2) also decreased the conversion of BzOH but to a lesser extent (e.g., from 51.0% to 38.0% over BiOBr-sMS in Figure 8), supporting partial contribution of ·O2 to the oxidation of BzOH. However, when tert-butanol was added to trap hydroxyl radicals [39,40], no obvious influence on the BzOH conversion was observed (e.g., from 51.0% to 47.2% over BiOBr-sMS in Figure 8), suggesting a negligible contribution of hydroxyl radicals to the oxidation of BzOH in our photocatalytic reaction systems. The same reactive species (h+ and ·O2) responsible for the oxidation of BzOH to BzH were also found in other photocatalytic systems (e.g., Ni-OTiO2 [9]).
EPR spectra were collected with DMPO a spin-trapping agent to identify the generation of ·O2. Strong characteristic signals of DMPO–·O2 adduct were observed over BiOBr-sMS and BiOBr-MS under Xenon illumination [3,19], confirming the formation of superoxide radicals (Figure 9). In contrast, no EPR signal was detected in darkness (Figure 9). As compared with BiOBr-MS, the signal intensities of DMPO–·O2 adduct generated over BiOBr-sMS were stronger (the difference curve between BiOBr-sMS and BiOBr-MS displayed in Figure 9), indicating the formation of more ·O2 over BiOBr-sMS, due to its more negative CB position, which also contributed to its enhanced photocatalytic activity.

3. Materials and Methods

3.1. Chemicals Reagents

Bismuth nitrate pentahydrate (Bi(NO3)3·5H2O) and benzaldehyde (BzH) were purchased from Aladdin (Shanghai, China). KBrO3, triethylene glycol (TEG), acetonitrile (CH3CN), benzyl alcohol (BzOH), octane (C8H18), triethanolamine (TEA) and p-quinone (BQ) were provided by Sinopharm Chemical Reagent, China. Tert-butanol (TBA) was obtained from Shanghai Macklin Biochemical, China. Poly(vinylpyrrolidone) (PVP, Mw: 55,000) and 5,5-dimethyl-1-pyrrolidine N-oxide (DMPO) were supplied by Sigma-Aldrich (Shanghai, China). All chemicals were of analytical reagents and used as received. All solutions were prepared in deionized (DI) water.

3.2. Solvothermal Synthesis of BiOBr-(s)MS

BiOBr-sMS was made as follows: 0.4851 g (1.0 mmol) of Bi(NO3)3·5H2O was first dissolved in 25 mL of TEG upon sonication and magnetic stirring. Then, 0.4 g of PVP was added and stirred for 30 min, followed by the addition of 0.1670 g (1.0 mmol) of KBrO3. After being stirred for another 1 h, the whole reaction mixture was transferred to a 50-mL autoclave and heated at 150 °C for 3 h. The BiOBr-sMS particles were recovered by centrifugation, washed five times with DI water and dried in a 60 °C vacuum for 12 h.
BiOBr-MS was prepared following the same procedure except that no PVP was added in initial reaction mixture.

3.3. Characterization

The crystal structure was determined by powder X-ray diffractometer (XRD, Bruker AXS D8 Advance, Karlsruhe, Germany) with Cu Kα (1.5406 Å) as the X-ray source in the 2θ range of 10°~80° at a scanning speed of 5°/min. The morphology was observed by a scanning electron microscope (SEM, Hitachi SU8010, Tokyo, Japan). UV-vis diffuse reflectance spectra (DRS, Tpkyo, Japan) were collected on the Hitachi UH4150 UV-vis spectrometer with BaSO4 as a reference, and converted from reflection to absorbance by the Kubelka-Munk method. Photoluminescence (PL) spectra were performed via the Hitachi F-4600 fluorescence spectrophotometer with an excitation wavelength at 320 nm. N2 adsorption–desorption isotherms were measured on Micromeritics ASAP 2020 analyzer (USA). All samples were degassed at 150 °C for 4 h under vacuum before measurement. The specific surface area was calculated using the BET method (SBET) from adsorption data in a relative pressure range from 0.05 to 0.30. The pore volume (Vp) was assessed from the adsorbed amount at a relative pressure of 0.99.

3.4. Photocatalytic Aerobic Oxidation of BzOH over BiOBr Nanostructures

In a typical process, 20 mg of BiOBr-sMS was first stirred in 10 mL of BzOH (50 mM) solution in CH3CN in a closed double-wall quartz cell in the dark at room temperature for 30 min to reach the adsorption equilibrium. The reaction temperature was controlled by circulating water. The quartz cell was connected to a balloon full of molecular oxygen (Scheme 2). Then, the suspension was illuminated from the top for 2–4 h with a 225 W Xe lamp (CEL-PF300-T8, Beijing China Education AuLight Technology, Beijing, China) to simulate sunlight. At certain intervals, about 1.5 mL of suspension was sampled, centrifuged and filtered through a filter membrane. The product was analyzed with n-octane as an internal standard on the Fuli 9720 gas chromatograph (China) fitted with a FID detector. The catalytic experiments were repeated with the standard deviation of about 5%.

3.5. Scavenging and Spin-Trapping Tests

TEA (66 μL), BQ (0.0541 g) and TBA (48 μL) were added to scavenge h+, ·O2 and ·OH, respectively, during photocatalytic aerobic oxidation of BzOH, as described in Section 3.4. In spin-trapping studies, DMPO was used as a trapping agent to detect ·O2 by electron paramagnetic resonance (EPR) spectra, which were recorded on the Bruker EMXplus Spectrometer at room temperature.

4. Conclusions

Porous hierarchical BiOBr microspheres (BiOBr-MS) of 2.57 ± 0.49 μm were first self-assembled without the assistance of a template via a facile solvothermal synthesis in triethylene glycol (TEG) at 150 °C for 3 h. The slow release of bromide ions upon reduction of KBrO3 in TEG facilitated the controlled growth and self-assembly of primary BiOBr nanoplates. The addition of PVP during solvothermal synthesis of BiOBr-MS reduced the particle size by nearly three-fold to generate BiOBr sub-microspheres (BiOBr-sMS) of 0.87 ± 0.25 μm. BiOBr-sMS exhibited remarkably improved photocatalytic activity than BiOBr-MS for aerobic photooxidation of benzyl alcohol (BzOH) to benzaldehyde (BzH) under simulated sunlight irradiation (the conversions of BzOH (50 mM) over BiOBr-sMS and BiOBr-MS were, respectively, 51.3% and 29.6% with 100% selectivity to BzH after Xenon illumination for 2 h at 25 °C). Scavenging tests and DMPO-O2 spin-trapping EPR spectra supported that photogenerated holes (h+) and ·O2 were the main reactive species for the BzOH oxidation over BiOBr spheres. The higher photocatalytic activity of BiOBr-sMS was attributed to its more open hierarchical structure, efficient charge separation, more negative conduction-band position and the generation of larger amounts of ·O2 species.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13060958/s1, Figure S1: Time-dependent aerobic photo-oxidation of BzOH over BiOBr nanostructures.

Author Contributions

Z.W.: Resource, Formal analysis, Writing—Original Draft. C.L.: Investigation, Methodology, Validation. F.C.: Conceptualization, Writing—Review and Editing. R.C.: Funding acquisition, Writing—Review and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (22076149) and Special Project of State Key Laboratory of New Textile Materials & Advanced Processing Technologies from Wuhan Science and Technology Bureau (2022013988065204).

Data Availability Statement

Data are available within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brühne, F.; Wright, E. Benzaldehyde; Wiley-VCH Verlag GmbH & Co. KGaA: Hoboken, NJ, USA, 2000. [Google Scholar]
  2. Rak, M.J.; Lerro, M.; Moores, A. Hollow iron oxide nanoshells are active and selective catalysts for the partial oxidation of styrene with molecular oxygen. Chem. Commun. 2014, 50, 12482–12485. [Google Scholar] [CrossRef] [PubMed]
  3. Xiao, X.; Jiang, J.; Zhang, L. Selective oxidation of benzyl alcohol into benzaldehyde over semiconductors under visible light: The case of Bi12O17Cl2 nanobelts. Appl. Catal. B Environ. 2013, 142–143, 487–493. [Google Scholar] [CrossRef]
  4. Schultz, D.M.; Yoon, T.P. Solar synthesis: Prospects in visible light photocatalysis. Science 2014, 343, 1239176. [Google Scholar] [CrossRef] [PubMed]
  5. Ma, B.; Huang, E.; Wu, G.; Dai, W.; Guan, N.; Li, L. Fabrication of WO2.72/RGO nano-composites for enhanced photocatalysis. RSC Adv. 2017, 7, 2606–2614. [Google Scholar] [CrossRef]
  6. Li, P.; Zhao, H.; Yan, X.; Yang, X.; Li, J.; Gao, S.; Cao, R. Visible-light-driven photocatalytic hydrogen production coupled with selective oxidation of benzyl alcohol over CdS@MoS2 heterostructures. Sci. China Mater. 2020, 63, 2239–2250. [Google Scholar] [CrossRef]
  7. Karimian, D.; Yadollahi, B.; Mirkhani, V. Harvesting visible light for aerobic oxidation of alcohols by a novel and efficient hybrid polyoxometalate. Dalton Trans. 2015, 44, 1709–1715. [Google Scholar] [CrossRef]
  8. Su, Y.; Han, Z.; Zhang, L.; Wang, W.; Duan, M.; Li, X.; Zheng, Y.; Wang, Y.; Lei, X. Surface hydrogen bonds assisted meso-porous WO3 photocatalysts for high selective oxidation of benzylalcohol to benzylaldehyde. Appl. Catal. B Environ. 2017, 217, 108–114. [Google Scholar] [CrossRef]
  9. She, H.; Zhou, H.; Li, L.; Wang, L.; Huang, J.; Wang, Q. Nickel-doped excess oxygen defect titanium dioxide for efficient selective photocatalytic oxidation of benzyl alcohol. ACS Sustain. Chem. Eng. 2018, 6, 11939–11948. [Google Scholar] [CrossRef]
  10. Lam, S.-M.; Jaffari, Z.H.; Sin, J.-C.; Zeng, H.; Lin, H.; Li, H.; Mohamed, A.R. Insight into the influence of noble metal decorated on BiFeO3 for 2,4-dichlorophenol and real herbicide wastewater treatment under visible light. Colloids Surf. A Physicochem. Eng. Asp. 2021, 614, 126138. [Google Scholar] [CrossRef]
  11. Chin, Y.-H.; Sin, J.-C.; Lam, S.-M.; Zeng, H.; Lin, H.; Li, H.; Huang, L.; Mohamed, A.R. 3-D/3-D Z-scheme heterojunction composite formed by marimo-like Bi2WO6 and mammillaria-like ZnO for expeditious sunlight photodegradation of dimethyl phthalate. Catalysts 2022, 12, 1427. [Google Scholar] [CrossRef]
  12. Li, J.; Yu, Y.; Zhang, L. Bismuth oxyhalide nanomaterials: Layered structures meet photocatalysis. Nanoscale 2014, 6, 8473–8488. [Google Scholar] [CrossRef]
  13. Chen, J.; Guan, M.; Cai, W.; Guo, J.; Xiao, C.; Zhang, G. The dominant {001} facet-dependent enhanced visible-light photoactivity of ultrathin BiOBr nanosheets. Phys. Chem. Chem. Phys. 2014, 16, 20909–20914. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, F.J.; Gu, Y.Y.; Yang, Z.Y.; Xie, Y.Y.; Zhang, J.J.; Shang, X.T.; Zhao, H.B.; Zhang, Z.Z.; Wang, X.X. The effect of halogen on BiOX (X = Cl, Br, I)/Bi2WO6 heterojunction for visible-light-driven photocatalytic benzyl alcohol selective oxidation. Appl. Catal. A-Gen. 2018, 567, 65–72. [Google Scholar] [CrossRef]
  15. Bisht, N.S.; Mehta, S.P.S.; Sahoo, N.G.; Dandapat, A. The room temperature synthesis of a CuO-Bi-BiOBr ternary Z-scheme photocatalyst for enhanced sunlight driven alcohol oxidation. Dalton Trans. 2021, 50, 5001–5010. [Google Scholar] [CrossRef]
  16. Wang, X.; Zhang, Y.; Xu, H.; Ji, X.; Gan, L.; Zhang, R. Sugar-regulated bismuth oxybromide flowers with active imprinting sites for efficient photooxidative ability. J. Alloys Compd. 2021, 879, 160374. [Google Scholar] [CrossRef]
  17. Bisht, N.S.; Pancholi, D.; Sahoo, N.G.; Melkani, A.B.; Mehta, S.P.S.; Dandapat, A. Effect of Ag–Fe–Cu tri-metal loading in bismuth oxybromide to develop a novel nanocomposite for the sunlight driven photocatalytic oxidation of alcohols. Catal. Sci. Technol. 2019, 9, 3923–3932. [Google Scholar] [CrossRef]
  18. Sin, J.C.; Lam, S.M.; Zeng, H.; Lin, H.; Li, H.; Huang, L.; Liaw, S.J.; Mohamed, A.R.; Lim, J.W. Construction of visible light-driven Eu-doped BiOBr hierarchical microflowers for ameliorated photocatalytic 2,4-dichlorophenol and pathogens decomposition with synchronized hexavalent chromium reduction. Mater. Today Sustain. 2023, 22, 100340. [Google Scholar] [CrossRef]
  19. Wang, H.; Yong, D.; Chen, S.; Jiang, S.; Zhang, X.; Shao, W.; Zhang, Q.; Yan, W.; Pan, B.; Xie, Y. Oxygen-vacancy-mediated exciton dissociation in BiOBr for boosting charge-carrier-involved molecular oxygen activation. J. Am. Chem. Soc. 2018, 140, 1760–1766. [Google Scholar] [CrossRef]
  20. Dai, R.; Zhang, L.; Ning, J.; Wang, W.; Wu, Q.; Yang, J.; Zhang, F.; Wang, J.-A. New insights into tuning BiOBr photocatalysis efficiency under visible-light for degradation of broad-spectrum antibiotics: Synergistic calcination and doping. J. Alloys Compd. 2021, 887, 161481. [Google Scholar] [CrossRef]
  21. Huo, Y.; Zhang, J.; Miao, M.; Jin, Y. Solvothermal synthesis of flower-like BiOBr microspheres with highly visible-light photocatalytic performances. Appl. Catal. B Environ. 2012, 111–112, 334–341. [Google Scholar] [CrossRef]
  22. Juntrapirom, S.; Tantraviwat, D.; Anuchai, S.; Thongsook, O.; Channei, D.; Inceesungvorn, B. Boosting photocatalytic coupling of amines to imines over BiOBr: Synergistic effects derived from hollow microsphere morphology. J. Environ. Chem. Eng. 2021, 9, 106732. [Google Scholar] [CrossRef]
  23. Ahmad, A.; Meng, X.C.; Yun, N.; Zhang, Z.S. Preparation of hierarchical BiOBr microspheres for visible light-induced photocatalytic detoxification and disinfection. J. Nanomater. 2016, 2016, 1373725. [Google Scholar] [CrossRef]
  24. Nan, Q.; Huang, S.; Zhou, Y.; Zhao, S.; He, M.; Wang, Y.; Li, S.; Huang, T.; Pan, W. Ionic liquid-assisted synthesis of porous BiOBr microspheres with enhanced visible light photocatalytic performance. Appl. Organomet. Chem. 2018, 32, e4596. [Google Scholar] [CrossRef]
  25. Grabs, I.-M.; Bradtmöller, C.; Menzel, D.; Garneting, G. Formation mechanisms of iron oxide nanoparticles in different nonaqueous media. Cryst. Growth Des. 2012, 12, 1469–1475. [Google Scholar] [CrossRef]
  26. Cai, W.; Wan, J. Facile synthesis of superparamagnetic magnetite nanoparticles in liquid polyols. J. Colloid Interface Sci. 2007, 305, 366–370. [Google Scholar] [CrossRef] [PubMed]
  27. Miguel-Sancho, N.; Bomati-Miguel, O.; Roca, A.G.; Martinez, G.; Arruebo, M.; Santamaria, J. Synthesis of magnetic nanocrystals by thermal decomposition in glycol media: Effect of process variables and mechanistic study. Ind. Eng. Chem. Res. 2012, 51, 8348–8357. [Google Scholar] [CrossRef]
  28. Zhou, S.; Wang, C.; Xia, Y. Polyol synthesis of Pd icosahedral nanocrystals: Insights into the growth mechanism and size control. Chem. Mater. 2022, 34, 5065–5073. [Google Scholar] [CrossRef]
  29. Guan, M.; Xiao, C.; Zhang, J.; Fan, S.; An, R.; Cheng, Q.; Xie, J.; Zhou, M.; Ye, B.; Xie, Y. Vacancy associates promoting solar-driven photocatalytic activity of ultrathin bismuth oxychloride nanosheets. J. Am. Chem. Soc. 2013, 135, 10411–10417. [Google Scholar] [CrossRef]
  30. Di, J.; Xia, J.; Ji, M.; Xu, L.; Yin, S.; Chen, Z.; Li, H. Bidirectional acceleration of carrier separation spatially via N-CQDs/atomically-thin BiOl nanosheets nanojunctions for manipulating active species in a photocatalytic process. J. Mater. Chem. A 2016, 4, 5051–5061. [Google Scholar] [CrossRef]
  31. Sing, K.S.W. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity (Recommendations 1984). Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  32. Rabah, M.A.; Yasri, N.G.; Alaya, M.N. Impact of aging on the structural, textural, and acid properties of WO3-SO42−-SnO2 solid acids. J. Alloys Compd. 2019, 790, 452–465. [Google Scholar] [CrossRef]
  33. Ben, T.; Ren, H.; Ma, S.; Cao, D.; Lan, J.; Jing, X.; Wang, W.; Xu, J.; Deng, F.; Simmons, J.M.; et al. Targeted synthesis of a porous aromatic framework with high stability and exceptionally high surface area. Angew. Chem. Int. Ed. 2009, 48, 9457–9460. [Google Scholar] [CrossRef] [PubMed]
  34. Ali, M.A.; Asaoka, S. Ni-Mo-titania-alumina catalysts with USY zeolite for low pressure hydrodesulfurization and hydrocracking. Pet. Sci. Technol. 2009, 27, 984–997. [Google Scholar] [CrossRef]
  35. Al-Omair, M.A.; El-Sharkawy, E.A. Removal of heavy metals via adsorption on activated carbon synthesized from solid wastes. Environ. Technol. 2007, 28, 443–451. [Google Scholar] [CrossRef]
  36. Yan, J.; Wu, G.; Guan, N.; Li, L. Nb2O5/TiO2 heterojunctions: Synthesis strategy and photocatalytic activity. Appl. Catal. B Environ. 2014, 152–153, 280–288. [Google Scholar] [CrossRef]
  37. Zhang, H.; Liu, L.; Zhou, Z. Towards better photocatalysts: First-principles studies of the alloying effects on the photocatalytic activities of bismuth oxyhalides under visible light. Phys. Chem. Chem. Phys. 2012, 14, 1286–1292. [Google Scholar] [CrossRef]
  38. Xiao, X.; Liu, C.; Hu, R.; Zuo, X.; Nan, J.; Li, L.; Wang, L. Oxygen-rich bismuth oxyhalides: Generalized one-pot synthesis, band structures and visible-light photocatalytic properties. J. Mater. Chem. 2012, 22, 22840–22843. [Google Scholar] [CrossRef]
  39. Chen, F.; Lv, H.; Chen, W.; Chen, R. Catalytic wet peroxide oxidation of anionic pollutants over fluorinated Fe3O4 microspheres at circumneutral pH values. Catalysts 2022, 12, 1564. [Google Scholar] [CrossRef]
  40. Guo, S.; Chen, M.; Huang, Y.; Wei, Y.; Ali, J.; Cai, C.; Wei, Q.S. Three-dimensionally printed zero-valent copper with hierarchically porous structures as an efficient Fenton-like catalyst for enhanced degradation of tetracycline. Catalysts 2023, 13, 446. [Google Scholar] [CrossRef]
Scheme 1. Structure models of BiOBr. (a) Three-dimensional projection, (b) [001] direction and (c) [110] direction.
Scheme 1. Structure models of BiOBr. (a) Three-dimensional projection, (b) [001] direction and (c) [110] direction.
Catalysts 13 00958 sch001
Figure 1. XRD patterns of BiOBr-sMS and BiOBr-MS.
Figure 1. XRD patterns of BiOBr-sMS and BiOBr-MS.
Catalysts 13 00958 g001
Figure 2. SEM images (ac) and particle size histogram (d) of BiOBr-MS.
Figure 2. SEM images (ac) and particle size histogram (d) of BiOBr-MS.
Catalysts 13 00958 g002
Figure 3. SEM (ac), particle size histogram (d) and TEM images (e,f) of BiOBr-sMS.
Figure 3. SEM (ac), particle size histogram (d) and TEM images (e,f) of BiOBr-sMS.
Catalysts 13 00958 g003
Figure 4. N2 adsorption–desorption isotherms of BiOBr-sMS and BiOBr-MS.
Figure 4. N2 adsorption–desorption isotherms of BiOBr-sMS and BiOBr-MS.
Catalysts 13 00958 g004
Figure 5. Photocatalytic performances of BiOBr-(s)MS. Reaction conditions: 20 mg catalyst, O2 balloon and 50 mM BzOH in 10 mL CH3CN, illuminated with Xenon lamp for 2 h.
Figure 5. Photocatalytic performances of BiOBr-(s)MS. Reaction conditions: 20 mg catalyst, O2 balloon and 50 mM BzOH in 10 mL CH3CN, illuminated with Xenon lamp for 2 h.
Catalysts 13 00958 g005
Figure 6. Time-dependent photocatalytic aerobic oxidation of BzOH over BiOBr-sMS. Reaction conditions: 20 mg BiOBr-sMS, O2 balloon and 50 mM BzOH in 10 mL CH3CN, illuminated with Xenon lamp.
Figure 6. Time-dependent photocatalytic aerobic oxidation of BzOH over BiOBr-sMS. Reaction conditions: 20 mg BiOBr-sMS, O2 balloon and 50 mM BzOH in 10 mL CH3CN, illuminated with Xenon lamp.
Catalysts 13 00958 g006
Figure 7. UV-vis spectra (a) and photoluminescence spectra (b) of BiOBr-sMS. Inset of (a) is the corresponding Tauc plots of (αhν)1/2–(hν).
Figure 7. UV-vis spectra (a) and photoluminescence spectra (b) of BiOBr-sMS. Inset of (a) is the corresponding Tauc plots of (αhν)1/2–(hν).
Catalysts 13 00958 g007
Figure 8. Trapping experiments of photocatalytic oxidation of BzOH over BiOBr-(s)MS.
Figure 8. Trapping experiments of photocatalytic oxidation of BzOH over BiOBr-(s)MS.
Catalysts 13 00958 g008
Figure 9. DMPO-O2 spin-trapping EPR spectra of BiOBr-(s)MS.
Figure 9. DMPO-O2 spin-trapping EPR spectra of BiOBr-(s)MS.
Catalysts 13 00958 g009
Scheme 2. Illustration of the photoreaction system used in this work.
Scheme 2. Illustration of the photoreaction system used in this work.
Catalysts 13 00958 sch002
Table 1. Comparison of catalytic performances of various photocatalysts under Xe lamp illumination.
Table 1. Comparison of catalytic performances of various photocatalysts under Xe lamp illumination.
EntryPhotocatalystConditionsConvn/Sel/Yield (%)Ref.
1BiOBr-sMS20 mg catalyst, 0.5 mmol BzOH, O2 balloon, 10 mL CH3CN, 25 °C, 2 h51.3%/100%/51.3%
(96.5%/67.0%/64.7% at 4 h)
This work
2WO2.720.1 g catalyst, 25 mmol BzOH, O2 (20 mL/min), 60 mL BTF, 25 °C, 8 h9.5%/91.0%/8.6%[5]
3WO2.72/rGO0.1 g catalyst, 25 mmol BzOH, O2 (20 mL/min), 60 mL BTF, 25 °C, 8 h37.6%/92.0%/34.6%[5]
4Fluorinated mesoporous WO3 0.1 g catalyst, 0.025 mmol BzOH, 50 mL H2O, 25 °C, 4 h.~57%/~99%/56% [8]
5Hybrid-SiW11O3940 mg catalyst, 1 mmol BzOH, 3 mL CH3CN purged with O2, 25 °C, 4 h80% yield[7]
60.8Br-BiOBr
/Bi2WO6
20 mg catalyst, 0.2 mmol BzOH, 2.5 mL BTF saturated with O2, 25 °C, 4 h30.9%/100%/30.9%[14]
7TiO2 0.3 g catalyst, 25 mmol BzOH, 27 mL BTF, air bubbled at 20 mL/min, 5 h, (temperature unavailable) 21.6%/91.4%/19.7%[36]
8Nb2O5/TiO2 0.3 g catalyst, 25 mmol BzOH, 27 mL BTF, air bubbled at 20 mL/min, 5 h, (temperature unavailable)64.3%/85.1%/54.7%[36]
9Ni(1%)-OTiO280 mg catalyst, 0.5 mmol BzOH, 5 mL BTF saturated with O2 (2atm) for 5 min, 1 h (temperature unavailable)93%/99%/88%[9]
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

Wang, Z.; Liu, C.; Chen, F.; Chen, R. Self-Assembly of Porous Hierarchical BiOBr Sub-Microspheres for Efficient Aerobic Photooxidation of Benzyl Alcohol under Simulated Sunlight Irradiation. Catalysts 2023, 13, 958. https://doi.org/10.3390/catal13060958

AMA Style

Wang Z, Liu C, Chen F, Chen R. Self-Assembly of Porous Hierarchical BiOBr Sub-Microspheres for Efficient Aerobic Photooxidation of Benzyl Alcohol under Simulated Sunlight Irradiation. Catalysts. 2023; 13(6):958. https://doi.org/10.3390/catal13060958

Chicago/Turabian Style

Wang, Zhigang, Cheng Liu, Fengxi Chen, and Rong Chen. 2023. "Self-Assembly of Porous Hierarchical BiOBr Sub-Microspheres for Efficient Aerobic Photooxidation of Benzyl Alcohol under Simulated Sunlight Irradiation" Catalysts 13, no. 6: 958. https://doi.org/10.3390/catal13060958

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