Next Article in Journal
Low-Temperature Solution-Processed HfZrO Gate Insulator for High-Performance of Flexible LaZnO Thin-Film Transistor
Next Article in Special Issue
Experimental and Theoretical Investigation of Gadolinium Oxyhydride (GdHO) Thin Films: Optical, Photocatalytic, and Electronic Properties
Previous Article in Journal
Increasing the Accuracy of the Characterization of a Thin Semiconductor or Dielectric Film on a Substrate from Only One Quasi-Normal Incidence UV/Vis/NIR Reflectance Spectrum of the Sample
Previous Article in Special Issue
Construction of 2D/2D Mesoporous WO3/CeO2 Laminated Heterojunctions for Optimized Photocatalytic Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bi/Mn-Doped BiOCl Nanosheets Self-Assembled Microspheres toward Optimized Photocatalytic Performance

1
Shandong Provincial Key Laboratory of Molecular Engineering, School of Chemistry and Chemical Engineering, Qilu University of Technology (Shandong Academy of Sciences), Jinan 250353, China
2
Key Laboratory of Functional Inorganic Material Chemistry, Ministry of Education of the People’s Republic of China, School of Chemistry and Materials Science, Heilongjiang University, Harbin 150080, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2023, 13(17), 2408; https://doi.org/10.3390/nano13172408
Submission received: 10 July 2023 / Revised: 20 August 2023 / Accepted: 23 August 2023 / Published: 25 August 2023

Abstract

:
Doping engineering of metallic elements is of significant importance in photocatalysis, especially in the transition element range where metals possess empty ‘d’ orbitals that readily absorb electrons and increase carrier concentration. The doping of Mn ions produces dipole interactions that change the local structure of BiOCl, thus increasing the specific surface area of BiOCl and the number of mesoporous distributions, and providing a broader platform and richer surface active sites for catalytic reactions. The combination of Mn doping and metal Bi reduces the forbidden bandwidth of BiOCl, thereby increasing the absorption in the light region and strengthening the photocatalytic ability of BiOCl. The degradation of norfloxacin by Bi/Mn-doped BiOCl can reach 86.5% within 10 min. The synergistic effect of Mn doping and Bi metal can change the internal energy level and increase light absorption simultaneously. The photocatalytic system created by such a dual-technology combination has promising applications in environmental remediation.

1. Introduction

Semiconductor photocatalysts have drawn considerable attention because of their widespread application in the degradation of organic pollutants in water [1,2,3]. BiOCl, as a classical material with a layered structure, has a promising future in the fields of optics, electricity, and magnetism because of its internal electric field and electron-hole separation resulting from the alternating combination of [Bi2O2]2+ and [Cl] layers [4,5]. However, the performance of BiOCl is limited because it can only be activated by UV light. Therefore, the wider bandgap of BiOCl (about 3.4 eV) necessitates additional methods to improve light utilization. The development of modification techniques for BiOCl groups has been extensively studied, including defect control, construction of heterojunctions, metal deposition, and heteroatom doping. These techniques can also increase the carrier concentration and accelerate the charge transfer rate [6,7,8,9,10].
Among these methods, metal deposition is the compounding of metals on the surface of photocatalysts. It uses the better electrical conductivity of metals to promote electron transfer and effectively prevent electron-hole recombination [11,12]. In addition, due to the different directions of electron transfer, the Schottky barrier or plasma resonance effect will occur, respectively. When the Fermi energy level of metal is lower than that of the semiconductor, electrons tend to transfer from the semiconductor surface to the metal surface. As a consequence, the electron-hole domain is fixed on the metal and semiconductor, respectively, resulting in a Schottky barrier, and thus effectively suppressing the charge recombination [13,14]. When the size of the metal particles is small enough, it is easy to excite hot electrons under visible light irradiation, thus transferring electrons to the semiconductor conduction band position. The collective electron oscillation in the semiconductor conduction band under light irradiation could trigger the surface plasmon resonance effect (SPR) [15,16]. Such a phenomenon can occur not only with precious metals but also with the metal Bi, which has great cost savings and a good overall performance compared to precious metals. Indeed, metal Bi nanoparticle loading has been extensively utilized to improve the photocatalytic performance of semiconductors. For instance, Bi/Bi2O3 [17], Bi/Bi2MoO6 [18], Bi/C3N4 [19], and Bi/CdS [3], etc. photocatalyst systems have been reported. Chang et al. [20] used in situ electron microscopy and theoretical studies to explain in detail the interfacial transport pathways of carriers during the SPR effect induced by metallic bismuth particles in the heterojunction system. Moreover, the metal Bi triggers an increase in the light absorption range after the bandgap change, which is favorable for the photoreactivity of the semiconductor. These studies contribute to a better understanding of the reasons for the increased activity of various bismuth-based photocatalysts.
Heteroatom doping is also a general and effective photocatalyst modification strategy. Because transition metal atoms have unsaturated ‘d’ orbitals, the energy level structure of the semiconductor can be regulated by transition metal atom doping for improved light absorption and photogenerated carrier generation. It can reduce the energy required for electron leap and promote charge separation, thus improving the photocatalytic activity of the photocatalyst [21,22,23,24,25,26]. Currently, numerous researchers have certified that the photocatalytic activity of BiOCl in the degradation of organic pollutants can be improved effectively by metal doping [27,28,29,30]. Manganese (Mn), a common transition metal, is often used for doping into semiconductors because of the diverse valence states of the transition and its low price. Recently, Mn doping has been used to improve the photocatalytic degradation performance of BiOCl. Cen et al. degraded metronidazole using Mn-BiOCl and achieved a removal efficiency of 91.6% after 60 min of treatment under simulated solar light [30]. Pare et al. reported that Mn-BiOCl could remove 98% of malachite green after 120 min under visible light irradiation [31]. In addition, Lin et al. [32] improved the photocatalytic CO2 reduction reaction efficiency by doping manganese ions (Mn2+) in CsPbBr3 halide chalcogenide nanoplates and applying an external magnetic field. Mn doping generates spin-polarized electrons and increases the number of photogenerated carriers using the synergistic effect of Mn doping and the applied magnetic field, which prolongs the carrier lifetime and inhibits charge recombination.
In this work, metal Bi composite and metal Mn doping were used to optimize the bandgap of BiOCl semiconductors. A one-step solvothermal method is used in situ to reduce metal Bi and dope transition metal Mn, both of which modify BiOCl to form Bi/Mn doped BiOCl (MBB) structures. This heterojunction takes advantage of the tight coupling between metallic Bi and BiOCl to shorten the charge transfer distance. Meanwhile, Mn doping changes the internal structure of BiOCl and creates a new bonding energy linkage with it. MBB possesses a higher specific surface area and a more suitable bandgap for light absorption, which enhances the photoresponse and increases the photogenerated carrier concentration and electron-hole separation efficiency. The deposition of metallic Bi and Mn doping modification will further enhance the photocatalytic activity and strengthen the degradation rate of norfloxacin by MBB, thereby constructing a highly active photocatalytic system.

2. Experimental Section

2.1. Chemicals

Bismuth nitrate pentahydrate (Bi(NO3)3·5H2O) was purchased from Shanghai Dibai Biotechnology Co., Ltd. (Shanghai, China), potassium chloride (KCl) was purchased from Tianjin Guangfu Technology Development Co., Ltd. (Tianjin, China), manganese chloride tetrahydrate (MnCl2·4H2O) was purchased from Aladdin Reagent Co., Ltd. (Shanghai, China), and N,N-dimethylformamide (DMF) was purchased from Tianjin Fuyu Fine Chemical Co., Ltd. (Tianjin, China). Norfloxacin (C16H18FN3O3) was purchased from Shanghai XianDing Biotechnology Co., Ltd. (Shanghai, China), and sodium sulfate (Na2SO4) was purchased from Aladdin Reagent Co., Ltd. (Shanghai, China). All the above chemicals are analytical grade and used without further purification.

2.2. Synthesis

The Bi/Mn-doped BiOCl microspheres were prepared by a solvothermal method. A total of 6 mmol Bi(NO3)3·5H2O and 6 mmol KCl were added to 32 mL DMF. After stirring for 30 min, a certain amount of MnCl2·4H2O (molar ratio Mn:Bi = 1:20, 1:30, 1:40) was added and stirred until complete dissolution. The mixture was then transferred to a 50 mL polytetrafluoroethylene (PTFE)-lined stainless steel autoclave for 3 h at 160 °C. The reaction was cooled down and the catalyst was washed with water and ethanol. Finally, the catalyst was dried at 60 °C. The collected sample was denoted as MBB (the molar ratio of Mn:Bi = 1:30 was optimal and this ratio was used for subsequent tests).
Next, 6 mmol Bi(NO3)3·5H2O and 6 mmol KCl were added to 32 mL DMF. After stirring for 30 min, the solution was transferred to a 50 mL PTFE-lined stainless steel autoclave for continuous reaction at 160 °C for 12 h. The collected sample was denoted as Bi/BiOCl.
Then, 6 mmol Bi(NO3)3·5H2O and 6 mmol KCl were added to 32 mL distilled water and the solution was stirred for 30 min. A certain amount of MnCl2·4H2O (molar ratio Mn:Bi = 1:30) was added, stirred until complete dissolution, and then the mixture was transferred to a 50 mL PTFE-lined stainless steel autoclave for 3 h at 160 °C. The collected sample was denoted as MB.
Finally, 6 mmol Bi(NO3)3·5H2O and 6 mmol KCl were added to 32 mL distilled water, the solution was stirred for 30 min, and the mixed solution was transferred to a 50 mL PTFE-lined stainless steel autoclave for continuous reaction at 160 °C for 12 h. The collected sample was denoted as BiOCl.

3. Results and Discussion

3.1. Morphology and Microstructure

The scanning electron microscope (SEM) image of Figure 1a demonstrates many distributed BiOCl nanosheets. The nanosheet structure of MB is shown in Figure S1, and the uniform distribution of Mn, Bi, O, and Cl elements in MB nanosheets is shown in Figure S2. The aggregation of nanosheets forms the microsphere structure of Bi/BiOCl as shown in Figure 1b. Figure 1c,d shows the structural features of MBB. It is observed that MBB is formed by the self-assembly of many nanosheets into tightly bound microspheres, and this structure minimizes the agglomeration of nanosheets [22]. The surface free energy of the thinner nanosheets allows them to be uniformly dispersed and combined into a microsphere structure [33], which maximizes the exposure of the specific surface area and increases the active sites for photocatalysis.
Meanwhile, the mapping images (Figure 1e–h) show that each element is uniformly distributed. The presence of Mn elements can be observed in the mapping images [34]. The energy dispersive X-ray (EDX) of Figure 1i demonstrates the distribution of the elements, with the four elements Mn, Bi, O, and Cl co-existing in the MBB structure [35]. The transmission electron microscopy (TEM) of Figure 1j clearly shows the ultrathin nanosheet structure of MBB, and two lattice stripes are observed by further high-resolution TEM (HRTEM) images (Figure 1k). Between these stripes, a lattice spacing of 0.324 nm was obtained by exposing the metallic Bi (012) crystal plane. There is also a lattice spacing of 0.273 nm obtained by exposing the BiOCl (110) crystal plane, which is slightly reduced compared to the original crystal plane spacing, resulting from the effect of lattice distortion caused by Mn doping.
The X-ray diffraction (XRD) patterns of Figure 2a clearly show the diffraction peaks of BiOCl and Bi in MBB. The crystallographic planes corresponding to each diffraction peak are marked in detail in Figure 2a. Among them, BiOCl in MBB is in the tetragonal structured crystalline phase, corresponding to the standard card JCPDS no. 73-2060 [27,28]. The diffraction peaks corresponding to the (110) and (011) crystallographic planes are the most significant, indicating that the crystallographic planes where the microsphere nanosheets are heavily exposed are probably the (110) crystallographic planes. The presence of in situ reduced metal Bi in the system was also demonstrated by XRD. There are obvious diffraction peaks of singlet Bi at 2θ = 27.2, 37.9, and 39.6°, corresponding to the crystallographic planes (012), (104), and (110), respectively (JCPDS no. 85-1329) [17]. The diffraction peaks of metallic Bi have been marked by orange shading in Figure 2a. Meanwhile, the magnification of the yellow shaded part shows that the diffraction peak of MBB is shifted to a higher angle than that of Bi/BiOCl. This change is due to the replacement of Bi3+ (with a larger ionic radius) by Mn2+ (which has a smaller ionic radius), occupying the interstitial sites and causing the lattice to contract. The ultraviolet-visible diffuse reflectance (UV-vis) image in Figure 2b reflects the optical properties of the MBB catalyst. As can be seen from Figure 2b, MBB exhibits higher absorption intensity than Bi/BiOCl, MB, and BiOCl in the 400–800 nm range, indicating that the Mn doping and metal Bi together enhance the sensitivity of BiOCl to light. The bandgap values of 2.84, 3.05, 3.33, and 3.37 eV for MBB, Bi/BiOCl, MB, and BiOCl, respectively [36], are seen in the bandgap diagram of Figure 2c. This indicates that MBB has a forbidden bandwidth more suitable for sunlight absorption for two reasons. Firstly, Mn doping creates an intermediate energy level in the energy band structure of BiOCl. This energy level can be used as a bridge for electron leap to achieve the reduced bandgap effect of BiOCl. Secondly, the metal Bi has a larger absorption rate of light and can enhance the utilization of UV and visible light by BiOCl. The above reasons make MBB absorb a broader range of visible light [37].
Figure 2d–f shows the characterization of the modulated MBB (Mn:Bi = 1:20, 1:30, 1:40) series. It is observed in the XRD plots of Figure 2d that there is almost no change in XRD although the amounts of doped Mn are different. However, the presence of metallic Bi in MBB is confirmed by three strong Bi diffraction peaks. In the UV-vis plot of Figure 2e, the light absorption intensity of the three ratios of MBB does not differ by much, but the bandgap values (Figure 2f) are slightly different. The bandgaps of MBB (1:20, 1:30, 1:40) samples are 2.79, 2.84, and 2.90 eV, respectively [36]. Moreover, the bandgap values become smaller as the amount of Mn doping gradually increases. These results indicate that Mn doping can effectively regulate the light absorption property of BiOCl.
Figure 3a shows the Raman spectra to determine the vibrational modes of the molecules inside the BiOCl structure. It can be seen from Figure 3a that the characteristic peak of MBB at 144 cm−1 caused by the interlayer vibration of A1g is significantly weakened and blue-shifted. This difference indicates that the Mn doping in BiOCl may be chemically coordinated with the internal bonds, thus shifting the vibrational peak [38]. The Bi–Cl bond stretching within Eg occurs at 199 cm−1 while the O-atom stretching vibration within B1g appears at 395 cm−1. Both the above characteristic peaks almost disappear for MBB, indicating that the doping technique and metal composite are effective in modifying the internal structure and surface state of the material [39]. The electron paramagnetic resonance (EPR) pattern in Figure 3b also further demonstrates the doping of Mn elements in BiOCl, showing the stronger signal response of MBB compared to BiOCl, MB, and Bi/BiOCl due to the six EPR vibrational peaks splitting from the interaction between the nuclear spins of Mn2+ ions (S = 5/2) and their electron spins. These six splitting peaks indicate that the dipole interaction of Mn2+ ions changes the local structure of BiOCl, and indirectly proves the Mn doping in BiOCl [32]. The nitrogen adsorption–desorption isotherm profiles in Figure 3c show typical type IV isotherms for BiOCl, MB, Bi/BiOCl, and MBB, indicating that all three catalysts have a mesoporous structure and MBB exhibits a higher nitrogen adsorption effect. As seen in Table S1, the specific surface area of MBB (21.9 m2/g) is not only 1.2 times higher than that of Bi/BiOCl (18.4 m2/g), but also 22 times higher than that of BiOCl (1.0 m2/g) and 3.4 times higher than that of MB (6.5 m2/g). In addition, the pore size pore capacity of MBB is higher than that of Bi/BiOCl, MB, and BiOCl materials. The mesopore distribution curves in Figure 3d also show that MBB has better mesopore distribution peaks, a large number of mesopore structures, and a high specific surface area that can better trap charges, accelerate carrier separation, and provide more active sites for catalytic reactions [40,41].
Figure 4a–d shows the X-ray photoelectron spectra (XPS) of the MBB, showing the surface chemical states of the four elements. All elements are calibrated in reference to the binding energy of C 1s (284.8 eV). Figure 4a shows the spectrum of the element Bi in MBB, and the two peaks at 158.95 and 164.33 eV belong to the characteristic peaks of Bi3+, corresponding to the Bi 4f7/2 and Bi 4f5/2 tracks, respectively [42]. In addition, two small peaks at 158.59 and 163.86 eV were separated within the characteristic peak of Bi3+, and this peak belongs to the Bi0 characteristic peak, which proves the presence of metallic Bi on the surface of the MBB structure. Figure 4b shows the O elemental spectrum in MBB with distinct peaks at 527.07 and 531.82 eV due to metal-O bonding inside the structure and hydroxyl groups on the sample surface, respectively [43]. However, unlike the previous Bi-O characteristic peaks around 529 eV, the MBB binding energy is blue-shifted, probably due to the substitution of the original lattice oxygen in Mn-doped BiOCl to form Mn-O bonds. Figure 4c shows the elemental spectrum of Cl in MBB, and the characteristic peaks at 198.23 and 199.88 eV belong to the Cl 2p3/2 and Cl 2p1/2 orbitals, respectively. Figure 4d shows the spectrum of the Mn element in MBB. The 640.34 and 652.08 eV split peaks belong to the Mn 2p3/2 and Mn 2p1/2 orbitals. The combined analysis of XRD, EPR, and XPS speculates that the Mn doping in MBB possesses two forms: one is the replacement of some high-valent Bi3+ ions by low-valent Mn2+ ions to form Mn-O bonds, which easily induces the vibration of O atoms in MBB, and another is the grafting of Mn2+ on the surface of BiOCl to form manganese oxygen clusters, which excites the motion of photogenerated carriers [44,45]. Finally, in the full XPS spectrum of Figure 4e, it can be seen more clearly that the MBB sample contains four elements, Bi, O, Cl, and Mn, thus confirming the doping of Mn. The XPS valence band spectrum energy of MBB is known to be 2.24 eV in Figure 4f. The energy level structure produced by Mn doping can both promote electron leap and block the recombination of photogenerated electrons and holes, thus inhibiting charge recombination.
Figure 5a shows the electrochemical impedance spectrum (EIS) to analyze the resistance ability of the material to the AC current. It can be seen that MBB has a smaller radius of curvature and weaker resistance to current than Bi/BiOCl, MB, and BiOCl, indicating that MBB heterojunction possesses stronger charge transfer ability and diffusion ability [41,46]. The energy level change of MBB can be inferred from the Mott–Schottky curve in Figure 5b, in which the slope of the curves for MBB, Bi/BiOCl, MB, and BiOCl can visually be seen to be positive, reflecting that the materials are all n-type semiconductors. In addition, the extension of the Schottky curve and the intersection of the X-axis determine that the flat-band potentials of MBB, Bi/BiOCl, MB, and BiOCl are −1.23, −1.07, −1.01, and −0.99 V, respectively. The carrier concentration (Nd) in the material can also be derived from Equation (1):
N d = ( 2 e 0 ε ε 0 ) ( d 1 C 2 d V ) 1
The vacuum dielectric constant (ε0) of BiOCl is taken as 55 [47]. The carrier concentrations of MBB, Bi/BiOCl, MB, and BiOCl were calculated to be 3.41 × 1019, 2.66 × 1019, 1.53 × 1019, and 1.22 × 1019, respectively. MBB exhibits the highest carrier concentrations, showing that it has better conductivity to accelerate the charge flow. Finally, the conduction band positions of the four materials are derived from the flat-band potential, as calculated in Equation (2):
E (RHE) = E (Ag/AgCl) + 0.0591pH + 0.197
The calculated CB values for MBB, Bi/BiOCl, MB, and BiOCl are estimated to be −0.62, −0.46, −0.40, and −0.38 V, respectively. Thus, the calculated VB values are estimated to be 2.22, 2.57, 2.93, and 2.99 V based on the equation ECB = EVB − Eg [48,49]. Consequently, the involved samples’ energy band structures are determined and shown in Figure 5c.

3.2. Photocatalytic Activity and Mechanism

To further investigate the photocatalytic activity of MBB, photocatalytic degradation of norfloxacin antibiotic was performed to explore its photocatalytic ability. Figure 6a shows the transient time-varying spectra of MBB degradation of norfloxacin. It can be seen that the intensity of the absorbance curve gradually decreases with the increase of illumination time, which indicates that MBB has a significant degradation ability of norfloxacin. The maximum absorbance wavelength of norfloxacin is taken at 264 nm [50]. To investigate the difference in the degradation ability of MBB, Bi/BiOCl, MB, and BiOCl, we clearly show the comparison of the degradation performance of the four materials in Figure 6b. The degradation did not change significantly in the dark treatment. Still, in only 10 min under light, MBB degraded norfloxacin by 86.5%, which was 11.3% (1.15 times) higher than Bi/BiOCl, 56.2% (2.85 times) higher than MB, and 72.1% (6.01 times) higher than BiOCl, respectively. The reason for such a significant photocatalytic performance possessed by MBB is attributed to the interfacial engineering of metal Bi with BiOCl and the Mn-doped structure defect engineering.
Meanwhile, the rate curves of MBB, Bi/BiOCl, MB, and BiOCl conformed to the first-order kinetic equation (Figure 6c), and the rate constants of MBB, Bi/BiOCl, MB, and BiOCl (Figure 6d) were estimated to be 0.150, 0.102, 0.022, and 0.010 min−1, respectively. The rate constant of MBB was calculated to be 1.47 times, 6.82 times, and 15 times higher than those of Bi/BiOCl, MB, and BiOCl, respectively. This result more obviously shows that the degradation activity of MBB is higher, which once again proves that Mn doping and metal composite increase the utilization of light, accelerate charge separation and transfer, and improve the degradation rate. The unsaturated ‘d’ orbitals of Mn absorb a large number of electrons, thereby increasing the carrier concentration on the semiconductor surface to further oxidize the contaminant [37,51]. The in situ deposition of metallic Bi and the tightly coupled interface accelerates charge separation. Meanwhile, it further proves that MBB has great potential in environmental remediation, and such a dual technology combination of surface modification and internal structure doping also provides good ideas for the future development of photocatalysts.
In order to study the stability of the MBB photocatalyst, a long-performance test was executed on photocatalytic norfloxacin degradation. As can be seen in Figure S3a, no obvious decrease in degradation rate is observed after four cycles. Furthermore, XRD patterns of MBB hybrid before and after four cycles of photocatalytic degradation reaction were recorded and shown in Figure S3b. It can be clearly seen that the crystal structure of the MBB hybrid does not display a significant change after the long-performance test. These results demonstrate the good stability of the MBB photocatalysts.
To understand the effect of Mn doping concentration on the photocatalytic degradation rate of MBB, we investigated different doping concentrations of MBB (Mn:Bi = 1:20, 1:30, 1:40) for the photodegradation of norfloxacin antibiotics. As shown in Figure 7a, the degradation rates of the three samples were 80.7%, 86.5%, and 77.5%, respectively. It can be seen that the best photocatalytic degradation performance was obtained for MBB (1: 30) with different ratios of modulated Mn doping. Such peak changes were also reflected in the rate curves and rate constants. The rate curves in Figure 7b demonstrate that the compound first-order kinetic curve changes for all three samples. It can be estimated that the kinetic constants for MBB (Mn:Bi = 1:20, 1:30, 1:40) were 0.113, 0.150, and 0.102 min−1 (Figure 7c). In summary, MBB (1: 30) has the most suitable Mn doping concentration and metal Bi complex, which allows MBB to degrade norfloxacin at optimal performance. Usually, the photocatalytic degradation process occurs with three reactive species OH, O2, and h+ in the degradation system, as shown in Figure 7d. Each of the three substances is added to play an inhibitory role in the degradation process to identify the radicals that play a major and minor degradation role in the reaction system, with the capture of hydroxyl groups, hole, electrons, and superoxide by tert-butanol (t-BuOH), potassium iodide (KI), silver nitrate (AgNO3), and benzoquinone (BQ), respectively [52]. Figure 7d shows that adding AgNO3 and BQ has a stronger inhibitory effect on the reaction system, indicating that the magnitude of radical activity in the MBB degradation system is O2 > h+ > OH. The reaction system is characterized by O2 is the main active species; h+ and OH are the secondary active species.
As the mechanism diagram presented in Figure 8, MBB is a microsphere formed by a large number of nanosheets because of the surface free energy assembly. A Mn 3d-O 2p intermediate energy level formed near the top of the valence band of BiOCl in its energy level via Mn doping. The introduced Mn 3d-O 2p intermediate energy level narrowed the bandgap of BiOCl, optimizing light absorption and modulating the photoelectric properties of BiOCl. Moreover, abundant OVs caused by Mn-doping would serve as electron traps to promote the separation of photogenerated carriers [53,54]. The tightly coupled interface of metal Bi and the Mn doping within the structure together broadens the light absorption range, enhances the carrier lifetime, and accelerates the carrier transfer, which shows the degradation of highly toxic norfloxacin; sunlight irradiation can convert the antibiotics into CO2, H2O, and other non-toxic and harmless small molecules. Meanwhile, the energy level structure of MBB excites charge transfer from VB to CB under illumination, leaving holes at the VB position. The surface engineering of MBB is demonstrated by the interfacial coupling of metal Bi, which allows electrons from the CB of BiOCl to leap to the Bi surface to increase the electron concentration. Electrons concentrated on the surface of metal Bi can convert the O2 adsorbed on the catalyst surface into •O2 [37]. Furthermore, the internal engineering of MBB is achieved by Mn doping to change the energy level position. The holes in the VB position can convert OH in water to •OH. In summary, both tight interfacial coupling techniques and energy-level structure engineering are important in creating high-performance catalysts. Both in situ reduced metal Bi and internally doped Mn contribute to the high catalytic activity of MBB and build up the complete photocatalyst system.

4. Conclusions

In conclusion, the MBB microsphere structure was constructed by a simple solvothermal method, using in situ reduced metallic Bi to form a tight heterogeneous interface and reduce the charge separation resistance. Replacing high-valent Bi atoms with low-valent Mn to create Mn doping can build an intermediate energy level and accelerate electron transfer. The metal Bi and the charged defect generated by Mn doping together promote the utilization of the light by BiOCl. The specific surface area of MBB and active sites were also increased to promote the photocatalytic degradation of norfloxacin. The degradation rate of MBB was 1.15 times that of Bi/BiOCl, 2.85 times that of MB, and 6 times that of BiOCl. Furthermore, the rate constant of MBB was 1.47 times, 6.82 times, and 15 times higher than those of Bi/BiOCl, MB, and BiOCl, respectively. Such efficient photocatalytic activity can be ascribed to in situ anchored metal Bi and Mn doping with unsaturated ‘d’ orbitals. These modifications greatly promote the charge transfer within the MBB structure and prolong the carrier lifetime. Superoxide anions are the main active substances confirmed by the capture experiments. The combination of two structural engineering techniques can be used to create efficient photocatalysts to provide a broader range of ideas for future catalyst modification.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13172408/s1, Figure S1: SEM image of MB; Figure S2: Elemental mappings of Mn (a), Bi (b), O (c), and Cl (d) in MB; Figure S3: Cycling runs for the photocatalytic degradation of norfloxacin in MBB nanocomposite suspensions. (a); (b) XRD patterns of MBB before and after photocatalytic degradation reaction for four cycles (b); Table S1: The specific surface areas, pore diameters, and pore volumes for BiOCl, MB, Bi/BiOCl and MBB, respectively.

Author Contributions

Conceptualization, M.L.; formal analysis, S.W. and D.S.; funding acquisition, B.W.; investigation, S.W., D.S., L.L. and Z.L.; methodology, M.L.; project administration, W.Z.; resources, L.L.; software, B.W.; supervision, L.L.; writing—original draft, S.W. and D.S.; writing—review and editing, M.L. and W.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (52172206), the Basic Research Project of Science, Education, and Production Integration Pilot Project (2023PY015), the Heilongjiang Province Natural Science Foundation (LH2021B021), the Shandong Province Natural Science Foundation (ZR2022QD062).

Data Availability Statement

No additional data are available.

Acknowledgments

We gratefully acknowledge the support of the funding and the Development Plan of Youth Innovation Team in Colleges and Universities of Shandong Province.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Li, F.; Wang, P.; Li, M.; Zhang, T.; Li, Y.; Zhan, S. Efficient Photo-Fenton Reaction for Tetracycline and Antibiotic Resistant Bacteria Removal Using Hollow Fe-doped In2O3 Nanotubes: From Theoretical Research to Practical Application. Water Res. 2023, 240, 120088. [Google Scholar] [CrossRef] [PubMed]
  2. Xu, M.-L.; Jiang, X.-J.; Li, J.-R.; Wang, F.-J.; Li, K.; Cheng, X. Self-Assembly of a 3D Hollow BiOBr@Bi-MOF Heterostructure with Enhanced Photocatalytic Degradation of Dyes. ACS Appl. Mater. Interfaces 2021, 13, 56171–56180. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, B.; Feng, W.; Zhang, L.; Zhang, Y.; Huang, X.; Fang, Z.; Liu, P. In situ Construction of a Novel Bi/CdS Nanocomposite with Enhanced Visible Light Photocatalytic Performance. Appl. Catal. B Environ. 2017, 206, 510–519. [Google Scholar] [CrossRef]
  4. Gibson, Q.D.; Manning, T.D.; Zanella, M.; Zhao, T.; Murgatroyd, P.A.E.; Robertson, C.M.; Jones, L.A.H.; McBride, F.; Raval, R.; Cora, F.; et al. Modular Design via Multiple Anion Chemistry of the High Mobility van der Waals Semiconductor Bi4O4SeCl2. J. Am. Chem. Soc. 2020, 142, 847–856. [Google Scholar] [CrossRef] [PubMed]
  5. Shi, Y.; Li, J.; Mao, C.; Liu, S.; Wang, X.; Liu, X.; Zhao, S.; Liu, X.; Huang, Y.; Zhang, L. Van Der Waals Gap-rich BiOCl Atomic Layers Realizing Efficient, Pure-water CO2-to-CO Photocatalysis. Nat. Commun. 2021, 12, 5923. [Google Scholar] [CrossRef]
  6. Dong, X.; Cui, Z.; Shi, X.; Yan, P.; Wang, Z.; Co, A.C.; Dong, F. Insights into Dynamic Surface Bromide Sites in Bi4O5Br2 for Sustainable N2 Photofixation. Angew. Chem. Int. Ed. Engl. 2022, 61, 202200937. [Google Scholar] [CrossRef] [PubMed]
  7. Wu, L.; Su, F.; Liu, T.; Liu, G.Q.; Li, Y.; Ma, T.; Wang, Y.; Zhang, C.; Yang, Y.; Yu, S.H. Phosphorus-Doped Single-Crystalline Quaternary Sulfide Nanobelts Enable Efficient Visible-Light Photocatalytic Hydrogen Evolution. J. Am. Chem. Soc. 2022, 144, 20620–20629. [Google Scholar] [CrossRef]
  8. Li, H.; Chen, S.; Shang, H.; Wang, X.; Yang, Z.; Ai, Z.; Zhang, L. Surface Hydrogen Bond Network Spatially Confined BiOCl Oxygen Vacancy for Photocatalysis. Sci. Bull. 2020, 65, 1916–1923. [Google Scholar] [CrossRef]
  9. Wei, S.; Zhong, H.; Wang, H.; Song, Y.; Jia, C.; Anpo, M.; Wu, L. Oxygen Vacancy Enhanced Visible Light Photocatalytic Selective Oxidation of Benzylamine over Ultrathin Pd/BiOCl Nanosheets. Appl. Catal. B Environ. 2022, 305, 121032. [Google Scholar] [CrossRef]
  10. Zhong, Y.; Wu, C.; Chen, D.; Zhang, J.; Feng, Y.; Xu, K.; Hao, W.; Ding, H.; Lv, G.; Du, Y.; et al. Design of Lateral and Vertical Bi4O5I2/BiOCl Heterojunctions with Different Charge Migration Pathway for Efficient Photoredox Activity. Appl. Catal. B Environ. 2023, 329, 122554. [Google Scholar] [CrossRef]
  11. Chung, M.; Jin, K.; Zeng, J.S.; Ton, T.N.; Manthiram, K. Tuning Single-Atom Dopants on Manganese Oxide for Selective Electrocatalytic Cyclooctene Epoxidation. J. Am. Chem. Soc. 2022, 144, 17416–17422. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, X.; Fu, N.; Liu, J.C.; Yu, K.; Li, Z.; Xu, Z.; Liang, X.; Zhu, P.; Ye, C.; Zhou, A.; et al. Atomic Replacement of PtNi Nanoalloys within Zn-ZIF-8 for the Fabrication of a Multisite CO2 Reduction Electrocatalyst. J. Am. Chem. Soc. 2022, 144, 23223–23229. [Google Scholar] [CrossRef]
  13. Jiang, Z.; Song, S.; Zheng, X.; Liang, X.; Li, Z.; Gu, H.; Li, Z.; Wang, Y.; Liu, S.; Chen, W.; et al. Lattice Strain and Schottky Junction Dual Regulation Boosts Ultrafine Ruthenium Nanoparticles Anchored on a N-Modified Carbon Catalyst for H2 Production. J. Am. Chem. Soc. 2022, 144, 19619–19626. [Google Scholar] [CrossRef] [PubMed]
  14. Sun, Z.X.; Sun, K.; Gao, M.L.; Metin, Ö.; Jiang, H.L. Optimizing Pt Electronic States through Formation of a Schottky Junction on Non-reducible Metal–Organic Frameworks for Enhanced Photocatalysis. Angew. Chem. Int. Ed. Engl. 2022, 61, 202206108. [Google Scholar] [CrossRef] [PubMed]
  15. An, X.; Wei, T.; Ding, P.; Liu, L.M.; Xiong, L.; Tang, J.; Ma, J.; Wang, F.; Liu, H.; Qu, J. Sodium-Directed Photon-Induced Assembly Strategy for Preparing Multisite Catalysts with High Atomic Utilization Efficiency. J. Am. Chem. Soc. 2023, 145, 1759–1768. [Google Scholar] [CrossRef]
  16. Devasenathipathy, R.; Wang, J.Z.; Xiao, Y.H.; Rani, K.K.; Lin, J.D.; Zhang, Y.M.; Zhan, C.; Zhou, J.Z.; Wu, D.Y.; Tian, Z.Q. Plasmonic Photoelectrochemical Coupling Reactions of para-Aminobenzoic Acid on Nanostructured Gold Electrodes. J. Am. Chem. Soc. 2022, 144, 3821–3832. [Google Scholar] [CrossRef]
  17. Liu, X.; Cao, H.; Yin, J. Generation and Photocatalytic Activities of Bi@Bi2O3 Microspheres. Nano Res. 2011, 4, 470–482. [Google Scholar] [CrossRef]
  18. Zhao, Z.; Zhang, W.; Sun, Y.; Yu, J.; Zhang, Y.; Wang, H.; Dong, F.; Wu, Z. Bi Cocatalyst/Bi2MoO6 Microspheres Nanohybrid with SPR-Promoted Visible-Light Photocatalysis. J. Phys. Chem. C 2016, 120, 11889–11898. [Google Scholar] [CrossRef]
  19. Dong, F.; Zhao, Z.; Sun, Y.; Zhang, Y.; Yan, S.; Wu, Z. An Advanced Semimetal–Organic Bi Spheres–g-C3N4 Nanohybrid with SPR-Enhanced Visible-Light Photocatalytic Performance for NO Purification. Environ. Sci. Technol. 2015, 49, 12432–12440. [Google Scholar] [CrossRef]
  20. Chang, X.; Xie, L.; Sha, W.E.I.; Lu, K.; Qi, Q.; Dong, C.; Yan, X.; Gondal, M.A.; Rashid, S.G.; Dai, Q.I.; et al. Probing the Light Harvesting and Charge Rectification of Bismuth Nanoparticles Behind the Promoted Photoreactivity onto Bi/BiOCl Catalyst by (in-situ) Electron Microscopy. Appl. Catal. B Environ. 2017, 201, 495–502. [Google Scholar] [CrossRef]
  21. Cao, L.M.; Hu, C.G.; Li, H.H.; Huang, H.B.; Ding, L.W.; Zhang, J.; Wu, J.X.; Du, Z.Y.; He, C.T.; Chen, X.M. Molecule-Enhanced Electrocatalysis of Sustainable Oxygen Evolution Using Organoselenium Functionalized Metal-Organic Nanosheets. J. Am. Chem. Soc. 2023, 145, 1144–1154. [Google Scholar] [CrossRef] [PubMed]
  22. Li, L.; Bu, L.; Huang, B.; Wang, P.; Shen, C.; Bai, S.; Chan, T.S.; Shao, Q.; Hu, Z.; Huang, X. Compensating Electronic Effect Enables Fast Site-to-Site Electron Transfer over Ultrathin RuMn Nanosheet Branches toward Highly Electroactive and Stable Water Splitting. Adv. Mater. 2021, 33, 2105308. [Google Scholar] [CrossRef] [PubMed]
  23. Matsuzaki, K.; Tsunoda, N.; Kumagai, Y.; Tang, Y.; Nomura, K.; Oba, F.; Hosono, H. Hole-Doping to a Cu(I)-Based Semiconductor with an Isovalent Cation: Utilizing a Complex Defect as a Shallow Acceptor. J. Am. Chem. Soc. 2022, 144, 16572–16578. [Google Scholar] [CrossRef] [PubMed]
  24. Xiao, Y.; Fan, Z.; Nakabayashi, M.; Li, Q.; Zhou, L.; Wang, Q.; Li, C.; Shibata, N.; Domen, K.; Li, Y. Decoupling Light Absorption and Carrier Transport via Heterogeneous Doping in Ta3N5 Thin Film Photoanode. Nat. Commun. 2022, 13, 7769. [Google Scholar] [CrossRef]
  25. Zhang, K.; Zhao, J.; Hu, Q.; Yang, S.; Zhu, X.; Zhang, Y.; Huang, R.; Ma, Y.; Wang, Z.; Ouyang, Z.; et al. Room-Temperature Magnetic Field Effect on Excitonic Photoluminescence in Perovskite Nanocrystals. Adv. Mater. 2021, 33, 2008225. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Hou, T.; Xu, Q.; Wang, Q.; Bai, Y.; Yang, S.; Rao, D.; Wu, L.; Pan, H.; Chen, J.; et al. Dual-Metal Sites Boosting Polarization of Nitrogen Molecules for Efficient Nitrogen Photofixation. Adv. Sci. 2021, 8, 2100302. [Google Scholar] [CrossRef]
  27. Shahid, M.Z.; Mehmood, R.; Athar, M.; Hussain, J.; Wei, Y.; Khaliq, A. BiOCl Nanoplates Doped with Fe3+ Ions for the Visible-Light Degradation of Aqueous Pollutants. ACS Appl. Nano Mater. 2021, 4, 746–758. [Google Scholar] [CrossRef]
  28. Guan, C.; Hou, T.; Nie, W.; Zhang, Q.; Duan, L.; Zhao, X. Sn4+ Doping Enhanced inner Electric Field for Photocatalytic Performance Promotion of BiOCl Based Nanoflowers. Appl. Surf. Sci. 2022, 604, 154498. [Google Scholar] [CrossRef]
  29. Wang, Y.; Zhang, J.; Wang, K.; Pei, N.; Huang, X.; Chen, J.; Zhang, X.; Yu, Q. Enhanced Photocatalytic Degradation of Bisphenol A by Co-doped BiOCl Nanosheets under Visible Light Irradiation. Appl. Catal. B Environ. 2018, 221, 320–328. [Google Scholar] [CrossRef]
  30. Cao, J.; Li, J.; Chu, W.; Cen, W. Facile Synthesis of Mn-doped BiOCl for Metronidazole Photodegradation: Optimization, Degradation Pathway, and Mechanism. Chem. Eng. J. 2020, 400, 125813. [Google Scholar] [CrossRef]
  31. Pare, B.; Sarwan, B.; Jonnalagadda, S.B. Photocatalytic Mineralization Study of Malachite Green on the Surface of Mn-doped BiOCl Activated by Visible Light under Ambient Condition. Appl. Surf. Sci. 2011, 258, 247–253. [Google Scholar] [CrossRef]
  32. Lin, C.C.; Liu, T.R.; Lin, S.R.; Boopathi, K.M.; Chiang, C.H.; Tzeng, W.Y.; Chien, W.C.; Hsu, H.S.; Luo, C.W.; Tsai, H.Y.; et al. Spin-Polarized Photocatalytic CO2 Reduction of Mn-Doped Perovskite Nanoplates. J. Am. Chem. Soc. 2022, 144, 15718–15726. [Google Scholar] [CrossRef]
  33. Baimanov, D.; Wu, J.; Chu, R.; Cai, R.; Wang, B.; Cao, M.; Tao, Y.; Liu, J.; Guo, M.; Wang, J.; et al. Immunological Responses Induced by Blood Protein Coronas on Two-Dimensional MoS2 Nanosheets. ACS Nano 2020, 14, 5529–5542. [Google Scholar] [CrossRef] [PubMed]
  34. Cortecchia, D.; Mróz, W.; Neutzner, S.; Borzda, T.; Folpini, G.; Brescia, R.; Petrozza, A. Defect Engineering in 2D Perovskite by Mn(II) Doping for Light-Emitting Applications. Chem 2019, 5, 2146–2158. [Google Scholar] [CrossRef]
  35. Hou, F.; Yao, Q.; Zhou, C.S.; Ma, X.M.; Han, M.; Hao, Y.J.; Wu, X.; Zhang, Y.; Sun, H.; Liu, C.; et al. Te-Vacancy-Induced Surface Collapse and Reconstruction in Antiferromagnetic Topological Insulator MnBi2Te4. ACS Nano 2020, 14, 11262–11272. [Google Scholar] [CrossRef]
  36. Makuła, P.; Pacia, M.; Macyk, W. How To Correctly Determine the Band Gap Energy of Modified Semiconductor Photocatalysts Based on UV–Vis Spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [Google Scholar] [CrossRef]
  37. Wang, A.; Zheng, Z.; Wang, H.; Chen, Y.; Luo, C.; Liang, D.; Hu, B.; Qiu, R.; Yan, K. 3D Hierarchical H2-reduced Mn-doped CeO2 Microflowers Assembled from Nanotubes as a High-performance Fenton-like Photocatalyst for Tetracycline Antibiotics Degradation. Appl. Catal. B Environ. 2020, 277, 119171. [Google Scholar] [CrossRef]
  38. Huang, T.; Peng, H.; Wei, Q.; Peng, C.; Tian, Y.; Yao, S.; Han, X.; Zou, B. Magnetic Polaronic and Bipolaronic Excitons in Mn(II) doped (TDMP)PbBr4 and Their High Emission. Nano Energy 2022, 93, 106863. [Google Scholar] [CrossRef]
  39. Rao, R.R.; Corby, S.; Bucci, A.; Garcia-Tecedor, M.; Mesa, C.A.; Rossmeisl, J.; Gimenez, S.; Lloret-Fillol, J.; Stephens, I.E.L.; Durrant, J.R. Spectroelectrochemical Analysis of the Water Oxidation Mechanism on Doped Nickel Oxides. J. Am. Chem. Soc. 2022, 144, 7622–7633. [Google Scholar] [CrossRef]
  40. Pan, U.N.; Paudel, D.R.; Kumar Das, A.; Singh, T.I.; Kim, N.H.; Lee, J.H. Ni-nanoclusters Hybridized 1T–Mn–VTe2 Mesoporous Nanosheets for Ultra-low Potential Water Splitting. Appl. Catal. B Environ. 2022, 301, 120780. [Google Scholar] [CrossRef]
  41. Wu, Z.; Ye, F.; Liu, Q.; Pang, R.; Liu, Y.; Jiang, L.; Tang, Z.; Hu, L. Simultaneous Incorporation of V and Mn Element into Polyanionic NASICON for High Energy-Density and Long-Lifespan Zn-Ion Storage. Adv. Energy Mater. 2022, 12, 2200654. [Google Scholar] [CrossRef]
  42. Wang, W.; Favaro, M.; Chen, E.; Trotochaud, L.; Bluhm, H.; Choi, K.S.; van de Krol, R.; Starr, D.E.; Galli, G. Influence of Excess Charge on Water Adsorption on the BiVO4(010) Surface. J. Am. Chem. Soc. 2022, 144, 17173–17185. [Google Scholar] [CrossRef] [PubMed]
  43. Chen, P.; Zhang, P.; Kang, X.; Zheng, L.; Mo, G.; Wu, R.; Tai, J.; Han, B. Efficient Electrocatalytic Reduction of CO2 to Ethane over Nitrogen-Doped Fe2O3. J. Am. Chem. Soc. 2022, 144, 14769–14777. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, B.; Zou, J.; Shen, X.; Yang, Y.; Hu, G.; Li, W.; Peng, Z.; Banham, D.; Dong, A.; Zhao, D. Nanocrystal Supracrystal-derived Atomically Dispersed Mn-Fe Catalysts with Enhanced Oxygen Reduction Activity. Nano Energy 2019, 63, 103851. [Google Scholar] [CrossRef]
  45. Mi, Y.; Wen, L.; Wang, Z.; Cao, D.; Xu, R.; Fang, Y.; Zhou, Y.; Lei, Y. Fe(III) Modified BiOCl Ultrathin Nanosheet towards High-efficient Visible-light Photocatalyst. Nano Energy 2016, 30, 109–117. [Google Scholar] [CrossRef]
  46. Sun, Y.; Shin, H.; Wang, F.; Tian, B.; Chiang, C.W.; Liu, S.; Li, X.; Wang, Y.; Tang, L.; Goddard, W.A., 3rd; et al. Highly Selective Electrocatalytic Oxidation of Amines to Nitriles Assisted by Water Oxidation on Metal-Doped α-Ni(OH)2. J. Am. Chem. Soc. 2022, 144, 15185–15192. [Google Scholar] [CrossRef]
  47. Nalawade, Y.; Pepper, J.; Harvey, A.; Griffin, A.; Caffrey, D.; Kelly, A.G.; Coleman, J.N. All-Printed Dielectric Capacitors from High-Permittivity, Liquid-Exfoliated BiOCl Nanosheets. ACS Appl. Electron. Mater. 2020, 2, 3233–3241. [Google Scholar] [CrossRef]
  48. Li, J.; Pan, W.; Liu, Q.; Chen, Z.; Chen, Z.; Feng, X.; Chen, H. Interfacial Engineering of Bi19Br3S27 Nanowires Promotes Metallic Photocatalytic CO2 Reduction Activity under Near-Infrared Light Irradiation. J. Am. Chem. Soc. 2021, 143, 6551–6559. [Google Scholar] [CrossRef]
  49. Shiraishi, Y.; Hashimoto, M.; Chishiro, K.; Moriyama, K.; Tanaka, S.; Hirai, T. Photocatalytic Dinitrogen Fixation with Water on Bismuth Oxychloride in Chloride Solutions for Solar-to-Chemical Energy Conversion. J. Am. Chem. Soc. 2020, 142, 7574–7583. [Google Scholar] [CrossRef]
  50. Zhuang, Y.; Luan, J. Improved Photocatalytic Property of Peony-like InOOH for Degrading Norfloxacin. Chem. Eng. J. 2020, 382, 122770. [Google Scholar] [CrossRef]
  51. Wang, L.C.; Chang, L.C.; Chen, W.Q.; Chien, Y.H.; Chang, P.Y.; Pao, C.W.; Liu, Y.F.; Sheu, H.S.; Su, W.P.; Yeh, C.H.; et al. Atomically Dispersed Golds on Degradable Zero-valent Copper Nanocubes Augment Oxygen Driven Fenton-like Reaction for Effective Orthotopic Tumor Therapy. Nat. Commun. 2022, 13, 7772. [Google Scholar] [CrossRef] [PubMed]
  52. Xian, C.; He, J.; He, Y.; Nie, J.; Yuan, Z.; Sun, J.; Martens, W.N.; Qin, J.; Zhu, H.-Y.; Zhang, Z. High Nitrile Yields of Aerobic Ammoxidation of Alcohols Achieved by Generating •O2 and Br• Radicals over Iron-Modified TiO2 Photocatalysts. J. Am. Chem. Soc. 2022, 144, 23321–23331. [Google Scholar] [CrossRef] [PubMed]
  53. Jiang, J.; Zhang, L.; Li, H.; He, W.; Yin, J.J. Self-doping and Surface Plasmon Modification Induced Visible Light Photocatalysis of BiOCl. Nanoscale 2013, 5, 10573–10581. [Google Scholar] [CrossRef]
  54. Di, J.; Xia, J.; Yin, S.; Xu, H.; Xu, L.; Xu, Y.; He, M.; Li, H. One-pot Solvothermal Synthesis of Cu-modified BiOCl via a Cu-Containing Ionic Liquid and its Visible-light Photocatalytic Properties. RSC Adv. 2014, 4, 14281–14290. [Google Scholar] [CrossRef]
Figure 1. SEM images of BiOCl (a), Bi/BiOCl (b), MBB (c,d). Elemental mappings of Mn (e), Bi (f), O (g), and Cl (h) in MBB. EDX (i), TEM (j), and HRTEM (k) of MBB.
Figure 1. SEM images of BiOCl (a), Bi/BiOCl (b), MBB (c,d). Elemental mappings of Mn (e), Bi (f), O (g), and Cl (h) in MBB. EDX (i), TEM (j), and HRTEM (k) of MBB.
Nanomaterials 13 02408 g001
Figure 2. XRD patterns (a), UV-vis diffuse reflectance spectra (b), and the bandgap (c) of BiOCl, MB, Bi/BiOCl, and MBB, respectively. XRD patterns (d), UV-vis diffuse reflectance spectra (e), and the bandgap (f) of MBB (Mn:Bi = 1:20, 1:30, 1:40).
Figure 2. XRD patterns (a), UV-vis diffuse reflectance spectra (b), and the bandgap (c) of BiOCl, MB, Bi/BiOCl, and MBB, respectively. XRD patterns (d), UV-vis diffuse reflectance spectra (e), and the bandgap (f) of MBB (Mn:Bi = 1:20, 1:30, 1:40).
Nanomaterials 13 02408 g002
Figure 3. Raman spectra (a), EPR (b), and N2 adsorption–desorption isotherms (c) of BiOCl, MB, Bi/BiOCl, and MBB. Pore size distribution curves (d) of Bi/BiOCl and MBB.
Figure 3. Raman spectra (a), EPR (b), and N2 adsorption–desorption isotherms (c) of BiOCl, MB, Bi/BiOCl, and MBB. Pore size distribution curves (d) of Bi/BiOCl and MBB.
Nanomaterials 13 02408 g003
Figure 4. XPS spectra of Bi 4f (a), O 1s (b), Cl 2p (c), and Mn 2p (d). Full-scale XPS spectra (e) and valence band spectra (f) of MBB.
Figure 4. XPS spectra of Bi 4f (a), O 1s (b), Cl 2p (c), and Mn 2p (d). Full-scale XPS spectra (e) and valence band spectra (f) of MBB.
Nanomaterials 13 02408 g004
Figure 5. EIS (a), Mott–Schottky plots (b), and band energy diagram vs. reversible hydrogen electrode (c) of BiOCl, MB, Bi/BiOCl, and MBB, respectively.
Figure 5. EIS (a), Mott–Schottky plots (b), and band energy diagram vs. reversible hydrogen electrode (c) of BiOCl, MB, Bi/BiOCl, and MBB, respectively.
Nanomaterials 13 02408 g005
Figure 6. Time-varying absorption spectra of MBB for degradation norfloxacin. The dotted arrow have specified the meaning of the lines from top to bottom. They are in order as follows: 0 min, 2.5 min, 5 min, 7.5 min and 10 min. (a). Photocatalytic degradation curves of norfloxacin (b), rate curves (c), and rate constants (d) of BiOCl, MB, Bi/BiOCl, and MBB, respectively.
Figure 6. Time-varying absorption spectra of MBB for degradation norfloxacin. The dotted arrow have specified the meaning of the lines from top to bottom. They are in order as follows: 0 min, 2.5 min, 5 min, 7.5 min and 10 min. (a). Photocatalytic degradation curves of norfloxacin (b), rate curves (c), and rate constants (d) of BiOCl, MB, Bi/BiOCl, and MBB, respectively.
Nanomaterials 13 02408 g006
Figure 7. Histograms of photocatalytic degradation rates (a), rate curves (b), and rate constants (c) of MBB (Mn:Bi = 1:20, 1:30, 1:40). The inhibition of norfloxacin degradation by free radical scavengers in the MBB system (d).
Figure 7. Histograms of photocatalytic degradation rates (a), rate curves (b), and rate constants (c) of MBB (Mn:Bi = 1:20, 1:30, 1:40). The inhibition of norfloxacin degradation by free radical scavengers in the MBB system (d).
Nanomaterials 13 02408 g007
Figure 8. The proposed photocatalytic mechanism of MBB heterojunction assemblies.
Figure 8. The proposed photocatalytic mechanism of MBB heterojunction assemblies.
Nanomaterials 13 02408 g008
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, S.; Song, D.; Liao, L.; Wang, B.; Li, Z.; Li, M.; Zhou, W. Bi/Mn-Doped BiOCl Nanosheets Self-Assembled Microspheres toward Optimized Photocatalytic Performance. Nanomaterials 2023, 13, 2408. https://doi.org/10.3390/nano13172408

AMA Style

Wang S, Song D, Liao L, Wang B, Li Z, Li M, Zhou W. Bi/Mn-Doped BiOCl Nanosheets Self-Assembled Microspheres toward Optimized Photocatalytic Performance. Nanomaterials. 2023; 13(17):2408. https://doi.org/10.3390/nano13172408

Chicago/Turabian Style

Wang, Shijie, Dongxue Song, Lijun Liao, Bo Wang, Zhenzi Li, Mingxia Li, and Wei Zhou. 2023. "Bi/Mn-Doped BiOCl Nanosheets Self-Assembled Microspheres toward Optimized Photocatalytic Performance" Nanomaterials 13, no. 17: 2408. https://doi.org/10.3390/nano13172408

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