Next Article in Journal
Carbon Nanotube Coated Fibrous Tubes for Highly Stretchable Strain Sensors Having High Linearity
Next Article in Special Issue
Heteroatom-Doped Hierarchically Porous Biochar for Supercapacitor Application and Phenol Pollutant Remediation
Previous Article in Journal
Design of Waveguide Polarization Convertor Based on Asymmetric 1D Photonic Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Dispersion Cu2O QDs Decorated Bi2WO6 S-Scheme Heterojunction for Enhanced Photocatalytic Water Oxidation

1
Key Laboratory of Resources Conversion and Pollution Control of the State Ethnic Affairs Commission, College of Resources and Environmental Science, South-Central Minzu University, Wuhan 430074, China
2
School of Chemistry and Environmental Engineering, Wuhan Institute of Technology, Wuhan 430073, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(14), 2455; https://0-doi-org.brum.beds.ac.uk/10.3390/nano12142455
Submission received: 2 July 2022 / Revised: 13 July 2022 / Accepted: 15 July 2022 / Published: 18 July 2022
(This article belongs to the Special Issue Nanotechnology in Electrooxidation)

Abstract

:
Developing suitable photocatalysts for the oxygen evolution reaction (OER) is still a challenging issue for efficient water splitting due to the high requirements to create a significant impact on water splitting reaction kinetics. Herein, n-type Bi2WO6 with flower-like hierarchical structure and p-type Cu2O quantum dots (QDs) are coupled together to construct an efficient S-scheme heterojunction, which could enhance the migration efficiency of photogenerated charge carriers. The electrochemical properties are investigated to explore the transportation features and donor density of charge carriers in the S-scheme heterojunction system. Meanwhile, the as-prepared S-scheme heterojunction presents improved photocatalytic activity towards water oxidation in comparison with the sole Bi2WO6 and Cu2O QDs systems under simulated solar light irradiation. Moreover, the initial O2 evolution rate of the Cu2O QDs/Bi2WO6 heterojunction system is 2.3 and 9.7 fold that of sole Bi2WO6 and Cu2O QDs systems, respectively.

Graphical Abstract

1. Introduction

Sunlight provides an abundant renewable energy source to overcome the energy crisis that humans face in the future. Among all the strategies, solar energy conversion from sunlight into chemical energy has shown up as a sustainable and efficient route utilizing semiconductor photocatalysts [1,2]. As we know, water oxidation to dioxygen is a multi-electron transfer reaction in a photocatalytic water splitting process, which is a critical step and involves the difficult breaking of the O–H bond as well as the formation of an O–O bond [3,4]. Continuous efforts have been dedicated to the development of efficient water oxidation catalysts (WOCs), consisting of desirable semiconductor photocatalysts and cocatalysts with proper band structure and electrophilic ability, which could improve the light absorption capability and charge transportation with overall promoted photocatalytic performance for water oxidation [5,6].
Among various semiconductor photocatalysts, ternary metal oxide, n-type Bi2WO6, as one of the simplest members of the Aurivillius family, is comprised of accumulated layers of perovskite-like [WO4]2− octahedral sheets and [Bi2O2]2+ sheets [7,8,9]. Density functional theory (DFT) calculations show that the conduction band (CB) of Bi2WO6 is comprised of W 5d orbitals; the valence band (VB) mainly originates from hybridizing O 2p with Bi 6s orbitals, which not only enables the VB to be highly dispersed, but also facilitates the migration of photogenerated holes for specific oxidation reactions. In addition, the band gap of Bi2WO6 is about 2.8 eV, and the valence band edge is at +2.95 V vs. NHE (normal hydrogen electrode), which is high enough to trigger the water oxidation reaction for oxygen production. These unique properties reveal that Bi2WO6 can be utilized as a visible-light-driven photocatalyst for organic synthesis, CO2 reduction, and environmental remediation [10,11,12]. Nevertheless, similar to many semiconductors, the poor utilization efficiency of solar energy and high recombination rate of pure Bi2WO6 give rise to depressed photocatalytic activity and thereby cannot meet the rising demand of commercial applications [13,14,15].
Compared with mono-component photocatalysts, the hybrid heterojunction photocatalysts that hybridize at least two different functional catalysts into one system have attracted increasing attention in recent decades. In particular, the advanced Z- or S-scheme heterojunctions have been extensively investigated and reported; they synchronously realize efficient separation, transportation, and utilization of photoinduced charge with strong redox abilities by means of recombining weak electrons and holes at low potentials between the two semiconductors [16,17]. Therefore, to enhance the photocatalytic efficiency of Bi2WO6, it is feasible to couple Bi2WO6 with other cocatalysts for constructing an S-scheme heterojunction system [18,19,20]. For example, Liu et al. constructed Bi2WO6/CoAl-LDHs (layered double hydroxides) S-scheme heterojunction to obtain enhanced photo-Fenton-like catalytic performance, which profited from the synergistic effect of an internal electric field and S-scheme heterojunction [20]. Recently, quantum dots-modified semiconductor functional materials have received tremendous attention [21,22,23]. The quantum dots (QDs) can significantly increase the photon conversion efficiency by generating multiple excitons from a single photon owing to their unique quantum effect, but also easily match well with the band alignment of the host semiconductor [24]. Taking carbon QDs as an example, Kang et al. utilized carbon QDs to decorate Bi2WO6 for constructing the desirable band structure conditions induced by compensatory photo-electronic effects, thereby realizing overall water photo-splitting [25]. Moreover, the high specific surface area (SSA) supplies numerous active sites which are favorable for the adsorption of reactants and thus enhancing the observed photocatalytic activity. A major merit of these QDs decorated semiconductors is that more micro-heterojunction and a faster charge transfer process can be sustained due to the intimately contacted nature of the interface and the short charge-carrier transport paths. Among numerous semiconductors, Cu2O QDs have shown up as a good candidate for tailoring photo-response and promoting charge carrier migration properties because of the well-aligned overlapped band structures of Bi2WO6 and Cu2O [26]. In fact, Cu2O is widely applied as an effective co-catalyst in photocatalytic or electrocatalytic systems for a hydrogen evolution reaction (HER) owing to its high conduction band potential, exhibiting good photocatalytic H2 production activity [27,28].
In this study, we successfully decorated Cu2O QDs onto the surface of Bi2WO6 micro-flowers (MFs) with a uniform dispersion to form multiple S-scheme micro-heterojunctions for enhancing the efficiencies of solar light utilization and photogenerated charge migration. The incorporation of Cu2O QDs improved the adsorption ability of visible light and effectively facilitated the transportation of photoinduced charge carriers, and thus enhanced the photocatalytic activity for oxygen production under simulated solar light irradiation. This work suggests that the coupling of nanosized p-type Cu2O QDs and the three-dimensional Bi2WO6 MFs has a great potential for application in photocatalytic water oxidation.

2. Materials and Methods

2.1. Synthesis of Flower-Like Hierarchical Bi2WO6 MFs

In a typical procedure, 1.32 g of Na2WO4·2H2O was dissolved into 40 mL of purified water to form a transparent solution. Meanwhile, 1.96 g of Bi(NO3)3·5H2O was firstly mixed with 80 mL of HNO3 (0.3 M). After that, the Na2WO4 solution was dropped into the Bi(NO3)3 solution with vigorous magnetic stirring, and a white precipitate was formed quickly. Subsequently, 20 mL of NaOH solution (0.2 M) was added dropwise with stirring for 12 h. Finally, the mixture was transferred to a Teflon-lined autoclave and kept at 160 °C for 8 h. A light-yellow precipitate Bi2WO6 MFs was centrifuged, washed by purified water and dried in air at 60 °C.

2.2. Synthesis of Cu2O QDs/Bi2WO6 Heterojunction

Firstly, 0.025 g of hexadecyl trimethyl ammonium bromide (CTAB) was dissolved into 20 mL of purified water to form transparent solution. Then, 0.1 g of the as-prepared Bi2WO6 sample was added into the above CTAB solution with stirring for 30 min. Meanwhile, 0.008 g of copper acetate (Cu(Ac)2) and 0.016 g of ethylenediaminetetraacetic acid disodium (EDTA-Na) were dissolved into 5 mL of purified water. Subsequently, the Cu solution was mixed with the Bi2WO6 solution. Then, 10 mL of NaOH solution (0.05 M) was added dropwise into the mixed solution with stirring for 30 min. Afterwards, 10 mL of ascorbic acid (AA) solution (0.33 g) was dropped into the above solution with vigorous stirring for 1 h. The generated Cu2O/Bi2WO6 was washed with absolute ethanol and distilled water several times to remove the surfactant, and dried overnight in a vacuum oven. The final products were named 1.5 wt% Cu2O/Bi2WO6, 3wt% Cu2O/Bi2WO6, and 6 wt% Cu2O/Bi2WO6, where the 1.5, 3 and 6 wt% were the mass ratios of Cu2O to Bi2WO6 in the mixed solution according to the theoretical stoichiometric ratio of added copper and bismuth elements. For comparison, a control sample was prepared without the addition of Bi2WO6 and labeled as Cu2O.

2.3. Characterizations

X-ray diffraction (XRD) patterns of the prepared heterojunctions were performed using a Bruker D8 diffractometer (Billerica, MA, USA). The morphology and microstructure of the obtained catalysts were observed using a JSM5510LV (Tokyo, Japan) field emission scanning electron microscopy (SEM) and a JEOL 2100 (Tokyo, Japan) transmission electron microscopy (TEM). Raman spectra were recorded on an ISA dispersive Raman spectroscopy at 514 nm. Fourier transform infrared spectra (FTIR) were determined using a Bruker spectrometer (Billerica, MA, USA) with an ATR correction mode. X-ray photoelectron spectroscopy (XPS) was examined by a Thermo Escalab 250 instrument (Waltham, MA, USA) with Al-Kα radiation to determine the surface chemical species. UV–vis absorption spectra were conducted by a Cary 4000 UV-vis spectrometer (Waltham, MA, USA). Electron paramagnetic resonance (EPR) analyses were carried out using a Bruker EMS-plus instrument (Billerica, MA, USA) to detect the free radicals by using 5,5-dimethyl-1-pyrroline (DMPO) as a spin-trapping agent.

2.4. Photoelectrochemical Tests

Photoelectrochemical measurements were conducted using a CHI660E electrochemical workstation (Shanghai, China) with a three-electrode system in 0.05 M Na2SO4 electrolyte (20 mL, pH = 6.8). A catalyst deposited fluorine-doped tin oxide (FTO) electrode was served as a photoanode, while a Pt wire and a saturated calomel electrode (SCE) were applied as the counter electrode and reference electrode, respectively. For the photoanode preparation, 40 mg of the prepared photocatalysts were added into 2 mL of ethanol with 40 μL Nafion solution (5 wt%) and mixed homogeneously using a vortex oscillator. After that, the resulting mixture was dip-coated onto the prewashed FTO glass to obtain a film electrode with a controlled electrode area of 1 cm2. The solar light source (I0 = 100 mW cm−2) was simulated using a 200 W Xenon lamp coupled with an AM 1.5G filter. Electrochemical impedance spectroscopy (EIS) tests were measured at a scan frequency range of 0.1 to 100 kHz under a voltage amplitude of 10 mV and a potential bias of 0.298 V vs. SCE.

2.5. Photocatalytic Activities

The photocatalytic reactions were performed in a Teflon lining reactor under the simulated solar light. 0.05 g of samples were added into 200 mL of the solution with La2O3 (0.2 g) and AgNO3 (0.03 M). Before irradiation, the mixture was stirred for 30 min in the dark and then purged with N2 to removal O2. The concentration of O2 in the reactor was measured by using gas chromatograph (Tet, GC-2030,Tokyo, Japan) with a thermal conductivity at an interval of 30 min.

3. Results and Discussion

Figure 1a displays a possible formation procedure of Cu2O QDs/Bi2WO6 heterojunction through a facile hydrothermal and deposition route. Firstly, when the cationic surfactant CTAB is introduced, the CTAB can be adsorbed and anchored at the surface of Bi2WO6 MFs. The characteristic flower-like hierarchical Bi2WO6 with high SSA provides a structural framework for the uniform growth of nanoparticles on the sheets slowly with directed high-density. On the other hand, the EDTA and Cu(Ac)2 are mixed with the purified water to form a blue Cu complex. Subsequently, the mixture is added dropwise into the Bi2WO6/CTAB solution. As a result, the Cu complex is deposited on the surface of flower-like hierarchical Bi2WO6. With the addition of NaOH, Cu(II) ions from the Cu complex are slowly released to generate Cu(OH)2 nanoparticles. As expected, the negatively charged nanoparticles could be attracted and grafted by the positive CTAB to restrain the agglomeration effect. When the weak reductive AA is added, the formed Cu(OH)2 nanoparticles can be reduced to Cu2O QDs on the surface of Bi2WO6 MFs, which further maintains the stability of the nanosized Cu2O QDs without apparent aggregation. In Figure 1b, the XRD patterns of Bi2WO6 with different contents of Cu2O QDs are present. As displayed, the XRD pattern of the as-prepared Bi2WO6 is in good agreement with the standard diffraction pattern of orthorhombic Bi2WO6 (JCPDS No. 73-2020) [29], where the obvious peaks at 28.3°, 32.9°, 47.2°, 55.9°, 58.6°, 69.1°, 76.1°, 78.5°, and 87.7° can be indexed to the (113), (020), (220), (313), (226), (040), (333), (046), and (246) crystal planes, respectively. Moreover, the patterns for Cu2O/Bi2WO6 heterojunctions are similar to those of pure Bi2WO6, while no characteristic peaks belong to Cu2O are observed, which is ascribed to the low loading mass and high dispersion of Cu2O QDs in the Bi2WO6 matrix.
SEM images of the bare Bi2WO6 MFs are displayed in Figure 2a,b, where the uniform flower-like hierarchical Bi2WO6 with 2–3 μm diameter are observed clearly. It is found that the hierarchical structure of Bi2WO6 is assembled by ultrathin sheets with 40 nm of thickness, as present in Figure 2c,d, inferring high porosity and huge surface area, which benefits the exposure of more active sites.
After introducing the Cu2O QDs, as shown in Figure 3a,b, it is clearly observed that the size of the Bi2WO6 hierarchical flowers displays a negligible change, while the nanosheets comprised of the flowers are mechanically exfoliated and the surface of the flower-like hierarchical structure becomes smoother, which is possibly due to the vigorous stirring during the Cu2O QDs deposition process. Meanwhile, with the increasing of Cu initial amount, the Cu2O nanoparticles are observed and anchored at the surface of the hierarchical Bi2WO6 MFs. As displayed in Figure 3c, the 3 wt% Cu2O QDs are uniformly deposited on the surface of Bi2WO6 MFs, while once the amount of Cu(II) precursor reaches to 6 wt%, large Cu2O nanoparticles are detected in Figure 3d,e, which indicates that the excess Cu(II) precursor is harmful for the dispersion of Cu2O QDs and causes the aggregation.
TEM and HRTEM images of the Cu2O QDs/Bi2WO6 heterojunction are presented in Figure 4. The micro-size Bi2WO6 MFs with 2–3 μm diameter is observed, which is agreement with the results of SEM, as displayed in Figure 4a, where the large thickness of the sample hampers the penetration of electron beams, leading to the black area. In general, quantum dots are defined as semiconductor nanocrystals with particle sizes ranging from 1 to 20 nm, which possess unique electronic properties owing to the apparent quantum confinement effect. It can be clearly observed that the Cu2O nanoparticles with ~20 nm of diameter are uniformly dispersed at the surface of Bi2WO6 MFs in Figure 4b,c. Owing to the smaller size, the Cu2O QDs can easily anchor at the surface of micro-sized Bi2WO6 to form micro-heterojunctions, which shorten the charge-carrier transfer pathways through the intimately contacted interface. The clear lattice fringe of 0.307 nm ascribed to the (110) crystal facet of Cu2O is detected in Figure 4d. These results demonstrate the successful construction of heterojunctions between Bi2WO6 and Cu2O.
FTIR spectra of Bi2WO6 MFs, Cu2O, and Cu2O/Bi2WO6 are displayed in Figure 5a. The peaks at 818 and 703 cm−1 are attributed to the symmetric and asymmetric vibration of W–O, respectively [30]. The peaks centered at 1599, 2924 and 2845 cm−1 are due to the stretching vibration of O–H and C–H, respectively, which could be because of the usage of organic surfactants (CTAB, EDTA) during the synthesis procedure of the heterojunction system [31]. Besides, the characteristic peak of Cu2O is not found in the samples of Cu2O/Bi2WO6. To further investigate the composition of samples, Raman spectroscopy of the samples was performed, as shown in the Figure 5b. The characteristic peaks at 796 and 827 cm−1 can be ascribed to the antisymmetric and symmetric Ag stretch modes of the O–W–O band, respectively [32,33]. The peak at 714 cm−1 is associated with the antisymmetric bridging mode of the tungstate chain. In addition, the obvious vibration peak at 308 cm−1 is assigned to translational modes involving simultaneous motions of WO66− and Bi3+ [34]. For the pure Cu2O, the intense peaks at low frequencies of 213 and 260 cm−1 originate from the stretching vibration of Cu2O, which is consistent with the previous reports [35,36]. In the case of Cu2O/Bi2WO6, the characteristic peak at 308 cm−1 shifted to 296 cm−1, and the two peaks at 796 and 827 cm−1 became a broad peak at 809 cm−1 due to the cover of Cu2O on the surface of the Bi2WO6 MFs.
The XPS spectra were conducted to detect the chemical environment of elements in the catalyst, and all characteristic peaks were calibrated using C 1s (binding energy at 284.6 eV) as a reference. In Figure 6a, elements of W 4f, Bi 4f, O 1s, and Cu 2p were detected in the full survey spectrum of the 3 wt% Cu2O/Bi2WO6, demonstrating the coexistence of these elements in the sample. As presented in Figure 6b, two distinct peaks located at 159.8 and 165.1 eV are assigned to the characteristic peaks of Bi 4f7/2 and Bi 4f5/2 in the trivalent oxidation state, respectively. In the previous report, the binding energy of Bi 4f7/2 in Bi2WO6 MFs locates in the range of 158 to 159 eV while that for Bi2O3 appears between 159 and 160 eV. Therefore, the peak located at 159.8 eV could be assigned to Bi3+ in Bi2WO6 MFs [37,38]. In Figure 6c, the high resolution deconvoluted W 4f spectrum reveals two broad peaks at 38.2 and 36.0 eV corresponding to W 4f5/2 and W 4f7/2, respectively, suggesting the valence state of W element is +6 in the sample of Cu2O/Bi2WO6 heterojunction [39]. Moreover, as seen from Figure 6d, there are two obvious characteristic peaks at 953.3 and 933.5 eV, attributed to Cu 2p1/2 and Cu 2p3/2, respectively, revealing the feature of Cu+ in Cu2O [40,41]. In contrast, the CuO state generally has a main characteristic peak locates at a binding energy of higher than 933 eV and characteristic shake-up satellite peaks at around 937–945 eV [42,43,44,45]. The shake-up peaks are often detected at around 9–10 eV higher than the main peaks, which results from the vigorous photoelectrons synchronously interacting with a valence electron and then being excited to a higher binding energy level [46]. However, in Figure 6d, the peak belonging to Cu2+ at 933.7 eV with the shake-up peaks at 937–945 eV is not observed, revealing that the copper species in Cu2O/Bi2WO6 hybrids are mainly presented as Cu(I) [47,48,49].
UV–vis absorption spectra of various heterojunctions and the corresponding band gap energies calculated from the Tauc’s plots by (αhν) = A(hν − Eg)1/2 are presented in Figure 7, which reveals the sunlight response and absorption capability of Cu2O, Bi2WO6 MFs, and various Cu2O/Bi2WO6 hybrids. The absorption edge of Bi2WO6 MFs is about 460 nm, which suggests that the pure Bi2WO6 can only absorb UV and near-visible light. However, the absorption spectrum of Cu2O sharply rises at the beginning of 650 nm, displaying strong visible light response ability, which makes it a desirable candidate for utilization of solar energy. When depositing Cu2O QDs on the surface of Bi2WO6, the obtained Cu2O/Bi2WO6 hybrid system exhibits improved absorption ability for visible light, as displayed in Figure 7a. The corresponding band gap energies are calculated and displayed in Figure 7b, where the band gap energy of Cu2O/Bi2WO6 hybrids decreases with the introduction of Cu2O. Meanwhile, it is observed that the band gap of the 6 wt% Cu2O/Bi2WO6 hybrid is narrowed to 2.05 eV, which is obviously different from those of the 1.5 wt% and 3 wt% Cu2O/Bi2WO6 hybrids. This result suggests that the excess amount of Cu precursor did not result in the formation of Cu2O QDs but Cu2O microstructures on the surface of Bi2WO6. It demonstrates that the optimal amount of Cu precursor exists in the formation of QDs-MFs micro-heterojunction structure. On the other word, the excessive Cu precursor leads to the enhancement of sunlight response property.
To investigate the transportation behavior and efficiency of photoinduced charge carriers at the heterojunction interface, the photoelectrochemical properties of these samples were investigated. In Figure 8a, electrochemical impedance spectroscopies (EIS) of these samples in the manner of a Nyquist diagram were recorded in the dark and under light irradiation. In general, the radius of each semicircle is correlated to charge-transfer resistance (Rct) at the interface of electrode/electrolyte; a smaller semicircle implies a lower Rct value [50,51,52]. As shown in Figure 8a, Cu2O exhibits significantly smaller Rct under light irradiation (l) in comparison with being in darkness (d), indicating that the electrical resistance at the electrode/electrolyte interface is decreased due to the production of photoinduced charge carriers. In the case of the flower-like Bi2WO6 MFs, a larger semicircle is recorded, suggesting that the Bi2WO6 possesses poor electrochemical performance in charge-transfer process [53,54]. With the formation of the Cu2O QDs/Bi2WO6 heterojunction, the Rct of Bi2WO6 is intensively reduced, which apparently improves the photoelectrochemical property of Bi2WO6 and is favorable for the transportation of the photogenerated charge carriers.
To gain deeper insights into the characteristics of the prepared heterojunctions, flat band potential and carrier concentrations are deduced from the Mott–Schottky (M–S) curves [55,56]. The electrode potentials vs. SCE are converted to the reversible hydrogen electrode (RHE) potentials based on the following Nernst equation [57]:
VRHE = VSCE + 0.059 × pH + V0SCE
where VSCE is the experimental potential measured against the SCE, VRHE represents the converted potential vs. RHE, and V0SCE = 0.245 V at 25 °C. The Mott–Schottky (M–S) plots are depicted in Figure 8b–d, in which the flat band potentials at the electrode/electrolyte interface are calculated according to Equation (2) [36]:
1/C2 = (2/εrε0eNdA2)[(VVfb) − kT/e]c
where C is the specific capacity, εr and ε0 are the dielectric constant of the samples and the electric permittivity of vacuum (8.85 × 10−12 N−1 C2 m−2), respectively; Nd represents the carrier density of the catalysts, A is the efficient area of electrode, V and Vfb are the applied working potential and the flat band potential, respectively; k is the Boltzmann constant, T donates the absolute temperature, and e is the electron charge (1.602 × 10−19 C). In Figure 8b, a positive slope of M–S plot is observed, inferring a n-type semiconductor of Bi2WO6. In contrast, the negative slope of the M-S plot indicates a p-type behavior of Cu2O in Figure 8c, which is consistent with the previous reports [36,58]. Meanwhile, the flat band potentials of Cu2O and Bi2WO6 are calculated to be 0.74 and −0.18 V vs. RHE at pH = 6.8, respectively. In Figure 8d, an inverted “V-shape” curve is detected in the M–S plot of Cu2O/Bi2WO6, which is attributed to a characteristic curve of the p-n junction. It demonstrates that two distinct electronic behaviors (p- and n-type) are exhibited in the Cu2O/Bi2WO6 photoelectrode. Moreover, a slight shift of x intercept in Cu2O/Bi2WO6 occurs, implying the band realignment of Cu2O and Bi2WO6.
The photocatalytic water oxidization performances of these prepared samples are presented in Figure 9. As shown in Figure 9a, the Cu2O QDs/Bi2WO6 heterojunctions display significantly enhanced O2 evolution activities in comparison with the sole Bi2WO6 and Cu2O QDs under simulated solar light irradiation. The incorporation of Cu2O QDs improves the adsorption ability for visible light (Figure 7) as well as electrical conductivity of the prepared Cu2O QDs/Bi2WO6 heterojunctions (Figure 8a), thereby resulting in the enhancement of photocatalytic activity towards water oxidation under solar light irradiation, as the 1.5 wt% Cu2O QDs/Bi2WO6 heterojunction shown in Figure 9a. Meanwhile, the 3 wt% Cu2O QDs/Bi2WO6 heterojunction exhibits the best photocatalytic water oxidation performance, up to 50 μmol/L within 3 h, which is 2.1 and 6.1 times higher than that of pure Bi2WO6 and Cu2O QDs, respectively. Furthermore, the initial O2 evolution rate of the 3 wt% Cu2O QDs/Bi2WO6 heterojunction reaches 329 μmol h−1 g−1, which is 2.3 and 9.7 fold that of sole Bi2WO6 and Cu2O QDs system, respectively (Figure 9b), and is also superior to the reports in the literature (Table 1). However, excessive Cu(II) dosage (6wt%) is harmful for the dispersion of Cu2O QDs and causes the aggregation, leading to deteriorated catalytic performance. For the stability of the heterojunction system, as the recycling tests shown in Figure 9c, the photocatalytic performance of the 3 wt% Cu2O QDs/Bi2WO6 hybrid fades to some extent due to the excess deposition of Ag+ ions at the surface of heterojunction, but it still maintains good long-term stability and reuse potentiality. As a result, in Figure 9d, the 3 wt% Cu2O QDs/Bi2WO6 hybrid exhibits a sustainable photocatalytic O2 production capacity from water splitting.
For the 3 wt% Cu2O QDs/Bi2WO6 S-scheme heterojunction, the EPR results are displayed in Figure 10, where the signals attributed to the hydroxyl radicals (·OH) and superoxide radicals (·O2) are detected. As shown in Figure 10a, the characteristic four peaks caused by the existence of DMPO–OH∙ adduct are observed, apparently, which demonstrates that water molecular adsorbed on the surface of photocatalyst could efficiently react with the photoinduced holes and form ·OH [69]. On the other hand, in Figure 10b, the characteristic six peaks are clearly found, which is ascribed to the superoxide radical [70]. It is demonstrated that both of ·OH and ·O2 can be efficiently produced over the Cu2O QDs/Bi2WO6 hybrids under the solar light irradiation.
Based on the above results, two types of II or S-scheme heterojunction can be built between Cu2O QDs and Bi2WO6. Once the type II heterojunction is constructed, the trend of photoinduced charge carriers is for photogenerated holes at the VB of Bi2WO6 to migrate to the VB of Cu2O; correspondingly, the photoinduced electrons at the CB of Cu2O transfer to the CB of Bi2WO6. Consequently, photoinduced holes and electrons gather at the CB of Bi2WO5 and VB of Cu2O, respectively. Unfortunately, the VB potential of Cu2O is situated at +0.83 eV, which is quite low and makes it hard to guarantee enough oxidative potential to oxidize water and produce gaseous O2 [71]. Therefore, it is concluded that the Cu2O QDs/Bi2WO6 hybrids might tend to construct a novel S-scheme band structure, as presented in Figure 11a. The photoinduced electrons at the CB of Bi2WO6 are likely to quench the holes at the VB of Cu2O. Subsequently, the stronger reductive electrons at the CB of Cu2O and oxidative holes at the VB of Bi2WO6 are efficiently retained simultaneously. As described in Figure 11b, the separated photoinduced holes at the VB of Bi2WO6 react with the adsorbed H2O at the surface of hybridized system to generate O2, and the retained electrons at the CB of Cu2O are quenched by Ag+ ions. Therefore, the construction of an S-scheme heterojunction is conducive to inhibiting the recombination efficiency of the photoinduced charge carriers, giving rise to more photogenerated holes taking part in the photocatalytic reactions, thereby enhancing the photocatalytic efficiency towards O2 production.

4. Conclusions

In summary, we successfully prepared Cu2O QDs/Bi2WO6 heterojunctions by coupling hierarchical Bi2WO6 MFs with Cu2O QDs to construct efficient S-scheme heterojunctions, which could facilitate the migration of photoinduced charge carriers. The electrochemical properties are investigated to explore the transportation performance and donor density of charge carriers in the S-scheme heterojunction system. The results indicate that the synthesized S-scheme heterojunction shows improved photocatalytic activity for water oxidation compared with the sole Bi2WO6 and Cu2O QDs systems under simulated solar light illumination. The initial O2 evolution rate of the heterojunction system is 2.3 and 9.7 fold that of sole Bi2WO6 and Cu2O QDs system, respectively. Furthermore, it is evidently demonstrated that both of ·OH and ·O2 can be generated efficiently over the Cu2O QDs/Bi2WO6 heterojunction under the simulated solar light illumination.

Author Contributions

Conceptualization, D.T.; methodology, D.T. and J.K.; investigation, D.T., D.X., Z.L. and Y.Z.; data curation, D.T., D.X. and Y.Z.; visualization, D.T., D.X. and Z.L.; writing—original draft preparation, D.T. and J.K.; writing—review and editing, D.T., J.K. and L.L.; supervision, D.T.; project administration, D.T. and J.K.; funding acquisition, D.T., J.K. and J.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported financially by the Key R&D Program of Hubei Province (2020BCB062), the National Natural Science Foundation of China (21501138, 21908252), the Innovative Team Program of Natural Science Foundation of Hubei Province (2021CFA032), the Natural Science Foundation of Hubei Province (2019CFB556), the Science Research Foundation of Wuhan Institute of Technology (K201939), and the Fundamental Research Funds for the Central Universities of South-Central Minzu University (CXY22009, CZQ21014).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study is available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Samanta, B.; Morales-Garcia, A.; Illas, F.; Goga, N.; Anta, J.A.; Calero, S.; Bieberle-Hutter, A.; Libisch, F.; Munoz-Garcia, A.B.; Pavone, M.; et al. Challenges of modeling nanostructured materials for photocatalytic water splitting. Chem. Soc. Rev. 2022, 51, 3794–3818. [Google Scholar] [CrossRef]
  2. Ke, J.; He, F.; Wu, H.; Lyu, S.; Liu, J.; Yang, B.; Li, Z.; Zhang, Q.; Chen, J.; Lei, L.; et al. Nanocarbon-Enhanced 2D Photoelectrodes: A New Paradigm in Photoelectrochemical Water Splitting. Nano Micro Lett. 2020, 13, 24. [Google Scholar] [CrossRef]
  3. Lin, S.; Huang, H.; Ma, T.; Zhang, Y. Photocatalytic Oxygen Evolution from Water Splitting. Adv. Sci. 2020, 8, 2002458. [Google Scholar] [CrossRef]
  4. Materna, K.L.; Crabtree, R.H.; Brudvig, G.W. Anchoring groups for photocatalytic water oxidation on metal oxide surfaces. Chem. Soc. Rev. 2017, 46, 6099–6110. [Google Scholar] [CrossRef]
  5. Huang, S.; Yi, J.; Pan, Y.; Wang, C.; Jin, Y.; Song, Y.; Xu, Y.; Lam, C.H.; Li, H.; Xu, H. Steering Hole Transfer from the Light Absorber to Oxygen Evolution Sites for Photocatalytic Overall Water Splitting. Adv. Mater. Interfaces 2021, 8, 2101158. [Google Scholar] [CrossRef]
  6. Ong, W.-J.; Shak, K.P.Y. 2D/2D Heterostructured Photocatalysts: An Emerging Platform for Artificial Photosynthesis. Solar RRL 2020, 4, 2000132. [Google Scholar] [CrossRef]
  7. Xiong, J.; Song, P.; Di, J.; Li, H.; Liu, Z. Freestanding ultrathin bismuth-based materials for diversified photocatalytic applications. J. Mater. Chem. A 2019, 7, 25203–25226. [Google Scholar] [CrossRef]
  8. Chu, Y.; Miao, B.; Zheng, X.; Su, H. Fabrication of flower-globular Bi2WO6/BiOI@Ag3PO4 photocatalyst for the degradation of bisphenol A and cefepime under sunlight: Photoelectric properties, degradation performance, mechanism and biodegradability enhancement. Sep. Purif. Technol. 2021, 272, 118866. [Google Scholar] [CrossRef]
  9. Luo, S.; Ke, J.; Yuan, M.; Zhang, Q.; Xie, P.; Deng, L.; Wang, S. CuInS2 quantum dots embedded in Bi2WO6 nanoflowers for enhanced visible light photocatalytic removal of contaminants. Appl. Catal. B Environ. 2018, 221, 215–222. [Google Scholar] [CrossRef]
  10. Zeng, Y.; Yin, Q.; Liu, Z.; Dong, H. Attapulgite-interpenetrated g-C3N4/Bi2WO6 quantum-dots Z-scheme heterojunction for 2-mercaptobenzothiazole degradation with mechanism insight. Chem. Eng. J. 2022, 435, 134918. [Google Scholar] [CrossRef]
  11. Chen, P.; Du, T.; Jia, H.; Zhou, L.; Yue, Q.; Wang, H.; Wang, Y. A novel Bi2WO6/Si heterostructure photocatalyst with Fermi level shift in valence band realizes efficient reduction of CO2 under visible light. Appl. Surf. Sci. 2022, 585, 152665. [Google Scholar] [CrossRef]
  12. Cheng, C.; Chen, D.; Li, N.; Li, H.; Xu, Q.; He, J.; Lu, J. Bi2WO6 quantum dots with oxygen vacancies combined with g-C3N4 for NO removal. J. Colloid. Interface Sci. 2022, 609, 447–455. [Google Scholar] [CrossRef]
  13. Kovalevskiy, N.; Cherepanova, S.; Gerasimov, E.; Lyulyukin, M.; Solovyeva, M.; Prosvirin, I.; Kozlov, D.; Selishchev, D. Enhanced Photocatalytic Activity and Stability of Bi2WO6-TiO2-N Nanocomposites in the Oxidation of Volatile Pollutants. Nanomaterials 2022, 12, 359. [Google Scholar] [CrossRef]
  14. Li, H.; Liang, L.; Niu, X.; Zhang, D.; Fan, H.; Wang, K. Construction of a Bi2WO6/TiO2 heterojunction and its photocatalytic degradation performance. New J. Chem. 2022, 46, 8185–8194. [Google Scholar] [CrossRef]
  15. Li, L.; Yang, J.; Yang, L.; Fu, F.; Xu, H.; Fan, X. Photocatalytic performance of TiO2/Bi2WO6 photocatalysts with trace Fe3+ dopant for gaseous toluene decomposition. J. Environ. Chem. Eng. 2022, 10, 107708. [Google Scholar] [CrossRef]
  16. Zhang, L.; Zhang, J.; Yu, H.; Yu, J. Emerging S-Scheme Photocatalyst. Adv. Mater. 2022, 34, e2107668. [Google Scholar] [CrossRef]
  17. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-Scheme Heterojunction Photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  18. Su, M.; Sun, H.; Tian, Z.; Zhao, Z.; Li, P. Z-scheme 2D/2D WS2/Bi2WO6 heterostructures with enhanced photocatalytic performance. Appl. Catal. A Gen. 2022, 631, 118485. [Google Scholar] [CrossRef]
  19. Song, R.; Yao, J.; Yang, M.; Ye, Z.; Xie, Z.; Zeng, X. Active site regulated Z-scheme MIL-101(Fe)/Bi2WO6/Fe(III) with the synergy of hydrogen peroxide and visible-light-driven photo-Fenton degradation of organic contaminants. Nanoscale 2022, 14, 7055–7074. [Google Scholar] [CrossRef]
  20. Shao, B.; Liu, Z.; Tang, L.; Liang, Q.; He, Q.; Wu, T.; Pan, Y.; Cheng, M.; Liu, Y.; Tan, X.; et al. Construction of Bi2WO6/CoAl-LDHs S-scheme heterojunction with efficient photo-Fenton-like catalytic performance: Experimental and theoretical studies. Chemosphere 2022, 291, 133001. [Google Scholar] [CrossRef]
  21. You, Z.; Yue, X.; Zhang, D.; Fan, J.; Xiang, Q. Construction 0D/2D heterojunction by highly dispersed Ag2S quantum dots (QDs) loaded on the g-C3N4 nanosheets for photocatalytic hydrogen evolution. J. Colloid. Interface Sci. 2022, 607, 662–675. [Google Scholar] [CrossRef]
  22. Prusty, D.; Paramanik, L.; Parida, K. Recent Advances on Alloyed Quantum Dots for Photocatalytic Hydrogen Evolution: A Mini-Review. Energ. Fuels 2021, 35, 4670–4686. [Google Scholar] [CrossRef]
  23. Meena, B.; Subramanyam, P.; Suryakala, D.; Biju, V.; Subrahmanyam, C. Efficient solar water splitting using a CdS quantum dot decorated TiO2/Ag2Se photoanode. Int. J. Hydrog. Energ. 2021, 46, 34079–34088. [Google Scholar] [CrossRef]
  24. Kang, Z.; Lin, E.; Qin, N.; Wu, J.; Bao, D. Bismuth Vacancy-Mediated Quantum Dot Precipitation to Trigger Efficient Piezocatalytic Activity of Bi2WO6 Nanosheets. ACS Appl. Mater. Interfaces 2022, 14, 11375–11387. [Google Scholar] [CrossRef]
  25. Nie, H.; Wei, K.; Li, Y.; Liu, Y.; Zhao, Y.; Huang, H.; Shao, M.; Liu, Y.; Kang, Z. Carbon dots/Bi2WO6 composite with compensatory photo-electronic effect for overall water photo-splitting at normal pressure. Chin. Chem. Lett. 2021, 32, 2283–2286. [Google Scholar] [CrossRef]
  26. Shen, H.; Liu, G.; Zhao, Y.; Li, D.; Jiang, J.; Ding, J.; Mao, B.; Shen, H.; Kim, K.-S.; Shi, W. Artificial all-solid-state system by RGO bridged Cu2O and Bi2WO6 for Z-scheme H2 production and tetracycline degradation. Fuel 2020, 259, 116311. [Google Scholar] [CrossRef]
  27. Ma, B.; Bi, J.; Lv, J.; Kong, C.; Yan, P.; Zhao, X.; Zhang, X.; Yang, T.; Yang, Z. Inter-embedded Au-Cu2O heterostructure for the enhanced hydrogen production from water splitting under the visible light. Chem. Eng. J. 2021, 405, 126709. [Google Scholar] [CrossRef]
  28. Wei, T.; Zhu, Y.-N.; An, X.; Liu, L.-M.; Cao, X.; Liu, H.; Qu, J. Defect Modulation of Z-Scheme TiO2/Cu2O Photocatalysts for Durable Water Splitting. ACS Catal. 2019, 9, 8346–8354. [Google Scholar] [CrossRef]
  29. Zhong, S.; Yang, J.; Zhou, H.; Li, C.; Bai, L. Performance and mechanism of adsorption during synergistic photocatalytic degradation of tetracycline in water under visible (solar) irradiation. Sol. Energ. Mat. Sol. Cells 2022, 238, 111646. [Google Scholar] [CrossRef]
  30. Geng, Y.; Zhang, P.; Kuang, S. Fabrication and enhanced visible-light photocatalytic activities of BiVO4/Bi2WO6 composites. RSC Adv. 2014, 4, 46054–46059. [Google Scholar] [CrossRef]
  31. Liu, L.; Ding, L.; Liu, Y.; An, W.; Lin, S.; Liang, Y.; Cui, W. Enhanced visible light photocatalytic activity by Cu2O-coupled flower-like Bi2WO6 structures. Appl. Surf. Sci. 2016, 364, 505–515. [Google Scholar] [CrossRef]
  32. Li, Y.; Liu, J.; Huang, X.; Yu, J. Carbon-modified Bi2WO6 nanostructures with improved photocatalytic activity under visible light. Dalton Trans. 2010, 39, 3420–3425. [Google Scholar] [CrossRef]
  33. Zhou, H.; Wen, Z.; Liu, J.; Ke, J.; Duan, X.; Wang, S. Z-scheme plasmonic Ag decorated WO3/Bi2WO6 hybrids for enhanced photocatalytic abatement of chlorinated-VOCs under solar light irradiation. Appl. Catal. B Environ. 2019, 242, 76–84. [Google Scholar] [CrossRef]
  34. Ma, H.; Shen, J.; Shi, M.; Lu, X.; Li, Z.; Long, Y.; Li, N.; Ye, M. Significant enhanced performance for Rhodamine B, phenol and Cr(VI) removal by Bi2WO6 nancomposites via reduced graphene oxide modification. Appl. Catal. B Environ. 2012, 121, 198–205. [Google Scholar] [CrossRef]
  35. Liu, Y.; Ren, F.; Shen, S.; Fu, Y.; Chen, C.; Liu, C.; Xing, Z.; Liu, D.; Xiao, X.; Wu, W.; et al. Efficient enhancement of hydrogen production by Ag/Cu2O/ZnO tandem triple-junction photoelectrochemical cell. Appl. Phys. Lett. 2015, 106, 123901. [Google Scholar] [CrossRef]
  36. Dubale, A.A.; Su, W.-N.; Tamirat, A.G.; Pan, C.-J.; Aragaw, B.A.; Chen, H.-M.; Chen, C.-H.; Hwang, B.-J. The synergetic effect of graphene on Cu2O nanowire arrays as a highly efficient hydrogen evolution photocathode in water splitting. J. Mater. Chem. A 2014, 2, 18383–18397. [Google Scholar] [CrossRef]
  37. Xiong, S.; Bao, S.; Wang, W.; Hao, J.; Mao, Y.; Liu, P.; Huang, Y.; Duan, Z.; Lv, Y.; Ouyang, D. Surface oxygen vacancy and graphene quantum dots co-modified Bi2WO6 toward highly efficient photocatalytic reduction of CO2. Appl. Catal. B Environ. 2022, 305, 121026. [Google Scholar] [CrossRef]
  38. Zhao, F.; Gao, D.; Zhu, X.; Dong, Y.; Liu, X.; Li, H. Rational design of multifunctional C/N-doped ZnO/Bi2WO6 Z-scheme heterojunction for efficient photocatalytic degradation of antibiotics. Appl. Surf. Sci. 2022, 587, 152780. [Google Scholar] [CrossRef]
  39. Yang, M.; Xu, T.; Jin, X.; Shen, Q.; Sun, C. Oxygen vacancies enriched Bi2WO6 for enhanced decabromodiphenyl ether photodegradation via C-Br bond activation. Appl. Surf. Sci. 2022, 581, 152439. [Google Scholar] [CrossRef]
  40. Li, J.; Mei, Z.; Liu, L.; Liang, H.; Azarov, A.; Kuznetsov, A.; Liu, Y.; Ji, A.; Meng, Q.; Du, X. Probing defects in nitrogen-doped Cu2O. Sci. Rep. 2014, 4, 7240. [Google Scholar] [CrossRef] [Green Version]
  41. Zhang, Y.; Wang, Q.; Liu, D.; Wang, Q.; Li, T.; Wang, Z. Cu2O-BiOI isotype (p-p) heterojunction: Boosted visible-light-driven photoelectrochemical activity for non-enzymatic H2O2 sensing. Appl. Surf. Sci. 2020, 521, 146434. [Google Scholar] [CrossRef]
  42. Prabhakar Reddy, K.; Choi, H.; Kim, D.; Ryoo, R.; Park, J.Y. Cu oxide deposited on shape-controlled ceria nanocrystals for CO oxidation: Influence of interface-driven oxidation states on catalytic activity. Catal. Sci. Technol. 2021, 11, 6134–6142. [Google Scholar] [CrossRef]
  43. Yuan, J.; Zhang, J.-J.; Yang, M.-P.; Meng, W.-J.; Wang, H.; Lu, J.-X. CuO Nanoparticles Supported on TiO2 with High Efficiency for CO2 Electrochemical Reduction to Ethanol. Catalysts 2018, 8, 171. [Google Scholar] [CrossRef] [Green Version]
  44. Liu, P.; Hensen, E.J.M. Highly Efficient and Robust Au/MgCuCr2O4 Catalyst for Gas-Phase Oxidation of Ethanol to Acetaldehyde. J. Am. Chem. Soc. 2013, 135, 14032–14035. [Google Scholar] [CrossRef]
  45. Zhang, H.; Tan, H.-R.; Jaenicke, S.; Chuah, G.-K. Highly efficient and robust Cu catalyst for non-oxidative dehydrogenation of ethanol to acetaldehyde and hydrogen. J. Catal. 2020, 389, 19–28. [Google Scholar] [CrossRef]
  46. Cheng, C.; Zhang, C.; Gao, X.; Zhuang, Z.; Du, C.; Chen, W. 3D Network and 2D Paper of Reduced Graphene Oxide/Cu2O Composite for Electrochemical Sensing of Hydrogen Peroxide. Anal. Chem. 2018, 90, 1983–1991. [Google Scholar] [CrossRef] [Green Version]
  47. Zeng, Z.; Yan, Y.; Chen, J.; Zan, P.; Tian, Q.; Chen, P. Boosting the Photocatalytic Ability of Cu2O Nanowires for CO2 Conversion by MXene Quantum Dots. Adv. Funct. Mater. 2019, 29, 1806500. [Google Scholar] [CrossRef]
  48. Pang, X.; Bai, H.; Zhao, Y.; Qu, L.; Xu, D.; Ding, J.; Fan, W.; Shi, W. Photoelectrochemical detection of 4-nitrophenol by sensitive Ni/Cu2O photocathode. Electrochim. Acta 2021, 367, 137453. [Google Scholar] [CrossRef]
  49. Li, Z.; Xin, Y.; Zhang, Z. New Photocathodic Analysis Platform with Quasi-Core/Shell-Structured TiO2@Cu2O for Sensitive Detection of H2O2 Release from Living Cells. Anal. Chem. 2015, 87, 10491–10497. [Google Scholar] [CrossRef]
  50. Tang, D.; Luo, Y.; Lei, W.; Xiang, Q.; Ren, W.; Song, W.; Chen, K.; Sun, J. Hierarchical porous carbon materials derived from waste lentinus edodes by a hybrid hydrothermal and molten salt process for supercapacitor applications. Appl. Surf. Sci. 2018, 462, 862–871. [Google Scholar] [CrossRef]
  51. Lei, W.D.; Yang, B.K.; Sun, Y.J.; Xiao, L.W.; Tang, D.Y.; Chen, K.; Sun, J.; Ke, J.; Zhuang, Y.A. Self-sacrificial template synthesis of heteroatom doped porous biochar for enhanced electrochemical energy storage. J. Power Sources 2021, 488, 229455. [Google Scholar] [CrossRef]
  52. Luo, Z.; Liu, M.; Tang, D.; Xu, Y.; Ran, H.; He, J.; Chen, K.; Sun, J. High H2O2 selectivity and enhanced Fe2+ regeneration toward an effective electro-Fenton process based on a self-doped porous biochar cathode. Appl. Catal. B Environ. 2022, 315, 121523. [Google Scholar] [CrossRef]
  53. Ke, J.; Liu, J.; Sun, H.; Zhang, H.; Duan, X.; Liang, P.; Li, X.; Tade, M.O.; Liu, S.; Wang, S. Facile assembly of Bi2O3/Bi2S3/MoS2 n-p heterojunction with layered n-Bi2O3 and p-MoS2 for enhanced photocatalytic water oxidation and pollutant degradation. Appl. Catal. B Environ. 2017, 200, 47–55. [Google Scholar] [CrossRef]
  54. Li, J.; Xu, X.; Zhang, B.; Hou, W.; Lv, S.; Shi, Y. Controlled synthesis and fine-tuned interface of NiS nanoparticles/Bi2WO6 nanosheets heterogeneous as electrocatalyst for oxygen evolution reaction. Appl. Surf. Sci. 2020, 526, 146718. [Google Scholar] [CrossRef]
  55. Zhang, X.; Zhang, H.; Yu, J.; Wu, Z.; Zhou, Q. Preparation of flower-like Co3O4 QDs/Bi2WO6 p-n heterojunction photocatalyst and its degradation mechanism of efficient visible-light-driven photocatalytic tetracycline antibiotics. Appl. Surf. Sci. 2022, 585, 152547. [Google Scholar] [CrossRef]
  56. Ke, J.; Zhou, H.; Liu, J.; Zhang, Z.; Duan, X.; Wang, S. Enhanced light-driven water splitting by fast electron transfer in 2D/2D reduced graphene oxide/tungsten trioxide heterojunction with preferential facets. J. Colloid. Interface Sci. 2019, 555, 413–422. [Google Scholar] [CrossRef]
  57. Li, M.; Zhang, L.; Fan, X.; Zhou, Y.; Wu, M.; Shi, J. Highly selective CO2 photoreduction to CO over g-C3N4/Bi2WO6 composites under visible light. J. Mater. Chem. A 2015, 3, 5189–5196. [Google Scholar] [CrossRef]
  58. Yang, J.; Wang, X.; Zhao, X.; Dai, J.; Mo, S. Synthesis of Uniform Bi2WO6-Reduced Graphene Oxide Nanocomposites with Significantly Enhanced Photocatalytic Reduction Activity. J. Phys. Chem. C 2015, 119, 3068–3078. [Google Scholar] [CrossRef]
  59. Chen, J.; Tao, X.; Li, C.; Ma, Y.; Tao, L.; Zheng, D.; Zhu, J.; Li, H.; Li, R.; Yang, Q. Synthesis of bipyridine-based covalent organic frameworks for visible-light-driven photocatalytic water oxidation. Appl. Catal. B Environ. 2020, 262, 118271. [Google Scholar] [CrossRef]
  60. Yan, J.; Liu, J.; Ji, Y.; Batmunkh, M.; Li, D.; Liu, X.; Cao, X.; Li, Y.; Liu, S.; Ma, T. Surface Engineering to Reduce the Interfacial Resistance for Enhanced Photocatalytic Water Oxidation. ACS Catal. 2020, 10, 8742–8750. [Google Scholar] [CrossRef]
  61. Shah, J.H.; Malik, A.S.; Idris, A.M.; Rasheed, S.; Han, H.; Li, C. Intrinsic photocatalytic water oxidation activity of Mn-doped ferroelectric BiFeO3. Chin. J. Catal. 2021, 42, 945–952. [Google Scholar] [CrossRef]
  62. Zhu, M.; Sun, Z.; Fujitsuka, M.; Majima, T. Z-Scheme Photocatalytic Water Splitting on a 2D Heterostructure of Black Phosphorus/Bismuth Vanadate Using Visible Light. Angew. Chem. Int. Ed. 2018, 57, 2160–2164. [Google Scholar] [CrossRef] [PubMed]
  63. Djatoubai, E.; Khan, M.S.; Haq, S.U.; Guo, P.; Shen, S. Rational design of BiFeO3 nanostructures for efficient charge carrier transfer and consumption for photocatalytic water oxidation. J. Alloys Compd. 2022, 911, 164920. [Google Scholar] [CrossRef]
  64. Liu, S.; Pan, J.; Kong, W.; Li, X.; Zhang, J.; Zhang, X.; Liu, R.; Li, Y.; Zhao, Y.; Wang, D.; et al. Synergetic Nanoarchitectonics of Defects and Cocatalysts in Oxygen-Vacancy-Rich BiVO4/reduced graphene oxide Mott–Schottky Heterostructures for Photocatalytic Water Oxidation. ACS Appl. Mater. Inter. 2022, 14, 12180–12192. [Google Scholar] [CrossRef]
  65. Li, X.; Xiong, J.; Gao, X.; Ma, J.; Chen, Z.; Kang, B.; Liu, J.; Li, H.; Feng, Z.; Huang, J. Novel BP/BiOBr S-scheme nano-heterojunction for enhanced visible-light photocatalytic tetracycline removal and oxygen evolution activity. J. Hazard. Mater. 2020, 387, 121690. [Google Scholar] [CrossRef]
  66. Di, J.; Chen, C.; Zhu, C.; Ji, M.; Xia, J.; Yan, C.; Hao, W.; Li, S.; Li, H.; Liu, Z. Bismuth vacancy mediated single unit cell Bi2WO6 nanosheets for boosting photocatalytic oxygen evolution. Appl. Catal. B Environ. 2018, 238, 119–125. [Google Scholar] [CrossRef]
  67. Akbari, S.S.; Unal, U.; Karadas, F. Photocatalytic Water Oxidation with a CoFe Prussian Blue Analogue–Layered Niobate Hybrid Material. ACS Appl. Energ. Mater. 2021, 4, 12383–12390. [Google Scholar] [CrossRef]
  68. Qi, R.; Liu, J.; Yuan, H.; Yu, Y. Sulfur-Doped BiOCl with Enhanced Light Absorption and Photocatalytic Water Oxidation Activity. Nanomaterials 2021, 11, 2221. [Google Scholar] [CrossRef]
  69. Huang, G.; Li, S.; Liu, L.; Zhu, L.; Wang, Q. Ti3C2 MXene-modified Bi2WO6 nanoplates for efficient photodegradation of volatile organic compounds. Appl. Surf. Sci. 2020, 503, 144183. [Google Scholar] [CrossRef]
  70. Liu, Y.; Zhou, Y.; Tang, Q.; Li, Q.; Chen, S.; Sun, Z.; Wang, H. A direct Z-scheme Bi2WO6/NH2-UiO-66 nanocomposite as an efficient visible-light-driven photocatalyst for NO removal. RSC Adv. 2020, 10, 1757–1768. [Google Scholar] [CrossRef] [Green Version]
  71. Hu, W.; Wu, F.; Liu, W. Construction of S-scheme heterojunction by doping Bi2WO6 into Bi2O3 for efficiently enhanced visible-light photocatalytic performance. J. Mater. Sci. 2022, 57, 4265–4282. [Google Scholar] [CrossRef]
Figure 1. (a) Preparation illustrator and (b) XRD patterns of Cu2O QDs/Bi2WO6 heterojunction.
Figure 1. (a) Preparation illustrator and (b) XRD patterns of Cu2O QDs/Bi2WO6 heterojunction.
Nanomaterials 12 02455 g001
Figure 2. FESEM images of the pure flower-like Bi2WO6 samples with (a,b) wide scope and (c,d) higher resolutions.
Figure 2. FESEM images of the pure flower-like Bi2WO6 samples with (a,b) wide scope and (c,d) higher resolutions.
Nanomaterials 12 02455 g002
Figure 3. FESEM images of the Cu2O QDs/Bi2WO6 heterojunctions with different Cu amounts: (a) 1.5 wt%, (b,c) 3 wt%, and (d,e) 6 wt%.
Figure 3. FESEM images of the Cu2O QDs/Bi2WO6 heterojunctions with different Cu amounts: (a) 1.5 wt%, (b,c) 3 wt%, and (d,e) 6 wt%.
Nanomaterials 12 02455 g003
Figure 4. TEM (a,b) and HRTEM (c,d) images of the 3 wt% Cu2O QDs/Bi2WO6 sample.
Figure 4. TEM (a,b) and HRTEM (c,d) images of the 3 wt% Cu2O QDs/Bi2WO6 sample.
Nanomaterials 12 02455 g004
Figure 5. (a) FTIR spectra and (b) Raman spectra of the as-prepared Cu2O/Bi2WO6 samples: a. Bi2WO6, b. 1.5 wt% Cu2O/Bi2WO6, c. 3 wt% Cu2O/Bi2WO6, d. 6 wt% Cu2O/Bi2WO6, and e. Cu2O QDs.
Figure 5. (a) FTIR spectra and (b) Raman spectra of the as-prepared Cu2O/Bi2WO6 samples: a. Bi2WO6, b. 1.5 wt% Cu2O/Bi2WO6, c. 3 wt% Cu2O/Bi2WO6, d. 6 wt% Cu2O/Bi2WO6, and e. Cu2O QDs.
Nanomaterials 12 02455 g005
Figure 6. XPS spectra of the 3 wt% Cu2O/Bi2WO6: full spectrum survey (a), Bi 4f (b), W 4f (c), and Cu 2p (d).
Figure 6. XPS spectra of the 3 wt% Cu2O/Bi2WO6: full spectrum survey (a), Bi 4f (b), W 4f (c), and Cu 2p (d).
Nanomaterials 12 02455 g006
Figure 7. UV–visible absorption curves (a) and Tauc’s plots (b) of the prepared Bi2WO6 and different Cu2O/Bi2WO6 heterojunctions.
Figure 7. UV–visible absorption curves (a) and Tauc’s plots (b) of the prepared Bi2WO6 and different Cu2O/Bi2WO6 heterojunctions.
Nanomaterials 12 02455 g007
Figure 8. (a) Nyquist plots, Mott–Schottky curves of (b) Bi2WO6, (c) Cu2O, and (d) Cu2O/Bi2WO6.
Figure 8. (a) Nyquist plots, Mott–Schottky curves of (b) Bi2WO6, (c) Cu2O, and (d) Cu2O/Bi2WO6.
Nanomaterials 12 02455 g008
Figure 9. (a) Photocatalytic water oxidization performance and (b) initial O2 evolution rate of these as-synthesized Cu2O/Bi2WO6 heterojunctions. (c) Recycling curves and (d) stability test of the 3 wt% Cu2O QDs/Bi2WO6 heterojunction.
Figure 9. (a) Photocatalytic water oxidization performance and (b) initial O2 evolution rate of these as-synthesized Cu2O/Bi2WO6 heterojunctions. (c) Recycling curves and (d) stability test of the 3 wt% Cu2O QDs/Bi2WO6 heterojunction.
Nanomaterials 12 02455 g009
Figure 10. EPR spectra of DMPO-OH∙ (a) and DMPO-O2 (b) of the 3 wt% Cu2O QDs/Bi2WO6 heterojunction.
Figure 10. EPR spectra of DMPO-OH∙ (a) and DMPO-O2 (b) of the 3 wt% Cu2O QDs/Bi2WO6 heterojunction.
Nanomaterials 12 02455 g010
Figure 11. Proposed mechanisms of (a) construction of the novel S-scheme band structure and (b) photocatalytic water oxidization on the Cu2O QDs/Bi2WO6 heterojunction under simulated sunlight irradiation.
Figure 11. Proposed mechanisms of (a) construction of the novel S-scheme band structure and (b) photocatalytic water oxidization on the Cu2O QDs/Bi2WO6 heterojunction under simulated sunlight irradiation.
Nanomaterials 12 02455 g011
Table 1. Comparison of photocatalytic O2 evolution performance between the 3 wt% Cu2O/Bi2WO6 heterojunction and literature reports.
Table 1. Comparison of photocatalytic O2 evolution performance between the 3 wt% Cu2O/Bi2WO6 heterojunction and literature reports.
CatalystsLight SourceO2 Evolution Rate in First Hour (μmol h−1 g−1)StabilityRef.
BpCo-COF-1300 W Xe lamp (λ > 420 nm)1524 h[59]
IrOx-am@TiO2LED-405 lamp143.64 h[60]
Mn-BiFeO3300 W Xe lamp (λ > 420 nm)2556 h[61]
BP/BiVO4300 W Xe lamp (λ > 420 nm)1023 runs, 9 h[62]
BiFeO3300 W Xe lamp (λ > 420 nm)82.25 h[63]
Ov-BiVO4/rGO300 W Xe lamp (λ > 420 nm)1803 runs, 15 h[64]
Sol-10BP/BiOBr300 W Xe lamp (λ > 420 nm)89.54 runs, 16 h[65]
VBi-rich Bi2WO6300 W Xe lamp (λ > 420 nm)100.139 h[66]
KCa2Nb3O10/CoFe-PB300 W Xe lamp (λ > 420 nm)894 runs, 12 h[67]
S-BiOCl200 W Xe lamp (λ > 420 nm)1425 runs, 25 h[68]
3 wt% Cu2O/Bi2WO6200 W Xe lamp (λ > 420 nm)3294 runs, 12 hThis work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tang, D.; Xu, D.; Luo, Z.; Ke, J.; Zhou, Y.; Li, L.; Sun, J. Highly Dispersion Cu2O QDs Decorated Bi2WO6 S-Scheme Heterojunction for Enhanced Photocatalytic Water Oxidation. Nanomaterials 2022, 12, 2455. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12142455

AMA Style

Tang D, Xu D, Luo Z, Ke J, Zhou Y, Li L, Sun J. Highly Dispersion Cu2O QDs Decorated Bi2WO6 S-Scheme Heterojunction for Enhanced Photocatalytic Water Oxidation. Nanomaterials. 2022; 12(14):2455. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12142455

Chicago/Turabian Style

Tang, Diyong, Desheng Xu, Zhipeng Luo, Jun Ke, Yuan Zhou, Lizhong Li, and Jie Sun. 2022. "Highly Dispersion Cu2O QDs Decorated Bi2WO6 S-Scheme Heterojunction for Enhanced Photocatalytic Water Oxidation" Nanomaterials 12, no. 14: 2455. https://0-doi-org.brum.beds.ac.uk/10.3390/nano12142455

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