无机材料学报, 2022, 38 (6): 701, 网络出版: 2023-10-17  

单晶WO3/红磷S型异质结的构建及光催化活性研究

Construction and Photocatalytic Activity of Monoclinic Tungsten Oxide/Red Phosphorus Step-scheme Heterojunction
作者单位
1 1.新疆师范大学 化学化工学院, 乌鲁木齐 830054
2 2.新疆师范大学 新疆储能与光电催化材料重点实验室, 乌鲁木齐 830054
摘要
S型异质结被广泛应用于光解水产氢和解决环境污染问题。本研究通过简单的水热法制备了单晶WO3/水热处理后的红磷(HRP)复合材料。XPS和EPR等表征结果证实单晶WO3/HRP复合材料形成了S型异质结。5%WO3/HRP异质结复合物在可见光下展现出最佳的光催化活性, 在4 min内对罗丹明B(RhB)的降解率高达97.6%。此外, 制氢速率可以达到870.69 μmol·h-1·g-1, 是纯HRP的3.62倍。这可归功于单晶WO3和HRP之间形成紧密的S型异质结, 使其光生载流子快速分离并提高氧化还原能力。本研究制备的RP基光催化剂为解决日益增长的清洁新能源和饮用水需求提供了参考。
Abstract
S-scheme heterojunction has been extensively investigated for hydrogen evolution and environmental pollution issues. In this study, a monoclinic WO3/hydrothermally treated red phosphorus (HRP) S-scheme composite was prepared by hydrothermal method. XPS and EPR characterization confirmed that the monoclinic WO3/HRP composite formed S-scheme heterojunction. 5%WO3/HRP composite displayed the optimal photocatalytic activity under visible light irradiation, and its degradation rate of Rhodamine B (RhB) reached 97.6% after 4 min of visible light irradiation, while its hydrogen evolution rate reached 870.69 μmol·h-1·g-1 which was 3.62 times of that of pure HRP. This could be ascribed to the tight interfacial bonding between WO3 and HRP, and the formation of S-scheme heterostructure, enabling rapid separation of photogenerated carriers and therefore improving the strong redox capacity. This study provided a promising RP-based photocatalyst to meet the demand for clean energy and drinking water.

Red phosphorous (RP) is one of the most abundant, inexpensive, less toxic, and easily available materials[1-3]. It possesses strong visible light absorption, and has great potential for photocatalytic applications[4]. However, the photogenerated electron and hole are easily recombined in photocatalytic efficiency, which greatly limits its application because of RP large particle size and agglomeration[5]. RP is combined with other semiconductor material to construct heterojunction, which can effectively overcome the above disadvantages and improve the photocatalytic activity[6], such as SnO2/RP[7], Bi2O3/RP[8], Bi2O2CO3/RP[9], g-C3N4/RP[10] and Bi2Fe4O9/RP[11]. Therefore, it is necessary to find suitable semiconductor matching RP to build heterojunction.

Tungsten trioxide (WO3), as a typical n-type semiconductor with narrow bandgap energy (2.4-2.8 eV)[12], was widely used in the management of environmental remediation[13]. Although remarkable advances have been made, the narrow band gap of WO3 also causes fast recombination rate of photo-induced charges, leading to poor photocatalytic activity[14-15]. Compared with pure WO3 and RP, WO3/RP composite showed better hydrogen evolution, which could reduce the recombination of charge carriers[16]. However, WO3/RP composites via a low temperature calcination method led to aggregation and decrease of the active sites of RP.

In this study, the WO3/HRP S-scheme heterojunction was constructed through hydrothermal treatment. The photocatalytic performance was estimated by photocatalytic organic degradation and hydrogen evolution. Finally, the S-scheme heterojunction photocatalytic mechanism of WO3/HRP composites was directly evidenced by in situ X-ray photoelectron spectroscopy (XPS) and electron paramagnetic resonance (EPR) results.

1 Experimental

1.1 Preparation of WO3/HRP composite

Preparation of WO3: Na2WO4·2H2O (60.63 mmol·L-1) was dissolved in 50 mL of ultrapure water, then 0.9 mL of lactic acid was dropped into the solution and stirred for 15 min. Subsequently, pH of solution was adjusted to 1.0 using 6.0 mol·L-1 HCl. The solution was stirred for 30 min, and heated at 120 ℃ for 12 h, centrifuged. The obtained samples were washed with deionized water, dried at 60 ℃ for 5 h, and then calcined at 500 ℃ for 2 h.

Preparation of WO3/HRP composites: HRP was treated according to Ref.[17]. The composite with HRP was obtained based on WO3 mass fraction (3%, 5% and 7%), and the solution was hydrothermally treated at 150 ℃ for 4 h, centrifuged and washed for several times, and freeze-dried for 18 h to obtain WO3/HRP composite and marked as 3%WO3/HRP, 5%WO3/HRP, and 7%WO3/HRP, respectively.

1.2 Photocatalytic pollutant degradation

A 300 W Xe lamp (UV light was cutoff-filtered, 140 mW/cm2) was used, and RhB was used as target pollutant. 5 mg photocatalyst was added into 20 mL RhB solution (10 mg·L-1), and then stirred in dark for 30 min to ensure absorption-desorption equilibrium. During the irradiation process, the suspension was removed every 1 min for centrifugation and the absorbance of the residual RhB solution was determined by UV-Vis spectrophotometer at the maximum absorption wavelength (λmax=554 nm).

1.3 Photocatalytic hydrogen evolution

Photocatalytic hydrogen production was carried out by using a 300 W Xe lamp (UV light was cutoff-filtered, 140 mW/cm2). The generated hydrogen was in situ detected periodically using an online gas chromatography (GC-7900) with a thermal conductivity detector. 50 mg photocatalyst was dispersed in 80 mL solution containing Na2SO3 (0.35 mol·L-1) and Na2S (0.15 mol·L-1) as the sacrifice agent, and the suspension was magnetically stirred under visible light irradiation.

2 Results and discussion

2.1 Photocatalyst characterization

The crystal structures of the HRP, WO3 and 5%WO3/HRP composite were analyzed by XRD patterns (Fig. 1). For pure HRP, the characteristic diffraction peak at 2θ=14.9° corresponded to (102) crystal plane, consistent with the reported results in the literature[17]. The characteristic peaks of WO3 indexed at 2θ=22.8°, 24.1°, and 33.9° correspond to the planes (002), (020) and (200) facets (JCPDS 83-0951), according to the lattice parameters of monoclinic tungsten oxide[18-20]. For 5%WO3/HRP composite, the diffraction peaks of the HRP and WO3 were obviously present.

图 1.

Fig. 1. XRD patterns of HRP, WO3 and 5%WO3/HRP composite

下载图片 查看所有图片

The surface chemical compositions of all samples were characterized by FT-IR spectroscopy (Fig. S1). For pure HRP, the vibration absorption of the P-P-O, P-O and P=O bond located at 1008, 1178 and 1641 cm-1, which was consistent with the characteristic absorption peaks of RP[21]. For pure WO3, the absorption peaks at 956, 823 and 753 cm-1 were referred to W=O, O-W-O and W-O-W bonds[22]. The characteristic peaks of WO3 and HRP functional groups were presented in the FT-IR spectrum of 5%WO3/HRP, indicating the coexistence of WO3 and HRP in the composite.

The morphologies of as-prepared samples were examined by SEM. HRP removed impurities from the surface of commercial RP, and its surface existed micropore (Fig. S2(a, d)). The typical morphology of WO3 was rectangular nanoplates with a certain thickness (Fig. S2(b, e)). 5%WO3/HRP composite revealed a large number of loose particles, which facilitates the improved light adsorption, and also provides abundant spaces for the anchoring of other particles (Fig. S2(c, f)). Furthermore, the basic elements of W, O, and P were present in the WO3/HRP composite (Fig. S2(g, h). Thus, HRP and WO3 coexisted in the WO3/HRP composites.

The crystal structure and lattice orientation of each sample were further investigated through TEM. HRP surface was amorphous, and tiny folded, which was favorable to its surface adsorption properties (Fig. 2(a, b)). The WO3 nanostructure exhibited regular sheet-like morphology structure, which was consistent with the SEM results (Fig. S2(c)). The lattice size was 0.37 nm that matched the distance (020) planes of monoclinic WO3 structure, according to the lattice parameters of monoclinic WO3 (JCPDS 83-0951)[18]. TEM images of 5%WO3/HRP showed that there was an additional layer of homogeneous WO3 on the surface of HRP and closely connected to each other, avoiding aggregation of HRP and forming a heterogeneous structure that facilitates interfacial charge transfer (Fig. 2(e, f)).

图 2.

Fig. 2. TEM images of (a, b) HRP, (c, d) WO3 and (e, f) 5%WO3/HRP composite

下载图片 查看所有图片

XPS survey spectra were conducted to investigate the chemical compositions and bonding configurations of the samples[23]. C, O, P and W were present in the composite (Fig. S3(a)). The high-resolution P2p spectrum of HRP (Fig. 3(a)) was deconvoluted into two peaks at 129.82 and 130.61 eV, corresponding to P2p3/2 and P2p1/2, respectively, and the binding energy peak at 134.19 eV corresponded to P-O bond, indicating the presence of phosphorus oxides on the surface of HRP. While in 5%WO3/HRP composite, the P-O bond disappeared and the binding energies of the P2p3/2 and P2p1/2 were shifted towards higher binding energy, 130.12 and 130.97 eV, respectively, indicating that combination can prevent further oxidation of HRP. The W4f region of WO3 (Fig. 3(b)) was divided into two peaks at 35.52 and 36.33 eV, assigned to W4f5/2 and W4f7/2 orbitals of W5+. Another two peaks at 37.66 and 38.32 eV were corresponded to the W4f5/2 and W4f7/2 orbitals of W6+. W5+ was derived from the unsaturated W-O bonds on the surface of WO3. The presence of W5+ indicated the exfoliation of bulk WO3 into ultrathin WO3 nanosheets. The binding energy at 41.63 eV was associated with W5f3/2[24]. While, the binding energies of W4f5/2 and W4f7/2 in the 5%WO3/HRP composite shifted to low binding energy, and the ratio of W6+/W5+ increased, which facilitated its energy conversion. Thus, when HRP was closely contacted with WO3, the electrons could flow into WO3 from HRP though the interface. Therefore, the heterojunction was formed between WO3 and HRP, leading to strong electronic interaction.

图 3.

Fig. 3. (a) P2p and (b) W4f XPS spectra of samples

下载图片 查看所有图片

2.2 Photocatalytic performance

2.2.1 Photocatalytic degradation

The photodegradation performance of these catalysts was tested under visible light irradiation with RhB as the target pollutant (Fig. 4(a)). The equilibrium state of adsorption-desorption between the photocatalyst and the target pollutant was achieved after 30 min of reaction in the dark. Obviously, WO3/HRP composites showed better adsorption capacity for RhB. Under visible light condition, the photodegradation performance of WO3/HRP composites were higher than those of HRP and WO3, and 5%WO3/HRP composite was demonstrated the highest photocatalytic activity, and its removal rate reached 97.6% after 4 min. However, with further increasing of the WO3/HRP ratio, the photocatalytic performance decreased since the excess WO3 hinder the transfer of photogenerated carriers and reduce the exposure of active sites.

图 4.

Fig. 4. (a) Photodegradation curves, (b) rate curves, and (c) hydrogen production rates of HRP, WO3, 3%WO3/HRP, 5%WO3/HRP, and 7%WO3/HRP composites

下载图片 查看所有图片

The photodegradation reactions of RhB in the synthesized samples conformed to the pseudo-first-order reaction model (Fig. 4(b)):

ln(Ct/C0′) = −kt

where k is the rate consistent (min−1), C0′ is the concentration of pollutants in the equilibrium of adsorption- desorption, and Ct is the concentration of contaminants for the remaining time after irradiation. The k of HRP, WO3, 3%WO3/HRP, 5%WO3/HRP and 7%WO3/HRP were 0.17, 0.01, 0.68, 0.75 and 0.69 min-1, respectively. Among them, the k of 5%WO3/HRP composite was highest, which was 4.5 and 75 times of those of HRP and WO3. Therefore, the formation of heterojunction could enhance the photocatalytic activity. Furthermore, after five cycles of photocatalytic degradation of RhB, 5%WO3/HRP composite still had high photocatalytic activity (90.4%), showing good stability with practical application potential.

2.2.2 Photocatalytic hydrogen evolution

The photoreduction ability of HRP, WO3 and WO3/HRP composites were examined by photocatalytic water splitting into hydrogen. HRP possessed a relatively low photocatalytic activity with hydrogen production rate of 240.5 μmol·h-1·g-1, and the pure WO3 was almost absent, whereas a sharp increase in the rate of hydrogen production was observed in the WO3/HRP composite (Fig. 4(c)). Fascinatingly, the hydrogen evolution rate of 5%WO3/HRP composite was 870.69 μmol·h-1·g-1, which was 3.62 times of that of pure HRP. The apparent quantum efficiency (AQE) was calculated according to the following formula[26]:

AQE= (2·NA·M)/(/(hc))×100%

Where NA is the Avogadro’s constant (6.02×1023 mol-1), M is the average H2 generation rate (mol·s-1), E is power of lamp source, h is the Plank’s constant (6.626×10-34 J·s), λ is the excitation wavelength, and c is the speed of light (3.0×108 m·s-1).

AQE of 5%WO3/HRP composite was calculated to be 11.61%. Further increasing the amount of WO3 beyond 5% lead to a decrease in photocatalytic hydrogen evolution rate, which could originate from shielding of the light absorption by excess amount of WO3. Therefore, appropriate WO3 content had a remarkable effect on the activity enhancement of WO3/HRP composite.

The above RhB photodegradation and photocatalytic hydrogen evolution results exhibited that the 5%WO3/ HRP composite possessed superior photocatalytic activity. The introduction of appropriate WO3 content inhibits the overgrowth and agglomeration of HRP, and the heterojunction between WO3 and HRP promotes the separation of electrons and holes and accelerates the carrier migration.

2.3 Catalyst mechanism analysis

The optical absorption properties of prepared samples were investigated by UV-Vis DRS (Fig. 5(a)). HRP had significant absorption of visible light, meaning it was typical visible light photocatalytic material. For pure WO3, the absorption edge was located at around 446 nm. Compared with pure WO3 and HRP, the 5%WO3/HRP composite had a red-shifted absorption edge, which illustrated enhancement of absorption in both UV and IR regions. The UV-visible light absorption performance of 5%WO3/HRP composite combined the advantages of two materials, which successfully formed a heterojunction, and enhanced light-trapping ability. The band gap (Eg) values of the pure HRP and WO3 were obtained through transformation with Kubella-Munk function, and were 1.9 and 2.78 eV, respectively (Fig. 5(b)).

图 5.

Fig. 5. (a) UV-Vis DRS spectra of HRP, WO3 and 5%WO3/HRP composite, (b) Tauc plots of HRP and WO3, and (c) I-t curves of HRP, WO3 and 5%WO3/HRP composite

下载图片 查看所有图片

The generation, migration, and recombination processes of photocarriers were investigated by PL spectra, transient photocurrent (I-t) curves and electrochemical impedance spectroscopy (EIS) measurements[29]. The 5%WO3/HRP composite had lower PL signal, indicating lower recombination of photoinduced electrons and holes (Fig. S3(c)). In addition, the photocurrent signal intensity of the 5%WO3/HRP composite was 3.3 μA/cm2, which was 4~5 times of those of HRP (0.8 μA/cm2) and WO3 (0.7 μA/cm2), respectively (Fig. 5(c)). These results implied that the lifetime of the photoexcited electron-hole pairs was significantly prolonged due to the more efficient separation, after combining HRP with WO3.

The arc radius of 5%WO3/HRP composite was smaller than those of HRP and WO3, indicating the minimum charge impendence and the fastest reaction speed (Fig. 6(a)). This finding was consistent with the higher photocatalytic activity of the 5%WO3/HRP composite due to the tight interface between WO3 and HRP, which promoted carrier separation migration. Similarly, the equivalent circuit was analyzed to provide more intuitive understanding of the internal charge and surface charge- transfer mechanism of the catalyst during the reaction (insert in Fig. 6(a)). Here, R1 and R2 were denoted as the electrolyte solution and charge transfer resistances, which included the resistances of the photocatalyst, ITO substrate, electrolyte solution, and wire connections throughout the circuit. CPE1 was the constant phase element that represents the bilayer capacitance of the charge transfer, and W1 was the resistance with interfacial diffusion. R1 of HRP, WO3 and 5%WO3/HRP were 10.1, 11.71 and 12.53 Ω, respectively. Under the same concentration, R2 of HRP, WO3, HRP and 5%WO3/HRP were 672, 167 and 93.4 Ω, respectively. Obviously, the interfacial charge resistances of heterojunction composites were lower, indicating easier charge transfer. Therefore, the superior photocatalytic activity of 5%WO3/HRP composite contributed to the highest photoelectric conversion efficiency, smaller interfacial transfer impendence and enhanced visible light absorption ability. The carrier lifetime (τe) of samples are based on the EIS bode plots (Fig. S4(a, b)), proving the carrier transfer process, according to the equation[30]:

τe =1/(2πfmax)

图 6.

Fig. 6. (a) EIS spectra of HRP, WO3 and 5%WO3/HRP composite, (b) Mott-Schottky curves of HRP and WO3, and (c) EPR spectra of HRP, WO3 and 5%WO3/HRP composite

下载图片 查看所有图片

where fmax is the maximum frequency peak position, the calculated τe of the HRP and 5%WO3/HRP composite were 1.68 and 2.7 ms, respectively. The carrier lifetime of 5%WO3/HRP heterostructure composite was much longer than those of pure HRP, suggesting that the formation of the heterostructure between HRP and WO3 could greatly prolong the lifetime of the photo-generated electrons, and enhanced the photocatalytic activity.

2.4 Photogenerated carrier analysis

The energy band structures of samples were revealed by Mott-Schott (M-S) measurements (Fig. 6(b)). The flat potentials (Efb) of HRP and WO3 were -0.8 and -0.66 V (vs. Ag/AgCl), respectively, and then converted to the hydrogen standard electrode potential based on the formula:

Efb (vs. NHE) =Efb (pH 0, vs. Ag/AgCl) + EAgCl + 0.059·pH

Where EAgCl is 0.197 V, and pH of the electrolyte was 6.8. Here, Efb (vs. NHE) of HRP and WO3 were -0.2 and -0.06 V, respectively. The slopes of WO3 and HRP curves were positive, both samples were n-type semiconductors, and Efb is 0.1-0.3 V higher than its conduction band potential (ECB). Therefore, the ECB of HRP and WO3 were -0.4 and -0.26 V (vs. NHE). In accordance with the formula: Eg = EVB-ECB, the EVB were 1.5 and 2.52 V, respectively.

The hydroxyl radicals (·OH) was important species in the photocatalytic reactions, EPR spectra were presented in Fig. 6(c). HRP had relatively weak signal because of the weak oxidation potential of photogenerated holes, WO3 displayed moderately strong signal due to its more positive EVB, and 5%WO3/HRP composite had strong DMPO-·OH signal, so the photogenerated holes remained in WO3 and did not transfer to the VB of HRP. Because the potential of OH-/·OH pair was +2.38 V (vs. NHE)[31], which was positive than that of HRP (+1.5 V (vs. NHE)) and negative than that of the WO3 (+2.52 V (vs. NHE), inferring that the ·OH radical was produced by WO3. These results indicated that the photogenerated electrons and holes in 5%WO3/HRP composite were present in the CB of HRP and VB of WO3, respectively, and the charge transfer belonged to S-scheme heterojunction.

Based on above discussions, the S-scheme mechanism was proposed in Fig. 7. HRP is a reducing photocatalyst with smaller work function (5.61 eV)[17] and higher Fermi level. WO3 is an oxidizing photocatalyst with large work function (6.23 eV)[32] and lower Fermi level (Fig. 7(a)). When the WO3 photocatalysts was in close contact with HRP, electrons spontaneously transferred from HRP to WO3 until their Ef reached the same level. During the migration of electrons, the interface region of WO3 possesses a positive charge due to the loss of electrons, which leads to the formation of electron depletion layer and the upward bending of the energy band. While interface region near the HRP is negatively charged due to the gain of electrons, which leads to the formation of an electron accumulation layer and the downward bending of the band edge. As a result, an internal electric field is formed at the interface of the WO3/HRP heterojunction, impeding the continuous flow of electrons from HRP to WO3 (Fig. 7(b)). Under visible light irradiation, the electrons were excited from VB to CB of WO3 and HRP. Driven by the internal electric field, band bending and Coulomb interaction, the photogenerated electrons in the CB of WO3 spontaneously slid toward HRP, and recombined with the holes on the VB of HRP. However, the useful electrons and holes of strong redox ability could be retained (Fig. 7(c)). Therefore, the improved photocatalytic performance of the WO3/HRP composite was mainly ascribed to the formation of S-scheme heterojunctions, which contribute to the strong redox capacity for the degradation of organic water pollutants and hydrogen generation.

图 7.

Fig. 7. Photocatalytic mechanism of the WO3/HRP composite(a) Before contact; (b) After contact in darkness; (c) S-scheme transfer process of photogenerated carriers under visible light irradiation

下载图片 查看所有图片

3 Conclusions

In this study, the S-scheme WO3/HRP heterojunction photocatalysts were prepared by the hydrothermal method. The heterojunctions displayed significantly enhanced photocatalytic RhB degradation and hydrogen evolution. The 5%WO3/HRP heterojunction photocatalysts degraded 97.6% RhB within 4 min. Meanwhile, the photocatalytic hydrogen evolution was almost 3.62 times of pure HRP. The enhanced photocatalytic performance was attributed to the S-scheme heterojunction between WO3 and HRP, which effectively transferred the photo-generated carriers, and suppressed the recombination of electron-hole pairs. This work could provide new prospect for design and construction of novel heterojunction photocatalyst.

参考文献

[1] TAO J, ZHANG M, GAO X, et al.. Photocatalyst Co3O4/red phosphorus for efficient degradation of malachite green under visible light irradiation[J]. Materials Chemistry and Physics, 2020, 240: 122185.

[2] LIU E, QI L, CHEN J, et al. In situ fabrication of a 2D Ni2P/red phosphorus heterojunction for efficient photocatalytic H2 evolution[J]. Materials Research Bulletin, 2019, 115: 27.

[3] LIANG Z, DONG X, HAN Y, et al.. In-situ growth of 0D/2D Ni2P quantum dots/red phosphorus nanosheets with p-n heterojunction for efficient photocatalytic H2 evolution under visible light[J]. Applied Surface Science, 2019, 484: 293.

[4] WU C, JING L, DENG J, et al.. Elemental red phosphorus-based photocatalysts for environmental remediation: a review[J]. Chemosphere, 2021, 274: 129793.

[5] WANG Z, BAI Y, LI Y, et al.. Bi2O2CO3/red phosphorus S-scheme heterojunction for H2 evolution and Cr(VI) reduction[J]. Journal of Colloid and Interface Science, 2022, 609: 320.

[6] ZHU Y, LI J, DONG C, et al.. Red phosphorus decorated and doped TiO2 nanofibers for efficient photocatalytic hydrogen evolution from pure water[J]. Applied Catalysis B: Environmental, 2019, 255: 117764.

[7] AIHEMAITI X, WANG X, LI Y, et al.. Enhanced photocatalytic and antibacterial activities of S-scheme SnO2/Red phosphorus photocatalyst under visible light[J]. Chemosphere, 2022, 296: 134013.

[8] ZHU E, MA Y, DU H, et al.. Three-dimensional bismuth oxide/red phosphorus heterojunction composite with enhanced photoreduction activity[J]. Applied Surface Science, 2020, 528: 146932.

[9] WANG Z, BAI Y, LI Y, et al.. Bi2O2CO3/red phosphorus S-scheme heterojunction for H2 evolution and Cr(VI) reduction[J]. Journal of Colloid and Interface Science, 2022, 609: 320.

[10] WANG W, LI G, AN T, et al.. Photocatalytic hydrogen evolution and bacterial inactivation utilizing sonochemical-synthesized g-C3N4/red phosphorus hybrid nanosheets as a wide-spectral- responsive photocatalyst: the role of type I band alignment[J]. Applied Catalysis B: Environmental, 2018, 238: 126.

[11] ZHU E, ZHAO S, DU H, et al.. Construction of Bi2Fe4O9/red phosphorus heterojunction for rapid and efficient photo-reduction of Cr(VI)[J]. Journal of the American Ceramic Society, 2021, 104: 5411.

[12] ZHOU L, LI Y, YANG S, et al.. Preparation of novel 0D/2D Ag2WO4/WO3step-scheme heterojunction with effective interfacial charges transfer for photocatalytic contaminants degradation and mechanism insight[J]. Chemical Engineering Journal, 2021, 420: 130361.

[13] ZHANG J, LIU Z, LIU Z. Novel WO3/Sb2S3 heterojunction photocatalyst based on WO3 of different morphologies for enhanced efficiency in photoelectrochemical water splitting[J]. ACS Applied Materials & Interfaces, 2016, 8(15): 9684.

[14] LING Y, DAI Y. Direct Z-scheme hierarchical WO3/BiOBr with enhanced photocatalytic degradation performance under visible light[J]. Applied Surface Science, 2020, 509: 145201.

[15] SONG C, WANG X, ZHANG J, et al.. Enhanced performance of direct Z-scheme CuS-WO3 system towards photocatalytic decomposition of organic pollutants under visible light[J]. Applied Surface Science, 2017, 425: 788.

[16] JIANG S, CAO J, GUO M, et al.. Novel S-scheme WO3/RP composite with outstanding overall water splitting activity for H2 and O2 evolution under visible light[J]. Applied Surface Science, 2021, 558: 149882.

[17] TUERHONG M, CHEN P, MA Y, et al.. Bi2MoO6/red phosphorus heterojunction for reducing Cr(VI) and mitigating Escherichia coli infection[J]. Journal of Solid State Chemistry, 2022, 315: 123468.

[18] PAN T, CHEN D, XU W, et al.. Anionic polyacrylamide-assisted construction of thin 2D-2D WO3/g-C3N4 step-scheme heterojunction for enhanced tetracycline degradation under visible light irradiation[J]. Journal of Hazardous Materials, 2020, 393: 122366.

[19] ZHANG N, LI X, YE H, et al.. Oxide defect engineering enables to couple solar energy into oxygen activation[J]. Journal of the American Chemical Society, 2016, 138(28): 8928.

[20] XIAO T, TANG Z, YANG Y, et al. In situ construction of hierarchical WO3/g-C3N4 composite hollow microspheres as a Z-scheme photocatalyst for the degradation of antibiotics[J]. Applied Catalysis B: Environmental, 2018, 220: 417.

[21] GUO C, DU H, MA Y, et al.. Visible-light photocatalytic activity enhancement of red phosphorus dispersed on the exfoliated kaolin for pollutant degradation and hydrogen evolution[J]. Journal of Colloid and Interface Science, 2020, 585: 167.

[22] ZHOU J, AN X, TANG Q, et al.. Dual channel construction of WO3photocatalysts by solution plasma for the persulfate-enhanced photodegradation of bisphenol A.[J]. Applied Catalysis B: Environmental, 2020, 277: 119221.

[23] FENG C, TANG L, DENG Y, et al.. Synthesis of branched WO3@W18O49 homojunction with enhanced interfacial charge separation and full-spectrum photocatalytic performance[J]. Chemical Engineering Journal, 2020, 389: 124474.

[24] ZHANG K, JIN B, PARK C, et al.. Black phosphorene as a hole extraction layer boosting solar water splitting of oxygen evolution catalysts[J]. Nature Communications, 2019, 10(1): 2001.

[25] ZOU X, DONG Y, KE J, et al.. Cobalt monoxide/tungsten trioxide p-n heterojunction boosting charge separation for efficient visible- light-driven gaseous toluene degradation[J]. Chemical Engineering Journal, 2020, 400: 125919.

[26] YUE X, YI S, WANG R, et al.. Well-controlled SrTiO3@Mo2C core-shell nanofiber photocatalyst: boosted photo-generated charge carriers transportation and enhanced catalytic performance for water reduction[J]. Nano Energy, 2018, 47: 463.

[27] AIHEMAITI X, WANG X, WANG Z, et al.. Effective prevention of charge trapping in red phosphorus with nanosized CdS modification for superior photocatalysis[J]. Journal of Environmental Chemical Engineering, 2021, 9(6): 106479.

[28] DONG C, YANG Y, HU X, et al.. Self-cycled photo-Fenton-like system based on an artificial leaf with a solar-to-H2O2 conversion efficiency of 1.46%[J]. Nature Communications, 2022, 13(1): 4982.

[29] GE H, XU F, CHENG B, et al.. S-scheme heterojunction TiO2/CdS nanocomposite nanofiber as H2-production photocatalyst[J]. ChemCatChem, 2019, 11(24): 6301.

[30] LIU Y, PAN D, XIONG M, et al. In-situ fabrication SnO2/SnS2 heterostructure for boosting the photocatalytic degradation of pollutants[J]. Chinese Journal of Catalysis, 2020, 41(10): 1554.

[31] JIA Y, WANG Z, QIAO X, et al.. A synergistic effect between S-scheme heterojunction and Noble-metal free cocatalyst to promote the hydrogen evolution of ZnO/CdS/MoS2photocatalyst[J]. Chemical Engineering Journal, 2021, 424: 130368.

[32] FU J, XU Q, LOW J, et al. Ultrathin 2D/2D WO3/g-C3N4 step-scheme H2-production photocatalyst[J]. Applied Catalysis B: Environmental, 2019, 243: 556.

吐尔洪·木尼热, 赵红刚, 马玉花, 齐献慧, 李钰宸, 闫沉香, 李佳文, 陈平. 单晶WO3/红磷S型异质结的构建及光催化活性研究[J]. 无机材料学报, 2022, 38(6): 701. Munire TUERHONG, Honggang ZHAO, Yuhua MA, Xianhui QI, Yuchen LI, Chenxiang YAN, Jiawen LI, Ping CHEN. Construction and Photocatalytic Activity of Monoclinic Tungsten Oxide/Red Phosphorus Step-scheme Heterojunction[J]. Journal of Inorganic Materials, 2022, 38(6): 701.

引用该论文: TXT   |   EndNote

相关论文

加载中...

关于本站 Cookie 的使用提示

中国光学期刊网使用基于 cookie 的技术来更好地为您提供各项服务,点击此处了解我们的隐私策略。 如您需继续使用本网站,请您授权我们使用本地 cookie 来保存部分信息。
全站搜索
您最值得信赖的光电行业旗舰网络服务平台!