Jing Wana,
Aseel Shaker Al-Baldawyb,
Shanzhi Qua,
Jinshen Lana,
Xiaofang Yea,
Yuchen Feia,
Jingtian Zhaoa,
Ziyun Wanga,
Rongdun Honga,
Shengshi Guoa,
Shengli Huang*a,
Shuping Li*a and
Junyong Kanga
aEngineering Research Center of Micro-nano Optoelectronic Materials and Devices, Ministry of Education, Fujian Key Laboratory of Semiconductor Materials and Applications, CI Center for OSED, Department of Physics, Jiujiang Research Institute, Xiamen University, Xiamen 361005, China. E-mail: lsp@xmu.edu.cn; huangsl@xmu.edu.cn
bDepartment of Food Science, Faculty of Agriculture, University of Kufa, Kufa, Najaf 054003, Iraq
First published on 26th September 2022
A ternary semiconductor ZnO/MoS2/Ag2S nanorod array in an intimate core–shell structure was synthesized on glass substrates. The physicochemical properties and photocatalytical performance of the specimen were characterized and compared with single ZnO and binary ZnO/Ag2S and ZnO/MoS2 nanorod arrays. It is found that the coating layers depressed the band edge emission of the ZnO core, improved light absorption in the visible range, reduced charge transfer resistance, and increased photocatalytic activity. The ternary heterojunction nanorod array possessed full solar absorption with an efficiency of 52.88% for the degradation of methylene blue under visible light in 30 min. The efficiency was higher than other arrays and was 7.6 times that of the ZnO array. Theory analysis revealed that the coating layer brought different band alignment in the heterojunctions for efficient charge separation and conduction, which was beneficial for the photocatalytic performance.
The semiconductors with good photovoltaic properties, high energy conversion efficiency and excellent physicochemical properties have received extensive attention, among which ZnO is a promising candidate. The ZnO has a direct bandgap of 3.37 eV at room temperature, which owns a strong ability to absorb ultraviolet (UV) light. In addition, the characteristics of non-toxicity, strong stability and low cost make ZnO nanomaterials widely used in photocatalysts, light-emitting diodes, solar cells, gas sensors, and varistors. However, the wide bandgap limits its photocatalytic activity under visible light, and its high exciton binding energy (60 meV) leads to a high electron–hole recombination rate.4,5 The key to improve photocatalytic activity of ZnO is to achieve efficient spatial separation and rapid transfer of photogenerated charges to surface reaction sites. Constructing an intimate electric field at the catalyst interface is an effective way to achieve maximum carrier separation. The photogenerated charges are transferred under the driving fore of the electric field, which greatly accelerates the separation of electron–hole pairs.6–8 As visible light occupies approximately 43% of the incoming solar energy, developing efficient photocatalysts capable of harvesting visible sunlight is of great significance from a practical point of view.9 Combining ZnO with narrow bandgap semiconductors to form heterojunction nanostructures may broaden the light absorption range, separate electrons and holes efficiently, and present excellent functionalities.10,11 Furthermore, in contrast to a single or binary compound, the assembly of multi-junctions with continuous change of band gap, band alignment, absorption coefficient and crystal structure may prominently improve photoelectric conversion and photocatalytic performance.12 Therefore, the ZnO can be coupled with other semiconductors in special characteristics, such as metal sulphide-based semiconductors with unique physical and chemical properties.13 For example, MoS2 has a two-dimensional layered structure similar to graphene and transition from an indirect to direct bandgap semiconductor by changing the layer number, in which the value of bandgap (Eg) can be turned from 1.29 to 1.90 eV.14 While Ag2S is a narrow bandgap semiconductor (Eg ≈ 1.0 eV) with broad and strong light absorption in the entire solar spectrum, and has attracted great attention as a potential photocatalyst or photosensitizer.15,16 The alliance of ZnO nanomaterials with the metal sulphides may bring special properties and functionalities.
With above consideration, ternary semiconductor heterojunction ZnO/MoS2/Ag2S nanorod array was synthesized on the conductive fluorine-doped tin oxide (FTO) glass substrates. The physicochemical properties and photocatalytic performance of the specimen were characterized and compared with single semiconductor ZnO and binary semiconductor heterojunction ZnO/MoS2 and ZnO/Ag2S nanorod arrays. It is found that the nanorods were well aligned vertically on the FTO substrate. The MoS2 nanosheets and Ag2S quantum dots adhered tightly on the ZnO nanorod surface, forming an intimate core–shell structure with a strong engineered chemical state for the binary and ternary semiconductor heterojunction arrays. The core–shell nanorod arrays presented weak bandedge emission of the ZnO, strong light absorption in the visible range, low electric resistance, and high photocatalysis under visible light, especially for the ZnO/MoS2/Ag2S arrays. The special properties were well analyzed by the band alignment.
Core-shell structure of the ZnO/MoS2/Ag2S nanorods was investigated by TEM, as displayed in Fig. 2(a) and (b). The intensity profile displays an obvious variation of light contrast of the edge area and dark contrast of the core area (see Fig. 2(a)), confirming formation of the core–shell structure. The shell layer comprises a thin film and lots of small particles in the outside. The film adheres tightly to the core rod without any interspace. The thickness of the film is 47.5 nm, while the diameter of the particles varies in a range of 30–60 nm. The HRTEM image in Fig. 2(b) presents distinct crystal lattice of the shell. The interplanar crystal spacing of the film varies with 0.35 and 0.62 nm, which corresponds to the (004) and (002) planes of the hexagonal MoS2 (JCPDS no. 37-1492). The lattice spacing of the particle is 0.20 nm, corresponding to the (200) plane of the monoclinic Ag2S (JCPDS no. 14-0072). Therefore, the ingredients of the specimen as tested by the HRTEM are in good coincident with the EDS in Fig. 1(c4). Fig. 2(c) displays SAED pattern of the shell (the area marked by a yellow circle in the inset), showing polycrystalline diffraction rings of MoS2 nanosheets and single crystal spots of Ag2S quantum dots. The SAED pattern in Fig. 2(d) exhibits diffraction lines of the core (the red circle area in the inset), suggesting single crystal of the ZnO. The overall results confirm the presence of a core–shell structure of the heterogeneous ZnO/MoS2/Ag2S nanocomposite instead of a physical mixture. The presence of this semiconductor heterojunction might be better for the separation and transferring of photoinduced charges for photocatalytic performance through interfacial bonding.
Elemental composition and chemical state of the specimens were characterized by XPS. Fig. 3(a) presents survey spectra, demonstrating the presence of corresponding elements in the specimens, in addition to carbon. The carbon content may come from the residual carbon in the hydrothermal solution or the adventitious carbon in the XPS instrument. Fig. 3(b) displays Zn 2p spectra of the specimens. For ZnO nanorods, the two peaks locating at 1022.00 and 1045.00 eV correspond to Zn 2p1/2 and Zn 2p3/2, respectively.20,21 The addition of Ag2S quantum dots results in a red shift of the peaks to 1021.75 and 1044.81 eV, while the coating of MoS2 sheets induces a blue shift of the peaks to 1022.70 and 1045.70 eV. The peak position of the ZnO/MoS2/Ag2S array is the same as that of the ZnO/MoS2 array, which may ascribe to the larger content of MoS2 than Ag2S. The engineered chemical state of the Zn 2p suggests the strong binding and interaction between the shell and the core. Fig. 3(c) shows XPS spectra of the O 1s. The resonant band can be deconvolved into two peaks, which locate at 530.40 and 531.50 eV for the ZnO array and can be attributed to the O–Zn bonds and oxygen vacancies (or C–O/CO bonds).22 After coating the quantum dots and/or nanosheets, the binding energy of the bond and vacancy takes a little change, while their relative intensity (Ibond/Ivacancy) decreases successively from ZnO/Ag2S to ZnO/MoS2 and to ZnO/MoS2/Ag2S, suggesting increasing amount of vacancy or carbon substance. Fig. 3(d) exhibits S 2p spectra of the arrays. Two peaks at 160.40 and 161.63 eV are observed in the ZnO/Ag2S array, which are in good agreement with S 2p3/2 and S 2p1/2 of the Ag2S quantum dots.23 The profile of the ZnO/MoS2 array can be fitted by two sets of doublets, in which a couple at lower energies (163.28 eV for S 2p1/2 and 162.08 eV for S 2p3/2) correspond to 1T-phase MoS2, and another couple at higher energies (164.14 eV for S 2p1/2 and 162.40 eV for S 2p3/2) correspond to 2H-phase MoS2.18,24 For ZnO/MoS2/Ag2S array, the resonant peaks for the Ag2S and MoS2 in 1T and 2H phases are also observed. However, the two peaks of the Ag2S quantum dots blue shift to 161.31 and 162.12 eV, respectively, and the two peaks of the 1T phase blue shift to 163.43 and 162.27 eV, respectively, while the two peaks of the 2H phase remain unchanged. The band intensity of the 1T phase is much stronger than the 2H phase, suggesting that the content of the former phase is much larger than the later phase in the as-prepared ZnO/MoS2 and ZnO/MoS2/Ag2S arrays. The Mo 3d spectra also present a couple of doublets in Fig. 3(e). For ZnO/MoS2 array, the couple of doublet peaks at 232.40 and 229.28 eV match with Mo 3d3/2 and 3d5/2of the 1T-phase MoS2, and another couple of doublet peaks at higher energies of 233.30 and 230.16 eV match with Mo 3d3/2 and 3d5/2 of the 2H-phase MoS2.18 Therefore, this observation reveals that the MoS2 sheets in the ZnO/MoS2 and ZnO/MoS2/Ag2S arrays are composed of mixed 1T and 2H phases, which is coincident with that found in Fig. 3(d). Compared with the ZnO/MoS2 array, the 1T phase is roughly blue-shifted by 0.2 eV and the 2H phase is stable for the ZnO/MoS2/Ag2S array. In addition, a peak at high binding energy of 226.50 eV can be assigned to S 2s, and a peak at 235.70 eV can be attributed to Mo6+ of MoO3 owing to small portion of Mo oxidation.18 The Ag 3d spectrum of the ZnO/Ag2S array in Fig. 3(f) shows two peaks at 367.33 eV and 373.33 eV, which can be ascribed to Ag 3d5/2 and Ag 3d3/2, respectively.23 The two peaks blue-shift to 368.23 and 374.25 eV for the ZnO/MoS2/Ag2S array, which suggests increasing binding energy of Ag2S quantum dots for the modulation of the MoS2 sheets, and the engineered chemical state for the strong interaction in the shell layer.
Fig. 3 XPS spectra of the specimens: (a) survey spectrum, (b) Zn 2p, (c) O 1s, (d) S 2p, (e) Mo 3d, and (f) Ag 3d. |
The crystal structure analysis was also carried out through XRD, as presented in Fig. 4(a). For the ZnO nanorod arrays, except the peaks at 26.61°, 37.95°, 51.78°, 61.87° and 65.94° (2θ) that can be attributed to the FTO substrate (JCPDS Card no. 41-1445), all the other diffraction peaks at 31.77°, 34.42°, 36.25°, 47.54°, 56.60°, 62.86° and 72.56° can be well indexed of the (100), (002), (101), (102), (110), (103) and (004) planes of the wurtzite ZnO crystal (JCPDS Card no. 36-1451). The other specimens possess the ZnO phase, but the XRD patterns do not reveal the existence of Ag2S and/or MoS2 phases, which may be due to the thin coating layer or the high content of 1T phase MoS2.25,26 Fig. 4(b) shows PL spectra of the samples at the excitation wavelength of 325 nm. The ZnO nanorod arrays display a UV emission at ∼376 nm (3.30 eV), a blue emission at 441 nm (2.81 eV) with a shoulder peak at 410 nm (3.02 eV), and a green emission at 547 nm (2.27 eV). The UV emission is much stronger than the others. As confirmed in the previous research,27–29 the UV emission originates from bandedge emission for the recombination of free excitons, the blue emissions at 410 and 441 nm derive from transitions from Zn interstitial and extended Zn interstitial states (Zni) to the valence band respectively, and the green emission comes from recombination of electrons trapped in the conduction band and deeply trapped holes in oxygen vacancies (Vo). For the heterogeneous nanorod arrays, the bandedge emission is significantly reduced, while the blue and green emissions are distinctly amplified. Nevertheless, the emission bands of the impurities and vacancies are depressed obviously at the excitation of 400 nm (see Fig. S1†). The weakened and vanished emission suggests limited exciton recombination in the heterogeneous specimens, which is advantageous for their potential photovoltaic application. Raman spectra in Fig. S2† display tow resonant bands at 380 and 408 cm−1 for ZnO/MoS2 and ZnO/MoS2/Ag2S nanorod arrays, which are the typical in-plane (E12g) and out of plane (A1g) vibration modes of individual Mo and S atoms.30,31 The existence of E12g and A1g modes reveals the presence of the MoS2 sheets.
UV-Vis absorption spectra were carried out to investigate the light response, as shown in Fig. 4(c). The pure ZnO displays a bandedge absorption at 380 nm with low absorption in the visible region. When the ZnO is coupled with Ag2S to form a binary heterojunction, the absorption edge gets a remarkable red-shift with an obvious enhanced light absorption toward visible range. Moreover, When the ZnO is coupled with MoS2 to form a binary heterojunction and further coupled with Ag2S to form a ternary heterojunction, the specimens exhibit a strong and full light absorption in 200–800 nm with an obvious declining in 900–1350 nm. The improved light absorption in visible range may lead to much electron–hole pair separation and enhance photocatalytic activity of the catalysts. The optical bandgap energy (Eg) of the semiconductor catalysts can be calculated from the Tauc plots using the formula:32
(αhν)1/n = A(hν − Eg) | (1) |
The electrochemical impedance spectroscopy (EIS) was conducted to evaluate conductivity of the samples, as shown in Fig. 4(d). The single and binary components arrays display a semicircle in the high-frequency region and a straight line in the low-frequency region, while the ternary components array show two semicircles. The impedance can be fitted by the equivalent circuit, where diagram (i) for ZnO, ZnO/Ag2S and ZnO/MoS2 arrays, and diagram (ii) for ZnO/MoS2/Ag2S array. The solid lines are the curves obtained after fitting, and the achieved resistances are shown in Table 1. For the heterojunction nanorod arrays, the solution resistance (R1) increases, while the charge transfer resistance (R2) decreases. In comparison with pure ZnO array, the R2 of ZnO/Ag2S and ZnO/MoS2/Ag2S arrays reduces to be 65% and 64%, respectively, while that of the ZnO/MoS2 array decreases to be only 41%. This suggests that the heterojunction can not only promote charge separation, but also improve the electrical conductivity, and the MoS2 sheets play an important role for the contribution.
ZnO | ZnO/Ag2S | ZnO/MoS2 | ZnO/MoS2/Ag2S | |
---|---|---|---|---|
R1 (Ω cm2) | 3.537 | 5.144 | 6.124 | 5.149 |
R2 (Ω cm2) | 11.67 | 7.591 | 4.789 | 7.485 |
R3 (Ω cm2) | — | — | — | 234.6 |
In order to evaluate photocatalytic activity of the arrays, MB aqueous solutions as catalyzed by different specimen ware examined under visible and UV light. Fig. 4(e) and (f) show absorption spectra of the degraded MB aqueous solutions as catalyzed by different specimens in 30 min. The peak intensity of the degraded solutions is reduced after the photocatalytic process, regardless in visible or UV light, suggesting catalytic ability of the semiconductor specimens. The photodegradation efficiency E can be defined by the following equation,33
E = (C0 − C)/C0 × 100% | (2) |
Light irradiation | ZnO | ZnO/Ag2S | ZnO/MoS2 | ZnO/MoS2/Ag2S |
---|---|---|---|---|
UV | 28.55% | 31.91% | 26.28% | 58.12% |
Visible | 6.94% | 17.24% | 22.74% | 52.88% |
From practical point, keeping high photochemical activity and stability is a critical requirement for the catalysts. Therefore, the photocatalytic performance of the nanowire arrays was measured in three successive reaction periods, as displayed in Fig. S4,† and the efficiency was also computed and supplied in Table S1.† It is found that there is little variation in the peak intensity of the degradation solutions as catalyzed by the single and binary components, while the peak intensity improves obviously with increasing cycle for the ternary components. The distinct reduced efficiency may result from the losing of MoS2 nanosheets and Ag2S nanoparticles for the weak binding, or the surface covering of the degraded products. Nevertheless, the photocatalytic efficiency of ZnO/MoS2/Ag2S array retains larger than others in the UV light and visible light in any cycle. The stability of the ternary array was further tested by the photocatalytic performance under visible light in different conditions in Fig. S5 and Table S2 as well as the structure images in Fig. S6.† The catalytic efficiency of the specimen is found to nearly keep stable after 3 cycles. The efficiency of 25.30% in the 5th cycle is also larger than that of other samples in the 1st cycle under visible light. The nanorod array remains normal to the substrate after photocatalysis for 5 cycles. There is not clear difference in the array structure before and after photocatalytic performance, indicating a good photocatalytic stability of ZnO/MoS2/Ag2S array. Moreover, the efficiency increases with extending period and reaches 90.09% in 4 h. The MB solution is speculated to be degraded completely by the ternary array in about 5.5 h.
To make clear photocatalytic activity of the specimens, band structure of the heterogeneous interfaces is depicted in Fig. 5 based on the defect states in Fig. 4(b), bandgap (Eg) in Fig. S3,† flat band potential (EFB) in Mott–Schottky curves in Fig. S7,† work function (Ef) and electron affinity (EC) of the semiconductor components in bulk state. For the ZnO–Ag2S interface (see Fig. 5(a1)), the CB and VB of Ag2S component straddle those of ZnO component, which forms a Type II heterojunction.35 Since the absolute value of EF for Ag2S is smaller than that for ZnO, the Fermi level of Ag2S is higher than that of ZnO. During formation of the internal electric field, the free electrons in Ag2S flow to ZnO until the Fermi levels are aligned. Since the surface of ZnO is negatively charged while that of Ag2S is positively charged, the direction of the internal electric field is from Ag2S to ZnO, as indicated by a red arrow line. Therefore, the transfer of holes from Ag2S to ZnO is favorable by the internal electric field, while the transfer of electrons in the same direction is prohibited. Meanwhile, the flow of photogenerated charge carriers in ZnO is opposite as that in Ag2S, which results in the recombination of electrons in the CB of ZnO and holes in the VB of Ag2S at the interface, as manifested by a blue arrow line. In addition, the shallow defects in Zni supply some seating states for the excited electrons. This particular Z-scheme energy diagram and extra states reduce charge recombination in the interior of the component semiconductors and improves redox ability of the heterojunction, which is beneficial for the photocatalytic performance. The Z-scheme diagram also exist in ZnO–MoS2 interface, as shown in Fig. 5(b1). But for MoS2–Ag2S interface in Fig. 5(c1), the CB and VB of Ag2S component seat in between those of MoS2 component, which forms a Type I heterojunction. The internal electric field is from Ag2S to MoS2 as the work function of MoS2 is larger than that of Ag2S. As a result, the photoexcited electrons in the CB of MoS2 flow to the CB of Ag2S, while the carriers in other bands are restricted or gather at the interface for the energy barrier.
Fig. 5 Band diagrams and photocatalytic mechanisms of the heterogeneous nanorods at the interfaces: (a1, a2) ZnO/Ag2S, (b1, b2) ZnO/MoS2, (c1, c2) MoS2/Ag2S. |
In light of the energy band structure, under light irradiation, the electrons in the VB of Ag2S and ZnO are potentially excited to the CB and defect states. This Z-scheme heterojunction facilitates the residual holes in ZnO component to react with water molecules or hydroxide anion in the solution to produce ·OH and h+ species, while the excited electrons in Ag2S component react with oxygen to generate ·O2− species, all of which give rise to the decomposition of MB dyes, as illustrated in Fig. 5(a2). This process also occurs at the interfaces of ZnO–MoS2 (see Fig. 5(b2)) heterojunctions. But for MoS2–Ag2S, the Type I heterojunction induces electrons in Ag2S component to react with oxygen and generate ·O2− species, while the holes in the two components bring ·OH and h+ species (see Fig. 5(c2)). All these lead to the dye degradation. In light of free radical trapping experiments (see Fig. S8 and Table S4†), the ·OH and h+ radicals play an important role in the degradation by the ternary ZnO/MoS2/Ag2S array. Therefore, the band alignment enhances photocatalytic performance of heterogeneous core–shell nanorod arrays. Especially for the ternary component ZnO/MoS2/Ag2S array, it contains both the staggered gap (Type II) of ZnO–MoS2 heterojunction in Fig. 5(b2) and straddling gap (Type I) of MoS2–Ag2S heterojunction in Fig. 5(c2), which extends the light absorption by the component semiconductors, reduces charge recombination efficiently by the energy barrier and band alignment, and possesses the highest active catalysis under visible light than ZnO/Ag2S and ZnO/MoS2 arrays. This may also have promising application in other fields, such as solar cells, photodiodes and photosensors.
Footnote |
† Electronic supplementary information (ESI) available: PL spectra at the excitation wavelength of 400 nm, Raman spectra, Tauc plots of the absorption spectra, absorption spectra of degraded MB aqueous solutions in different periods and different conditions, SEM images of ZnO/MoS2/Ag2S nanoarrays before and after photodegradation, Mott–Schottky plots, absorption spectra of the degraded MB solution with different scavengers, and photodegradation efficiency. See https://doi.org/10.1039/d2ra03940k |
This journal is © The Royal Society of Chemistry 2022 |