Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Molybdenum-modified ZnIn2S4 with a bifunctional cocatalyst for efficient photocatalytic overall water splitting under simulated sunlight

Yutong Chen ab, Rong Wu *ab, Guoan Lin de, Jianyong Yue ab, Dezheng Kong ab, Chen Zhang abc, Shunhang Wei *c and Xiaoxiang Xu *de
aXinjiang Key Laboratory of Solid State Physics and Devices, Xinjiang University, Urumqi, Xinjiang 830017, People's Republic of China
bSchool of Physics Science and Technology, Xinjiang University, Urumqi, Xinjiang 830017, People's Republic of China. E-mail: wurongxju@sina.com
cZhejiang Engineering Research Center of MEMS, Shaoxing University, Shaoxing 312000, China. E-mail: wsh@usx.edu.cn
dClinical and Central Lab, Putuo People's Hospital, Tongji University, Shanghai, 200060, China
eSchool of Chemical Science and Engineering, Tongji University, Shanghai, 200092, China. E-mail: xxxu@tongji.edu.cn

Received 9th February 2025 , Accepted 31st March 2025

First published on 1st April 2025


Abstract

Photocatalytic overall water splitting (POWS) is a promising technique for sustainable hydrogen production that can potentially tackle the current energy and environmental challenges. As a visible-light-active photocatalyst, ZnIn2S4 is an ideal candidate for POWS; however, it exhibits low efficiency due to poor charge separation and improper surface reactions. Here, ZnIn2S4 was modified by doping Mo and depositing a bifunctional cocatalyst for efficient POWS under sunlight. Charge separation and photodeposition of Pt and CrOx were accelerated by Mo introduction to promote surface POWS reactions and suppress reverse reactions. Under simulated sunlight irradiation, the modified ZnIn2S4 achieved POWS with an apparent quantum efficiency (AQE) of 0.17% at 420 ± 20 nm and a solar-to-hydrogen (STH) conversion efficiency of 0.016%. Theoretical calculations indicated that Mo dopants introduced spin-polarized states at both the conduction band bottom and the valence band top, which are spin-forbidden for charge recombination. This study served as a useful guideline for the design and development of efficient photocatalysts for sunlight-driven POWS.


1. Introduction

Developing hydrogen energy, particularly through sustainable routes, is essential to tackle the challenges of increasing energy needs and environmental degradation.1 This is typically exemplified by photocatalytic overall water splitting (POWS) using solar energy where hydrogen and oxygen can be produced at the surface of a particulate photocatalyst. Previous studies on POWS have predominantly focused on wide-bandgap semiconductors whose light absorption is limited to UV-light.2–6 The solar-to-hydrogen (STH) conversion efficiency thereby has intrinsic limitations (<5%) due to their inadequate utilization of solar energy. Developing narrow-bandgap semiconductors with visible light absorption is essential to achieve efficient POWS using solar energy.7

In this context, transition metal sulfides have garnered significant attention due to their simple structure and ability to absorb visible light photons.8–11 Among them, ZnIn2S4 is widely studied because of its promising photoelectric properties and catalytic performance.12–16 However, the rapid recombination of photogenerated electrons and holes in ZnIn2S4 poses a great challenge for POWS.17,18 Moreover, the surface of ZnIn2S4 is usually deficient in active reaction sites for redox reactions and is subject to photocorrosion due to the vulnerability of the S-terminated surface during water splitting, which is another factor that inhibits POWS reactions.19–21 Besides, the POWS activity hinges on the proper construction of both reduction and oxidation sites, which would otherwise induce reverse reactions for the product gases H2 and O2.22,23 Rational cocatalyst deposition is indispensable for POWS not only for the rapid collection of electrons and holes but also for the inhibition of the reverse reaction as well as photocorrosion.24–28

In this work, Mo has been uniformly doped into ZnIn2S4, i.e. ZnIn2S4:Mo, using a simple one-step hydrothermal method. Mo doping significantly enhances the capture efficiency of photogenerated holes, thereby facilitating charge separation. Meanwhile, Pt and CrOx cocatalysts with a core–shell structure have been deposited onto ZnIn2S4:Mo, which substantially increases the surface charge transfer efficiency and inhibits reverse reactions. Under optimal conditions, the photocatalyst exhibits H2- and O2-evolution rates of 122.8 μmol g−1 h−1 and 65.9 μmol g−1 h−1 under simulated sunlight, respectively.

2. Experimental section

2.1. Chemicals

Zinc acetate dihydrate (C4H6O4Zn·2H2O, Aladdin, A.R.), indium(III) chloride tetrahydrate (InCl3·4H2O, Macklin, A.R.), thioacetamide (CH3CSNH2, Aladdin, A.R.), citric acid (C6H8O7, Aladdin, A.R., ≥99.5%(T)), sodium molybdate dihydrate (Na2MoO4·2H2O, Aladdin; A.R., 99.0%), chromium(III)nitrate nonahydrate (Cr(NO3)3·9H2O, Aladdin, A.R., 99.0%), and chloroplatinic acid hexahydrate (H2PtCl6·6H2O, Aladdin, A.R., Pt ≥ 37.5%) were used. All chemicals were used as received without further purification, and deionized water was employed throughout the entire experimental procedure.

2.2 Sample preparation

2.2.1. Synthesis of ZnIn2S4. ZnIn2S4 was prepared through hydrothermal reactions according to the literature.29 Typically, 2 mmol of zinc acetate dihydrate, 4 mmol of indium(III) chloride tetrahydrate, 10 mmol of thioacetamide, and 1 g of citric acid were dissolved in 60 mL of deionized water with stirring for 30 min. The solution was transferred into a Teflon-lined autoclave (100 mL) and maintained at 180 °C for 24 h. After cooling to room temperature naturally, the obtained product was collected by centrifugation, washed with deionized water and absolute ethanol, and then dried at 80 °C for further characterization.
2.2.2. Synthesis of Mo doped ZnIn2S4 (ZnIn2S4:Mo). To prepare Mo-doped ZnIn2S4, sodium molybdate dihydrate was added to the reaction solution for hydrothermal synthesis at molar contents of 1%, 3%, and 5%. The reaction conditions were kept the same as those for the synthesis of ZnIn2S4. The products were denoted as ZnIn2S4:Mo(x) (x = 1, 3 and 5), where x represents the molar percent of Mo doped into ZnIn2S4.
2.2.3. Depositing cocatalysts onto ZnIn2S4:Mo (Pt/CrOx@ZnIn2S4:Mo). The deposition of Pt and CrOx onto ZnIn2S4:Mo was carried out via photo-reduction of H2PtCl6 and photooxidation of Cr(NO3)3 simultaneously over ZnIn2S4:Mo. Taking ZnIn2S4:Mo(3) as an example, 100 mg of ZnIn2S4:Mo was suspended in an aqueous solution containing appropriate amounts of H2PtCl6 and Cr(NO3)3. The resultant suspensions were irradiated with a 300 W Xe lamp for 1 h. The resultant products were filtered, washed with deionized water and absolute ethanol and dried in a vacuum oven at 60 °C. The resulting catalyst exhibited a deep yellow color and was designated as Pt/CrOx@ZnIn2S4:Mo(3) (Scheme 1).
image file: d5dt00316d-s1.tif
Scheme 1 The schematic representation of ZnIn2S4:Mo(3) nanospheres.

2.3. Catalyst characterization

The microstructure of the sample was examined by field emission scanning electron microscopy (FESEM, Hitachi SU8010, Japan) and transmission electron microscopy (TEM, JEOL JEM-F200, Japan). X-ray diffraction (XRD) patterns were recorded on an X-ray diffraction instrument (Bruker D8 Advanced, Germany) operating at 60 kV and 50 mA with Cu-Kα radiation. The surface state was investigated utilizing the X-ray photoelectron spectroscopy (XPS) technique with the Thermo Fisher Scientific ESCALAB 250Xi model. The nitrogen adsorption and desorption isotherms at 77 K were obtained using a nitrogen adsorption analyser (Micromeritics ASAP 2460, USA). Prior to measurement, all samples were outgassed under a vacuum at 120 °C for a minimum of 5 hours. UV-vis diffuse reflection spectra (DRS) were obtained using a UV-vis spectrophotometer (Shimadzu UV-3600 plus, Japan) equipped with an integrating sphere assembly. The photoluminescence (PL) spectra were recorded at room temperature with a fluorescence spectrophotometer (Edinburgh FLS1000) using 375 nm as the excitation wavelength. Decay curves were obtained using an FLS980 fluorescence spectrophotometer (Edinburgh Instruments, UK) under the excitation wavelength at 325 nm. The average lifetime (Ave. τ) was calculated according to τ = (A1·τ12 + A2·τ22)/(A1·τ1 + A2·τ2) (τi is the lifetime; Ai is the relative intensity). A Bruker EMXplus-6/1 (Germany) was used for the characterization of electron paramagnetic resonance (EPR) with a 300 W Xenon lamp at room temperature. The steady-state surface photovoltage (SPV) spectra were collected using a CEL-SPS1000 spectroscopic analysis system containing a 500 W Xenon lamp (CEL-S500).

2.4. Electrochemical measurements

The experiment was conducted using a CHI760E electrochemical workstation. The experimental setup consisted of a three-electrode cell, with a calomel electrode as the reference electrode, platinum foil as the counter electrode, and a conductive glass electrode coated with a photocatalyst as the working electrode. The electrolyte used was a 0.5 mol L−1 Na2SO4 solution. For the photoanode preparation, 24 mg of the photocatalyst, 3 mg of acetylene black, and 3 mg of polyvinylidene difluoride were mixed uniformly. Afterwards, an appropriate amount of N-methyl pyrrolidone was added through droplet-wise addition. Finally, the resultant suspensions were deposited onto conductive glass and dried at room temperature. Transient photocurrent response measurements were conducted using a 300 W xenon lamp coupled with an AM1.5G filter as the light source. Electrochemical impedance spectroscopy analysis was performed in the frequency range of 1 to 106 Hz. Mott–Schottky measurements were carried out at a frequency of 2000 Hz.

2.5. Computational details

The theoretical calculations of the band structures for ZnIn2S4 and ZnIn2S4:Mo(3) were performed using density functional theory (DFT) with a commercial Vienna ab initio simulation package. The calculations were performed using the generalized gradient approximation (GGA), Perdew–Burke–Ernzerhof (PBE) function, and projector augmented-wave pseudopotential. A 2 × 2 × 2 cubic unit cell (a = b = c = 8.26 Å, α = β = γ = 90°) was used as the structural model. The structures were fully relaxed following the criteria that the forces on each atom are less than 0.02 eV Å−1. An energy cutoff of 400 eV and a total energy of less than 10−5 eV were used for static electric potential calculations and geometry optimization, respectively. A Monkhorst–Pack k-points mesh of 5 × 5 × 5 was sampled for all structures.

2.6. Photocatalytic overall water splitting measurement

Photocatalytic tests were performed on a commercial photocatalytic analysis system (Labsolar-6A, Beijing Perfectlight) at ambient temperature. Typically, 50 mg of photocatalysts was dispersed in 100 mL of deionized water (pH = 7). Air in the reaction equipment and solution was vacuumed first. Subsequently, the reactor was exposed to a 300 W Xenon lamp source with an AM1.5G filter. Gas evolution was monitored by on-line gas chromatography (GC9790II, FULI). The stability test was performed under the same conditions over consecutive testing cycles (5 h). The reaction equipment and solution were vacuumed before each testing cycle. The apparent quantum efficiency (AQE) was evaluated using the same system under illumination with different bandpass filters. AQE was calculated according to the following equation:
image file: d5dt00316d-t1.tif

To obtain the STH efficiency, 100 mg of photocatalyst and a 300 W xenon lamp with an AM1.5G filter (100 mW cm−2) were used.

image file: d5dt00316d-t2.tif

3. Results and discussion

3.1. Microstructure characterization

The ZnIn2S4 and ZnIn2S4:Mo(x) samples were synthesized via a one-step hydrothermal method and were characterized by various analytical techniques. Taking ZnIn2S4:Mo(3) as an example, both pristine ZnIn2S4 (Fig. 1a and b) and ZnIn2S4:Mo(3) (Fig. 1c and d) exhibit hierarchical flower-like microstructures comprising nanosheets, being consistent with ZnIn2S4 samples synthesized via a hydrothermal method.30 Mo doping, therefore, can preserve the flower-like microstructures of ZnIn2S4. To further elucidate the microstructural characteristics, the samples were inspected under HRTEM conditions. The ultra-thin two-dimensional nanosheets were similar before and after Mo doping, as depicted in the figure of pristine ZnIn2S4 (Fig. 1e) and ZnIn2S4:Mo(3) (Fig. 1g). In the magnified images, a lattice spacing of 0.32 nm is observed, corresponding to the (102) crystal plane of hexagonal ZnIn2S4. The elemental mapping of ZnIn2S4:Mo(3) (Fig. 1i) reveals a uniform dispersion of Zn, In, S, and Mo, providing evidence of the homogeneous doping of Mo in ZnIn2S4. The molar ratio of Zn/In/S in the ZnIn2S4:Mo(3) nanosheets is determined to be 1[thin space (1/6-em)]:[thin space (1/6-em)]1.66[thin space (1/6-em)]:[thin space (1/6-em)]4.00, as detailed in Table S1. This ratio, in comparison with that of the pristine ZnIn2S4, indicates a significant reduction in In content, suggesting that Mo atoms likely substitute a fraction of the In atoms within the ZnIn2S4 lattice.
image file: d5dt00316d-f1.tif
Fig. 1 Scanning electron microscopy images of (a and b) ZnIn2S4 and (c and d) ZnIn2S4:Mo(3); transmission electron microscopy and high-resolution transmission electron microscopy images of ZnIn2S4 (e and f) and ZnIn2S4:Mo(3) (g and h); (i–m) elemental mapping images of ZnIn2S4:Mo(3).

3.2. Structural and surface analysis

The crystal structures of pristine ZnIn2S4 and ZnIn2S4:Mo(x) (where x = 1, 3, 5 atom%) were characterized via X-ray diffraction (XRD). As depicted in Fig. 2a and Fig. S1, pristine ZnIn2S4 and ZnIn2S4:Mo(x) showed XRD patterns that matched well with the hexagonal ZnIn2S4 standard pattern (JCPDS:00-065-2023). The characteristic reflections at 21.6°, 27.7°, 47.5°, 52.4°, and 55.6° correspond to the (006), (102), (110), (116), and (022) crystal planes of hexagonal ZnIn2S4, respectively. The light absorption capacity of Mo-modified samples was then assessed utilizing UV-vis absorption spectroscopy in Fig. 2b. The spectra indicate that the light absorption capacity of ZnIn2S4 has been significantly enhanced through Mo doping, as indicated by its increased absorption across the UV, visible, and near-infrared regions.
image file: d5dt00316d-f2.tif
Fig. 2 (a) X-ray diffraction patterns; (b) UV-vis absorption spectra; (c) XPS spectra of In 3d and (d) S 2p of ZnIn2S4 and ZnIn2S4:Mo(3).

The elemental states and chemical composition of ZnIn2S4 before and after Mo doping were investigated by XPS. All core-level binding energies were adjusted with reference to the adventitious C 1s peaks at 284.7 eV. It is noteworthy that the peaks at 443.9 eV and 451.5 eV of ZnIn2S4:Mo(3) correspond to the In 3d5/2 and In 3d3/2 states, exhibiting a slight shift towards higher binding energy compared to ZnIn2S4 (Fig. 2c). This is probably due to a higher electronegativity of Mo compared to In that tends to attract electrons. Similar results were also observed in the S 2p state, where the peak positions underwent a blue shift of ∼0.2 eV (Fig. 2d). The increase can be attributable to the high electronegativity of Mo that results in the electron cloud shift from S towards Mo. The peaks of ZnIn2S4 and ZnIn2S4:Mo(3) at 1020.9 eV and 1043.9 eV, respectively, correspond to the Zn 2p3/2 and Zn 2p1/2 states (Fig. S2a). The peaks observed at 228.1 eV and 231.4 eV in ZnIn2S4:Mo(3) correspond respectively to the Mo 3d5/2 and Mo 3d3/2 states, being indicative of Mo4+ (Fig. S2b).31 Furthermore, Fig. S3a and b present the N2 adsorption–desorption isotherms that reveal hysteresis loops for both pristine ZnIn2S4 and ZnIn2S4:Mo(3) nanosheets, indicating mesoporous structures (type-IV isotherms). Using the Brunauer–Emmett–Teller (BET) method, we calculated their specific surface area and pore volume. The pristine ZnIn2S4 nanosheets have a BET specific surface area of 107.2 m2 g−1 and a pore volume of 0.22 cm3 g−1. In contrast, the ZnIn2S4:Mo(3) nanosheets have a specific surface area of 117.8 m2 g−1 and a pore volume of 0.23 cm3 g−1. These results suggest that Mo doping has only trivial impacts on the surface area and pore volume of ZnIn2S4.

3.3. Photocatalytic performance characterization

The photocatalytic performance of these samples was assessed based on their activity for overall water splitting under simulated AM1.5G conditions. No H2 and O2 evolution was detected when using naked ZnIn2S4 and ZnIn2S4:Mo(3), implying that the surface of ZnIn2S4 is unsuitable for POWS reactions. To facilitate water redox reactions, Pt and CrOx cocatalysts have been deposited onto these samples, which are critical to achieve POWS activity.32 The successful deposition of these cocatalysts is confirmed by XPS and TEM analysis (Fig. 4a and b). The Pt 4f spectra are characterized by four overlapping peaks, being attributable to the Pt0/Pt2+ valence state. The surface of ZnIn2S4 is typically terminated with S anions, which can strongly interact with the metallic Pt cocatalysts by forming Pt–S bonds according to the literature.33 Similarly, the Cr 3d spectra exhibit two peaks with binding energies of 586.2 eV and 576.1 eV, corresponding to the 2p3/2 and 2p1/2 orbitals of Cr3+ from CrOx.34 The cocatalysts were further inspected by using TEM, HRTEM, and elemental mapping (Fig. 5). The Pt cocatalyst was characterized as small granules with a diameter of approximately 5 nm, while CrOx appears as an amorphous layer covering the Pt particles. The metallic Pt can be verified by its typical crystal plane spacing of 0.23 nm, corresponding to the Pt (111) crystal plane.35 Elemental mapping analysis confirms that the deposition of Pt and CrOx is uniform on the surface of ZnIn2S4:Mo(3) (Fig. S5c).
image file: d5dt00316d-f3.tif
Fig. 3 (a) Comparison of photocatalytic H2 and O2 evolution rates for ZnIn2S4, Pt/CrOx@ZnIn2S4, and Pt/CrOx@ZnIn2S4:Mo(3); (b) cycling tests of the POWS activity of Pt/CrOx@ZnIn2S4:Mo(3); (c) comparison of photocatalytic H2 and O2 evolution rates for Pt/CrOx@ZnIn2S4:Mo(x) (x = 0, 1, 2, 3); (d) UV-vis light absorption and corresponding wavelength-dependent AQE of Pt/CrOx@ZnIn2S4:Mo(3).

image file: d5dt00316d-f4.tif
Fig. 4 (a and b) XPS spectra of Pt 4f and Cr 2p for Pt@CrOx/ZnIn2S4:Mo(3); (c) comparison of photocatalytic H2 and O2 evolution rates for Pt@ZnIn2S4:Mo, CrOx@ZnIn2S4:Mo(3) and Pt/CrOx@ZnIn2S4:Mo(3) nanosheets; (d) reverse reaction of Pt/CrOx@ZnIn2S4:Mo(3) and ZnIn2S4:Mo(3).

image file: d5dt00316d-f5.tif
Fig. 5 (a and b) Transmission electron microscopy and elemental mapping images of Pt/CrOx@ZnIn2S4:Mo(3); (c) Transmission electron microscopy and elemental mapping images of Pt/CrOx@ZnIn2S4:Mo(3).

By depositing the Pt and CrOx cocatalysts, both pristine ZnIn2S4 and ZnIn2S4:Mo(3) exhibit POWS activity under simulated sunlight irradiation. The H2 to O2 ratio approaches approximately 2 to 1, suggesting successful overall water splitting for these samples.

As illustrated in Fig. 3a, compared with pristine Pt/CrOx@ZnIn2S4, a significant enhancement in H2 and O2 evolution rates is observed for Pt/CrOx@ZnIn2S4:Mo(3), highlighting the importance of Mo doping. The doping content was also optimized to further enhance the POWS activity. As shown in Fig. S4, doping 3% Mo delivers the highest performance, with H2 and O2 evolution rates of 122.8 μmol g−1 h−1 and 65.9 μmol g−1 h−1, respectively. Further increasing the Mo content may lead to the formation of an impurity phase like MoS2, which serves to block light penetration and impairs the POWS activity (Fig. S5). ZnIn2S4 doped with 3% Mo was then used for further stability testing. Fig. 3b demonstrates that Pt/CrOx@ZnIn2S4:Mo(3) can stably produce H2 and O2 for four consecutive cycles, indicating good stability for POWS reactions. The importance of depositing both Pt and CrOx can be realized by several control experiments. First, the Pt content was varied from 0% to 3%. The activity of Pt/CrOx@ZnIn2S4:Mo(3) clearly shows an optimal point at 2% (Fig. 3c), suggesting that the presence of Pt modulates the photocatalytic activity. The role of CrOx can be elucidated through the control experiments with Pt and CrOx alone. As shown in Fig. 4c, depositing Pt alone on ZnIn2S4:Mo(3) substantially inhibits O2 evolution. This is probably due to the fact that Pt is effective in triggering reverse reactions between H2 and O2 or the oxygen reduction reaction (ORR). This is confirmed by the decrease in H2 and O2 levels in the presence of Pt loaded ZnIn2S4:Mo(3) (Pt@ZnIn2S4:Mo(3), Fig. 4d). This decrease is not observable when Pt is coated with CrOx. Therefore, CrOx has major contributions to the inhibition of reverse reactions. In addition, depositing CrOx alone (CrOx@ZnIn2S4:Mo(3)) shows a clear degradation of POWS activity, although the evolution of both H2 and O2 with the correct ratio is observed. This result indicates that Pt is more effective in promoting POWS reactions than CrOx. Combining these results, it is feasible to rationalize the roles of Pt and CrOx during POWS reactions. Upon the generation of photo-generated charges in ZnIn2S4:Mo(3) by sunlight, Pt efficiently collects charges for surface redox reactions. The product gases H2 and O2 are isolated from Pt due to the presence of CrOx which effectively inhibits reverse reactions and/or the ORR. Therefore, there is a collaborative effect between Pt and CrOx for improving the POWS activity.36 The POWS activity was further evaluated by collecting the action spectra, as shown in Fig. 3d. It is clear that the photocatalytic activity is photon-driven as the apparent quantum efficiency (AQE) of Pt/CrOx@ZnIn2S4:Mo(3) matches well with the light absorption spectra. The AQE at 420 ± 20 nm reaches 0.17% and the STH is determined to be 0.016%. To the best of our knowledge, these results are competitive with those of ZnIn2S4-based photocatalysts reported for POWS reactions (Table S2).

3.4. Photoelectrochemical analysis

To clarify the improved POWS activity after Mo doping, we performed photoelectrochemical measurements on ZnIn2S4, ZnIn2S4:Mo(3) and Pt/CrOx@ZnIn2S4:Mo(3). Electrochemical impedance spectroscopy (EIS) spectra are presented in Fig. 6a. The EIS results demonstrate that compared with pristine ZnIn2S4, ZnIn2S4:Mo (3) exhibits a lower charge transfer resistance (Rct), while a further reduced Rct is observed after cocatalyst deposition, indicating that Mo doping and cocatalyst deposition synergistically optimize the water redox reaction conditions. Furthermore, under chopped light irradiation, ZnIn2S4:Mo(3) exhibits a photocurrent density significantly higher than that of ZnIn2S4 yet slightly lower than that of Pt/CrOx@ZnIn2S4, which aligns well with the photocatalytic performance. These results validate that Mo doping and cocatalyst deposition synergistically accelerate charge separation in ZnIn2S4 (Fig. 6b). This deduction was further confirmed by surface photovoltage (SPV) testing that shows a significant increase in surface photovoltage in ZnIn2S4:Mo(3) and Pt/CrOx@ZnIn2S4:Mo(3) compared to ZnIn2S4 (Fig. 6c). All these results jointly confirm that the synergistic interaction between Mo doping and the cocatalyst contributes to enhanced charge separation, with the separated charges subsequently migrating to the catalyst surface to drive water redox reactions.
image file: d5dt00316d-f6.tif
Fig. 6 (a) Electrochemical impedance spectroscopy spectra of ZnIn2S4, ZnIn2S4:Mo(3) and Pt/CrOx@ZnIn2S4:Mo(3); (b) photocurrent measurements of ZnIn2S4, ZnIn2S4:Mo(3) and Pt/CrOx@ZnIn2S4:Mo(3); (c) surface photovoltage of ZnIn2S4, ZnIn2S4:Mo(3) and Pt/CrOx@ZnIn2S4:Mo(3); (d) photoluminescence spectra of ZnIn2S4, ZnIn2S4:Mo(3) and Pt/CrOx@ZnIn2S4:Mo(3); (e) time-resolved fluorescence spectra of ZnIn2S4, ZnIn2S4:Mo(3) and Pt/CrOx@ZnIn2S4:Mo(3); amplitude-weighted average lifetime is shown in the inset.

Photoluminescence (PL) is commonly employed to assess the ability of separating and recombining electrons and holes. Fig. 6d shows the photoluminescence spectra of pristine ZnIn2S4, ZnIn2S4:Mo(3), and Pt/CrOx@ZnIn2S4:Mo(3). An emission peak attributable to the band gap transition of ZnIn2S4 nanosheets is evident at 500 nm. The fluorescence intensity of the ZnIn2S4:Mo(3) nanosheets is significantly reduced compared to that of pristine ZnIn2S4. This observation confirms the enhanced separation efficiency of photogenerated electrons and holes by Mo doping. Moreover, Pt/CrOx@ZnIn2S4 shows further reduced PL intensity relative to ZnIn2S4:Mo(3), which highlights that cocatalyst deposition improves photocatalytic activity through additional charge separation promotion. The separation dynamics of photoinduced carriers were studied using time-resolved transient photoluminescence (TRPL) decay spectra as shown in Fig. 6e. A two-exponential model was used to fit the TRPL profiles (Table S3). The prolonged average lifetime (τ) of ZnIn2S4:Mo(3) compared to pristine ZnIn2S4 supports more effective electron–hole pair separation in the doped material, while the cocatalyst deposition further enhances charge separation efficiency through optimized carrier migration pathways.

3.5. Theoretical calculations

The impact of Mo doping on the electronic structures of ZnIn2S4 was investigated using density functional theory (DFT). As depicted in Fig. 7, the band structures, density of states (DOS), and projected density of states (PDOS) of pristine ZnIn2S4 and ZnIn2S4:Mo(3) were calculated based on DFT. Both ZnIn2S4 and ZnIn2S4:Mo(3) are inherently semiconductors, with calculated bandgaps of 1.136 and 0.671 eV. The band gap values derived from the generalized gradient approximation (GGA) method are known to be underestimated when compared to experimental measurements.37,38 However, the calculated results can be used for qualitative analysis.39 It can be seen from the band structure that Mo doping slightly reduces the bandgap of ZnIn2S4. The bandgap reduction arises from the fact that Mo doping introduces spin-polarized states both at the bottom of the conduction band and at the top of the valence band. These states are mainly contributed by the Mo 4d orbitals that are also strongly hybridized with In 5s and S 3p orbitals according to the PDOS. The behaviour of photo-generated charges, in either the conduction or the valence band, is modulated by these Mo-dominated states that are probably the origin of improved charge separation. For instance, the photo-generated electrons and holes could have opposite spins, making their recombination spin-forbidden according to the selection rules, resulting in ameliorated charge separation.
image file: d5dt00316d-f7.tif
Fig. 7 Band structures, density of states (DOS) and projected density of states (PDOS) for (a) ZnIn2S4:Mo(3) and (b) ZnIn2S4. The Fermi level is marked by the dotted lines.

3.6. Band structure analysis

For better understanding the POWS process, the band edge positions of ZnIn2S4 and ZnIn2S4:Mo(3) were determined by Mott–Schottky (MS) analysis in conjunction with X-ray photoelectron spectroscopy (XPS) valence band scans. The flat-band potential (Vfb) was calculated based on the Mott–Schottky plot (Fig. 8a) using the following formula:
image file: d5dt00316d-t3.tif
where C and A are the capacitance at the interface and electrode area, respectively, V is the bias applied, kB is Boltzmann's constant, ND is the donor concentration, T is the absolute temperature, ε is the dielectric constant, and e is the electric charge. It can be seen from the MS plot that both samples exhibit positive slopes, indicative of n-type semiconductivity. By examining the intercepts of the MS curves with the energy axis, the flat band potentials of ZnIn2S4 and ZnIn2S4:Mo(3) are determined to be 0.50 and 0.36 V vs. RHE, respectively.40,41 As illustrated in Fig. 8b, the onsets of the valence band of ZnIn2S4 and ZnIn2S4:Mo(3) are found to be 1.51 eV and 1.36 eV, respectively. Utilizing the Kubelka–Munk method, the bandgap (Eg) energies were derived from the UV-Vis diffuse reflectance spectra, as depicted in Fig. 8c. The calculated bandgap for ZnIn2S4:Mo(3) is determined to be approximately 2.40 eV, which is lower than that of ZnIn2S4 (2.50 eV), being consistent with theoretical calculations. Combining these results, the band edge positions of these two compounds can be deduced, as illustrated in Fig. 8d. ZnIn2S4:Mo(3) exhibits a slightly reduced band gap and negatively shifted band edge positions compared to ZnIn2S4 (Fig. 8).

image file: d5dt00316d-f8.tif
Fig. 8 (a) Mott–Schottky plots of ZnIn2S4 and ZnIn2S4:Mo(3); (b) XPS VB spectra of ZnIn2S4 and ZnIn2S4:Mo(3); (c) Tauc plot of ZnIn2S4 and ZnIn2S4:Mo(3); (d) band structures of ZnIn2S4 and ZnIn2S4:Mo(3).

3.7. Photocatalytic mechanism analysis

Based on the above findings and analysis, the mechanism of POWS of Pt/CrOx@ZnIn2S4:Mo(3) under simulated sunlight is illustrated in Scheme 2. Mo doping can effectively modulate the electronic structure of ZnIn2S4, facilitating charge separation. This improved charge separation is responsible for the high photocatalytic activity. Meanwhile, photodeposition of co-catalysts (Pt and CrOx) plays a vital role during POWS reactions. Pt is mainly responsible for the collection of photo-generated charges and promotes water redox reactions. CrOx serves as a functional layer that prevents H2 and O2 reverse reactions and the ORR. The synergistic effects between Mo doping, Pt and CrOx facilitate a smooth POWS reaction at the surface of ZnIn2S4, with stoichiometric H2- and O2- evolution, under simulated sunlight.
image file: d5dt00316d-s2.tif
Scheme 2 Mechanism for POWS in the Pt/CrOx@ZnIn2S4:Mo(3) system under simulated sunlight irradiation.

4. Conclusions

In this work, Mo is doped into ZnIn2S4 using a simple one-step hydrothermal method, followed by photo-deposition of Pt and CrOx cocatalysts for POWS reactions. The Mo doping introduces spin-polarized states at both the conduction band bottom and the valence band top, which are spin-forbidden for charge recombination and beneficial for charge separation. Depositing Pt facilitates charge collection and surface redox reactions. CrOx serves as a functional layer covering Pt, which is critical for inhibiting water-splitting reverse reactions and the ORR. Under optimal conditions, the photocatalytic systems achieve average H2- and O2- evolution rates of 122.8 μmol g−1 h−1 and 65.9 μmol g−1 h−1, respectively, under simulated sunlight irradiation. An AQE of 0.17% at 420 ± 20 nm and an STH of 0.016% have been obtained. This study proposes a novel strategy for POWS through the synergistic effect of element doping and co-catalyst deposition, offering a promising guideline for the development of photocatalysts for efficient solar-driven water-splitting reactions.

Author contributions

Yutong Chen: investigation, conceptualization, data curation, methodology, formal analysis, and writing – original draft. Rong Wu: supervision and writing – review & editing. Guoan Lin: theoretical calculations. Jianyong Yue: investigation, methodology, and validation. Dezheng Kong: investigation, methodology, and validation. Chen Zhang: investigation, methodology, and validation. Shunhang Wei: supervision and writing – review & editing. Xiaoxiang Xu: supervision and writing – review & editing. All authors have given approval to the final version of the manuscript.

Data availability

The data supporting this article have been included as part of the ESI.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (22066024, 52302291 and 52172225), the Special Project for Key Research and Development Program of Xinjiang Autonomous Region (2022B01033); the Key Projects of the Natural Science Program in Xinjiang Autonomous Region (2023D01D03).

References

  1. S. Chen, T. Takata and K. Domen, Particulate photocatalysts for overall water splitting, Nat. Rev. Mater., 2017, 2, 17050 CrossRef CAS.
  2. W. Lei, Y. Wang, H. Wang, N. Suzuki, C. Terashima and A. Fujishima, Gelation-induced controlled synthesis of TiO2 with tunable phase transition for efficient photocatalytic hydrogen evolution, Inorg. Chem. Front., 2024, 11, 2178–2186 RSC.
  3. C. Li, H. Che, P. Huo, Y. Yan, C. Liu and H. Dong, Confinement of ultrasmall CoFe2O4 nanoparticles in hierarchical ZnIn2S4 microspheres with enhanced interfacial charge separation for photocatalytic H2 evolution, J. Colloid Interface Sci., 2021, 581, 764–773 CrossRef CAS PubMed.
  4. X.-J. Lu, I. Ullah, J.-H. Li, S. Chen, C.-Z. Yuan and A.-W. Xu, A bimetallic CoZn metal–organic-framework derived CoZnS@NSC Co-catalyst loaded on g-C3N4 for significantly augmented photocatalytic H2 evolution, Inorg. Chem. Front., 2024, 11, 3435–3445 RSC.
  5. F. Xing, J. Li, C. Wang, S. Jin, H. Jin and J. Li, Efficient photocatalytic hydrogen evolution of g-C3N4/Vs-SnS2/CdS through a sulfur vacancy-rich SnS2 induced charge storage effect, Inorg. Chem. Front., 2024, 11, 2884–2893 RSC.
  6. S. Zhang, X. Liu, C. Liu, S. Luo, L. Wang, T. Cai, Y. Zeng, J. Yuan, W. Dong, Y. Pei and Y. Liu, MoS2 Quantum Dot Growth Induced by S Vacancies in a ZnIn2S4 Monolayer: Atomic-Level Heterostructure for Photocatalytic Hydrogen Production, ACS Nano, 2018, 12, 751–758 CrossRef CAS PubMed.
  7. X. Liu, L. Yan, W. Li, K. Chen, F. Wang, J. Xiao, T. Hisatomi, T. Takata and K. Domen, Enhancing the Photocatalytic Activity of CaTaO2N for Overall Water Splitting through Surface Nitride Ion Enrichment, ACS Catal., 2024, 14, 10561–10567 CrossRef CAS.
  8. P. Kulkarni, S. K. Nataraj, R. G. Balakrishna, D. H. Nagaraju and M. V. Reddy, Nanostructured binary and ternary metal sulfides: synthesis methods and their application in energy conversion and storage devices, J. Mater. Chem. A, 2017, 5, 22040–22094 RSC.
  9. Y. Li, H. Dong, L. Li, L. Tang, R. Tian, R. Li, J. Chen, Q. Xie, Z. Jin, J. Xiao, S. Xiao and G. Zeng, Recent advances in waste water treatment through transition metal sulfides-based advanced oxidation processes, Water Res., 2021, 192, 116850 CrossRef CAS PubMed.
  10. R. Kalia, B. M. Pirzada, R. K. Kunchala and B. S. Naidu, Noble metal free efficient photocatalytic hydrogen generation by TaON/CdS semiconductor nanocomposites under natural sunlight, Int. J. Hydrogen Energy, 2023, 48, 16246–16258 CrossRef CAS.
  11. R. Kalia, P. Yadav, A. Verma, J. Yadav and B. S. Naidu, Natural sunlight induced photocatalytic hydrogen evolution by SrTaO2N/CdS nanocomposites, J. Alloys Compd., 2025, 1015, 178790 CrossRef CAS.
  12. S. Shen, J. Chen, X. Wang, L. Zhao and L. Guo, Microwave-assisted hydrothermal synthesis of transition-metal doped ZnIn2S4 and its photocatalytic activity for hydrogen evolution under visible light, J. Power Sources, 2011, 196, 10112–10119 CrossRef CAS.
  13. R. Yang, L. Mei, Y. Fan, Q. Zhang, R. Zhu, R. Amal, Z. Yin and Z. Zeng, ZnIn2S4-Based Photocatalysts for Energy and Environmental Applications, Small Methods, 2021, 5, 2100887 CrossRef CAS PubMed.
  14. T. Zhang, T. Wang, F. Meng, M. Yang and S. Kawi, Recent advances in ZnIn2S4-based materials towards photocatalytic purification, solar fuel production and organic transformations, J. Mater. Chem. C, 2022, 10, 5400–5424 RSC.
  15. X. Zheng, Y. Song, Y. Liu, Y. Yang, D. Wu, Y. Yang, S. Feng, J. Li, W. Liu, Y. Shen and X. Tian, ZnIn2S4-based photocatalysts for photocatalytic hydrogen evolution via water splitting, Coord. Chem. Rev., 2023, 475, 214898 CrossRef CAS.
  16. P. Yadav and B. S. Naidu, in Green Hydrogen Economy for Environmental Sustainability. Volume 1: Fundamentals and Feedstocks, American Chemical Society, 2024, ch. 6, vol. 1473, pp. 119–146 Search PubMed.
  17. W. Guan, L. Zhang, P. Wang, Y. Wang, H. Wang, X. Dong, M. Meng, L. Sui, Z. Gan, L. Dong and L. Yu, Highly Efficient Photocatalytic Hydrogen Evolution over Mo-Doped ZnIn2S4 with Sulfur Vacancies, Nanomaterials, 2022, 12, 3980 CrossRef CAS PubMed.
  18. H. Su, C. Rao, L. Zhou, Y. Pang, H. Lou, D. Yang and X. Qiu, Mo-Doped/Ni-supported ZnIn2S4-wrapped NiMoO4 S-scheme heterojunction photocatalytic reforming of lignin into hydrogen, Green Chem., 2022, 24, 2027–2035 RSC.
  19. F. Huang, Z. Li, Y. Xu, A. Yan, T. Zhang, Q. Wang, S. Li, S. Lu, W. Zhao, Y. Gao and J. Zhang, Excellent anti-photocorrosion and hydrogen evolution activity of ZnIn2S4-based photocatalysts: In situ design of photogenerated charge dynamics, Chem. Eng. J., 2023, 473, 145430 CrossRef CAS.
  20. H. Su, H. Lou, Z. Zhao, L. Zhou, Y. Pang, H. Xie, C. Rao, D. Yang and X. Qiu, In situ Mo doped ZnIn2S4 wrapped MoO3 S-scheme heterojunction via Mo-S bonds to enhance photocatalytic HER, Chem. Eng. J., 2022, 430, 132770 CrossRef CAS.
  21. F. Xing, Q. Liu and C. Huang, Mo-Doped ZnIn2S4 Flower-Like Hollow Microspheres for Improved Visible Light-Driven Hydrogen Evolution, Sol. RRL, 2020, 4, 1900483 CrossRef CAS.
  22. A. S. Alotabi, T. D. Small, Y. Yin, D. J. Osborn, S. Ozaki, Y. Kataoka, Y. Negishi, K. Domen, G. F. Metha and G. G. Andersson, Reduction and Diffusion of Cr-Oxide Layers into P25, BaLa4Ti4O15, and Al:SrTiO3 Particles upon High-Temperature Annealing, ACS Appl. Mater. Interfaces, 2023, 15, 14990–15003 CAS.
  23. R. Li, T. Takata, B. Zhang, C. Feng, Q. Wu, C. Cui, Z. Zhang, K. Domen and Y. Li, Criteria for Efficient Photocatalytic Water Splitting Revealed by Studying Carrier Dynamics in a Model Al-doped SrTiO3 Photocatalyst, Angew. Chem., Int. Ed., 2023, 62, e202313537 CrossRef CAS PubMed.
  24. K. Chen, J. Xiao, J. J. M. Vequizo, T. Hisatomi, Y. Ma, M. Nakabayashi, T. Takata, A. Yamakata, N. Shibata and K. Domen, Overall Water Splitting by a SrTaO2N-Based Photocatalyst Decorated with an Ir-Promoted Ru-Based Cocatalyst, J. Am. Chem. Soc., 2023, 145, 3839–3843 CrossRef CAS PubMed.
  25. L. Lin, P. Kaewdee, V. Nandal, R. Shoji, H. Matsuzaki, K. Seki, M. Nakabayashi, N. Shibata, X. Tao, X. Liang, Y. Ma, T. Hisatomi, T. Takata and K. Domen, Flux-Assisted Synthesis of Y2Ti2O5S2 for Photocatalytic Hydrogen and Oxygen Evolution Reactions, Angew. Chem., Int. Ed., 2023, 62, e202310607 CrossRef CAS PubMed.
  26. Y. Qi, B. Zhang, G. Zhang, Z. Zheng, T. Xie, S. Chen, G. Ma, C. Li, K. Domen and F. Zhang, Efficient overall water splitting of a suspended photocatalyst boosted by metal-support interaction, Joule, 2024, 8, 193–203 CrossRef CAS.
  27. J. Xiao, M. Nakabayashi, T. Hisatomi, J. J. M. Vequizo, W. Li, K. Chen, X. Tao, A. Yamakata, N. Shibata, T. Takata, Y. Inoue and K. Domen, Sub-50 nm perovskite-type tantalum-based oxynitride single crystals with enhanced photoactivity for water splitting, Nat. Commun., 2023, 14, 8030 CrossRef CAS PubMed.
  28. W. Li, W. Duan, G. Liao, F. Gao, Y. Wang, R. Cui, J. Zhao and C. Wang, 0.68% of solar-to-hydrogen efficiency and high photostability of organic-inorganic membrane catalyst, Nat. Commun., 2024, 15, 6763 CrossRef CAS PubMed.
  29. H. Jing, G. Xu, B. Yao, J. Ren, Y. Wang, Z. Fang, Q. Liang, R. Wu and S. Wei, Sulfur Vacancy-Enriched Rhombohedral ZnIn2S4 Nanosheets for Highly Efficient Photocatalytic Overall Water Splitting under Visible Light Irradiation, ACS Appl. Energy Mater., 2022, 5, 10187–10195 CrossRef CAS.
  30. H. Jing, J. Ren, J. Yue, S. Liu, Q. Liang, R. Wu, Y. Wang, Z. Fang, H. Li and S. Wei, ZnIn2S4 with oxygen atom doping and surface sulfur vacancies for overall water splitting under visible light irradiation, Catal. Sci. Technol., 2023, 13, 226–232 RSC.
  31. Y. Liu, Y. Li, F. Peng, Y. Lin, S. Yang, S. Zhang, H. Wang, Y. Cao and H. Yu, 2H- and 1T- mixed phase few-layer MoS2 as a superior to Pt co-catalyst coated on TiO2 nanorod arrays for photocatalytic hydrogen evolution, Appl. Catal., B, 2019, 241, 236–245 CrossRef CAS.
  32. T. Ma, W. Li, J. Li, W. Duan, F. Gao, G. Liao, J. Li and C. Wang, Multisite Cocatalysis: Single atomic Pt2+/Pt0 active sites synergistically improve the simulated sunlight driven H2O-to-H2 conversion performance of Sb2S3 nanorods, J. Colloid Interface Sci., 2024, 658, 476–486 CrossRef CAS PubMed.
  33. R. Wang, L. Shi, E. Y. Konysheva and X. Xu, Selective Photocatalysis of Benzyl Alcohol Valorization by Cocatalyst Engineering Over Zn2In2S5 Nanosheets, Adv. Funct. Mater., 2024, 2418074 Search PubMed.
  34. T. Higashi, K. Seki, Y. Sasaki, Y. Pihosh, V. Nandal, M. Nakabayashi, N. Shibata and K. Domen, Mechanistic Insights into Enhanced Hydrogen Evolution of CrO/Rh Nanoparticles for Photocatalytic Water Splitting, Chem. – Eur. J., 2023, 29, e202204058 CrossRef CAS PubMed.
  35. Q. Li, C. Cui, H. Meng and J. Yu, Visible-Light Photocatalytic Hydrogen Production Activity of ZnIn2S4 Microspheres Using Carbon Quantum Dots and Platinum as Dual Co-catalysts, Chem. – Asian J., 2014, 9, 1766–1770 CrossRef CAS PubMed.
  36. T. Takata, J. Jiang, Y. Sakata, M. Nakabayashi, N. Shibata, V. Nandal, K. Seki, T. Hisatomi and K. Domen, Photocatalytic water splitting with a quantum efficiency of almost unity, Nature, 2020, 581, 411–414 CrossRef CAS PubMed.
  37. B. Neupane, H. Tang, N. K. Nepal, S. Adhikari and A. Ruzsinszky, Opening band gaps of low-dimensional materials at the meta-GGA level of density functional approximations, Phys. Rev. Mater., 2021, 5, 063803 CrossRef CAS.
  38. Y.-N. Zhang, M. Zhang, P. Zhang and W.-B. Hu, Investigation of electronic structure and optoelectronic properties of Si-doped β-Ga2O3 using GGA + U method based on first-principle, Acta Phys. Sin., 2024, 73, 017102-017101–017102-017110 Search PubMed.
  39. H. Xiao, J. Tahir-Kheli and W. A. Goddard III, Accurate Band Gaps for Semiconductors from Density Functional Theory, J. Phys. Chem. Lett., 2011, 2, 212–217 CrossRef CAS.
  40. G. Lin and X. Xu, Ba-Modified LaTiO2N as an Efficient Visible Light Active Photocatalyst for Water Oxidation, ACS Sustainable Chem. Eng., 2020, 8, 9641–9649 CrossRef CAS.
  41. Y. Wang, S. Jin, G. Pan, Z. Li, L. Chen, G. Liu and X. Xu, Zr doped mesoporous LaTaON2 for efficient photocatalytic water splitting, J. Mater. Chem. A, 2019, 7, 5702–5711 RSC.

Footnote

Electronic supplementary information (ESI) available: XRD, XPS, BET, TEM-EDX, photocatalytic water splitting performance, element content, comparisons of photocatalytic performance and fitting data of TRPL. See DOI: https://doi.org/10.1039/d5dt00316d

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.