Next Article in Journal
Catalytic Characterization of Synthetic K+ and Na+ Sodalite Phases by Low Temperature Alkali Fusion of Kaolinite during the Transesterification of Spent Cooking Oil: Kinetic and Thermodynamic Properties
Next Article in Special Issue
MIL-53(Fe)@perylene Diimide All-Organic Heterojunctions for the Enhanced Photocatalytic Removal of Pollutants and Selective Oxidation of Benzyl Alcohol
Previous Article in Journal
The Catalytic Performance of Nanorod Nickel Catalyst in the Hydrolysis of Lithium Borohydride and Dimethylamine Borane
Previous Article in Special Issue
A Critical Review of the Use of Bismuth Halide Perovskites for CO2 Photoreduction: Stability Challenges and Strategies Implemented
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

One Stone Two Birds: Utilization of Solar Light for Simultaneous Selective Phenylcarbinol Oxidation and H2 Production over 0D/2D-3D Pt/In2S3 Schottky Junction

1
Key Laboratory of Green and Precise Synthetic Chemistry and Applications, Ministry of Education, College of Chemistry and Materials Science, Huaibei Normal University, Huaibei 235000, China
2
State Key Laboratory of Photocatalysis on Energy and Environment, Fuzhou University, Fuzhou 350116, China
3
College of Material and Chemical Engineering, Pingxiang University, Pingxiang 337055, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2023, 13(3), 461; https://doi.org/10.3390/catal13030461
Submission received: 30 January 2023 / Revised: 15 February 2023 / Accepted: 20 February 2023 / Published: 22 February 2023

Abstract

:
Precise regulation and control solar-light-driven charges photoexcited on photocatalysts for separation-transfer and target redox reactions is an attractive and challenging pathway toward sustainability. Herein, 0D/2D-3D Pt/In2S3 Schottky junction was fabricated for simultaneous selective phenylcarbinol conversion into value-added aldehydes and production of clean energy H2 by directly utilizing photoexcited holes and electrons in one reaction system under mild reaction conditions. In contrast to pure water splitting and pure In2S3, the reaction thermodynamics and kinetics of H2 evolution on the Pt/In2S3 were significantly enhanced. The optimized 0.3% Pt/In2S3 exhibited the highest and most stable photocatalytic activity with 22.1 mmol g−1 h−1 of H2 production rate and almost 100% selectivity of benzaldehyde production. Notably, this dual-function photocatalysis also exhibited superiority in contrast to sacrificial-agent H2 evolution reactions such as lactic acid, Na2S, methanol and triethanolamine. The turnover frequency (TOF) could reach up to ~2394 h−1. The Pt clusters anchored at the electron location and strong metal-support interactions (SMSI) between Pt and In2S3 synergistically improved the spatial charge separation and directional transportation (~90.1% of the charge transport efficiency could be achieved over the Pt/In2S3 hybrid), and thus result in significant enhancement of photocatalytic H2 evolution with simultaneous benzaldehyde production.

1. Introduction

Since Fujishima and Honda reported electrochemical photolysis of water for hydrogen (H2) production at a TiO2 electrode in 1972 [1], photocatalytic H2 production (PHP), as one of the most promising strategies to address the severe issues of environment and energy, has attracted extensive and ongoing attention [2,3,4,5], because PHP can be driven by inexhaustible solar energy and the reaction conditions are not as rigorous as traditional industrial methods such as coal gasification and electrolytic processes [3,4,5,6,7]. For instance, 15,364 scientists from 184 countries made a joint appeal to humans in 2017: “World Scientists’ Warning to Humanity: A Second Notice”. One of the noteworthy appeals was sustainable development [8]. Clean energy instead of fossil fuels is ineluctable in the future. More recently, European Union and other countries have made incentive schemes for green hydrogen fuel. However, PHP faces many challenges for practical application [9,10,11,12]. Two critical points are the design of efficient photocatalysts with high and stable quantum efficiency and the enhancement of output-input ratio. For photocatalytic overall water splitting into hydrogen (H2) and oxygen (O2), a lot greater than zero of the Gibbs free energy change and the sluggish oxidation half-reaction both make PHP hard in terms of thermodynamics and kinetics [12,13,14]. Although sacrificial-reagent PHP is in favor of improvements of both reaction thermodynamics and kinetics, sacrificial reagents simultaneously bring waste of the photoexcited holes, the increase in cost and the burden on the environment such as emissions of greenhouse-gas, inorganic salt and organic pollutants [14,15,16]. Moreover, the charge carrier recombination is still one of the challenging limitations for the photocatalysis technologies [17,18,19,20,21,22,23,24]. Recently, PHP coupled with organics transformation has held great attention [25,26,27,28,29,30]. In this dual-function photoredox reaction system, not only can the photoexcited electrons be utilized for reducing H+/H2O into H2, the photoexcited holes can also be used for oxidizing organics into fine chemicals. For instance, aromatic aldehydes and H2 can be simultaneously obtained by the photocatalytic splitting of aromatic alcohols in one reaction system [31,32,33]. However, the photocatalysts suitable for this dual-function photoredox reaction system with efficient reaction kinetics still need to be explored.
To drive this dual-function photoredox reaction, choosing photocatalysts with a proper band gap and suitable band positions is the initial step. Among various photocatalysts, low-dimensional metal sulfides showed tremendous potential for this dual-function photoredox reaction because of their appealing optical–electrical characteristics and appropriate band structures [34,35,36]. However, metal sulfides used as photocatalysts still face various problems: low utilization of visible light, photocorrosion and recommendation of photoexcited charge carriers, which significantly inhibit its reaction kinetics and stability [37,38,39]. To address the problems, many approaches have been developed such as doping [40,41], noble-metal deposition [30,42,43], cocatalysts [5,31,44] and heterojunction composites [45,46,47]. After modification, the photocatalytic activity and stability of the pristine metal sulfides (Zn3In2S6, CdS, ZnxCd1-xS and ZnIn2S4) both are improved. However, precise regulation and control solar-light-driven charges photoexcited on photocatalysts for separation-transfer and target redox reactions is still a challenge.
Herein, 0D/2D-3D Pt/In2S3 heterostructure was prepared by sequential hydrothermal-photodeposition methods and was applied for PHP with simultaneously selective phenylcarbinol conversion under simulated sunlight irradiation. In the previous study, In2S3 exhibited potential applications in photocatalytic pollutant degradation [48], selective oxidation [49,50], H2 production [22,23,24], etc. It may have been an alternative photocatalyst for this dual-function photoredox reaction. In addition, In2S3 possesses smaller band-gap energy (~2.0 eV) [51] than Zn3In2S6 (~2.9 eV) [14], CdS (~2.4 eV) [52,53], Zn0.5Cd0.5S (~2.6 eV) [45] and ZnIn2S4 (~2.4 eV) [31,54], suggesting more and broad light absorption. The 0D/2D-3D Pt/In2S3 hierarchical structure has the following advantages: 2D nanosheets and 3D spheres of In2S3 hierarchical structure facilitate light harvesting via multi-layer reflection, 0D Pt deposition and close contact, and sedimentary separation from the reaction system. On the other hand, the exposed 0D Pt clusters can make full use of Pt atoms and save costs. Therefore, the 0D/2D-3D hierarchical structure is significant for PHP. Moreover, it has been demonstrated that the Schottky junction can improve charge separation [55]. In this study, 0D Pt clusters were anchored at the separated electron location of In2S3 by an in situ photoreduction process. The formed Pt/In2S3 Schottky junction coupled with strong metal–support interactions (SMSI) between 0D Pt clusters and 2D In2S3 nanosheets can improve the electron separation and transportation from In2S3 into Pt for PHP and reserve the holes at In2S3 for selective oxidation of phenylcarbinol, and thus result in significant enhancement of PHP with almost 100% selectivity of benzaldehyde production. Notably, benzaldehyde is important for chemical raw material, methylene reagents, perfume, herbicide intermediates, etc. In addition, the as-prepared 0D/2D-3D Pt/In2S3 heterostructure exhibits superiority for PHP coupled with phenylcarbinol in contrast to sacrificial agents such as lactic acid, Na2S, methanol and triethanolamine. Moreover, the photocatalytic mechanism was also studied profoundly by several recognized techniques such as the photoelectrochemical (PEC) test, in situ electron paramagnetic resonance (EPR) and in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFT), etc.

2. Results and Discussion

2.1. Catalysts Characterization

The micro-structures of 2D-3D In2S3 and 0D/2D-3D Pt/In2S3 were characterized by scanning electron microscope (SEM), transmission electron microscope (TEM) and high-resolution TEM (HRTEM). As shown in Figure 1a–c, the In2S3 presented spherical-like morphology. The 3D sphere was further composed of many crisscross nanosheets. Clearly, the as-prepared In2S3 possessed a 2D-3D hierarchical structure. Compared to the pervious reported 2D In2S3 nanosheets [49] and 0D In2S3 nanocrystal [48], the 2D-3D In2S3 can facilitate light harvesting via multi-layer reflection. It also has advantages with respect to the previous method of 2D-3D In2S3 preparation [50], in which an amino acid (aspartic acid, serine or glycine) was necessary for assisting formation of 2D-3D hierarchical structure. Notably, the 2D-3D nanosheet-sphere structure of In2S3 was kept after loading Pt (Figure 1d–f). The Pt nanoparticles were not observed on the SEM image, which may have been caused by the small size of Pt. The micro-structures of In2S3 and Pt/In2S3 were further investigated by TEM and HRTEM images. As displayed in Figure 2a,b, the nanosheet-sphere structure of 2D-3D In2S3 can be clearly observed and the nanosheets of In2S3 were uniform and homogeneous. Moreover, the unambiguous lattice fringes with the d-spacing value of 0.62 nm correspond to the (111) crystal plane of cubic In2S3 (Figure 2c,d). After Pt photo-deposited on In2S3, the morphology of Pt/In2S3 was very similar to that of In2S3 (Figure 2e). In addition, small and distinguishable Pt clusters with a mean size of about 0.8 nm were observed on the nanosheets of In2S3 (Figure 2f). The d-spacing value of the distinct lattice fringes was also 0.62 nm, which was assigned to the (111) crystal plane of cubic In2S3 (Figure 2g,h). The energy-dispersive X-ray spectrometer (EDX) results indicate that Pt/In2S3 was composed of Pt, In and S elements, and Pt was uniformly dispersed on the In2S3 (Figure 2i,j).
The crystal phase, chemical composition and state were studied by powder X-ray diffraction (PXRD) and X-ray photoelectron spectroscopy (XPS). The Pt/In2S3 hybrid exhibited the similar PXRD pattern to the pristine In2S3 (Figure 3a), and the diffraction peaks of In2S3 and Pt/In2S3 both can be indexed to cubic β-In2S3 with Fd-3m(227) space group (JCPDS No. 65-0459). No Pt diffraction peaks were observed in the PXRD pattern of Pt/In2S3. It was expected because of the cluster state of Pt, i.e., due to the dispersion and low content of Pt. The ICP-OES indicated that the practical weight ratio of Pt in 1% Pt/In2S3 was about 0.7%. The PXRD peaks located at about 14.2, 23.3, 27.4, 28.7, 33.2, 36.3, 41.0, 43.6, 47.7, 50.0, 55.9, 56.6, 59.3, 66.6, 69.7, 77.1 and 79.5° were attributed to the diffraction of the (111), (220), (311), (222), (400), (331), (422), (511), (440), (531), (533), (622), (444), (731), (800), (662) and (840) crystal planes of cubic β-In2S3 (JCPDS No. 65-0459, a = b = c = 10.77 Å), respectively. The PXRD results were consistent with the above HRTEM analysis (Figure 2c,g). In the light of the above results, it can be seen that the 2D-3D morphology and crystal phase of In2S3 did not change after the deposition of Pt clusters. These results also indicate the stability of Pt/In2S3 because Pt/In2S3 was obtained in the PHP process of In2S3 by reducing [PtCl6]2−. Figure 3b–d presents the high-resolution XPS spectra of In 3d, S 2p and Pt 4f, respectively. For the pure In2S3, two peaks of In 3d observed at 445.2 and 452.7 eV were attributed to In 3d5/2 and In 3d3/2 (Figure 3b), and two peaks of S 2p located at 161.9 and 163.1 eV belonged to S 2p3/2 and S 2p1/2 (Figure 3c), respectively.
Moreover, the spin-orbit separations of In 3d and S 2p were 7.5 and 1.2 eV, respectively. These results demonstrate that the chemical states of In and S in the as-prepared In2S3 were In3+ and S2−. For the Pt/In2S3 hybrid, the XPS peaks of In 3d and S2p were similar to that of the pure In2S3. The Pt 4f exhibited two group peaks at 71.9 and 72.1 eV (4f7/2), which corresponded to the Pt0 and Pt2+, respectively (Figure 3d). Of note, the binding energies of In 3d and S 2p of the Pt/In2S3 hybrid were shifted to high energy (0.1–0.2 eV) with respect to the pure In2S3. It was demonstrated that the binding energy shift was derived from the electronic interaction between two contacted nanomaterials, and the positive and negative shifts mean electrons were lost and gathered, respectively [31,56,57]. Thus, the strong metal-support interactions (SMSI) occurred in the Pt/In2S3 hybrid. Specifically, after Pt photo-deposited on In2S3, the majority carriers (electrons for n-type semiconductors) of In2S3 were migrated into Pt, and the Pt was electron enriched. The strong metal-support interactions could result in photoexcited charge separation and H2 evolution conveniently. This is discussed further below.
The Brunauer–Emmett–Teller (BET) surface areas, optical properties and band-energy positions of In2S3 and Pt/In2S3 were studied by the nitrogen adsorption–desorption method, UV-vis diffuse reflectance spectroscopy (UV-vis DRS) and Mott–Schottky (M-S) measurements. Both In2S3 and Pt/In2S3 presented type-IV isotherms with an H3 hysteresis loop (Figure 4a). This meant that the presence of porous structures resulted from the 2D-3D In2S3 hierarchical structure. Correspondingly, the average pore diameters of In2S3 and Pt/In2S3 were about 13.91 and 13.64 nm, respectively. The BJH cumulative volume of pores of In2S3 and Pt/In2S3 were approximately 0.41 and 0.30 cm3 g−1, respectively. The BET surface areas of In2S3 and Pt/In2S3 were approximately 63.1 ± 2.1 and 63.3 ± 2.9 m2 g−1, respectively. Evidently, after Pt clusters were loaded on 2D-3D In2S3, the surface area change was negligible, while the pore volume and pore diameter were decreased. It is normal to observe these results because the 2D-3D In2S3 hierarchical structure was not altered when Pt clusters were loaded on the surfaces of In2S3 nanosheets. The UV-vis DRS spectra indicate that both In2S3 and Pt/In2S3 possessed well visible light absorption below 600 nm (Figure 4b). Based on DRS spectra, the band-gap energy (Eg) was determined by the Kubelka–Munk function: (αhν)2 = A(hν − Eg), where A, h, α, and ν are proportionality constant, Planck constant, absorption coefficient and frequency, respectively [33,58]. Compared to the pure In2S3, the light absorption of Pt/In2S3 diminished (Figure 4b), while the Eg of Pt/In2S3 showed no noticeable change (both about 2.1 eV, Figure 4c). Moreover, the positive slops of M-S plots were observed on both In2S3 (Figure 4d) and Pt/In2S3 (Figure 4e), indicating that the n-type property of In2S3 semiconductor had not changed after Pt deposition. Interestingly, the flat potential of In2S3 was negatively shifted from −0.25 to −0.51 V (vs NHE) after Pt deposition. Generally, for n-type semiconductors, the flat potential lies beneath the conduction band (CB) at about −0.1 eV [56]. Therefore, the CB of In2S3 and Pt/In2S3 was located at −0.35 and −0.61 eV, respectively. According to the function: Eg = EVB − ECB (EVB and ECB were the valence band energy and the CB energy, respectively), the valence band (VB) of In2S3 and Pt/In2S3 lies at 1.75 and 1.49 eV, respectively. Evidently, with respect to the pure In2S3 (Figure 4f), VB and CB of Pt/In2S3 were uplifted by 0.26 and 0.26 eV, respectively, demonstrating that the photoexcited holes showed faster mobility, thus facilitating PHP reaction [31,59]. The strong metal–support interactions between Pt and In2S3 resulted in electron migration from In2S3 into Pt. Thus, the energy bands of In2S3 swept upward when In2S3 was contacted with Pt.
To gain more insights into the charge carrier transportation between Pt and In2S3 over the Pt/In2S3 hybrid, the work function of In2S3 was measured by an ultraviolet photoelectron spectroscopy (UPS), As presented in Figure 5a, the work function of In2S3 was 4.66 eV (21.22 − (16.77 − 0.21) = 4.66). It was smaller than the work function of Pt (5.65 eV) [14]. Moreover, the carrier densities (ND) of In2S3 and Pt-modified In2S3 (Pt/In2S3) were detected from M-S plots via the function: ND = (2/eεε0)[dUFL/d(1/C2)] = (2/eεε0) (1/kM-S). Here, e, ε0, ε, kM-S, UFL and C are elementary charge, vacuum permittivity, relative permittivity (8.4 for In2S3 [23]), the slope of the M-S curve, Fermi level potential and capacitance, respectively. Evidently, after Pt deposition on In2S3, the carrier density of In2S3 was always reduced under different frequencies (Figure 5b). Based on the results of work function and carrier density, the metal–support interactions and consequential electron transportation between Pt and In2S3 are illustrated in Figure 5c–e. The In2S3 possessed a higher Fermi level (EF = Evac − WF, where WF, Evac and EF are work function, vacuum level and Fermi level, respectively) than Pt (Figure 5c). Thus, the electrons were transported from In2S3 into Pt, which resulted in a ND decrease in In2S3 and the formation of the Schottky junction (Figure 5d). The Schottky barrier with height of 0.99 eV would facilitate photoexcited electrons transportation from In2S3 into Pt and inhibit the backflow of electrons from Pt into In2S3 again (the barrier height is the difference of Fermi levels of In2S3 and Pt (−4.66 − (−5.65) = 0.99)). Simultaneously, the photoexcited holes left at In2S3 (Figure 5e). Consequently, the photoexcited electron-hole pairs of In2S3 were separated spatially through Pt/In2S3 Schottky junctions. In the light of the above analyses, the Pt/In2S3 hybrid may be fit for PHP.

2.2. Evalution of PHP Activity

The PHP activity was evaluated by photocatalytic selective oxidation of phenylcarbinol (PhCH2OH) under simulated sunlight. The two control groups (with photocatalyst in the dark and without photocatalyst under light irradiation) were firstly performed and showed no H2 production. Then, we detected PHP activities of In2S3 and Pt/In2S3 composites with different content of Pt (0.1%, 0.3%, 0.5% and 1%). As depicted in Figure 6a, the pure In2S3 exhibited low PHP activity (H2: 0.9 mmol g−1 h−1). However, the PHP activity of the In2S3 was significantly enhanced by loading a low amount of Pt. The photocatalytic H2 evolution rates of 0.1%, 0.3%, 0.5% and 1% Pt/In2S3 hybrids were about 4.1, 22.1, 17.1 and 14.6 mmol g−1 h−1, respectively. The PHP activities of these Pt/In2S3 hybrids appeared to have a volcano-like distribution. The 0.3% Pt/In2S3 hybrid exhibited the highest PHP activity, which was approximately 24.56 times that of the pure In2S3. It indicates that the apparent PHP kinetics of In2S3 was meaningfully improved by loading Pt clusters. The as-synthesized Pt/In2S3 also exhibited a higher H2 production rate (7.97 mmol g−1 h−1) under visible light irradiation than the reported Pt/Zn3In2S6 [14], Pt/CdS [42], etc [21,22,23,24]. (Table 1). Moreover, negligible PHP activity was detected for pure water splitting over 0.3% Pt/In2S3 because of the sluggish oxidation half-reaction and significant Gibbs free energy change (H2O = H2 + 1/2O2, ΔG ≈ 238 kJ mol−1 >> 0). Compared to PHP through overall water splitting, the thermodynamics of PHP was also remarkably ameliorated (PhCH2OH = PhCHO + H2, ΔG ≈ 28 kJ mol−1). It should also be noted that the oxidized products of phenylcarbinol over the 0.3% Pt/In2S3 hybrid were almost entirely benzaldehyde (the selectivity was nearly 100%). It may have been caused by the suitable oxidative potential of Pt/In2S3 for selectively oxidizing PhCH2OH into PhCHO. Evidently, in the dual-function photoredox reaction system, Pt/In2S3 not only can selectively oxidize phenylcarbinol into fine value-added chemicals (benzaldehyde) with high selectivity but also can obtain clean energy (H2) simultaneously. Moreover, the turnover number (TON) based on the amount of Pt was calculated to be about 9576 after 4 h. Notably, the corresponding turnover frequency (TOF) was about 2394 h−1, which is comparable with the traditional thermal catalytic system (1109 h−1) [27,60].
Figure 6b shows the PHP activity of 0.3% Pt/In2S3 under monochromatic light with different wavelengths. The 380 nm-light and 500 nm-light driven PHP activities were higher than 400 nm. It indicates that the PHP activity of Pt/In2S3 was not only dependent on its light absorption spectrum. It is comprehensible because the PHP activity was an overall effect of light absorption, incident light intensity and light energy. Pt/In2S3 was inactive under 600 nm light because it was longer than the excitation wavelength of In2S3 (<590 nm). Nevertheless, the PHP activity of Pt/In2S3 under 500 nm light could still reach up to 3.2 mmol g−1 h−1 with 3.72% of apparent quantum efficiency. In addition, the 0.3% Pt/In2S3 hybrid demonstrated good photocatalytic stability with little H2 production decrease (<0.5%) after five recycles (Figure 6c). To compare this dual-function photoredox reaction system with the sacrificial agent PHP, the classic sacrificial agents: lactic acid (LA), Na2S (NS), triethanolamine (TA) and methanol (MA) were chosen [20]. Figure 6d depicts PHP comparisons between sacrificial agents and PhCH2OH over 0.3% Pt/In2S3 under the same reaction conditions. Specifically, the PHP rates of 0.3% Pt/In2S3 only reached 1.1, 0.6, 0.5 and 0.003 mmol g−1 h−1 when LA, NS, TA and MA were added into the reaction system, respectively. Clearly, the superior PHP rate of 0.3% Pt/In2S3 was achieved through using phenylcarbinol, which was about 20.1, 36.8, 44.2 and 7366.7 times higher than that using LA, Na2S and TEOA, respectively. These results confirm the superiority of photocatalytic selective conversion organics coupled with photocatalytic H2 production, in which organics were selectively transformed into high value-added chemicals and simultaneous H2 with the enhanced production rate that could be obtained.

2.3. Photocatalytic Mechanism

The photocatalytic activity demonstrated that it was mainly influenced by three factors: light absorption, active sites (likely surface area) and photoexcited charges separation and transportation [61,62,63]. Specifically, the photocatalysts are excited by the incident light to produce electron-hole pairs, which are then separated and transferred onto the active sites for redox reactions. Based on the above characterizations, after Pt clusters deposition on the 2D-3D In2S3, the light absorption did not become stronger (Figure 4b), and the surface area underwent a negligible change (Figure 4a). The surface area was not the main factor for the enhanced PHP activity, which is consistent with the reported Pt-loaded photocatalysts [64,65]. Therefore, light absorption and surface area are not the main factors for the boosted PHP activity. However, the electron circulating and the uplifted energy bands were observed on Pt/In2S3 hybrid as the result of the strong metal–support interactions between In2S3 and Pt. To understand the reasons behind the enhanced photocatalytic activity, the photoexcited charge behaviors were investigated. As shown in Figure 7a, the photocurrent of the Pt/In2S3 improved 2.08 times in contrast to the pure In2S3 under simulated sunlight illumination, suggesting efficient charge separation and transfer [66,67]. In addition, the photocurrent of the Pt/In2S3 still increased 1.85 times when MVCl2 was added into the bath solution as an electron scavenger (Figure 7b). Moreover, the charge transport efficiency (ηtra) can be evaluated by the function: ηtra = JH2O/JMVCl2 (JH2O and JMVCl2 are the photocurrent densities of the sample with and without MVCl2, respectively) [68,69,70]. As expected, 90.1% of the charge transport efficiency could be achieved over the Pt/In2S3 hybrid, which was approximately 1.13 times of the pure In2S3. These results indicate that the separation and transportation of In2S3 can be improved by loading Pt clusters. To further evaluate the impact of strong metal–support interactions on charge separation and transportation, the electrochemical impedance spectroscopy (EIS) [71] and linear sweep voltammetry (LSV) tests were carried out [5,14]. Pt/In2S3 exhibited a smaller arc radius than In2S3 (Figure 7c), and the charge transport resistance of Pt/In2S3 (30.5 Ω) was weaker than that of In2S3 (33.9 Ω). This means that the loaded Pt clusters can speed charge separation and transportation of In2S3. This result is in line with the photocurrent analysis and can be further confirmed by LSV curves. As displayed in Figure 7d, compared to the pure In2S3, the current density of Pt/In2S3 exhibited a visible enhancement under light irradiation. In addition, the H2 evolution overpotential of Pt/In2S3 (−0.64 V) was 0.34 V lower than that of In2S3 (−0.98 V), which is conducive to H2 evolution. Consequently, it can be concluded that the improved charge separation-transportation and the reduced H2 evolution overpotential contribute to the efficient PHP activity of Pt/In2S3.
To understand the in-depth information behind these results, the utilization rates of photoexcited electrons (e) and holes (h+) were studied through in situ electron paramagnetic resonance (EPR) measurements [68]. As the control group, photolysis refers to the reaction system without photocatalysts (Figure 7e,f). In other words, the signal of photolysis is the intrinsic signal of the active TEMPO. The EPR signal intensity is reduced when TEMPO is captured by photoexcited electrons or holes [68]. When In2S3 or Pt/In2S3 was added into the reaction system, the EPR signals were both reduced for detecting electrons (Figure 7e) and holes (Figure 7f). This indicates that the photoexcited electrons and holes can be separated and transported on the surfaces of In2S3 and Pt/In2S3. Notably, the EPR signal of Pt/In2S3 for photoexcited electrons was significantly lower than that of In2S3 (Figure 7e). This suggests efficient electron separation and transportation from In2S3 to Pt for reducing water/protons to H2. In addition, the weaker EPR signal of TEMPO on Pt/In2S3 was also observed than that on In2S3 in the presence of PhCH2OH under simulated sunlight illumination (Figure 7f). This indicates efficient hole transportation from the In2S3 component of Pt/In2S3 to reactive molecules of PhCH2OH. Thus, the efficient separation and transportation of the photogenerated holes and electrons contribute to the enhanced PHP activity of the Pt/In2S3 hybrid. The photoexcited holes can be fleetly consumed by PhCH2OH to produce PhCHO. Simultaneously, the photoexcited electrons were spent by H+/H2O to produce H2.
To further inspect the conversion process of PhCH2OH in this dual-function photocatalysis system, the in situ EPR with the addition of DMPO and in situ DRIFT were carried out [5,31,72,73,74]. As presented in Figure 8a, PhCH2OH with DMPO under light irradiation could not produce EPR signals (photolysis). However, sextet peaks belonging to carbon-centered radicals (.CH(OH)Ph) [5,68] were observed on In2S3 and Pt/In2S3. This means that the conversion process of PhCH2OH is a free radical reaction. Moreover, the EPR intensity of Pt/In2S3 was more intense than In2S3, implying efficient charge separation-transportation and fast PhCH2OH dehydrogenation on Pt/In2S3. In addition, one peak at 1703 cm−1C=O) fell to the carbonyl group (C=O) of benzaldehyde (PhCHO) and doublet peaks at 2873 and 2935 cm−1C-H), attributed to the carbon–hydrogen bond (C-H) of the aldehyde group, were clearly observed on Pt/In2S3 under simulated sunlight (Figure 8b). These results indicate that PhCH2OH is selectively oxidized into PhCHO via a carbon-centered radical process. The effects of the reactive species on PHP were also investigated by the trapping experiments (Figure 8c). Triethanolamine (TA) and carbon tetrachloride (CTC) were used as trapping agents for photoexcited holes and electrons, respectively. When TA was added into the reaction system, the H2 production rate decreased. This indicates that the dehydrogenation of PhCH2OH to H2 production is restrained by TA. For the trapping agent CCl4, a relatively large decrease was observed in the H2 production rate. This indicates that the photogenerated electrons are major reductive species for the reduction in protons to H2. These results suggest that synergistic effect occurred between PhCH2OH dehydrogenation and H2 production.
From the above analysis, the photocatalytic mechanism was proposed as depicted in Figure 8c. Under simulated sunlight irradiation, the photoexcited electrons and holes were generated on the In2S3 nanosheets. Then, on the one hand, PhCH2OH was oxidized into .CH(OH)Ph, and the .CH(OH)Ph free radical was further oxidized into PhCHO by the photoexcited holes located at In2S3. On the other hand, the photoexcited electrons were first consumed by [PtCl6]2− to produce Pt clusters and were then separated and transported from In2S3 into Pt clusters efficiently to produce H2 by reducing H+/H2O. During this redox process, [PtCl6]2− was reduced by the photoexcited electrons on the surfaces of In2S3, and Pt0 was anchored at the separated electron location. Thus, the as-synthesized Pt/In2S3 heterostructure would facilitate the electron transportation from In2S3 into Pt and improve PHP activity. Due to the competing reactions of H2 evolution, the Pt cluster was in the optimized state for H2 production. During the coupled redox reaction of PhCH2OH oxidation and H2 evolution, one molecule of PhCH2OH was oxidized into one molecule of PhCHO by consuming two photoexcited holes. Simultaneously, one molecule of H2 was produced by expending two electrons. Thus, the efficient, stable and atom-economic dual-function photocatalytic reaction system was achieved on the 0D/2D-3D Pt/In2S3 heterostructures.

3. Experiments and Methods

3.1. Materials

5,5-dimethyl-1-pyrroline-N-oxide (DMPO) and 2,2,6,6-Tetramethylpiperidin-1-oxyl (TEMPO) for EPR-spectroscopy were purchased from Sigma-Aldrich. Indium chloride tetrahydrate (InCl3·4H2O, 99.9%), thioacetamide (C2H5NS, ≥99.0%), chloroplatinic acid hexahydrate (H2PtCl6·6H2O, Pt ≥35.7%), potassium ferrocyanide trihydrate (K4[Fe(CN)6]·3H2O, ≥99.5%), phenylcarbinol (C7H8O, ≥99.0%), lactic acid (C3H6O3, 85%), triethanolamine (C6H15NO3, ≥99.0%), sodium sulfate (Na2SO4, ≥98%), potassium chloride (KCl, 99.8%), potassium ferricyanide (K3[Fe(CN)6], 99%) and other used reagents were all analytical reagents and were used directly with no further purification.

3.2. Preparation of 2D-3D In2S3 and 0D/2D-3D Pt/In2S3

The 0D/2D-3D Pt/In2S3 heterostructure was prepared as depicted in Figure 9. Briefly, InCl3 was wholly dissolved into acid solution and reacted with thioacetamide (TAA) to form [In(TAA)4]3+ and [In(TAA)6]3+ complexes via In-S bonds [75]. Then, these complexes underwent the hydrothermal process to produce 2D-3D In2S3. In2S3 was easily formed because the solubility product constant (Ksp) of In2S3 was very small (5.7 × 10−74) [75]. Finally, Pt/In2S3 heterostructure was obtained by an in situ photodeposition process. Namely, the Pt/In2S3 was synthesized in the process of PHP coupled with simultaneously selective phenylcarbinol conversion.
In a typical synthesis, 1 mmol InCl3·4H2O was dissolved in deionized water, and the pH of the InCl3 solution was adjusted to 1.0 by adding HCl to prevent InCl3 hydrolysis. Then, 2.5 mmol C2H5NS was gradually added into the above solution and constantly stirred. The pH of the above solution was adjusted to 3.0 again by adding H2O. The above solution was transferred to a 100 mL Teflon-lined stainless-steel reactor and held at 180 °C for 24 h. After natural cooling to 25 °C, the orange precipitate was collected and washed with distilled water and anhydrous ethanol several times. Finally, the sediments were dried in a vacuum oven at 60 °C for 2 h. The pure 2D-3D In2S3 was obtained. The 0D/2D-3D Pt/In2S3 heterostructure was obtained by an in situ photodeposition method in the process of photocatalytic selective conversion of phenylcarbinol into benzaldehyde and H2. The details are presented in the following section.

3.3. Photocatalytic Activity Test

The photocatalytic H2 generation was carried out in a gas-tight Pyrex reactor. The 300 W Xenon lamp (PLS-SXE300D, Perfect Light Co., Beijing, China) was used as the simulated solar light. Typically, 10 mg In2S3 powders were dispersed in 10 mL phenylcarbinol solution and then different amounts of H2PtCl6·6H2O were added. After bubbling argon to remove dissolved oxygen, the suspension was irradiated for photocatalytic H2 production. After irradiation for 2 h, the H2 was quantified using a gas chromatograph spectrometer (GC 9790II, Fuli, Wenling, China) equipped with a molecular sieve 5A column. The reaction liquor was detected by high performance liquid chromatography (HP-LC, watersE2695, MA, USA). The detector of the HP-LC was PDA 2998. The mobile phase consisted of 40% deionized water and 60% acetonitrile with a flow rate of 1 mL min−1. Finally, the precipitate (Pt//In2S3) after light exposure was collected, washed with ethanol and dried at 60 °C for 2 h. Catalysts with different Pt content added were rewritten as x% Pt/In2S3 (x is a weight ratio of Pt in the Pt/In2S3 composite, x = 0.1, 0.3, 0.5, 1.0). For comparison, the pure In2S3 was also quantitatively analyzed for H2 production without adding H2PtCl6·6H2O. The apparent quantum efficiency (AQE) for H2 evolution was obtained by the following equation: AQE = (2 × NH/Np) × 100%, where NH and Np are the numbers of evolved H2 molecules and incident photons, respectively. Turnover number (TON) was calculated based on the quantity of H2 and Pt: TON = NH/Npt, where Npt is the number of Pt. The turnover frequency (TOF) was measured via TON divided by reaction time. Benzaldehyde selectivity was calculated by the equation: Selectivity = [CCHO/(C0-COH] × 100%, where C0, COH and CCHO are the concentrations of phenylcarbinol, the residual phenylcarbinol and the corresponding aldehydes, respectively.

3.4. Characterization

Powder X-ray diffraction (PXRD) pattern of the sample was determined by a Bruker D8 X-ray powder diffractometer using Ni-filtered Cu Kα radiation. The microstructure and morphologies of the prepared samples were carried out by scanning electron microscope (SEM, Regulus 8200, Hitachi Limited, Tokyo, Japan) and transmission electron microscope (TEM, JEM2100, JEOL, Akishima-shi, Japan). Elemental mappings were measured using an energy-dispersive X-ray spectrometer (EDX). X-ray photoelectron spectroscopy (XPS, Thermo Scientific, Massachusetts, America) measurements were performed on Thermo Scientific ESCA Lab250 spectrometer. All the binding energies were referred to the C 1s peak at 284.6 eV of the surface adventitious carbon. UV-vis diffuse reflectance spectra (DRS) of the powders were obtained on a UV-vis spectrophotometer (Shimadzu UV-3600, Kyoto, Japan), with BaSO4 used as a reference. Brunauer–Emmett–Teller (BET) surface areas were tested on a Micromeritics ASAP 2460 instrument. In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFT) measurements were detected on a Nicolet 8700 FTIR spectrometer. The carbon-centered radicals and the photoexcited charges (electrons and holes) were in situ seen on an electron paramagnetic resonance (EPR, A300, Karlsruhe, Bruker, Germany) by DMPO and TEMPO as trapping agents, respectively. The work function of the In2S3 was obtained on an ultraviolet photoelectron spectroscopy (UPS, Thermo ESCALAB 250XI, (Waltham, MA, USA). The actual Pt content in the Pt/In2S3 sample was measured by an inductively coupled plasma optical emission spectrometry (ICP-OES, Agilent 5110, Santa Clara, CA, USA).

3.5. Photoelectrochemical Property Test

The photoelectrochemical (PEC) tests were carried out on a CHI-660E electrochemical workstation (CH Instruments, Bee Cave, TX, USA). An Ag/AgCl and a Pt wire were used as the reference electrode and the counter electrode, respectively. The sample powder was deposited on the FTO (50 mm × 50 mm) as a working electrode. Typically, a uniform solution was obtained by ultrasonically dispersing 5 mg samples into 400 μL deionized water. Then, 20 μL of the above solution were deposited on the FTO substrate. The working electrode was obtained after drying at room temperature. The transient photocurrent responses, linear sweep voltammetry (LSV) plots and Mott–Schottky (M-S) plots were detected in a 0.2 M Na2SO4 aqueous solution. Electrochemical impedance spectroscopy (EIS) Nyquist plots were detected in 0.1 M KCl solution containing 0.1 M K3[Fe(CN)6]/K4[Fe(CN)6].

4. Conclusions

In summary, a 2D-3D In2S3 hierarchical structure decorated by 0D Pt clusters was successfully fabricated by the sequential hydrothermal process and in situ photodeposition. The strong metal–support interactions (SMSI) of the Pt/In2S3 hybrid improved the charge separation and transportation. and thus. resulted in the significant enhancement of photocatalytic H2 production. The optimized 0.3% Pt/In2S3 exhibited the highest and stable photocatalytic activity with 22.1 mmol g−1 h−1 of H2 evolution rate and almost 100% selectivity of benzaldehyde production. In addition, the turnover frequency of 0.3% Pt/In2S3 reached up to approximately 2394 h−1, and 3.72% of apparent quantum efficiency was achieved under 500 nm light irradiation. Coupling phenylcarbinol conversion with H2 evolution was superior to the traditional sacrificial agents. The H2 production using phenylcarbinol was approximately 20.1, 36.8, 44.2 and 7366.7 times higher than that using lactic acid, Na2S, triethanolamine and methanol as sacrificial agents under the same reaction condition, respectively. Notably, in this dual-function photocatalysis, the photoexcited holes located at the In2S3 were utilized for selective oxidizing phenylcarbinol into value-added fine chemicals benzaldehyde; conversely, the photoexcited electrons on the In2S3 were used firstly for reducing [PtCl6]2- to fabricate Pt clusters anchored at the separated electron location and then transported from the In2S3 to the Pt clusters for H2 production. The Pt clusters were stable, and the charge transport efficiency of In2S3 reached up to approximately 90.1% by the modification of the Pt clusters. Moreover, the synergistic effect occurred between PhCH2OH dehydrogenation and H2 production. This work is expected to aid the design of efficient and stable photocatalysts to simultaneously utilize photoexcited holes and electrons, thereby gaining the value-added fine chemicals and clean energy in one reaction system.

Author Contributions

H.Z.: investigation, data curation, writing—original draft preparation. P.X.: formal analysis, resources, writing—original draft preparation. B.L.: resources, methodology. Q.L.: software. X.Z.: resources, formal analysis, writing—review and editing. S.Z.: resources. Z.R.: validation. S.C.: funding acquisition, formal analysis, writing—review and editing, supervision. S.M.: Conceptualization, formal analysis, writing—review and editing, supervision, project administration, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (NSFC, 52002142, 51972134, 52272297 and 52002142), the Foundation of Anhui Province for Distinguished Young Scholars (2022AH020038), the Foundation of Anhui Province for Outstanding Young Graduate-student Advisors (2022yjsds036), Open Project Program of the State Key Laboratory of Photocatalysis on Energy and Environment (SKLPEE-KF201804 and SKLPEE-KF202202) in Fuzhou University, the Natural Science Foundation of Anhui Province (2108085MB43), the University Natural Science Research Project of Anhui Province (KJ2021A0524).

Data Availability Statement

Data only available upon request from corresponding author.

Conflicts of Interest

The authors declared that there is no conflict of interest.

References

  1. Fujishima, A.; Honda, K. Electrochemical Photocatalysis of Water at a Semiconductor Electtode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  2. Nishiyama, H.; Yamada, T.; Nakabayashi, M.; Maehara, Y.; Yamaguchi, M.; Kuromiya, Y.; Tokudome, H.; Akiyama, S.; Watanabe, T.; Narushima, R.; et al. Photocatalytic solar hydrogen production from water on a 100-m2 scale. Nature 2021, 598, 304–307. [Google Scholar] [CrossRef]
  3. Liu, L.; Du, S.; Guo, X.; Xiao, Y.; Yin, Z.; Yang, N.; Bao, Y.; Zhu, X.; Jin, S.; Feng, Z.; et al. Water-Stable Nickel Metal-Organic Framework Nanobelts for Cocatalyst-Free Photocatalytic Water Splitting to Produce Hydrogen. J. Am. Chem. Soc. 2022, 144, 2747–2754. [Google Scholar] [CrossRef]
  4. Rasool, M.A.; Sattar, R.; Anum, A.; Al-Hussain, S.A.; Ahmad, S.; Irfan, A.; Zaki, M.E. An Insight into Carbon Nanomaterial-Based Photocatalytic Water Splitting for Green Hydrogen Production. Catalysts 2023, 13, 66. [Google Scholar] [CrossRef]
  5. Meng, S.; Chen, C.; Gu, X.; Wu, H.; Meng, Q.; Zhang, J.; Lei, W. Efficient Photocatalytic H2 Evolution, CO2 Reduction and N2 Fixation Coupled with Organic Synthesis by Cocatalyst and Vacancies Engineering. Appl. Catal. B Environ. 2021, 285, 119789. [Google Scholar] [CrossRef]
  6. Cai, M.D.; Cao, S.Y.; Zhuo, Z.Z.; Wang, X.; Shi, K.Z.; Cheng, Q.; Xue, Z.M.; Du, X.; Shen, C.; Liu, X. Fabrication of Ni2P Cocatalyzed CdS Nanorods with a Well-Defined Heterointerface for Enhanced Photocatalytic H2 Evolution. Catalysts 2022, 12, 417. [Google Scholar] [CrossRef]
  7. Battula, V.R.; Jaryal, A.; Kailasam, K. Visible light-driven simultaneous H2 production by water splitting coupled with selective oxidation of HMF to DFF catalyzed by porous carbon nitride. J. Mater. Chem. A 2019, 7, 5643–5649. [Google Scholar] [CrossRef]
  8. Ripple, W.J.; Wolf, C.; Newsome, T.M.; Galetti, M.; Alamgir, M.; Crist, E.; Mahmoud, M.I.; Laurance, W.F. World Scientists’ Warning to Humanity: A Second Notice. BioScience 2017, 67, 1026–1028. [Google Scholar] [CrossRef] [Green Version]
  9. Xu, D.; Zhang, S.-N.; Chen, J.-S.; Li, X.-H. Design of the Synergistic Rectifying Interfaces in Mott-Schottky Catalysts. Chem. Rev. 2023, 123, 1–30. [Google Scholar] [CrossRef]
  10. Oshima, T.; Nishioka, S.; Kikuchi, Y.; Hirai, S.; Yanagisawa, K.; Eguchi, M.; Maeda, K. An Artificial Z-Scheme Constructed from Dye-Sensitized Metal Oxide Nanosheets for Visible Light-Driven Overall Water Splitting. J. Am. Chem. Soc. 2020, 142, 8412–8420. [Google Scholar] [CrossRef]
  11. Jin, X.X.; Wang, R.Y.; Zhang, L.X.; Si, R.; Shen, M.; Wang, M.; Tian, J.J.; Shi, J.L. Electron Configuration Modulation of Nickel Single Atoms for Elevated Photocatalytic Hydrogen Evolution. Angew. Chem. Int. Ed. 2020, 59, 6827–6831. [Google Scholar] [CrossRef]
  12. Zhu, Q.H.; Xu, Q.; Du, M.M.; Zeng, X.F.; Zhong, G.F.; Qiu, B.C.; Zhang, J.L. Recent progress of metal sulfide photocatalysts for solar energy conversion. Adv. Mater. 2022, 34, 2202929. [Google Scholar] [CrossRef]
  13. Zhang, F.; Li, J.M.; Wang, H.F.; Li, Y.P.; Liu, Y.; Qian, Q.; Zhang, G. Realizing Synergistic Effect of Electronic Modulation and Nanostructure Engineering over Graphitic Carbon Nitride for Highly Efficient Visible-Light H2 Production Coupled with Benzyl Alcohol Oxidation. Appl. Catal. B Environ. 2020, 269, 118772. [Google Scholar] [CrossRef]
  14. Meng, S.; Ye, X.; Zhang, J.; Fu, X.; Chen, S. Effective use of photogenerated electrons and holes in a system: Photocatalytic selective oxidation of aromatic alcohols to aldehydes and hydrogen production. J. Catal. 2018, 367, 159–170. [Google Scholar] [CrossRef]
  15. Zong, X.; Yan, H.H.; Wu, G.P.; Ma, G.J.; Wen, F.Y.; Wang, L.; Li, C. Enhancement of photocatalytic H2 evolution on CdS by loading MoS2 as cocatalyst under visible light irradiation. J. Am. Chem. Soc. 2008, 130, 7176–7177. [Google Scholar] [CrossRef]
  16. Fazil, M.; Ahmad, T. Pristine TiO2 and Sr-Doped TiO2 Nanostructures for Enhanced Photocatalytic and Electrocatalytic Water Splitting Applications. Catalysts 2023, 13, 93. [Google Scholar] [CrossRef]
  17. AlSalka, Y.; Al-Madanat, O.; Hakki, A.; Bahnemann, D.W. Boosting the H2 production efficiency via photocatalytic organic reforming: The role of additional hole scavenging system. Catalysts 2021, 11, 1423. [Google Scholar] [CrossRef]
  18. Al-Madanat, O.; Alsalka, Y.; Curti, M.; Dillert, R.; Bahnemann, D.W. Mechanistic insights into hydrogen evolution by photocatalytic reforming of naphthalene. ACS Catal. 2020, 10, 7398–7412. [Google Scholar] [CrossRef]
  19. Al-Madanat, O.; AlSalka, Y.; Ramadan, W.; Bahnemann, D.W. TiO2 photocatalysis for the transformation of aromatic water pollutants into fuels. Catalysts 2021, 11, 317. [Google Scholar] [CrossRef]
  20. Schneider, J.; Bahnemann, D.W. Undesired role of sacrificial reagents in photocatalysis. J. Chem. Phys. Lett. 2013, 4, 3479–3483. [Google Scholar] [CrossRef]
  21. Zhou, P.; Lv, F.; Li, N.; Zhang, Y.L.; Mu, Z.J.; Tang, Y.H.; Guo, S. Strengthening reactive metal-support interaction to stabilize high-density Pt single atoms on electron-deficient g-C3N4 for boosting photocatalytic H2 production. Nano Energy 2019, 56, 127–137. [Google Scholar] [CrossRef]
  22. Ma, X.H.; Li, W.J.; Li, H.D.; Dong, M.; Li, X.Y.; Geng, L.; Wang, T. Fabrication of novel and noble-metal-free MoP/In2S3 Schottky heterojunction photocatalyst with efficient charge separation for enhanced photocatalytic H2 evolution under visible light. J. Colloid Interf. Sci. 2022, 617, 284–292. [Google Scholar] [CrossRef]
  23. Yang, L.F.; Li, A.Q.; Dang, T.; Wang, Y.F.; Liang, L.; Tang, J.; Zhang, Z. S-scheme In2S3/Zn3In2S6 microsphere for efficient photocatalytic H2 evolution with simultaneous photodegradation of bisphenol A. Appl. Surf. Sci. 2023, 612, 155848. [Google Scholar] [CrossRef]
  24. Zhang, R.Y.; Jia, X.W.; Li, Y.R.; Yu, X.D.; Xing, Y. Oxidation co-catalyst modified In2S3 with efficient interfacial charge transfer for boosting photocatalytic H2 evolution. Int. J. Hydrogen Energ. 2022, 47, 25300–25308. [Google Scholar] [CrossRef]
  25. Lin, Q.C.; Li, Z.H.; Lin, T.J.; Li, B.L.; Liao, X.C.; Yu, H.Q.; Yu, C.L. Controlled preparation of P-doped g-C3N4 nanosheets for efficient photocatalytic hydrogen production. Chin. J. Chem. Eng. 2020, 28, 2677–2688. [Google Scholar] [CrossRef]
  26. Chen, X.B.; Shen, S.H.; Guo, L.J.; Mao, S.S. Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 2010, 110, 6503–6570. [Google Scholar] [CrossRef]
  27. Chai, Z.G.; Zeng, T.T.; Li, Q.; Lu, L.Q.; Xiao, W.J.; Xu, D.S. Efficient visible light-driven splitting of alcohols into hydrogen and corresponding carbonyl compounds over a Ni-modified CdS photocatalyst. J. Am. Chem. Soc. 2016, 138, 10128–10131. [Google Scholar] [CrossRef]
  28. Xiong, Z.; Hou, Y.D.; Yuan, R.S.; Ding, Z.X.; Ong, W.J.; Wang, S.B. Hollow NiCo2S4 nanospheres as a cocatalyst to support ZnIn2S4 nanosheets for visible-light-driven hydrogen production. Acta. Phys.-Chim. Sin. 2022, 38, 2111021. [Google Scholar]
  29. Li, S.C.; Shi, M.Y.; Yu, J.H.; Li, S.J.; Lei, S.L.; Lin, L.G.; Wang, J.J. Two-dimensional blue-phase CX (X = S, Se) monolayers with high carrier mobility and tunable photocatalytic water splitting capability. Chin. Chem. Lett. 2021, 32, 1977–1982. [Google Scholar] [CrossRef]
  30. Yuan, L.; Li, Y.H.; Tang, Z.R.; Gong, J.L.; Xu, Y.J. Defect-promoted visible light-driven CC coupling reactions pairing with CO2 reduction. J. Catal. 2020, 390, 244–250. [Google Scholar] [CrossRef]
  31. Meng, S.; Wu, H.; Cui, Y.; Zheng, X.; Wang, H.; Chen, S.; Fu, X. One-step synthesis of 2D/2D-3D NiS/Zn3In2S6 hierarchical structure toward solar-to-chemical energy transformation of biomass-relevant alcohols. Appl. Catal. B Environ. 2020, 266, 118617. [Google Scholar] [CrossRef]
  32. Qi, M.Y.; Conte, M.; Anpo, M.; Tang, Z.R.; Xu, Y.J. Cooperative coupling of oxidative organic synthesis and hydrogen production over semiconductor-based photocatalysts. Chem. Rev. 2021, 121, 13051–13085. [Google Scholar] [CrossRef]
  33. Fu, X.; Zhang, L.; Liu, L.; Li, H.; Meng, S.; Ye, X.; Chen, S. In situ photodeposition of MoSx on CdS nanorods as a highly efficient cocatalyst for photocatalytic hydrogen production. J. Mater. Chem. A 2017, 5, 15287–15293. [Google Scholar] [CrossRef]
  34. Shen, R.C.; Ren, D.D.; Ding, Y.N.; Guan, Y.T.; Ng, Y.H.; Zhang, P.; Li, X. Nanostructured CdS for efficient photocatalytic H2 evolution: A review. Sci. China Mater. 2020, 63, 2153–2188. [Google Scholar] [CrossRef]
  35. Li, L.; Guo, C.F.; Ning, J.Q.; Zhong, Y.J.; Chen, D.L.; Hu, Y. Oxygen-vacancy-assisted construction of FeOOH/CdS heterostructure as an efficient bifunctional photocatalyst for CO2 conversion and water oxidation. Appl. Catal. B Environ. 2021, 293, 120203. [Google Scholar] [CrossRef]
  36. Han, G.Q.; Jin, Y.H.; Burgess, R.A.; Dickenson, N.E.; Cao, X.M.; Sun, Y.J. Visible-Light-Driven Valorization of Biomass Intermediates Integrated with H2 Production Catalyzed by Ultrathin Ni/CdS Nanosheets. J. Am. Chem. Soc. 2017, 139, 15584–15587. [Google Scholar] [CrossRef]
  37. Liu, M.C.; Chen, Y.B.; Su, J.Z.; Shi, J.W.; Wang, X.X.; Guo, L.J. Photocatalytic Hydrogen Production using Twinned Nanocrystals and an Unanchored NiSx Co-Catalyst. Nat. Energy 2016, 1, 16151. [Google Scholar] [CrossRef]
  38. Li, S.J.; Cai, M.J.; Liu, Y.P.; Wang, C.C.; Yan, R.Y.; Chen, X.B. Constructing Cd0.5Zn0.5S/Bi2WO6 S-scheme heterojunction for boosted photocatalytic antibiotic oxidation and Cr(VI) reduction. Adv. Powder Mater. 2023, 2, 100073. [Google Scholar] [CrossRef]
  39. Dong, Y.J.; Han, Q.; Hu, Q.Y.; Xu, C.J.; Dong, C.Z.; Peng, Y.; Lan, Y. Carbon quantum dots enriching molecular nickel polyoxometalate over CdS semiconductor for photocatalytic water splitting. Appl. Catal. B 2021, 293, 120214. [Google Scholar] [CrossRef]
  40. Ye, H.F.; Shi, R.; Yang, X.; Fu, W.F.; Chen, Y. P-doped ZnxCd1-xS solid solutions as photocatalysts for hydrogen evolution from water splitting coupled with photocatalytic oxidation of 5-hydroxymethylfurfural. Appl. Catal. B Environ. 2018, 233, 70–79. [Google Scholar] [CrossRef]
  41. Yang, Y.; Ren, W.; Zheng, X.; Meng, S.; Cai, C.; Fu, X.; Chen, S. Decorating Zn0.5Cd0.5S with C, N Co-Doped CoP: An Efficient Dual-Functional Photocatalyst for H2 Evolution and 2,5-Diformylfuran Oxidation. ACS Appl. Mater. Inter. 2022, 14, 54649–54661. [Google Scholar] [CrossRef]
  42. Zhu, Z.; Zhang, S.; Chen, G.; Meng, S.; Zheng, X.; Chen, S.; Zhang, F. Minimized Pt deposition on CdS simultaneously maximizes the performance of hydrogen production and aromatic alcohols oxidation. Appl. Surf. Sci. 2021, 564, 150446. [Google Scholar] [CrossRef]
  43. Shi, X.W.; Dai, C.; Wang, X.; Hu, J.Y.; Zhang, J.Y.; Zheng, L.; Zhu, M. Protruding Pt single-sites on hexagonal ZnIn2S4 to accelerate photocatalytic hydrogen evolution. Nat. Commun. 2022, 13, 1287. [Google Scholar] [CrossRef]
  44. Andreou, E.K.; Koutsouroubi, E.D.; Vamvasakis, I.; Armatas, G.S. Ni2P-modified P-Doped Carbon Nitride Hetero-Nanostructures for Efficient Photocatalytic Aqueous Cr(VI) Reduction. Catalysts 2023, 13, 437. [Google Scholar] [CrossRef]
  45. Yang, Y.; Zheng, X.Z.; Liu, J.F.; Qi, Z.L.; Su, T.Y.; Cai, C.; Chen, S. Efficient H2 evolution on Co3S4/Zn0.5Cd0.5S nanocomposites by photocatalytic synergistic reaction. Inorg. Chem. Front. 2022, 9, 1943–1955. [Google Scholar] [CrossRef]
  46. Shen, R.C.; Ding, Y.N.; Li, S.B.; Zhang, P.; Xiang, Q.J.; Ng, Y.H.; Li, X. Constructing low-cost Ni3C/twin-crystal Zn0.5Cd0.5S heterojunction/ homojunction nanohybrids for efficient photocatalytic H2 evolution. Chin. J. Catal. 2021, 42, 25–36. [Google Scholar] [CrossRef]
  47. Li, K.; Chai, B.; Peng, T.Y.; Mao, J.; Zan, L. Preparation of AgIn5S8/TiO2 heterojunction nanocomposite and its enhanced photocatalytic H2 production property under visible light. ACS Catal. 2013, 3, 170–177. [Google Scholar] [CrossRef]
  48. He, Y.; Li, D.; Xiao, G.; Chen, W.; Chen, Y.; Sun, M.; Huang, H.; Fu, X. A New Application of Nanocrystal In2S3 in Efficient Degradation of Organic Pollutants under Visible Light Irradiation. J. Phys. Chem. C 2009, 113, 5254–5262. [Google Scholar] [CrossRef]
  49. Sun, X.; Luo, X.; Zhang, X.; Xie, J.; Jin, S.; Wang, H.; Zheng, X.; Wu, X.; Xie, Y. Enhanced Superoxide Generation on Defective Surfaces for Selective Photooxidation. J. Am. Chem. Soc. 2019, 141, 3797–3801. [Google Scholar] [CrossRef]
  50. Li, T.; Zhang, S.; Meng, S.; Ye, X.; Fu, X.; Chen, S. Amino acid-assisted synthesis of In2S3 hierarchical architectures for selective oxidation of aromatic alcohols to aromatic aldehydes. RSC Adv. 2017, 7, 6457–6466. [Google Scholar] [CrossRef] [Green Version]
  51. Meng, S.; Ye, X.; Ning, X.; Xie, M.; Fu, X.; Chen, S. Selective oxidation of aromatic alcohols to aromatic aldehydes by BN/metal sulfide with enhanced photocatalytic activity. Appl. Catal. B Environ. 2016, 182, 356–368. [Google Scholar] [CrossRef]
  52. Meng, S.; Cui, Y.; Wang, H.; Zheng, X.; Fu, X.; Chen, S. Noble metal-free 0D-1D NiSx/CdS nanocomposites toward highly efficient photocatalytic contamination removal and hydrogen evolution under visible light. Dalton T. 2018, 47, 12671–12683. [Google Scholar] [CrossRef]
  53. Cheng, T.T.; Gao, H.J.; Liu, G.R.; Pu, Z.S.; Wang, S.F.; Yi, Z.; Yang, H. Preparation of core-shell heterojunction photocatalysts by coating CdS nanoparticles onto Bi4Ti3O12 hierarchical microspheres and their photocatalytic removal of organic pollutants and Cr (VI) ions. Colloid. Surf. A 2022, 633, 127918. [Google Scholar] [CrossRef]
  54. Li, W.J.; Lin, Z.Y.; Yang, G.W. A 2D self-assembled MoS2/Znln2S4 heterostructure for efficient photocatalytic hydrogen evolution. Nanoscale 2017, 9, 18290–18298. [Google Scholar] [CrossRef]
  55. Pasupuleti, K.S.; Reddeppa, M.; Park, B.-G.; Oh, J.-E.; Kim, S.-G.; Kim, M.-D. Efficient Charge Separation in Polypyrrole/GaN-Nanorod-Based Hybrid Heterojunctions for High-Performance Self-Powered UV Photodetection. Phys. Status Solidi-Rapid Res. Lett. 2021, 15, 2000518. [Google Scholar] [CrossRef]
  56. Wu, H.; Meng, S.; Zhang, J.; Zheng, X.; Wang, Y.; Chen, S.; Fu, X. Construction of two-dimensionally relative p-n heterojunction for efficient photocatalytic redox reactions under visible light. Appl. Surf. Sci. 2020, 505, 144638. [Google Scholar] [CrossRef]
  57. Deng, H.Z.; Fei, X.G.; Yang, Y.; Fan, J.J.; Yu, J.G.; Cheng, B.; Zhang, L.Y. S-scheme heterojunction based on p-type ZnMn2O4 and n-type ZnO with improved photocatalytic CO2 reduction activity. Chem. Eng. J. 2021, 409, 127377. [Google Scholar] [CrossRef]
  58. Liu, D.N.; Chen, D.Y.; Li, N.J.; Xu, Q.F.; Li, H.; He, J.H.; Lu, J.M. Surface engineering of g-C3N4 by stacked BiOBr sheets rich in oxygen vacancies for boosting photocatalytic performance. Angew. Chem. Int. Ed. 2020, 59, 4519–4524. [Google Scholar] [CrossRef]
  59. Yang, W.L.; Zhang, L.; Xie, J.F.; Zhang, X.D.; Liu, Q.H.; Yao, T.; Xie, Y. Enhanced photoexcited carrier separation in oxygen-doped ZnIn2S4 nanosheets for hydrogen evolution. Angew. Chem. Int. Edit. 2016, 55, 6716–6720. [Google Scholar] [CrossRef]
  60. Nielsen, M.; Kammer, A.; Cozzula, D.; Junge, H.; Gladiali, S.; Beller, M. Efficient hydrogen production from alcohols under mild reaction conditions. Angew. Chem., Int. Ed. 2011, 50, 9593. [Google Scholar] [CrossRef]
  61. Zheng, X.Z.; Zhang, Z.; Meng, S.; Wang, Y.X.; Li, D. Regulating charge transfer over 3D Au/ZnO hybrid inverse opal toward efficiently photocatalytic degradation of bisphenol A and photoelectrochemical water splitting. Chem. Eng. J. 2020, 393, 124676. [Google Scholar] [CrossRef]
  62. Khan, I.; Saeed, K.; Ali, N.; Khan, I.; Zhang, B.; Sadiq, M. Heterogeneous photodegradation of industrial dyes: An insight to different mechanisms and rate affecting parameters. J. Environ. Chem. Eng. 2020, 8, 104364. [Google Scholar] [CrossRef]
  63. Ahmad, S.; Almehmadi, M.; Janjuhah, H.T.; Kontakiotis, G.; Abdulaziz, O.; Saeed, K.; Ahmad, H.; Allahyani, M.; Aljuaid, A.; Alsaiari, A.A.; et al. The Effect of Mineral Ions Present in Tap Water on Photodegradation of Organic Pollutants: Future Perspectives. Water 2023, 15, 175. [Google Scholar] [CrossRef]
  64. Al-Madanat, O.; AlSalka, Y.; Dillert, R.; Bahnemann, D.W. Photocatalytic H2 production from naphthalene by various TiO2 photocatalysts: Impact of Pt loading and formation of intermediates. Catalysts 2021, 11, 107. [Google Scholar] [CrossRef]
  65. Al-Madanat, O.; Curti, M.; Günnemann, C.; AlSalka, Y.; Dillert, R.; Bahnemann, D.W. TiO2 photocatalysis: Impact of the platinum loading method on reductive and oxidative half-reactions. Catal. Today 2021, 380, 3–15. [Google Scholar] [CrossRef]
  66. Pasupuleti, K.S.; Chougule, S.S.; Jung, N.; Yu, Y.J.; Oh, J.-E.; Kim, M.-D. Plasmonic Pt nanoparticles triggered efficient charge separation in TiO2/GaN NRs hybrid heterojunction for the high performance self-powered UV photodetectors. Appl. Surf. Sci. 2022, 594, 153474. [Google Scholar] [CrossRef]
  67. Pasupuleti, K.S.; Reddeppa, M.; Park, B.-G.; Peta, K.R.; Oh, J.-E.; Kim, S.-G.; Kim, M.-D. Ag nanowire-plasmonic-assisted charge separation in hybrid heterojunctions of Ppy-PEDOT: PSS/GaN nanorods for enhanced UV photodetection. ACS Appl. Mater. Interfaces 2020, 12, 54181–54190. [Google Scholar] [CrossRef]
  68. Wan, J.; Liu, L.; Wu, Y.; Song, J.R.; Liu, J.Q.; Song, R.; Xiong, Y. Exploring the polarization photocatalysis of ZnIn2S4 material toward hydrogen evolution by integrating cascade electric fields with hole transfer vehicle. Adv. Funct. Mater. 2022, 32, 2203252. [Google Scholar] [CrossRef]
  69. Wan, J.; Yang, W.J.; Liu, J.Q.; Sun, K.L.; Liu, L.; Fu, F. Enhancing an internal electric field by a solid solution strategy for steering bulk-charge flow and boosting photocatalytic activity of Bi24O31ClxBr10-x. Chin. J. Catal. 2022, 43, 485–496. [Google Scholar] [CrossRef]
  70. Hu, Z.F.; Yuan, L.Y.; Liu, Z.F.; Shen, Z.R.; Yu, J.C. An Elemental Phosphorus Photocatalyst with a Record High Hydrogen Evolution Efficiency. Angew. Chem. Int. Ed. 2016, 55, 9793. [Google Scholar] [CrossRef] [Green Version]
  71. Pasupuleti, K.S.; Ghosh, S.; Jayababu, N.; Kang, C.J.; Cho, H.D.; Kim, S.-G.; Kim, M.-D. Boron doped g-C3N4 quantum dots based highly sensitive surface acoustic wave NO2 sensor with faster gas kinetics under UV light illumination. Sensor. Actuat. B Chem. 2023, 378, 133140. [Google Scholar] [CrossRef]
  72. Nowicka, E.; Hofmann, J.P.; Parker, S.F.; Sankar, M.; Lari, G.M.; Kondrat, S.A.; Knight, D.W.; Bethell, D.; Weckhuysen, B.M.; Hutchings, G.J. In situ spectroscopic investigation of oxidative dehydrogenation and disproportionation of benzyl alcohol. Phys. Chem. Chem. Phys. 2013, 15, 12147–12155. [Google Scholar] [CrossRef] [Green Version]
  73. Sun, Z.; Yang, X.; Yu, X.-F.; Xia, L.; Peng, Y.; Li, Z.; Zhang, Y.; Cheng, J.; Zhang, K.; Yu, J. Surface oxygen vacancies of Pd/Bi2MoO6-x acts as “Electron Bridge” to promote photocatalytic selective oxidation of alcohol. Appl. Catal. B Environ. 2021, 285, 119790. [Google Scholar] [CrossRef]
  74. Al-Madanat, O.; Nunes, B.N.; AlSalka, Y.; Hakki, A.; Curti, M.; Patrocinio, A.O.T.; Bahnemann, D.W. Application of EPR spectroscopy in TiO2 and Nb2O5 photocatalysis. Catalysts 2021, 11, 1514. [Google Scholar] [CrossRef]
  75. Chen, Z.X.; Li, D.Z.; Zhang, W.J.; Chen, C.; Li, W.J.; Sun, M.; Fu, X. Low-temperature and template-free synthesis of ZnIn2S4 microspheres. Inorg. Chem. 2008, 47, 9766–9772. [Google Scholar] [CrossRef]
Figure 1. SEM images of (ac) In2S3 and (df) Pt/In2S3.
Figure 1. SEM images of (ac) In2S3 and (df) Pt/In2S3.
Catalysts 13 00461 g001
Figure 2. (a,b) TEM, (c) HRTEM and (d) corresponding FFT images of In2S3. (e,f) TEM and (g) HRTEM and (h) corresponding FFT images of Pt/In2S3 (the Pt clusters are shown in the purple circles). (i) EDX-mapping images and (j) EDX spectrum of Pt/In2S3.
Figure 2. (a,b) TEM, (c) HRTEM and (d) corresponding FFT images of In2S3. (e,f) TEM and (g) HRTEM and (h) corresponding FFT images of Pt/In2S3 (the Pt clusters are shown in the purple circles). (i) EDX-mapping images and (j) EDX spectrum of Pt/In2S3.
Catalysts 13 00461 g002
Figure 3. (a) XRD patterns of In2S3 and Pt/In2S3. XPS spectra of (b) In 3d, (c) S 2p and (d) Pt 4f.
Figure 3. (a) XRD patterns of In2S3 and Pt/In2S3. XPS spectra of (b) In 3d, (c) S 2p and (d) Pt 4f.
Catalysts 13 00461 g003
Figure 4. (a) Nitrogen adsorption–desorption isotherms, (b) UV-vis DRS spectra and (c) band-gap energies of In2S3 and Pt/In2S3. M-S plots of (d) In2S3 and (e) Pt/In2S3. (f) The relationships of band energy positions between In2S3 and Pt/In2S3.
Figure 4. (a) Nitrogen adsorption–desorption isotherms, (b) UV-vis DRS spectra and (c) band-gap energies of In2S3 and Pt/In2S3. M-S plots of (d) In2S3 and (e) Pt/In2S3. (f) The relationships of band energy positions between In2S3 and Pt/In2S3.
Catalysts 13 00461 g004
Figure 5. (a) UPS spectrum of In2S3. (b) Carrier densities of In2S3 before and after Pt modification. Energy level diagrams for In2S3 and Pt (c) before and (d) after interfacing. (e) Illustration of photoexcited electrons transfers from In2S3 into Pt over Pt/In2S3 interface under light irradiation.
Figure 5. (a) UPS spectrum of In2S3. (b) Carrier densities of In2S3 before and after Pt modification. Energy level diagrams for In2S3 and Pt (c) before and (d) after interfacing. (e) Illustration of photoexcited electrons transfers from In2S3 into Pt over Pt/In2S3 interface under light irradiation.
Catalysts 13 00461 g005
Figure 6. (a) The PHP activity of In2S3 and Pt/In2S3 with different Pt weight ratios for photocatalytic selective oxidation of phenylcarbinol and H2 production under simulated sunlight irradiation. (b) The PHP activity of 0.3% Pt/In2S3 under monochromatic light irradiation with different wavelengths. (c) The PHP activity of 0.3% Pt/In2S3 under simulated sunlight irradiation for 5 recycles. (d) Comparison of 0.3% Pt/In2S3 PHP activity in different reaction systems (LA, NS, TA, MA and PhCH2OH presents aqueous solutions of lactic acid, Na2S, triethanolamine, methanol and phenylcarbinol, respectively) under simulated sunlight irradiation.
Figure 6. (a) The PHP activity of In2S3 and Pt/In2S3 with different Pt weight ratios for photocatalytic selective oxidation of phenylcarbinol and H2 production under simulated sunlight irradiation. (b) The PHP activity of 0.3% Pt/In2S3 under monochromatic light irradiation with different wavelengths. (c) The PHP activity of 0.3% Pt/In2S3 under simulated sunlight irradiation for 5 recycles. (d) Comparison of 0.3% Pt/In2S3 PHP activity in different reaction systems (LA, NS, TA, MA and PhCH2OH presents aqueous solutions of lactic acid, Na2S, triethanolamine, methanol and phenylcarbinol, respectively) under simulated sunlight irradiation.
Catalysts 13 00461 g006
Figure 7. Transient photocurrent responses of catalysts (a) without and (b) with methyl viologen dichloride (MVCl2). (c) EIS plots, (d) LSV curves, (e) TEMPO-e EPR spectra and (f) TEMPO-h+ EPR spectra of In2S3 and Pt/In2S3.
Figure 7. Transient photocurrent responses of catalysts (a) without and (b) with methyl viologen dichloride (MVCl2). (c) EIS plots, (d) LSV curves, (e) TEMPO-e EPR spectra and (f) TEMPO-h+ EPR spectra of In2S3 and Pt/In2S3.
Catalysts 13 00461 g007
Figure 8. (a) EPR spectra of DMPO-CH(OH)Ph over different photocatalysts. (b) In situ DRIFTS spectra of the 0.3% Pt/In2S3 hybrid with the existence of PhCH2OH under simulated sunlight. (c) The effect of the trapping agents on the PHP over the 0.3% Pt/In2S3 hybrid. (d) Illustration of the formation and photocatalytic mechanism of the 0D/2D-3D Pt/In2S3 heterostructure.
Figure 8. (a) EPR spectra of DMPO-CH(OH)Ph over different photocatalysts. (b) In situ DRIFTS spectra of the 0.3% Pt/In2S3 hybrid with the existence of PhCH2OH under simulated sunlight. (c) The effect of the trapping agents on the PHP over the 0.3% Pt/In2S3 hybrid. (d) Illustration of the formation and photocatalytic mechanism of the 0D/2D-3D Pt/In2S3 heterostructure.
Catalysts 13 00461 g008
Figure 9. The preparation procedure of the 0D/2D-3D Pt/In2S3 heterostructure.
Figure 9. The preparation procedure of the 0D/2D-3D Pt/In2S3 heterostructure.
Catalysts 13 00461 g009
Table 1. Comparison of photocatalytic performance over different photocatalysts for photocatalytic H2 production.
Table 1. Comparison of photocatalytic performance over different photocatalysts for photocatalytic H2 production.
PhotocatalystLight SourceReagentsH2 Evolution
(mmol g−1 h−1)
Ref.
Pt/In2S3λ ≥ 420 nmPhCH2OH7.97This Work
Pt/CdSλ > 420 nmPhCH2OH4.9[42]
Pt/Zn3In2S6λ ≥ 420 nmPhCH2OH0.9[14]
Pt/g-C3N4λ > 420 nmTEOA3.02[21]
MoP/In2S3λ ≥ 420 nmLactic acid0.5[22]
Zn3In2S6In2S3λ > 400 nmbisphenol A0.08[23]
PdS/In2S3λ > 420 nmNa2S/Na2SO33.6[24]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, H.; Xiao, P.; Meng, S.; Long, B.; Liu, Q.; Zheng, X.; Zhang, S.; Ruan, Z.; Chen, S. One Stone Two Birds: Utilization of Solar Light for Simultaneous Selective Phenylcarbinol Oxidation and H2 Production over 0D/2D-3D Pt/In2S3 Schottky Junction. Catalysts 2023, 13, 461. https://doi.org/10.3390/catal13030461

AMA Style

Zhang H, Xiao P, Meng S, Long B, Liu Q, Zheng X, Zhang S, Ruan Z, Chen S. One Stone Two Birds: Utilization of Solar Light for Simultaneous Selective Phenylcarbinol Oxidation and H2 Production over 0D/2D-3D Pt/In2S3 Schottky Junction. Catalysts. 2023; 13(3):461. https://doi.org/10.3390/catal13030461

Chicago/Turabian Style

Zhang, Huijun, Peipei Xiao, Sugang Meng, Baihua Long, Qing Liu, Xiuzhen Zheng, Sujuan Zhang, Zhaohui Ruan, and Shifu Chen. 2023. "One Stone Two Birds: Utilization of Solar Light for Simultaneous Selective Phenylcarbinol Oxidation and H2 Production over 0D/2D-3D Pt/In2S3 Schottky Junction" Catalysts 13, no. 3: 461. https://doi.org/10.3390/catal13030461

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop