DOI:
10.1039/D4QI01826E
(Research Article)
Inorg. Chem. Front., 2024,
11, 6970-6980
Photoredox catalysis enabled by atomically precise metal nanoclusters†
Received
23rd July 2024
, Accepted 31st August 2024
First published on 2nd September 2024
Abstract
Atomically precise metal nanoclusters (NCs), distinguished by their unique electronic structures, quantum confinement effects, and enriched active sites, have been considered highly promising photosensitizers for light harvesting and conversion. However, the ultra-short carrier lifetime and poor stability of metal NCs remarkably retard their widespread applications in photocatalysis. In this study, we achieved the modulation of carrier separation over metal NCs via heterostructure engineering by smartly integrating atomically precise silver NCs [Ag16(GSH)9] with transition metal chalcogenides (TMCs). The favorable energy level alignment between metal NCs and TMCs facilitates the electron transfer from the metal NCs to the TMCs, leading to a significantly prolonged charge lifetime and considerably enhanced photoactivity toward the selective reduction of nitro compounds to amino derivatives under visible light. The photocatalytic mechanism of these composite photosystems is elucidated herein. This work advances our fundamental understanding of charge transfer mechanisms over atomically precise metal NCs for solar energy conversion.
1. Introduction
Atomically precise metal nanoclusters (NCs) usually consist of a few to several hundreds of metal atoms with diameters comparable to the Fermi wavelength of electrons (<2 nm) and thus exhibit unique physicochemical properties and energy band structures that are quite distinct from those of conventional metal nanocrystals.1–4 These metal NCs feature discrete energy band structures characterized by a transition gap between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO).5–8 In previous studies,9 metal NCs have been used as photosensitizers for constructing heterostructured photocatalysts towards versatile visible-light-driven heterogeneous photocatalysis, including photocatalytic hydrogen generation, CO2 reduction, environmental remediation, and selective organic transformation.10–12 Intriguingly, the peculiar atom-stacking mode and abundant active sites of metal NCs endow the heterostructure with significantly enhanced photoactivities. However, to date, few metal NCs have been exploited for photocatalysis, with charge transfer characteristics and photocatalytic mechanism remaining elusive, which stems mainly from a scarcity of metal NCs with photosensitization capacity, the ultra-fast charge recombination rate of metal NCs, and the poor stability of metal NCs, markedly retarding the exploration of high-performance metal NC-based photocatalytic systems.13
Transition metal chalcogenides (TMCs) consist of anionic sulfur atoms bonded to transition metals and have been widely recognized as promising photocatalysts due to their suitable band gap, favorable energy band position, and abundance of active sites.14–16 Nevertheless, the exploration of robust and stable TMC photocatalysts is often hindered by rapid photoinduced electron–hole recombination, photo-corrosion, and short carrier lifetimes. The introduction of an elegant interface,17,18 modulation of the nanoarchitecture19,20 and construction of heterostructures21–23 are effective methods to solve the problems of TMCs. Co-catalyst engineering has also been proven to be an effective strategy to reinforce the charge separation of TMCs. For example, Jin et al. introduced graphdiyne into the NiCo2O4 photosystem through reasonable design, which is innovative and has the potential to attract the interest of readers.24 Intriguingly, the construction of heterostructured photocatalysts provides a convenient, easily accessible, and efficacious strategy to finely tune the interfacial charge transport pathway by fully harnessing the synergy of building blocks for improved photocatalytic efficiency.25 Hence, we speculate that the construction of heterostructures by rationally integrating atomically precise metal NCs with TMCs can optimally harness the photosensitization effect of metal NCs and physicochemical merits of TMCs. The TMCs feature suitable energy level positions and abundant catalytically active sites, which enable them to form favorable energy level alignments with metal NCs to foster interfacial charge transfer, as well as their synergistic cooperation, ultimately giving rise to the significantly enhanced photocatalytic activities of the composite photosystems.
Herein, atomically precise metal NCs [Ag16/(GSH)9] are self-assembled on the TMCs (CdIn2S4, CIS) to produce heterostructured photocatalysts in which metal NCs serve as light-harvesting antennas for photosensitizing TMCs. The favorable energy level alignment and intimate interfacial contact enable effective charge transfer between metal NCs and TMCs. We have found that the self-assembled CIS/Ag16(GSH)9 heterostructure shows significantly boosted photoactivities toward the reduction of aromatic nitro compounds to amino derivatives under visible light irradiation. The results imply that electrons photogenerated over Ag16/(GSH)9 can rapidly transfer from the LUMO level to the conduction band (CB) of CIS, thereby enhancing the charge separation of Ag16/(GSH)9 and prolonging the charge lifetime. Our work is anticipated to offer a new idea for precisely mediating the charge transport behaviors of atomically precise metal NCs for solar energy conversion.
2. Experimental section
2.1. Materials
Indium chloride tetrahydrate (InCl3·4H2O), cadmium chloride (CdCl2·2.5H2O), thiourea (Tu), ethylenediamine, ethanol (C2H5OH), sodium borohydride (NaBH4), reduced L-glutathione (GSH), sodium hydroxide (NaOH), silver nitrate (AgNO3), sodium sulfite anhydrous (Na2SO3), 4-nitroaniline (4-NA), 3-nitroaniline (3-NA), 2-nitroaniline (2-NA), 1-bromo-4-nitrobenzene, 1-chloro-4-nitrobenzene, 4-nitroanisole, 4-nitrotoluene (4-NT), nitrobenzene (NB), O-nitroacetophenone, and deionized water (DI H2O, Millipore, 18.2 MΩ cm resistivity) were used. All glassware was washed with aqua regia and rinsed with ethanol and ultrapure water. All the materials were of analytical grade and used as received without further purification.
2.2. Preparation of CdIn2S4 (CIS)26
Appropriate amounts of analytical grade CdCl2·2.5H2O and thiourea (Tu) were added to a Teflon-lined stainless-steel autoclave of 50 mL capacity, which was filled with ethylenediamine up to 90% of the capacity. The autoclave was maintained at 160 °C for 12 h and then air-cooled to room temperature. The precipitate was filtered off and washed with distilled water and absolute ethanol to remove residual impurities. After being dried in a vacuum at 70 °C for 3 h, the collected products were characterized as CdS. The prepared CdS and InCl3·4H2O in a molar ratio of 1:2 and excess Tu were placed in an autoclave that was then filled with distilled water up to 90% of the total volume. The autoclave was sealed and heated under autogenerated pressure at 180 °C for 10 h, and other procedures similar to those used to prepare CdS were applied. A final orange-yellow powder was obtained for characterization.
2.3. Preparation of Ag16(GSH)9 NCs27
Aqueous solutions of AgNO3 (20 mM) and GSH (50 mM) were prepared with DI H2O. A fresh aqueous solution of NaBH4 (112 mM) was prepared by dissolving 43 mg of NaBH4 in 2 mL of 1 M NaOH solution, followed by the addition of 8 mL of DI H2O. NaOH was added to the NaBH4 solution to improve the stability of borohydride ions against hydrolysis. In a typical experiment to synthesize Ag16(GSH)9 NCs, aqueous solutions of AgNO3 (125 μL, 20 mM) and GSH (150 μL, 50 mM) were first mixed in DI H2O (4.85 mL) under vigorous stirring to form thiolate–AgI complexes, followed by the addition of an aqueous solution of NaBH4 (50 μL, 112 mM). A deep-red solution of Ag16(GSH)9 NCs (ca. 5 mL) was collected after 5 min. This Ag16(GSH)9 solution was then incubated at room temperature for 3 h and the deep-red solution gradually decomposed into a colorless solution, forming thiolate–AgI complexes. Subsequently, a certain amount of 112 mM NaBH4 (50 μL) was introduced into this colorless solution under vigorous stirring, giving rise to the formation of a light-brown Ag16(GSH)9 solution after 15 min. Without stirring, this light-brown Ag16(GSH)9 solution was incubated at room temperature for 8 h. A strong red or green emission was then observed in the aqueous phase. The Ag16(GSH)9 NCs were collected without purification and stored at 4 °C for further use.
2.4. Preparation of CIS/Ag16(GSH)9 heterostructures
CIS/Ag16(GSH)9 heterostructures were fabricated via a self-assembly method under ambient conditions. The Ag16(GSH)9 NCs aqueous solution was slowly added dropwise into the CIS colloid aqueous solution under vigorous stirring for 2 h. The liquid volume percentage of Ag16(GSH)9 NCs in the nanocomposites was controlled to be 5%, 10%, 20%, 30% and 40% with the samples denoted as CIS/xAg16(GSH)9 (x = 0.05, 0.1, 0.2, 0.3, 0.4). Finally, the mixture was centrifuged and dried in an oven at 60 °C.
2.5. Characterization
Fourier transform infrared (FTIR) spectroscopy measurements were performed on a Nicolet IS50 spectrometer. The powder X-ray diffraction (XRD) patterns were recorded using a MiniFlex 600 X-ray powder diffractometer. X-ray photoelectron spectroscopy (XPS) measurements were performed on a Thermo Fisher Scientific ESCALAB 250 X-ray instrument equipped with a standard and monochromatic source (Al Kα, 1486.6 eV), and the binding energies were normalized to the signal for C 1s at 284.8 eV. HRTEM investigations were performed on a TECNAI G2F20 (FEI). Elemental mapping analysis was conducted using EDX. The ultraviolet-visible (UV-vis) absorption spectra of the Ag16(GSH)9 were acquired on a Cary 5000 instrument (Agilent, USA). UV–Vis diffuse (UV-Vis DRS) reflectance spectra were recorded on a UV-VIS-NIR spectrometer (Cary 5000). Raman spectra were recorded using an InVia Reflex. Photoluminescence (PL) spectra were acquired on a Varian Cary Eclipse spectrometer with an excitation wavelength of 350 nm.
2.6. Photocatalytic reduction performances
For the photoreduction reaction, a 300 W Xe lamp (PLS-SXE300D, Beijing Perfect Light Co. Ltd, China) equipped with a 420 nm cut-off filter (λ > 420 nm) was used as the irradiation source. Here, 10 mg of catalyst and 40 mg of sodium sulfite anhydrous (Na2SO3) were added to 30 mL of 4-NA aqueous solution (10 ppm) and saturated with bubbling N2 under ambient conditions. Before visible light illumination, the above suspension was stirred in the dark for 30 min to ensure the establishment of adsorption–desorption equilibrium between the sample and reactant. During the process of the reaction, 2 mL of sample solution was collected at a certain time interval and centrifuged to completely remove the catalyst (8000 rpm, 15 min), and the supernatant was analyzed using a UV-vis absorption spectrophotometer. The procedure for the photoreduction of other aromatic nitro compounds is similar to that of 4-NA. The photoactivities of the samples were defined as follows.
2.7. Photoredox performances28,29
(a) Photocatalytic reduction of aromatic nitro compounds.
PEC water splitting measurements were performed using a conventional three-electrode configuration on an electrochemical workstation (CHI660E, CHI Shanghai, Inc.). The samples (1 cm × 1 cm) were directly employed as the working electrodes, and Pt foil, as well as the Ag/AgCl electrode, served as the counter and reference electrodes, respectively. The electrolyte was composed of 0.5 M Na2SO4 (pH = 6.69) aqueous solution. PEC water splitting performances were evaluated under simulated sunlight irradiation from a 300 W Xe lamp (PLS FX300HU, Beijing Perfect Light Co. Ltd, China) equipped with an AM 1.5 filter or visible light (λ > 420 nm). The potentials of the electrodes were calibrated against the normalized hydrogen electrode (NHE) based on the following formula: | | (1) |
(b) Photocatalytic degradation of organic pollutants.
Typically, 20 mg of catalyst powder was suspended in 50 mL methyl orange (MO) aqueous solution (20 mg L−1). Before light irradiation, the mixtures were magnetically stirred for 1 h in the dark to reach the adsorption–desorption equilibrium. Then, a 300 W Xe lamp (λ > 420 nm) was used as the visible light source to irradiate the suspensions under vigorous stirring. At a certain interval (20 min), 1 mL of the suspension was collected, followed by the addition of 2 mL of water, and centrifuged to get rid of the catalysts. The supernatant was monitored by measuring the maximum absorbance at 464 nm for MO using a UV-Vis spectrophotometer. The removal efficiency is expressed by the ratio of Ct/C0, where Ct is the reaction concentration and C0 is the initial concentration. | | (2) |
3. Results and discussion
3.1. Characterization of the CIS/Ag16(GSH)9 heterostructure
As displayed in Scheme 1, the CIS/Ag16(GSH)9 heterostructure was fabricated by self-assembling Ag16(GSH)9 NCs on the CIS matrix under ambient conditions. It should be emphasized that the preparation of Ag16(GSH)9 NCs is based on the strong binding between the GSH ligand and the Ag core, which forms the monolayer of GSH molecules on the Ag16(GSH)9 NCs surface.27 Note that the various polar functional groups such as carbonyl and amide bonds on the GSH ligand cause Ag16(GSH)9 NCs to interact with CIS via intermolecular forces by which Ag16(GSH)9 can be spontaneously and uniformly anchored to the surface of CIS to form the heterostructure.
|
| Scheme 1 A schematic illustration depicting the self-assembly of the CIS/Ag16(GSH)9 heterostructure. | |
The morphology and size of Ag16(GSH)9 NCs were analyzed by transmission electron microscopy (TEM). Fig. S1a† shows Ag16(GSH)9 NCs with an ultra-small size of ca. 1.9 nm (Fig. S1b†). In the UV-visible absorption spectrum (Fig. S1c†), Ag16(GSH)9 NCs showed an absorption peak at 480 nm, which is different from the surface plasmon resonance absorption of conventional silver nanoparticles, indicative of the distinct electronic structure of Ag16(GSH)9 NCs.30–33 As shown in the SEM image (Fig. 1a), pristine CIS exhibits a nanorod-like structure. The TEM image of CIS (Fig. 1c) confirmed the one-dimensional structure and revealed a lattice spacing of about 0.330 nm (Fig. 1d) corresponding to the (311) crystalline facet of CIS. Note that no Ag16(GSH)9 NCs were observed in the SEM image of the CIS/Ag16(GSH)9 heterostructure (Fig. 1b), which is because the size of Ag16(GSH)9 NCs (ca. 1.9 nm) is too small to be discerned. However, the TEM image indicates that Ag16(GSH)9 NCs are uniformly distributed on the surface of CIS (Fig. 1e). This can be strongly verified by the SEM mapping (Fig. S2†) and EDS results (Fig. S3†), which confirmed the deposition of Ag16(GSH)9 NCs on the CIS surface. Elemental mapping results from TEM measurements indicated the uniform deposition of Ag16(GSH)9 NCs on the CIS substrate, as displayed in Fig. 1(f–j), where the Ag signal from NCs can be observed. The Fourier transform infrared spectroscopy (FTIR) results for CIS and CIS/Ag16(GSH)9 are provided in Fig. S4 and Table S1,† in which the characteristic peaks at 1038, 1142 and 2951 cm−1 correspond to the stretching vibration modes of C–N and C–H bonds on the CIS, and a distinct stretching vibration mode is observed at 2358 cm−1, which corresponds to the N–H bond of Ag16(GSH)9 NCs.34–36
|
| Fig. 1 SEM images of (a) CIS and (b) the CIS/Ag16(GSH)9 heterostructure. TEM image of (c) CIS. (d) TEM and HRTEM (inset) images of CIS. (e) HRTEM image of the CIS/Ag16(GSH)9 heterostructure. (f) The elemental mapping result of the CIS/Ag16(GSH)9 heterostructure for (g) S, (h) In, (i) Cd, and (j) Ag signals. | |
Crystal structures of the samples were analyzed by X-ray diffraction (XRD) as displayed in Fig. 2a.26 XRD peaks at 14.13°, 23.18°, 27.24°, 28.48°, 33.00°, 40.73°, 43.31°, and 47.41°, correspond to the (111), (220), (311), (222), (400), (422), (511) and (440) crystal facets of CIS (JCPDS no. 27-0060). The sharp diffraction peaks of CIS were observed, which indicated the good crystallization of the samples. It should be noted that no peaks attributable to Ag16(GSH)9 NCs were observed in the XRD results of the CIS/Ag16(GSH)9 heterostructure, which may be due to the relatively low loading amount of Ag16(GSH)9 NCs or its generic amorphous properties. As shown in Fig. 2b, the light absorption band edges of CIS and CIS/Ag16(GSH)9 in the UV diffuse reflectance spectra (DRS) results are all located at ca. 650 nm, which is due to the band gap photoexcitation of the CIS matrix. Notably, the absorption of the CIS/Ag16(GSH)9 heterostructure is slightly enhanced when Ag16(GSH)9 NCs are loaded onto the CIS matrix, which is due to the photosensitization effect of Ag16(GSH)9 NCs.37,38 This can also be demonstrated by the color change of the CIS and CIS/Ag16(GSH)9 heterostructure (Fig. 2b, inset). According to the Kubelka–Munk function versus photon energy, the bandgap (Eg) values of CIS and the CIS/Ag16(GSH)9 heterostructure (Fig. 2c) were roughly determined as ca. 2.26 and 2.27 eV, respectively. Ag16(GSH)9 NCs deposition does not substantially change the bandgap of the CIS substrate, which is due to the shielding absorption of the NCs and CIS substrate. Fig. 2e shows three bands at ca. 250, 309, and 367 cm− in the Raman spectra of CIS and the CIS/Ag16(GSH)9 heterostructure, which correspond to the Eg, A1g, and F1u modes of the CIS substrate, respectively.39,40 Similarly, no peaks of Ag16(GSH)9 NCs were observed in the Raman results, probably due to the low deposition amount or amorphous properties of NCs. The specific surface areas of CIS and the CIS/Ag16(GSH)9 heterostructure were determined as 15.6227 m2 g−1 and 15.0465 m2 g−1, respectively, as shown in Fig. 2d and Table S1.† The results indicate that the loading of Ag16(GSH)9 NCs on the surface of the CIS substrate does not drastically change the specific surface area of the catalyst, which suggests that the specific surface area is not a key factor affecting the photocatalytic performances of the samples.
|
| Fig. 2 (a) XRD patterns and (b) DRS results of CIS and the CIS/Ag16(GSH)9 heterostructure, with (c) transformed plots calculated by the Kubelka–Munk function vs. the energy of light along with sample color in the inset [left: CIS; right: CIS/Ag16(GSH)9]. (d) N2 adsorption–desorption isotherms of CIS and the CIS/Ag16(GSH)9 heterostructure, with pore size distribution curves in the inset. (e) Raman spectra of CIS and the CIS/Ag16(GSH)9 heterostructure. High-resolution (f) S 2p, (g) In 3d, (h) Cd 3d, and (i) Ag 3d spectra of (I) CIS and (II) CIS/Ag16(GSH)9. | |
Fig. S5† shows the survey X-ray photoelectron spectra (XPS) of CIS and CIS/Ag16(GSH)9 with the signals of C, Cd, In, and S. Specifically, as displayed in Fig. S5a,† a clear Ag signal was detected in the survey XPS spectrum of the CIS/Ag16(GSH)9 heterostructure, which suggests that Ag16(GSH)9 NCs were successfully loaded on the CIS surface. The high-resolution N 1s and C 1s spectra (Fig. S5b & c†) also indicate the loading of Ag16(GSH)9 NCs on the CIS substrate. As shown in Table S2† and Fig. 2f (I), the binding energies (BEs) of the high-resolution S 2p spectrum of CIS are about 161.57 eV (S 2p3/2) and 162.76 eV (S 2p1/2), and these peaks correspond to the S2− species.41 As shown in Fig. 2g (I), the BEs of the high-resolution In 3d spectrum of CIS are about 444.95 eV (In 3d5/2) and 452.53 eV (In 3d3/2), and these peaks mainly originate from the In3+ species.42 As shown in Fig. 2h (I), the main peaks in the high-resolution Cd 3d spectrum of CIS are located at 405.33 eV (Cd 3d5/2) and 412.06 eV (Cd 3d3/2), which correspond to the Cd2+ species.42 Compared with the CIS substrate, as shown in Fig. 2f (II), g (II) and h (II), the peak positions of the CIS/Ag16(GSH)9 heterostructure were not shifted, indicating that the surface valence states of Ag16(GSH)9 NCs and CIS were not altered by NCs deposition. Notably, double-peak bands were observed in the high-resolution Ag 3d spectrum (Fig. 2i) of the CIS/Ag16(GSH)9 heterostructure, wherein the peaks at 367.82 eV (Ag 3d5/2) &373.79 eV (Ag 3d5/2) and 368.29e V (Ag 3d5/2) & 374.34 eV (Ag 3d5/2) correspond to the Ag+ and Ag0 species, respectively. This confirmed the successful loading of Ag16(GSH)9 NCs on the CIS surface. Table S3† summarizes the relationship between the chemical bonding species and BEs for the CIS substrate and CIS/Ag16(GSH)9 heterostructure.
3.2. Photocatalytic activities
To investigate the photocatalytic activities of the CIS/Ag16(GSH)9 heterostructure, the selected probe reaction was the anaerobic selective photoreduction of nitroaromatic compounds to amino compounds under visible light irradiation (>420 nm) and ambient conditions. Here, 4-nitroaniline (4-NA) was used as the representative nitro compound in the photocatalytic reaction; according to the optical absorption spectrum of 4-NA, its absorbance at 380 nm gradually decreased with the addition of a hole trapping agent (Na2SO3) and the continuous bubbling of nitrogen. This confirmed the successful reduction of 4-NA to 4-phenylenediammonium (4-PDA).43 We ascertained that it was indeed a photocatalytic reaction based on the control experiments without catalyst or light irradiation (Fig. S6†). The influence of Ag16(GSH)9 NCs loading percentage on the photocatalytic activities was then probed (Fig. S7†). In our work, the loading percentage of Ag16(GSH)9 NCs on the CIS was tuned by varying the volume of Ag16(GSH)9 NCs. The results show that the photoactivity of CIS/Ag16(GSH)9 gradually increased with increasing the Ag16(GSH)9 NCs loading amount and reached the maximum value when x = 0.3 (x refers to the volume ratio of NCs), and then it decreased when the Ag16(GSH)9 NCs loading amount was further increased. This may be because active sites on the CIS surface and light absorption of CIS are shielded by the metal NCs, resulting in decreased carrier density and reduced photoactivity; therefore, the optimal loading ratio of the CIS/Ag16(GSH)9 heterostructure can be determined.
The photocatalytic activity of the CIS/Ag16(GSH)9 heterostructure toward 4-NA reduction increased by ca. 65% in conversion as compared with the blank CIS substrate, confirming that Ag16(GSH)9 NCs plays a great role in boosting the photocatalytic reaction. In addition to 4-NA (Fig. 3a), the photocatalytic activities of the CIS/Ag16(GSH)9 heterostructure for the selective reduction of other nitro compounds were also explored and interestingly, similar results were observed. As shown in Fig. 3b–i, the CIS/Ag16(GSH)9 heterostructure exhibited a more enhanced photocatalytic activity relative to the blank CIS for the selective photoreduction of 3-NA (Fig. 3b), 2-NA (Fig. 3c), 1-chloro-4-nitrobenzene(Fig. 3d), 1-bromo-4-nitrobenzene (Fig. 3e), o-nitroacetophenone (Fig. 3f), 4-nitroanisole (Fig. 3g), NB (Fig. 3h), and 4-NT (Fig. 3i) under visible light irradiation. Notably, the photocatalytic performances of the CIS/Ag16(GSH)9 heterostructure were generally enhanced by a factor of 2–3 relative to the pristine CIS substrate, which strongly implies the synergistic effect of CIS and Ag16(GSH)9 NCs in boosting the photocatalytic activity. Besides the photoreduction reaction, the photocatalytic oxidative capacity of the CIS/Ag16(GSH)9 heterostructure was also probed by the oxidative degradation of methyl orange (MO) under visible light irradiation. Fig. S8† indicates that the CIS/Ag16(GSH)9 heterostructure demonstrated enhanced degradation efficiency in comparison with the blank CIS.
|
| Fig. 3 Photoactivities of CIS and the CIS/Ag16(GSH)9 heterostructure toward the selective photoreduction of nitroaromatics under visible irradiation (λ > 420 nm) with the addition of hole scavengers and N2 purging, including (a) 4-NA, (b) 3-NA, (c) 2-NA, (d) 1-chloro-4-nitrobenzene, (e) 1-bromo-4-nitrobenzene, (f) 2-nitroacetophenone, (g) 4-nitroanisole, (h) NB, and (i) 4-nitrotoluene together with the reaction model. | |
As shown in Fig. S9,† the photocatalytic performances of the CIS/Ag16(GSH)9 heterostructure with and without adding an electron trapping agent (K2S2O8) were probed. The results show that the photocatalytic activity of the CIS/Ag16(GSH)9 heterostructure was remarkably reduced with adding K2S2O8 to the reaction system, thus suggesting that photogenerated electrons play a decisive role in the photocatalytic reduction reaction of aromatic nitro compounds. Of particular note, we found that the photocatalytic activity of the CIS/Ag16(GSH)9 heterostructure was markedly reduced without adding the hole-trapping agent (Na2SO3), thus proving that the quenching of holes by Na2SO3 is essential in the photocatalytic reduction reaction. The addition of hole-trapping agents ensures that the electrons function as the sole active species and can efficiently photoreduce nitroaromatic compounds to amino derivatives, thus fulfilling the photoreduction catalysis. The stability of the CIS/Ag16(GSH)9 heterostructure was explored by a cyclic reaction, which is of paramount importance for practical applications. It was found that the stability of the CIS/Ag16(GSH)9 heterostructure was not substantially decreased after the cyclic test, which may be due to the rapid transfer of holes from CdIn2S4 to the metal NCs, which reduced the charge separation and retarded the photo-corrosion of CdIn2S4. The relatively good photocatalytic activity of the CIS/Ag16(GSH)9 heterostructure was observed after the cyclic reaction (Fig. S10†), which proved the favorable stability of the CIS/Ag16(GSH)9 heterostructure. Alternatively, XRD and XPS analyses of the CIS/Ag16(GSH)9 heterostructure after the cyclic reaction were also performed to further evaluate its photostability. As shown in Fig. S11 & S12,† the crystal structure and surface elemental valence states of the CIS/Ag16(GSH)9 heterostructure were not changed after the cyclic reaction, and the analytical results were the same as those of the newly prepared catalysts, which once again proved its good stability.
3.3. PEC performances
Photoelectrochemical (PEC) measurements were carried out to probe the charge separation efficiency of the CIS/Ag16(GSH)9 heterostructure. As shown in Fig. 4a, the transient photocurrent of the CIS/Ag16(GSH)9 heterostructure is significantly higher than that of the CIS substrate, indicating that the CIS/Ag16(GSH)9 heterostructure demonstrates a more efficient charge separation efficiency relative to the blank CIS. Mott–Schottky (M–S) results were probed to reveal the carrier density (Fig. 4b). The carrier density of the CIS/Ag16(GSH)9 heterostructure was calculated based on the (M–S) results (Fig. 4b), which far exceeded that of the CIS substrate (Fig. 4c). Electrochemical impedance spectroscopy (EIS) results suggest that the CIS/Ag16(GSH)9 heterostructure demonstrated a smaller semicircular arc radius relative to CIS, which implies its lower interfacial charge transfer resistance (Fig. 4d). As shown in Fig. 4e, the CIS/Ag16(GSH)9 heterostructure exhibited a larger photovoltage than CIS, which also confirmed the more efficient charge separation over the CIS/Ag16(GSH)9 heterostructure compared with CIS. This is consistent with other PEC results. As shown in Fig. 4f, the photoluminescence (PL) intensity of the CIS/Ag16(GSH)9 heterostructure is lower than that of CIS, confirming that the CIS/Ag16(GSH)9 heterostructure shows more enhanced charge migration efficiency compared with the CIS substrate. The charge carrier density was calculated using the following equation: | | (3) |
where ε is the dielectric constant (εCIS = 6.60), ε0 is the vacuum permittivity (8.86 × 10−12 F m−1), e0 is the unit of charge (1.6 × 10−19 C), and V is the potential.
|
| Fig. 4 (a) On–off transient photocurrents, (b) M–S results, (c) carrier density, (d) EIS results, (e) open-circuit potential decay plots, and (f) PL spectra of CIS and the CIS/Ag16(GSH)9 heterostructure (excitation wavelength: 350 nm). | |
3.4. Photocatalytic mechanism
Scheme 2 presents the photocatalytic mechanism of the CIS/Ag16(GSH)9 heterostructure according to the above analysis. First, the energy levels of the CIS substrate were determined by M–S and DRS results. As shown in Fig. S13,† the flat-band potential of the CIS was determined as −0.3 V vs. NHE. Considering its bandgap value of Eg = 2.27 eV, the VB of CIS was calculated as 1.97 V vs. NHE based on the formula Eg = EVB − ECB. The HOMO–LUMO levels of Ag16(GSH)9 NCs were determined from the CV and UV-vis absorption spectrum (Fig. S13†). As shown in Fig. S14a,† carbon paper without loading Ag16(GSH)9 NCs demonstrated no obvious peak, whereas an apparent peak at 0.78 V vs. NHE appeared after depositing Ag16(GSH)9NCs, and the peak intensity was boosted with increasing the amount of Ag16(GSH)9 NCs. Therefore, the HOMO and LUMO levels of Ag16(GSH)9 NCs were determined as −1.8 V vs. NHE and 0.78 V vs. NHE, respectively. Ag16(GSH)9 NCs and CIS form the type-II energy level alignment, which is beneficial for interfacial charge transfer.44 Thus, when CIS and Ag16(GSH)9 NCs are simultaneously photoexcited under visible light irradiation to produce electron–hole charge carriers, the electrons in the LUMO level of Ag16(GSH)9 NCs migrate to the CB of CIS, while the holes in the VB of CIS are transferred to the HOMO level of Ag16(GSH)9 NCs,45 promoting the carrier separation and boosting the photocatalytic activity. Consequently, the electrons participate in the photocatalytic reduction of nitroaromatics to amino compounds. It should be noted that the holes accumulating in the LUMO level of Ag16(GSH)9 NCs are effectively captured by hole scavengers or participate in the dye oxidative degradation reaction.
|
| Scheme 2 Schematic illustration of the photocatalytic mechanism of the CIS/Ag16(GSH)9 heterostructure. | |
4. Conclusions
To conclude, the CIS/Ag16(GSH)9 heterostructure was constructed by the self-assembly of atomically precise Ag16(GSH)9 NCs on the CIS substrate under ambient conditions. The self-assembled CIS/Ag16(GSH)9 heterostructure demonstrated remarkably enhanced and stable photocatalytic performances in the selective reduction of nitroaromatic compounds to amino derivatives under visible light irradiation in an anaerobic environment, far surpassing CIS under the same experimental conditions. This is mainly ascribed to the efficient charge transport between Ag16(GSH)9 NCs and the CIS substrate, thereby effectively preventing the recombination of photogenerated electron–hole pairs. The favorable charge transfer between Ag16(GSH)9 NCs and CIS was determined. Our work provides a typical paradigm for fabricating semiconductor–metal NC heterostructures for solar energy conversion.
Author contributions
Junyi Zhang: Writing the original draft, editing, formal analysis. Linjian Zhan: Methodology. Boyuan Ning: Investigation, validation. Yunhui He: Instrument testing, Zhixin Chen: Instrumental analysis, Guangcan Xiao and Fang-Xing Xiao: Supervision.
Data availability
The data supporting this article have been included as part of the ESI.†
Conflicts of interest
The authors declare no conflict of interest.
Acknowledgements
This work was financially supported by the National Natural Science Foundation of China (No. 21703038, 22072025). The financial support from the State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Science, is acknowledged (No. 20240018).
References
- F. Zheng, W. Zhang, Q. Guo, B. Yu, D. Wang and W. Chen, Metal clusters confined in porous nanostructures: synthesis, properties and applications in energy catalysis, Coord. Chem. Rev., 2024, 502, 215603 CrossRef CAS.
- T. Kawawaki, T. Okada, D. Hirayama and Y. Negishi, Atomically precise metal nanoclusters as catalysts for electrocatalytic CO2 reduction, Green Chem., 2024, 26, 122–163 RSC.
- X. Zou, X. Kang and M. Zhu, Recent developments in the investigation of driving forces for transforming coinage metal nanoclusters, Chem. Soc. Rev., 2023, 52, 5892–5967 RSC.
- T. Higaki, Y. Li, S. Zhao, Q. Li, S. Li, X. Du, S. Yang, J. Chai and R. Jin, Atomically tailored gold nanoclusters for catalytic application, Angew. Chem., Int. Ed., 2019, 58, 8291–8302 CrossRef CAS PubMed.
- M. Wang, Z. Wu, Z. Chu, J. Yang and C. Yao, Chemico-physical synthesis of surfactant- and ligand-free gold nanoparticles and their anti-galvanic reduction property, Chem. – Asian J., 2014, 9, 1006–1010 CrossRef CAS PubMed.
- Y. Negishi, T. Nakazaki, S. Malola, S. Takano, Y. Niihori, W. Kurashige, S. Yamazoe, T. Tsukuda and H. Häkkinen, A critical size for emergence of nonbulk electronic and geometric structures in dodecanethiolate-protected au clusters, J. Am. Chem. Soc., 2015, 137, 1206–1212 CrossRef CAS PubMed.
- Y. Li and F. Xiao, Tuning the photosensitization efficiency of atomically precise metal nanoclusters by super-efficient and exquisite interface modulation, J. Mater. Chem. A, 2023, 11, 589–599 RSC.
- Y. Xiao, Q. Mo, G. Wu, K. Wang, X. Ge, S. Xu, J. Li, Y. Wu and F.-X. Xiao, Charge modulation over atomically precise metal nanoclusters via non-conjugated polymers for photoelectrochemical water oxidation, J. Mater. Chem. A, 2023, 11, 2402–2411 RSC.
- X. Mai, J. Tang, J. Tang, X. Zhu, Z. Yang, X. Liu, X. Zhuang, G. Feng and L. Tang, Research progress on the environmental risk assessment and remediation technologies of heavy metal pollution in agricultural soil, J. Environ. Sci., 2025, 149, 1–20 CrossRef PubMed.
- X. Liu, Y. He, Z. Li, A. Cheng, Z. Song, Z. Yu, S. Chai, C. Cheng and C. He, Size transformation of au nanoclusters for enhanced photocatalytic hydrogen generation: interaction behavior at nanocluster/semiconductor interface, J. Colloid Interface Sci., 2023, 651, 368–375 CrossRef CAS PubMed.
- X. Yan, J. Dong, J. Zheng, Y. Wu and F.-X. Xiao, Customizing precise, tunable, and universal cascade charge transfer chains towards versatile photoredox catalysis, Chem. Sci., 2024, 15, 2898–2913 RSC.
- M. Miyauchi, H. Irie, M. Liu, X. Qiu, H. Yu, K. Sunada and K. Hashimoto, Visible-light-sensitive photocatalysts: nanocluster-grafted titanium dioxide for indoor environmental remediation, J. Phys. Chem. Lett., 2016, 7, 75–84 CrossRef CAS PubMed.
- X. Yan, X. Fu and F. Xiao, Filling the gap: atomically precise metal nanoclusters-induced z-scheme photosystem toward robust and stable solar hydrogen generation, Adv. Funct. Mater., 2023, 33, 2303737 CrossRef CAS.
- M. Chhowalla, H. S. Shin, G. Eda, L. Li, K. P. Loh and H. Zhang, The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets, Nat. Chem., 2013, 5, 263–275 CrossRef PubMed.
- Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman and M. S. Strano, Electronics and optoelectronics of two-dimensional transition metal dichalcogenides, Nat. Nanotechnol., 2012, 7, 699–712 CrossRef CAS PubMed.
- Y. Shiga, N. Umezawa, N. Srinivasan, S. Koyasu, E. Sakai and M. Miyauchi, A metal sulfide photocatalyst composed of ubiquitous elements for solar hydrogen production, Chem. Commun., 2016, 52, 7470–7473 RSC.
- W. Chong, B. Ng, L. Tan and S. Chai, Recent advances in nanoscale engineering of ternary metal sulfide-based heterostructures for photocatalytic water splitting applications, Energy Fuels, 2022, 36, 4250–4267 CrossRef CAS.
- L. Zhao, B. Yang, G. Zhuang, Y. Wen, T. Zhang, M. Lin, Z. Zhuang and Y. Yu, Thin in-plane In2O3/ZnIn2S4 heterostructure formed by topological-atom-extraction: optimal distance and charge transfer for effective CO2 photoreduction, Small, 2022, 18, 2201668 CrossRef CAS PubMed.
- C. Y. Toe, S. Zhou, M. Gunawan, X. Lu, Y. H. Ng and R. Amal, Recent advances and the design criteria of metal sulfide photocathodes and photoanodes for photoelectrocatalysis, J. Mater. Chem. A, 2021, 9, 20277–20319 RSC.
- Y. Liu, D. Huang, M. Cheng, Z. Liu, C. Lai, C. Zhang, C. Zhou, W. Xiong, L. Qin, B. Shao and Q. Liang, Metal sulfide/MOF-based composites as visible-light-driven photocatalysts for enhanced hydrogen production from water splitting, Coord. Chem. Rev., 2020, 409, 213220 CrossRef CAS.
- Q. Meng, Z. Jin and G. Liu, Sufficient and necessary conditions for oscillation of linear fractional-order delay differential equations, Adv. Differ. Equations, 2021, 2021, 89 CrossRef.
- X. Zhang, H. Liang, H. Li, Y. Xia, X. Zhu, L. Peng, W. Zhang, L. Liu, T. Zhao, C. Wang, Z. Zhao, C. Hung, M. M. Zagho, A. A. Elzatahry, W. Li and D. Zhao, Sequential chemistry toward core–shell structured metal sulfides as stable and highly efficient visible-light photocatalysts, Angew. Chem., Int. Ed., 2020, 59, 3287–3293 CrossRef CAS PubMed.
- T. Di, Q. Xu, W. Ho, H. Tang, Q. Xiang and J. Yu, Review on metal sulphide-based z-scheme photocatalysts, ChemCatChem, 2019, 11, 1394–1411 CrossRef CAS.
- H. Xie, K. Wang, D. Xiang, S. Li and Z. Jin, Enwrapping graphdiyne (g-CnH2n−2) on hollow NiCo2O4 nanocages derived from a prussian blue analogue as a p–n heterojunction for highly efficient photocatalytic hydrogen evolution, J. Mater. Chem. A, 2023, 11, 14971–14989 RSC.
- M. Ghosal Chowdhury, L. Sahoo, S. Maity, D. Bain, U. K. Gautam and A. Patra, Silver nanocluster/MoS2 heterostructures for hydrogen evolution, ACS Appl. Nano Mater., 2022, 5, 7132–7141 CrossRef CAS.
- J. Q. Hu, B. Deng, W. X. Zhang, K. B. Tang and Y. T. Qian, Synthesis and characterization of CdIn2S4 nanorods by converting CdS nanorods via the hydrothermal route, Inorg. Chem., 2001, 40, 3130–3133 CrossRef CAS PubMed.
- X. Yuan, M. I. Setyawati, A. S. Tan, C. N. Ong, D. T. Leong and J. Xie, Highly luminescent silver nanoclusters with tunable emissions: cyclic reduction–decomposition synthesis and antimicrobial properties, NPG Asia Mater., 2013, 5, e39 CrossRef CAS.
- Q. Zhang, J. Wang, X. Ye, Z. Hui, L. Ye, X. Wang and S. Chen, Self-assembly of CdS/CdIn2S4 heterostructure with enhanced photocascade synthesis of schiff base compounds in an aromatic alcohols and nitrobenzene system with visible light, ACS Appl. Mater. Interfaces, 2019, 11, 46735–46745 CrossRef CAS PubMed.
- Y. Xie, D. Liu, B. Wang, D. Li, Z. Yan, Y. Chen, J. Shen, Z. Zhang and X. Wang, Monolayer Bi2W1−xMoxO6 solid solutions for structural polarity to boost photocatalytic reduction of nitrobenzene under visible light, ACS Sustainable Chem. Eng., 2021, 9, 2465–2474 CrossRef CAS.
- E. Pulido Melián, O. González Díaz, J. M. Doña Rodríguez, G. Colón, J. A. Navío, M. Macías and J. Pérez Peña, Effect of deposition of silver on structural characteristics and photoactivity of TiO2-based photocatalysts, Appl. Catal., B, 2012, 127, 112–120 CrossRef.
- B. Xi, X. Chu, J. Hu, C. S. Bhatia, A. J. Danner and H. Yang, Preparation of Ag/TiO2/SiO2 films via photo-assisted deposition and adsorptive self-assembly for catalytic bactericidal application, Appl. Surf. Sci., 2014, 311, 582–592 CrossRef CAS.
- L. Zhang, H. Jiang, T. Huang, Y. Song, Y. Wang, F. Wang, L. Fan, X. Liu, L. Yang and H. Liu, Ag-lspr and molecular additive: a collaborative approach to improve the photovoltaic performance of perovskite solar cells, Chem. Eng. J., 2024, 481, 148572 CrossRef CAS.
- B. Tu, K. Yu, D. Fu, L. Zhou, R. Wang, X. Jiang, H. Liu, X. Cao, X. Gong, R. He, Y. Tang, T. Chen and W. Zhu, Amino-rich Ag-NWs/NH2-MIL-125(Ti) hybrid heterostructure via lspr effect for photo-assist uranium extraction from fluorine-containing uranium wastewater without sacrificial agents, Appl. Catal., B, 2023, 337, 122965 CrossRef CAS.
- Q. Zhang, Q. An, X. Luan, H. Huang, X. Li, Z. Meng, W. Tong, X. Chen, P. K. Chu and Y. Zhang, Achieving significantly enhanced visible-light photocatalytic efficiency using a polyelectrolyte: the composites of exfoliated titania nanosheets, graphene, and poly(diallyl-dimethyl-ammonium chloride), Nanoscale, 2015, 7, 14002–14009 RSC.
- Q. Mo, X. Lin, Z. Wei, X. Dai, S. Hou, T. Li and F.-X. Xiao, All-in-one: branched macromolecule-protected metal nanocrystals as integrated charge separation/motion centers for enhanced photocatalytic selective organic transformations, J. Mater. Chem. A, 2020, 8, 16392–16404 RSC.
- Z. Zeng, T. Li, Y. Li, X. Dai, M. Huang, Y. He, G. Xiao and F.-X. Xiao, Plasmon-induced photoelectrochemical water oxidation enabled by in situ layer-by-layer construction of cascade charge transfer channel in multilayered photoanode, J. Mater. Chem. A, 2018, 6, 24686–24692 RSC.
- Y. Wang, X. Liu, Q. Wang, M. Quick, S. A. Kovalenko, Q. Chen, N. Koch and N. Pinna, Insights into charge transfer at an atomically precise nanocluster/semiconductor interface, Angew. Chem., Int. Ed., 2020, 59, 7748–7754 CrossRef CAS PubMed.
- S. Biswas, S. Das and Y. Negishi, Progress and prospects in the design of functional atomically-precise Ag(i)-thiolate nanoclusters and their assembly approaches, Coord. Chem. Rev., 2023, 492, 215255 CrossRef CAS.
- Q. Mo, J. Li, S. Xu, K. Wang, X. Ge, Y. Xiao, G. Wu and F.-X. Xiao, Unexpected insulating polymer maneuvered solar CO2-to-syngas conversion, Adv. Funct. Mater., 2023, 33, 2210332 CrossRef CAS.
- W. K. Unger, B. Farnworth, J. C. Irwin and H. Pink, Raman and infrared spectra of CdIn2S4 and ZnIn2S4, Solid State Commun., 1978, 25, 913–915 CrossRef CAS.
- Y. Song, Y. Wu, S. Cao, Y. Zhang, D. Luo, J. Gao, Z. Li, L. Sun and J. Hou, Simultaneous photoelectrocatalytic oxidation and nitrite-ammonia conversion with
artificial photoelectrochemistry cells, Adv. Energy Mater., 2022, 12, 2201782 CrossRef CAS.
- F. Wang, W. Zhang, H. Liu, R. Cao and M. Chen, Roles of CeO2 in preparing Ce-doped CdIn2S4 with boosted photocatalytic degradation performance for methyl orange and tetracycline hydrochloride, Chemosphere, 2023, 338, 139574 CrossRef CAS PubMed.
- E. Deng, Y. Fan, H. Wang, Y. Li, C. Peng and J. Liu, Engineering a Z-scheme heterostructure on ZnIn2S4@NH2-MIL-125 composites for boosting the photocatalytic performance, Inorg. Chem., 2024, 63, 1449–1461 CrossRef CAS PubMed.
- Z. Wei, S. Hou, X. Lin, S. Xu, X. Dai, Y. Li, J. Li, F.-X. Xiao and Y. Xu, Unexpected boosted solar water oxidation by nonconjugated polymer-mediated tandem charge transfer, J. Am. Chem. Soc., 2020, 142, 21899–21912 CrossRef CAS PubMed.
- B. Tang, S. Zhu, H. Liang, S. Li, B. Liu and F.-X. Xiao, Tuning atomically precise metal nanocluster mediated photoelectrocatalysis via a non-conjugated polymer, J. Mater. Chem. A, 2022, 10, 4032–4042 RSC.
|
This journal is © the Partner Organisations 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.