Next Article in Journal
New Cu-Free Ti-Based Composites with Residual Amorphous Matrix
Previous Article in Journal
Lime as an Anti-Plasticizer for Self-Compacting Clay Concrete
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of AgInS2-xAg2S-yZnS-zIn6S7 (x, y, z = 0, or 1) Nanocomposites with Composition-Dependent Activity towards Solar Hydrogen Evolution

1
College of Science, China University of Petroleum, Qingdao 266580, China
2
State Key Laboratory of Heavy Oil Processing, College of Chemical Engineering, China University of Petroleum, Qingdao 266580, China
*
Authors to whom correspondence should be addressed.
Materials 2016, 9(5), 329; https://doi.org/10.3390/ma9050329
Submission received: 25 March 2016 / Revised: 20 April 2016 / Accepted: 26 April 2016 / Published: 29 April 2016
(This article belongs to the Section Advanced Composites)

Abstract

:
Metal sulfides-based nanomaterials have been used as a class of efficient solar driven photocatalysts. However, the H2-production rate observed over these photocatalysts remains problematic. Here, the AgInS2-xAg2S-yZnS-zIn6S7 (x, y, z = 0 or 1) nanocomposites with controlled compositions have been successfully prepared by a simple hydrothermal method with AgI polyhedrons as silver source. The obtained AgInS2-xAg2S-yZnS-zIn6S7 nanocomposites showed a composition-dependent activity for H2 evolution from aqueous solution under simulated sun-light irradiation. The results showed that the optimized product of AgInS2-Ag2S-ZnS nanoparticles synthesized with the precursor ratio of Ag:Zn = 1:1 exhibited the highest H2 evolution rate of 5.4 mmol·g−1·h−1. Furthermore, the catalyst can be used for 20 h without loss of activity, showing its high stability. It opens a new path to achieve highly efficient solar photocatalyst for H2 evolution from water splitting.

1. Introduction

Two major kinds of photocatalysts in terms of composition are mainly used to drive solar water-splitting [1]. One is single-component material possessing proper band structure aligned for hydrogen generation. The other is multi-component composite made of different photo-absorbers with the matched band alignments to split water. The bandgap of these materials should be relatively narrow, which can efficiently favor the absorption of solar light [2,3,4]. Additionally, the conduction band edge must be sufficiently more negative than the reduction potential of protons. Compared with single-component material, making a complex catalyst composite has several advantages. First, the potentials of the conduction and valence bands shift successively with composition. According to the band gap transition, the composite is able to absorb enough visible light. Second, photogenerated electrons and holes can move smoothly in the continuous valence and conduction bands, rather than in the discrete levels seen in the single photocatalyst. A series of ZnS-CuInS2-AgInS2 with a different percentage of the components have been successfully prepared, whose absorption edges were shifted monotonically to longer wavelengths as the ratio of MInS2 (M = Cu and Ag) to ZnS increases [5]. Moreover, optoelectronic materials with electronically coupled components usually possess interesting and unique optical and electronic properties, i.e., enlarged absorption region of the solar spectrum, facilitated charge transportation without increasing the rate of charge recombination, and enhanced charge collection efficiency [6,7]. These excellent properties make optoelectronic composites potential candidates for application in the fields of photocatalysis. The nanocomposite theoretically has the ability to provide a high photocatalytic water-splitting efficiency. The heterojunction with different energy levels may form an ideal system to cause a rapid photo-induced charge separation and decreased recombination chance of electron–hole pairs [8,9]. Therefore, under the irradiation of solar energy, a photoinduced electron would transfer efficiently in photochemically stable semiconductor nanocrystals by the synergetic effect of the multi-component system [10,11,12,13].
As is the case with metal oxide semiconductors, e.g., TiO2 and ZnO, I-III-VI2 (e.g., (Cu)AgInS2) [14,15,16,17,18] and II-VI (e.g., Zn(Cd)S) [19,20,21,22], groups have been developed as a new platform for photocatalytic applications. With the rapid growth of research in this area, synthesis strategies such as alloying, doping, and solid solution nanomaterials have attracted much attention with regard to band gap and electronic wave function engineering [23,24]. For example, Kudo and Peng synthesized (AgIn)xZn2(1-x)S2 and (CuIn)xZn2(1-x)S2 solid solutions with the capacity to reduce water to H2 under visible light [14,17]. Mao et al. investigated the tunable light absorption and photoluminescence properties of AgInS2/ZnS nanocrystals [25]. Michalska et al. synthesized core/shell CuInS2/ZnS quantum dots and studied the optical features and structure in detail [26]. However, there are no obvious heterointerfaces in solid solutions and the doping system, and it is difficult to achieve the morphology and size control. In semiconductor nanocomposites, alloying of the heterojunction interface can reduce the lattice mismatch between different components, where the mixed compositions allow for gradual strain release with improved optical properties [27,28]. The reaction chemistry of the AgInS2-ZnS system is complicated and very intriguing. It can favor the design of photocatalysts with widely tunable properties. However, it is still not well understood how to control the structure of the multi-component, and to tune the photochemistry properties of alloying semiconductors.
Herein, we have developed a facile hydrothermal method to prepare AgInS2-xAg2S-yZnS-zIn6S7 (x, y, z = 0 or 1) nanocomposites. The compositions of the resulting nanocomposites can be controlled through varying the ratios of feeding stocks, which are the main factors in determining their activities towards solar hydrogen evolution. The experimental results demonstrated that the heterojunction structures exhibited much higher photocatalytic activity than single AgInS2 and AgInS2-Ag2S for solar H2 evolution.

2. Results and Discussion

In order to obtain an optimized efficient solar photocatalyst for H2 production, the multi-component nanocomposites were synthesized at various molar ratios of Ag+ to Zn2+ ions. Figure 1 shows XRD patterns of the as-synthesized AgInS2-xAg2S-yZnS-zIn6S7 nanoheterojunctions produced at different Ag:Zn ratios. It shows that only single orthorhombic phase of AgInS2 is obtained when the starting molar ratio of Ag:Zn is 5:1 (Figure 1A). Compared with tetragonal AgInS2, the orthorhombic AgInS2 is usually crystallized at high temperature. In our work, the metastable orthorhombic AgInS2 can be obtained under a mild condition (180 °C), which may be due to the proper chemical environment and conditions used for the materials preparation. With decreasing the ratio to 2:1, monoclinic Ag2S appears in the product (Figure 1B). Further increasing the dosage of Zn2+ (Ag:Zn = 1:1), as shown in Figure 1C, the characteristic peak of hexagonal ZnS appears, meaning that AgInS2-Ag2S-ZnS nanocomposite formed. However, as shown in Figure 1D,E, further increasing of the amount of Zn2+ ions it results in the formation of In6S7, and the product thus switches to AgInS2-Ag2S-ZnS-In6S7. It is clear that no other characteristic peaks for external impurities are observed, indicating that the phase composition of the final products can be controlled through varying the addition of Zn2+ ions.
Figure 2 shows a series of SEM images of the as-synthesized samples and a comparison of their solar hydrogen evolution property correspondingly. As revealed in Figure 2A–E, different molar ratios of Ag:Zn result in the formation of products with diverse compositions and morphologies. The obtained AgInS2 nanoparticles with molar ratio of Ag:Zn at 5:1 display regular flowers with the size of 810 nm and are assembled by many nanosheets (Figure 2A). With the ratio decreased to 2:1, the produced AgInS2-Ag2S is mainly composed of nanoplates with a length of 300–400 nm and thickness of several tens of nanometer (Figure 2B). However, further increasing the dosage of zinc ions, i.e., Ag:Zn = 1:1, as shown in Figure 2C, the as-obtained AgInS2-Ag2S-ZnS shows aggregation of the nanoplates. When the mole number of Zn is higher than that of Ag, the product contains AgInS2-Ag2S-ZnS-In6S7. The particles reveal irregular and aggregated morphology (Figure 2D,E). Figure S1 shows diffuse reflectance spectra (DRS) of the prepared composites. Obviously, when the amount of Zn added in the preparation is higher than that of Ag (Ag:Zn = 1:5 or 1:2), the DRS spectra exhibit intense absorption bands with steep edges, which mainly depends on the UV absorption band of ZnS. With the decrease in the dosage of Zn(Ac)2, the composites have absorptions in the range of visible light (λ > 450 nm), whereas ZnS has little absorption in this range. The presence of AgInS2 and Ag2S contributes to the increased visible-light absorption. Notably, the absorption spectrum of AgInS2-xAg2S-yZnS-zIn6S7 composite photocatalysts can be easily controlled by modifying the dosage of Zn(Ac)2 continuously. The complexity of the composite in both structure and chemical composition may endow the materials with exceptional optical properties.
Figure 2F shows the photocatalytic activities of the photocatalysts. The hydrogen production mainly depends on the compositions regardless of their morphology. The AgInS2-Ag2S-ZnS with the feed ratio of Ag:Zn at 1:1 exhibits the highest activity with a hydrogen evolution rate of 5.4 mmol·g−1·h−1, which is higher than that of previous reported AgInS2, Ag2S, ZnS composite photoctalysts (comparison results is shown in Table S1). Moreover, the photocatalyst presents good stability even after 20 h duration (Supplementary Materials, Figure S2). The pristine AgInS2 nanoparticles possessed very weak photo-activity towards H2 evolution. Of course, Pt salt added in the photocatalytic reaction played an important role in promoting the photocatalytic activity. Without cocatalyst of Pt salt, the catalyst itself also achieved a comparable activity towards solar hydrogen evolution of a rate of 1.8 mmol·g−1·h−1 (Supplementary Materials, Figure S3). Figure S4 reveals the surface area of the samples, suggesting that surface characteristics do not dominate the photocatalytic activity for H2 evolution. In other words, activities of the nanocomposites are mainly determined by their compositions.
Typically, we further investigated the detailed structures of AgInS2-Ag2S-ZnS with the maximum activity by HRTEM measurement. Figure 3 clearly shows that the AgInS2-Ag2S-ZnS nanocomposite has clear interconnecting lattice fringes with interplaner spacing distance of 0.319, 0.285, and 0.295 nm, which correspond to (121) and (002) planes for orthorhombic AgInS2, (−112) plane for monoclinic Ag2S, and (101) plane for hexagonal ZnS, respectively. The results demonstrate that the photocatalyst composites have been successfully prepared through the simple hydrothermal method.
The photocurrent response is also used to compare the photo-activity on the as-synthesized samples. As shown in Figure 4, all the samples can be excited to produce photo-carriers under simulated sunlight illumination. The electrons transport from the photocatalysts to ITO glass surface and then external circuit to produce photocurrent [29]. Prompt and reproducible current responses during repeating on-off illumination cycles are observed from each electrode of the five samples. When the illumination stops, the photocurrent drops to zero immediately. The photocurrent density increases with the increase of ratio of Ag:Zn. For AgInS2-Ag2S-ZnS-In6S7 (Ag:Zn = 1:5), the photocurrent response is very weak. The result can be ascribed to the increased content of ZnS bearing a wide bandgap (3.55 eV), leading to the low utilization ratio of solar photons. In contrast, the AgInS2-Ag2S-ZnS-In6S7 produced at Ag:Zn = 1:2 exhibits an improved photocurrent. For the AgInS2-Ag2S-ZnS obtained at Ag:Zn = 1:1, a photocurrent peak was observed at the initial irradiation stage, followed by a slow decrease before reaching a constant value. It suggests that recombination of photo-generated electrons and holes occur in the process [30]. Meanwhile, the photocurrent of the other two samples obtained with Ag:Zn = 2:1 and Ag:Zn = 5:1, i.e., AgInS2-ZnS and AgInS2, are significantly higher than the above samples. Nevertheless, the current decreases along with the irradiation time, indicating the easy excitation and fast recombination of photo-generated carriers. The results are in accordance with previous reports [31], which suggest that the production of photocurrent is primarily determined by the transporting speed of excited electrons from semiconductor to ITO and the recombination at the electrolyte/film interface. Considering the photo-activity of photocurrent production and H2 evolution, it is deduced that the photocatalysts with stable and high photocurrent are favorable for the H2 generation. Therefore, the production and transfer rate of photo-generated electrons for the samples obtained with Ag:Zn of 1:1, 1:2, 1:5 are consisted with their hydrogen evolution activity. For the AgInS2 and AgInS2-ZnS nanoparticles obtained from Ag:Zn of 5:1 and 2:1, the hydrogen evolution rates are lower although their photocurrents are higher than the others.
Electrochemical impedance spectroscopy (EIS) is also a powerful technology to probe the charge transfer process at an electrode interface [32,33]. The semicircle at high frequencies represents the capacitance and the resistance of the solid-state interface layer, which is formed at highly charged states and results from the passivation reaction between the electrolyte and the surface of the electrode [33]. As shown in Figure 5, the semicircle diameters of AgInS2-Ag2S-ZnS (Ag:Zn = 1:1) and AgInS2-Ag2S-ZnS-In6S7 (Ag:Zn = 1:2) are smaller than those of the other samples, The higher hydrogen evolution activities would be attributed to faster interfacial electron transfer. The moderate amounts of ZnS and In6S7 added into AgInS2-Ag2S benefit the charge transfer and lower the recombination probability of photogenerated charges. Consequently, the ZnS and In6S7 nanoparticles can work as electron acceptors and transporters in the obtained composite, which significantly enhances the photocatalytic activity for H2 production.
It is known that Ag2S is a narrow band gap semiconductor and may have an extended absorption range, thus it can be easily excited to generate h+-e pairs and produce intense photo-current [34,35]. However, these generated electrons are unstable to reduce H2O into H2 due to their lower conduction band potential than H+/H2 pair. As shown in Figure 6, AgInS2 has an appropriate band structure for water splitting to evolve H2 [36]. Since the conduction band level is negative enough for a reduction potential of H2O/H2, ZnS with a wide bandgap (3.73 eV) is an ideal photocatalyst from the viewpoint of preeminent ability to generate H2 [37]. Herein, photocatalyst of AgInS2-Ag2S-ZnS possesses the best photocatalytic activity among the as-obtained samples with various ratios of Ag:Zn and excessive quantity of ZnS is unfavorable for solar light absorption to produce H2. However, regarding the feasible point of modifying by In6S7, more experimental and theoretical evidence is needed to clarify this in the future.

3. Materials and Methods

3.1. Materials

Silver acetate (AgAc), zinc acetate (Zn(Ac)2), potassium iodide (KI), polyvinylpyrrolidone (PVP) and ethylenediamine (En) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Indium trichloride (InCl3∙4H2O), sodium diethyldithiocarbamate trihydrate ((C2H5)2NCSSNa∙3H2O, Na(DDTC)), polyethylene glycol (PEG, molecular weight = 10,000 g∙mol−1) were purchased from Shanghai SSS Reagent Co., Ltd. (Shanghai, China). All the reagents were analytical grade and used as received without further purification.

3.2. Synthesis of AgI Polyhedrons

In a typical procedure, 0.05 g of PVP and 0.06 g of AgAc were dissolved in a mixture of 10 mL of deionized water and 1 mL of En under magnetic stirring at 60 °C. After they were totally dissolved, 5 mL of KI aqueous solution (0.084 mol/L) was introduced at a rate of 1 mL/min to the above solution with a syringe pump and the resulting mixture was stirred continuously for 10 min. SEM image of the obtained AgI precursor is given in Supplementary Materials, Figure S5.

3.3. Synthesis of In(DDTC)3

Fifty millilitres of ethanol solution of 0.1 mol/L of InCl3∙4H2O and 0.3 mol/L of Na(DDTC) were mixed under vigorous magnetic stirring for ~30 min. The obtained white product was rinsed with de-ionized water 3 times and collected via vacuum filtration.

3.4. Synthesis of AgInS2-Ag2S-ZnS Heterojunction Nanostructures

The various nanocomposites were synthesized by a simple hydrothermal method through varying the ratios of starting materials. Briefly, the freshly synthesized AgI polyhedrons (0.12 g), 0.3 g of In(DDTC)3, 0.08 g of Zn(Ac)2 (molar ratio of Ag:Zn = 1:1) and 25 mL of deionized water were sequentially transferred into a Teflon-lined autoclave with the capacity of 45 mL. The autoclave was sealed tightly and heated at 180 °C for 24 h. The product was washed with de-ionized water 3 times and collected by centrifugation.
For comparison, other samples with different compositions were also synthesized by simply changing the addition quantity of Zn(Ac)2 to 0.16 g (molar ratio of Ag:Zn = 1:2), 0.4 g (molar ratio of Ag:Zn = 1:5), 0.04 g (molar ratio of Ag:Zn = 2:1) and 0.016 g (molar ratio of Ag:Zn = 5:1).

3.5. Characterization

The crystalline phases of the samples were analyzed by X-ray powder diffraction (XRD) (PANalytical B.V., Almelo, The Netherlands) on a Philips X’Pert diffractometer with Cu Kα radiation (λ = 0.15418 nm). The morphologies and sizes were observed with a Hitachi S-4800 field emission scanning electronic microscopy (FESEM) (Tokyo, Japan). Transmission electron micrographs (TEM) were obtained using a JEM-2100UHR transmission microscope (JEOL, Tokyo, Japan). The optical absorption spectra were measured on a UV-vis spectrometer (Shimadu UV2600, Kyoto, Japan) over a range of 200 to1000 nm.

3.6. Evaluation of Photo-Electrochemical Performance

For photo-electrochemical measurements, AgInS2-xAg2S-yZnS-zIn6S7 heterojunctions were fabricated as the film electrodes. Typically, a glass substrate (ITO, 15 Ω/m2) was ultrasonically treated in a mixture of H2O, isopropanol, and acetone for 30 min. Then 30 mg of the as-synthesized samples and PEG aqueous solution (150 mg/mL) were mixed homogeneously. The obtained paste was spread onto the conducting glass substrate with a blade. Adhesive tapes were used as spacers. Finally, the resulted films with ~1 cm2 active area were annealed at 30 °C for 30 min under N2 atmosphere in order to achieve intimate contact between particles.
Photoelectrical response and electrochemical impedance spectroscopy (EIS) were measured on an electrochemical workstation (CHI 660E, CH Instruments, Chenhua, Shanghai, China) with a three-electrode cell. A Pt plate and saturated calomel electrode were used as the counter and reference electrodes, respectively. The as-prepared samples were employed as the working electrode with an active area of 1 × 1 cm2. The photocurrents were measured with the same three-electrode system under irradiation of a 300 W Xe lamp in 0.5 mol/L Na2SO4 aqueous solution. EIS was carried out with an open circuit potential (0.228 eV) from 1000 KHz–0.01 Hz in 0.50 M Na2SO4 containing equimolar [Fe(CN)6]3− and [Fe(CN)6]4− (2.5/2.5 mM) solution.

3.7. Evaluation of Photocatalytic Performance

The photocatalytic activity of the as-synthesized AgInS2-xAg2S-yZnS-zIn6S7 nanocomposites was evaluated by the photocatalytic reduction of water to produce H2 under simulated sunlight irradiation. The H2 evolution test was performed on a LabSolar-III AG reaction cell (Beijing Perfect Light Company, Beijing, China). In a typical experiment, 20 mg of AgInS2-xAg2S-yZnS-zIn6S7 nanoparticles were dispersed in an 60 mL of aqueous solution containing 2.6 g of Na2SO3 (0.33 mol∙L−1), 3.6 g of Na2S (0.74 mol∙L−1) as sacrificial reagents and 5 mL of 1 mg/mL Pt(NH4)2Cl6 as promoter. The aqueous solution was irradiated by a 300 W Xe arc lamp (PLS-SEX300/300UV, Beijing Perfect Light Company) to produce H2. The yield of H2 was determined with an on-line gas chromatograph (GC7900, Techcomp, Shanghai, China), which equipped with a molecular sieve 5A column and a thermal conductivity detector. N2 was used as the carrier gas.

4. Conclusions

In summary, we have synthesized a class of efficient solar-driven nanophotocatalysts made of AgInS2-xAg2S-yZnS-zIn6S7 composite, where the compositions can be controlled by varying the feed ratio of Ag+ to Zn2+ ions. The solar hydrogen evolution tests show that the as-synthesized AgInS2-Ag2S-ZnS obtained with Ag+:Zn2+ of 1:1 exhibits the highest activity with a rate of 5.4 mmol h−1·g−1. Compared with the pristine AgInS2, AgInS2-Ag2S-ZnS shows a significantly enhanced photocatalytic H2 production performance by a factor of 14 and even higher than AgInS2-Ag2S-ZnS-In6S7 under the same conditions. The results demonstrate that AgInS2 nanoparticles integrated with appropriate proportion of Ag2S and ZnS can be used as a promising photocatalyst for H2 production.

Supplementary Materials

The following are available online at www.mdpi.com/1996-1944/9/5/329/s1. Figure S1: Diffuse reflectance spectra of the as-prepared composite photocatalyhst with different ratio of Ag:Zn. Figure S2: The H2 evolved from water under solar light over AgInS2-Ag2S-ZnS obtained with Ag:Zn = 1:1. Figure S3: Photocatalytic H2 evolution under solar-light irradiation with the same sample as in Figure S2. It is showing that the addition of (NH4)2PtCl6 improved the production rate of H2 though the original catalyst also shows the good activity towards H2 evolution. Figure S4: BET surface area and H2 production of AgInS2-xAg2S-yZnS-zIn6S7 with different molar ratios of Ag+ to Zn2+ ions. Figure S5: An SEM image of the AgI precursor.

Acknowledgments

The authors gratefully acknowledge the financial support by National Natural Science Foundation of China (51402362 and 21471160), Shandong Natural Science Foundation (ZR2014EMQ012), Huangdao Key Science and Technology Programme (Contract No. 2014-1-50), Qingdao Science and Technology Program for Youth (14-2-4-34-jch) and the Fundamental Research Funds for the Central Universities (16CX05016A and 16CX05014A).

Author Contributions

Zhaojie Wang organized the project as the head of the working group; Shutao Wang and Junxue Liu performed prepared and characterized the photocatalysts; Wen Jiang wrote the paper; Yan Zhou contributed analysis tool of electrochemical workstation; Changhua An and Jun Zhang analyzed the data and the enhancement mechanism.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dahl, S.; Chorkendorff, I. Solar-fuel generation: Towards practical implementation. Nat. Mater. 2012, 11, 100–101. [Google Scholar] [CrossRef] [PubMed]
  2. Matsumura, M.; Furukawa, S.; Saho, Y.; Tsubomura, H. Cadmium Sulfide photocatalyzed hydrogen production from aqueous solutions of sulfite: Effect of crystal structure and preparation method of the catalyst. J. Phys. Chem. 1985, 89, 1327–1329. [Google Scholar] [CrossRef]
  3. Reber, J.F.; Rusek, M. Photochemical hydrogen production with platinized suspensions of cadmium sulfide and cadmium zinc sulfide modified by silver sulfide. J. Phys. Chem. 1986, 90, 824–834. [Google Scholar] [CrossRef]
  4. Kershaw, S.V.; Susha, A.S.; Rogach, A.L. Narrow bandgap colloidal metal chalcogenide quantum dots: Synthetic methods, heterostructures, assemblies, electronic and infrared optical properties. Chem. Soc. Rev. 2013, 42, 3033–3087. [Google Scholar] [CrossRef] [PubMed]
  5. Tsuji, I.; Kato, H.; Kudo, A. Photocatalytic Hydrogen Evolution on ZnS-CuInS2-AgInS2 Solid Solution Photocatalysts with Wide Visible Light Absorption Bands. Chem. Mater. 2006, 18, 1969–1975. [Google Scholar] [CrossRef]
  6. Mauser, C.; Da Como, E.; Baldauf, J.; Rogach, A.L.; Huang, J.; Talapin, D.V.; Feldmann, J. Spatio-temporal dynamics of coupled electrons and holes in nanosize CdSe-CdS semiconductor tetrapods. Phys. Rev. B 2010, 82, 081306. [Google Scholar] [CrossRef]
  7. Milliron, D.J.; Hughes, S.M.; Cui, Y.; Manna, L.; Li, J.; Wang, L.-W.; Alivisatos, A.P. Colloidal nanocrystal heterostructures with linear and branched topology. Nature 2004, 430, 190–195. [Google Scholar] [CrossRef] [PubMed]
  8. Laursen, A.B.; Kegnaes, S.; Dahl, S.; Chorkendorff, I. Molybdenum sulfides-efficient and viable materials for electro—And photoelectrocatalytic hydrogen evolution. Energy Environ. Sci. 2012, 5, 5577–5591. [Google Scholar] [CrossRef]
  9. Weber, M.F.; Dignam, M.J. Splitting water with semiconducting photoelectrodes—Efficiency considerations. Int. J. Hydrogen Energy 1986, 11, 225–232. [Google Scholar] [CrossRef]
  10. Yang, L.-Y.; Dong, S.-Y.; Sun, J.-H.; Feng, J.-L.; Wu, Q.-H.; Sun, S.-P. Microwave-assisted preparation, characterization and photocatalytic properties of a dumbbell-shaped ZnO photocatalyst. J. Hazard. Mater. 2010, 179, 438–443. [Google Scholar] [CrossRef] [PubMed]
  11. Zhang, Z.; Shao, C.; Li, X.; Zhang, L.; Xue, H.; Wang, C.; Liu, Y. Electrospun Nanofibers of ZnO−SnO2Heterojunction with High Photocatalytic Activity. J. Phys. Chem. C 2010, 114, 7920–7925. [Google Scholar] [CrossRef]
  12. Han, Z.; Qiu, F.; Eisenberg, R.; Holland, P.L.; Krauss, T.D. Robust photogeneration of H2 in water using semiconductor nanocrystals and a nickel catalyst. Science 2012, 338, 1321–1324. [Google Scholar] [CrossRef] [PubMed]
  13. Seger, B.; Pedersen, T.; Laursen, A.B.; Vesborg, P.C.K.; Hansen, O.; Chorkendorff, I. Using TiO2 as a Conductive Protective Layer for Photocathodic H2 Evolution. J. Am. Chem. Soc. 2013, 135, 1057–1064. [Google Scholar] [CrossRef] [PubMed]
  14. Tsuji, I.; Kato, H.; Kobayashi, H.; Kudo, A. Photocatalytic H2 Evolution Reaction from Aqueous Solutions over Band Structure-Controlled (AgIn)xZn2(1-x)S2 Solid Solution Photocatalysts with Visible-Light Response and Their Surface Nanostructures. J. Am. Chem. Soc. 2004, 126, 13406–13413. [Google Scholar] [CrossRef] [PubMed]
  15. Tsuji, I.; Kato, H.; Kudo, A. Visible-Light-Induced H2 Evolution from an Aqueous Solution Containing Sulfide and Sulfite over a ZnS–CuInS2–AgInS2 Solid-Solution Photocatalyst. Angew. Chem. Int. Ed. 2005, 117, 3631–3634. [Google Scholar] [CrossRef]
  16. Iwashina, K.; Iwase, A.; Ng, Y.H.; Amal, R.; Kudo, A. Z-Schematic Water Splitting into H2 and O2 Using Metal Sulfide as a Hydrogen-Evolving Photocatalyst and Reduced Graphene Oxide as a Solid-State Electron Mediator. J. Am. Chem. Soc. 2015, 137, 604–607. [Google Scholar] [CrossRef] [PubMed]
  17. Tsuji, I.; Kato, H.; Kobayashi, H.; Kudo, A. Photocatalytic H2 Evolution under Visible-Light Irradiation over Band-Structure-Controlled (CuIn)xZn2(1-x)S2 Solid Solutions. J. Phys. Chem. B 2005, 109, 7323–7329. [Google Scholar] [CrossRef] [PubMed]
  18. Xie, R.; Rutherford, M.; Peng, X. Formation of High-Quality I−III−VI Semiconductor Nanocrystals by Tuning Relative Reactivity of Cationic Precursors. J. Am. Chem. Soc. 2009, 131, 5691–5697. [Google Scholar] [CrossRef] [PubMed]
  19. Simon, T.; Bouchonville, N.; Berr, M.J.; Vaneski, A.; Adrović, A.; Volbers, D.; Wyrwich, R.; Döblinger, M.; Susha, A.S.; Rogach, A.L. Redox shuttle mechanism enhances photocatalytic H2 generation on Ni-decorated CdS nanorods. Nat. Mater. 2014, 13, 1013–1018. [Google Scholar] [CrossRef] [PubMed]
  20. Ma, C.; Moore, D.; Li, J.; Wang, Z.L. Nanobelts, Nanocombs, and Nanowindmills of Wurtzite ZnS. Adv. Mater. 2003, 15, 228–231. [Google Scholar] [CrossRef]
  21. Zong, X.; Yan, H.; Wu, G.; Ma, G.; Wen, F.; 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] [PubMed]
  22. Zhang, W.; Wang, Y.; Wang, Z.; Zhong, Z.; Xu, R. Highly efficient and noble metal-free NiS/CdS photocatalysts for H2 evolution from lactic acid sacrificial solution under visible light. Chem. Commun. 2010, 46, 7631–7633. [Google Scholar] [CrossRef] [PubMed]
  23. Mocatta, D.; Cohen, G.; Schattner, J.; Millo, O.; Rabani, E.; Banin, U. Heavily doped semiconductor nanocrystal quantum dots. Science 2011, 332, 77–81. [Google Scholar] [CrossRef] [PubMed]
  24. Lo, S.S.; Mirkovic, T.; Chuang, C.H.; Burda, C.; Scholes, G.D. Emergent Properties Resulting from Type-II Band Alignment in Semiconductor Nanoheterostructures. Adv. Mater. 2011, 23, 180–197. [Google Scholar] [CrossRef] [PubMed]
  25. Mao, B.; Chuang, C.; Lu, F.; Sang, L.; Zhu, J.; Burda, C. Study of the Partial Ag-to-Zn Cation Exchange in AgInS2/ZnS Nanocrystals. J. Phys. Chem. C 2013, 117, 648–656. [Google Scholar] [CrossRef]
  26. Michalska, M.; Aboulaich, A.; Medjahdi, G.; Mahiou, R.; Jurga, S.; Schneider, R. Amine ligands control of the optical properties and the shape of thermally grown core/shell CuInS2/ZnS quantum dots. J. Alloy. Compd. 2015, 645, 184–192. [Google Scholar] [CrossRef]
  27. Manna, L.; Scher, E.C.; Li, L.S.; Alivisatos, A.P. Epitaxial growth and photochemical annealing of graded CdS/ZnS shells on colloidal CdSe nanorods. J. Am. Chem. Soc. 2002, 124, 7136–7145. [Google Scholar] [CrossRef] [PubMed]
  28. Ivanov, S.A.; Piryatinski, A.; Nanda, J.; Tretiak, S.; Zavadil, K.R.; Wallace, W.O.; Werder, D.; Klimov, V.I. Type-II Core/Shell CdS/ZnSe Nanocrystals:  Synthesis, Electronic Structures, and Spectroscopic Properties. J. Am. Chem. Soc. 2007, 129, 11708–11719. [Google Scholar] [CrossRef] [PubMed]
  29. Jiang, L.C.; Zhang, W.-D. Charge transfer properties and photoelectrocatalytic activity of TiO2/MWCNT hybrid. Electrochim. Acta 2010, 56, 406–411. [Google Scholar] [CrossRef]
  30. Xiang, Q.; Yu, J.; Jaroniec, M. Preparation and Enhanced Visible-Light Photocatalytic H2-Production Activity of Graphene/C3N4 Composites. J. Phys. Chem. C 2011, 115, 7355–7363. [Google Scholar] [CrossRef]
  31. Bell, N.J.; Ng, Y.H.; Du, A.; Coster, H.; Smith, S.C.; Amal, R. Understanding the Enhancement in Photoelectrochemical Properties of Photocatalytically Prepared TiO2-Reduced Graphene Oxide Composite. J. Phys. Chem. C 2011, 115, 6004–6009. [Google Scholar] [CrossRef]
  32. Zhang, H.; Lv, X.; Li, Y.; Wang, Y.; Li, J. P25-Graphene Composite as a High Performance Photocatalyst. ACS Nano 2010, 4, 380–386. [Google Scholar] [CrossRef] [PubMed]
  33. He, B.-L.; Dong, B.; Li, H.-L. Preparation and electrochemical properties of Ag-modified TiO2 nanotube anode material for lithium–ion battery. Electrochem. Commun. 2007, 9, 425–430. [Google Scholar] [CrossRef]
  34. Chen, C.; Xie, Y.; Ali, G.; Yoo, S.H.; Cho, S.O. Improved conversion efficiency of Ag2S quantum dot-sensitized solar cells based on TiO2 nanotubes with a ZnO recombination barrier layer. Nanoscale Res. Lett. 2011, 6, 1–9. [Google Scholar] [CrossRef] [PubMed]
  35. Abdallah, H.B.; Bennaceur, R. Electronic structure of the hexaindium heptasulfide In6S7. Physica B: Condens. Matter 2006, 382, 181–188. [Google Scholar] [CrossRef]
  36. Peng, S.; Zhang, S.; Mhaisalkar, S.G.; Ramakrishna, S. Synthesis of AgInS2 nanocrystal ink and its photoelectrical application. Phys. Chem. Chem. Phys. 2012, 14, 8523–8529. [Google Scholar] [CrossRef] [PubMed]
  37. Kuzuya, T.; Tai, Y.; Yamamuro, S.; Hihara, T.; Peng, D.L.; Sumiyama, K. Synthesis of Zinc Sulfide Nanocrystals and Fabrication of Nanocrystal Superlattice. Mater. Trans. 2004, 45, 2650–2652. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of the as-synthesized AgInS2-xAg2S-yZnS-zIn6S7 nanojunctions with different molar ratios of Ag to Zn. AgInS2: JCPDS 25-1328, Ag2S JCPDS 14-0072, ZnS: JCPDS 36-1450, In6S7: JCPDS 19-0587.
Figure 1. X-ray diffraction patterns of the as-synthesized AgInS2-xAg2S-yZnS-zIn6S7 nanojunctions with different molar ratios of Ag to Zn. AgInS2: JCPDS 25-1328, Ag2S JCPDS 14-0072, ZnS: JCPDS 36-1450, In6S7: JCPDS 19-0587.
Materials 09 00329 g001
Figure 2. SEM images of AgInS2-xAg2S-yZnS-zIn6S7 composite nanoparticles with different molar ratios of Ag:Zn (A) 5:1; (B) 2:1; (C) 1:1; (D) 1:2; (E) 1:5 (The scale bars are all 400 nm); and a corresponding comparison chart of the photocatalytic H2 production in a reaction period of 12 h (F).
Figure 2. SEM images of AgInS2-xAg2S-yZnS-zIn6S7 composite nanoparticles with different molar ratios of Ag:Zn (A) 5:1; (B) 2:1; (C) 1:1; (D) 1:2; (E) 1:5 (The scale bars are all 400 nm); and a corresponding comparison chart of the photocatalytic H2 production in a reaction period of 12 h (F).
Materials 09 00329 g002
Figure 3. Transmission electron micrograph (TEM) (A) and HRTEM (B) images of the AgInS2-Ag2S-ZnS composite nanoparticles.
Figure 3. Transmission electron micrograph (TEM) (A) and HRTEM (B) images of the AgInS2-Ag2S-ZnS composite nanoparticles.
Materials 09 00329 g003
Figure 4. Photocurrent-density responses (I-t) of the samples synthesized with different ratios of Ag:Zn in 0.5 M Na2SO3 solution under simulated sunlight illumination.
Figure 4. Photocurrent-density responses (I-t) of the samples synthesized with different ratios of Ag:Zn in 0.5 M Na2SO3 solution under simulated sunlight illumination.
Materials 09 00329 g004
Figure 5. Nyquist plots of electrochemical impedance spectra (EIS) of the as-prepared samples (AgInS2-xAg2S-yZnS-zIn6S7). The EIS measurements were performed in the presence of a 2.5 mM K3[Fe(CN)6]/K4[Fe(CN)6] (1:1) mixture as a red-ox probe in 0.5 M Na2SO4 aqueous solution.
Figure 5. Nyquist plots of electrochemical impedance spectra (EIS) of the as-prepared samples (AgInS2-xAg2S-yZnS-zIn6S7). The EIS measurements were performed in the presence of a 2.5 mM K3[Fe(CN)6]/K4[Fe(CN)6] (1:1) mixture as a red-ox probe in 0.5 M Na2SO4 aqueous solution.
Materials 09 00329 g005
Figure 6. A diagrammatic scheme of energetic band structure of AgInS2-Ag2S-ZnS (Ag:Zn = 1:1).
Figure 6. A diagrammatic scheme of energetic band structure of AgInS2-Ag2S-ZnS (Ag:Zn = 1:1).
Materials 09 00329 g006

Share and Cite

MDPI and ACS Style

Wang, Z.; Wang, S.; Liu, J.; Jiang, W.; Zhou, Y.; An, C.; Zhang, J. Synthesis of AgInS2-xAg2S-yZnS-zIn6S7 (x, y, z = 0, or 1) Nanocomposites with Composition-Dependent Activity towards Solar Hydrogen Evolution. Materials 2016, 9, 329. https://doi.org/10.3390/ma9050329

AMA Style

Wang Z, Wang S, Liu J, Jiang W, Zhou Y, An C, Zhang J. Synthesis of AgInS2-xAg2S-yZnS-zIn6S7 (x, y, z = 0, or 1) Nanocomposites with Composition-Dependent Activity towards Solar Hydrogen Evolution. Materials. 2016; 9(5):329. https://doi.org/10.3390/ma9050329

Chicago/Turabian Style

Wang, Zhaojie, Shutao Wang, Junxue Liu, Wen Jiang, Yan Zhou, Changhua An, and Jun Zhang. 2016. "Synthesis of AgInS2-xAg2S-yZnS-zIn6S7 (x, y, z = 0, or 1) Nanocomposites with Composition-Dependent Activity towards Solar Hydrogen Evolution" Materials 9, no. 5: 329. https://doi.org/10.3390/ma9050329

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