DOI:
10.1039/C4RA11740A
(Paper)
RSC Adv., 2014,
4, 62223-62229
ZnS microsphere/g-C3N4 nanocomposite photo-catalyst with greatly enhanced visible light performance for hydrogen evolution: synthesis and synergistic mechanism study†
Received
3rd October 2014
, Accepted 12th November 2014
First published on 12th November 2014
Abstract
In this work, ZnS microsphere/g-C3N4 nanocomposites with different ZnS contents were synthesized by a facile precipitation route and characterized by a variety of techniques. The TEM results confirmed that ZnS microspheres are deposited on the surface of g-C3N4 nanosheets with intimate contact in the nanocomposites. Such type-II heterostructures formed between ZnS and g-C3N4 can unexpectedly improve the photocatalytic H2-production rate of g-C3N4 under visible light, which reaches an optimal value of up to 194 μmol h−1 g−1 at the ZnS content of 50 wt%. The synergistic effect of ZnS and g-C3N4 was proposed to be responsible for the efficient separation of the photogenerated charge carriers and, consequently, the enhancement of the visible light photocatalytic H2 production activity of the nanocomposites. This work highlights that the integration of g-C3N4 and ZnS will be an opportunity to obtain a g-C3N4-based nanocomposite photocatalyst with enhanced photocatalytic H2 production activity, and we hope our work may be helpful for the construction of photocatalysts with efficient visible light H2-production activities.
1. Introduction
Since the discovery by Honda and Fujishima of the photoelectron-chemical splitting of water over a TiO2 electrode,1 photocatalytic hydrogen production using semiconductor particles has been regarded as one of the most promising means of clean energy production from renewable sources.2,3 In the past decades, various semiconductor photocatalysts have been developed for the production of H2, including oxides, sulfides, oxynitrides and nanocomposites.4–10 Graphite-like carbon nitride (g-C3N4), a promising metal-free semiconductor photocatalyst, has attracted increasing attention owing to its visible-light response, suitable band position and stable, low toxicity, and low cost characteristics.7,11 However, the photocatalytic efficiency of pristine g-C3N4 is predominately constrained by the high recombination rate of its photogenerated electron–hole pairs. To restrain such undesirable recombination and achieve high activity, pristine g-C3N4 was often coupled with a variety of metal oxides or metal sulfides to construct binary or ternary hybrid photocatalysts with specially designed type-II heterostructures.12–32 These formed type-II heterostructures could accelerate the spatially separation of photogenerated charge carriers, and consequentially increase the photocatalytic activity of g-C3N4. Notably, in some of these photocatalysts, g-C3N4 performs as an effective visible-light sensitizer for the coupled wide-gap semiconductor (such as TiO2, ZnO and SrTiO3), whilst the combined wide band gap semiconductor could help to separate the photogenerated charge carriers. Utilization of this “synergistic effect” has been proven to feasible way to access enhanced photocatalytic H2 production activity for g-C3N4, but so far just very few examples were reported. Thus, combination of the g-C3N4 and wide-gap semiconductor to form hybrid photocatalyst for H2 production is highly desirable and challenging.
From the thermodynamic point of view, in the g-C3N4-based type-II heterostructures for the photocatalytic H2 production, the bottom level of the CB of the coupled semiconductor is generally more positive than that of g-C3N4, while at the same time this CB position should be more negative as much as possible than the redox potential of H+/H2 (0 V vs. NHE).33,34 When searching this semiconductor, we found that ZnS is in principle suitable. ZnS, with a bandgap of around 3.5 eV, is one of the most important metal sulfide materials in photocatalysis, due to the high CB position (−0.9 eV), good thermal stability, high electronic mobility and low toxicity.35–37 However, as a single component, the high recombination ratios of photoinduced electron–hole pairs and poor response to visible light greatly limiting its practical applications.38–41 It is expected that unite of g-C3N4 and ZnS will be an opportunity to obtain a novel g-C3N4-based composite photocatalyst with improved photocatalytic H2 production activity.
Here, we report the synthesis of ZnS microsphere/g-C3N4 (denoted as ZnS/g-C3N4) nanocomposite photocatalysts by a facile precipitation route. The ZnS/g-C3N4 nanocomposite exhibits distinctly enhanced H2 production efficiency under visible light irradiation as compared to pure g-C3N4 and ZnS. Based on the experimental results, a possible synergistic mechanism for the enhanced photocatalytic H2 evolution activity was proposed. This work for the first time shows a feasible example to enhance the photocatalytic H2 production activity of g-C3N4 by constructing type-II heterostructures with wide band gap ZnS. Moreover, the use of g-C3N4 as a wide band gap semiconductor sensitiser candidate in the construction of visible-light photocatalytic H2 production system was confirmed.
2. Experimental
2.1 Preparation of photocatalysts
All reagents were of analytical grade, purchased from Sinopharm Chemical Reagent Co., Ltd., China, and were used without further purification. The formation of ZnS microsphere/g-C3N4 composite involves a two-step process, and the schematic illustration of the formation process is described in Fig. 1. g-C3N4 was first prepared by thermal treatment of 10 g of urea in a crucible with a cover under ambient pressure in air, similar to reported methods. After dried at 80 °C for 24 h, the precursor was heated to 550 °C at a heating rate of 2.3 °C min−1 in a tube furnace for 4 h in air. The resulted final light yellow powder were washed with nitric acid (0.1 mol L−1), distilled water and absolute ethanol for three times, then dried at 60 °C for 12 h.
 |
| Fig. 1 Schematic illustration of the formation process of ZnS/g-C3N4 nanocomposites. | |
ZnS/g-C3N4 composite photocatalysts were synthesized by a facile precipitation method. Typically, 400 mL of a mixed solution of different amounts of g-C3N4, Zn(NO3)2 (0.05 M) and HNO3 (0.01 M) was heated to 70 °C; then an appropriate amount of thioacetamide (C2H5NS, TAA) was added. After that, the solution was immersed in a 70 °C water bath for 5 h. Then the reaction was quickly transferred to an ice water bath for quenching to below 5 °C. The product was collected by centrifugation, washed with distilled water and absolute ethanol, and dried in an oven at 60 °C in air. According to this method, different weight ratios of the ZnS to g-C3N4 samples were obtained and labeled as 20% ZnS/g-C3N4, 30% ZnS/g-C3N4, 40% ZnS/g-C3N4, 50% ZnS/g-C3N4, 60% ZnS/g-C3N4, and 70% ZnS/g-C3N4, respectively. Pure ZnS was also synthesized for comparison purpose in the absence of g-C3N4. In order to compare the photocatalytic activity with ZnS/g-C3N4 composites, a mechanically (physically) mixed ZnS/g-C3N4 sample with ZnS to g-C3N4 weight ratio of 50% (donated to m-50% ZnS/g-C3N4) was prepared. A measured amount of as-prepared bulk g-C3N4 and ZnS added in 50 mL methanol. After ultrasonically treated for 30 min, the homogeneous suspension stirred at room temperature until the residual methanol was removed. Then the obtained solid were collected and dried in a vacuum oven at 100 °C for 24 h. To improve the photocatalytic H2 evolution activity, 1 wt% Pt-dispersed photocatalysts were prepared by an incipient impregnation method using chloroplatinic acid as metal precursor.
2.2 Characterization
The phase purity and crystal structure of the obtained samples were examined by X-ray diffraction (XRD, Bruker D8 Advance X-ray diffractometer) with Cu-Kα radiation (λ = 1.5406 Å). Infrared spectra were obtained on KBr pellets on a Nicolet NEXUS470 FTIR in the range of 4000–500 cm−1. Field emission scanning electron microscopy (FESEM) measurement was performed by JSM-6700F instrument operated at an accelerating voltage of 10 kV. Transmission electron microscopy (TEM) was recorded on a JEOL-JEM-2010 (JEOL, Japan) operating at 200 kV. Surface analysis of the sample was examined by X-ray photoelectron spectroscopy (XPS) using a ESCA PHI500 spectrometer. The surface areas of the samples were measured by a TriStar II 3020-BET/BJH Surface Area. UV-vis diffuse reflectance spectra (DRS) were performed on a Shimadzu UV-2401 spectrophotometer equipped with spherical diffuse reflectance accessory. The photoluminescence (PL) spectra of the photocatalyst were obtained by a Varian Cary Eclipse spectrometer with an excitation wavelength of 370 nm. The electrochemical impedance spectroscopy (EIS) measurements were carried out in a conventional three-electrode, single-compartment quartz cell on an electrochemical station (CHI 660D).
2.3 Photocatalytic hydrogen production
The photocatalytic hydrogen production experiments were carried out in a 100 mL Pyrex flask connected to a closed gas circulation and evacuation system. Four low power UV-LEDs (3 W, 420 nm) (Shenzhen LAMPLIC Science Co. Ltd. China), used as light sources to trigger the photocatalytic reaction, were positioned 1 cm away from the reactor in four different directions. The focused intensity and areas on the flask for each UV-LED was ca. 80.0 mW cm−2 and 1 cm2, respectively. Typically, 50 mg of photocatalyst was dispersed by magnetic stirrer in 80 mL aqueous solution containing 20 mL methanol. Prior to irradiation, the suspension of the catalyst was dispersed by an ultrasonic bath for 10 min, and then purged with nitrogen for 40 min to completely remove the dissolved oxygen and ensure the reactor was in an anaerobic condition. A 0.4 mL gas was intermittently sampled through the septum, and the amount of H2 was determined by gas chromatography (GC-14C, Shimadzu, Japan, TCD, with nitrogen as a carrier gas and 5 Å molecular sieve column).
3. Results and discussion
3.1 Structure and morphology
The XRD patterns of the pure g-C3N4, ZnS, and ZnS/g-C3N4 composites are shown in Fig. 2. The distinct diffraction peaks at 13.0° and 27.4° in the g-C3N4 sample, corresponding to the in-plane structural packing motif and interlayer stacking of aromatic segments, can be indexed as the (100) and (002) peaks for graphitic materials, respectively. For pure ZnS sample, all the peaks can be well indexed to the cubic phase (JCPDS 65-0309). The main diffraction peaks at 28.7°, 33.3°, 47.7°, 56.5°, 69.7°, and 77.0° correspond to the (111), (200), (220), (311), (400) and (331) crystal planes, respectively. All the ZnS/g-C3N4 samples exhibit diffraction peaks corresponding to both g-C3N4 and ZnS, while the characteristic peaks of g-C3N4 (27.4°) and ZnS (28.7°) were very close and overlap with each other. There are no other impure peaks can be observed, reflecting these composites consist of two phases.
 |
| Fig. 2 XRD patterns of (a) g-C3N4, (b) 20% ZnS/g-C3N4, (c) 30% ZnS/g-C3N4, (d) 40% ZnS/g-C3N4, (e) 50% ZnS/g-C3N4, (f) 60% ZnS/g-C3N4, (g) 70% ZnS/g-C3N4, and (h) ZnS. | |
The structure and morphology of the samples were characterized by SEM and TEM. The pure g-C3N4 sample, synthesized by the polymerization reaction of urea, exhibits a layered and sheet-like surface morphology (Fig. S1a and b†), while the pure ZnS sample is composed of a large number of well-defined hierarchical microspheres with diameters ranging from 200 to 600 nm (Fig. S1c and d†). Taking the 50% ZnS/g-C3N4 nanocomposite sample as an example, we examine the morphology of the resulting nanocomposite photocatalysts. As shown in Fig. 3a and b, hierarchical ZnS microspheres are well loaded on or attached with the g-C3N4 nanosheets. From the magnified TEM image shown in Fig. 3c, we can further found that ZnS microspheres are closely deposited on the surface of g-C3N4 nanosheets, indicating the intimate interfacial contact between the ZnS microspheres and g-C3N4 nanosheets is readily obtained by such a simple precipitation approach. This intimate interfacial contact can be further confirmed by the high-resolution TEM image (Fig. 3d). The lattice fringes can be clearly observed in the inset of Fig. 3d, suggesting the well-defined crystal structure of ZnS; and the lattice spacing of ca. 0.31 nm can be assigned to the (111) plane of ZnS. It is important to note that the intimate contact (marked by the red dashed line in Fig. 3d) of ZnS and g-C3N4 could favor the transfer of charge carriers in the composite, therefore improving the charge separation and photocatalytic efficiency. Similarly, such intimate contact can be also found in the other composite samples, which are composed of hierarchical ZnS microspheres and sheet-like g-C3N4, but just with the clear differences in the ZnS content (Fig. S2†).
 |
| Fig. 3 (a and b) SEM, (c) TEM and (d) HRTEM images of the 50% ZnS/g-C3N4 sample. Inset in (d) shows the enlarged HRTEM image. | |
To further confirm the existence of g-C3N4 in the composite sample, FT-IR spectroscopy is performed. It could also be clearly seen that the main characteristic peaks of g-C3N4 appeared in ZnS/g-C3N4 nanocomposite photocatalysts. As shown in Fig. 4a–g, the peaks at 1637 cm−1 are attributable to C
N stretching vibration modes, while the 1252, 1323, 1407 and 1568 cm−1 can be ascribed to aromatic C–N stretching vibration modes.42 The peak at 810 cm−1 is related to characteristic breathing mode of s-triazine units.43 The broad band centered at 3187 cm−1 can be assigned to the stretching mode of the N–H bond.44 In the composites, the characteristic peaks of g-C3N4 did not change after combination with ZnS. With regard to the pure ZnS synthesized in the presence of TAA, the peaks at 1383 and 1615 cm−1 could be attributed to the TAA on the surface of ZnS.
 |
| Fig. 4 FTIR spectra of (a) pure g-C3N4, (b) 20% ZnS/g-C3N4, (c) 30% ZnS/g-C3N4, (d) 40% ZnS/g-C3N4, (e) 50% ZnS/g-C3N4, (f) 60% ZnS/g-C3N4, (g) 70% ZnS/g-C3N4, (h) pure ZnS. | |
XPS analyses were carried out on a typical 50% ZnS/g-C3N4 sample to investigate the chemical valence state of the elements. The high resolution XPS spectra of C 1s shows that there were two peaks located at 284.7 and 287.3 eV (Fig. 5a). The peak centered at 284.7 eV can be ascribed to graphitic or surface adventitious carbon which was usually observed on the XPS characterization of g-C3N4.45 The peak centered at 287.3 eV can be assigned to sp2-hybridized carbon in the aromatic ring. In the N 1s spectrum (Fig. 5b), several binding energies can be separated. The main signal showed occurrence of C–N–C groups (398.0 eV), tertiary nitrogen N–(C)3 groups (399.8 eV) and N–H groups (401.0 eV).46 The peak at 404.2 eV could be related to the charging effects.47 The peaks of S 2p3/2 and S 2p1/2, which were located at around 161.5 and 162.7 eV (Fig. 5c) were assigned to sulfur anions in the lattice of ZnS. Fig. 5d shows the peaks at 1021.8 and 1044.9 eV, which could be attributed to Zn 2p3/2 and Zn 2p1/2 binding energies of Zn2+ in ZnS/g-C3N4 nanocomposite. The XPS spectra of pure g-C3N4 and ZnS were shown in Fig. S3.† No obvious shift was observed in the Zn 2p and S 1s peak after the combination of ZnS, which indicates that there is no strong interfacial interaction between the ZnS and g-C3N4. Evidently, results from XRD, FT-IR and XPS revealed that the obtained composites contained g-C3N4 and ZnS.
 |
| Fig. 5 XPS spectra of (a) C 1s, (b) N 1s, (c) S 2p and (d) Zn 2p of the sample 50% ZnS/g-C3N4. | |
3.2 Optical absorption properties
The light absorption properties of the pristine ZnS, g-C3N4, and ZnS/g-C3N4 composites were characterized by UV-vis diffuse reflectance spectroscopy. The band gap energy of pristine ZnS and g-C3N4 can be calculated by the following formula: αhν = A(hν − Eg)n/2, where α, h, ν, Eg, and A are the absorption coefficient, Planck constant, the light frequency, the band gap, and a constant, respectively.48 Among them, n depends on the characteristics of the transition in a semiconductor (direct transition n = 1 and indirect transition n = 4). As shown in Fig. 6a, the pristine ZnS absorbs only ultraviolet light with wavelength shorter than around 365 nm, which can be assigned to a band gap of around 3.4 eV (Fig. 6b). The pristine g-C3N4 shows an absorption edge at about 455 nm (Fig. 6a), corresponding to band gap of 2.7 eV (Fig. 6b). After being coupled with g-C3N4, the optical absorption of the ZnS/g-C3N4 composites is enhanced distinctly in the visible light region, which allows for more efficient utilization of the solar spectrum to create photogenerated electrons and holes. The adsorption intensities increase gradually with the enhancement of the amount of g-C3N4 in the composites.
 |
| Fig. 6 (a) UV-vis diffuse reflectance spectra of pristine ZnS, g-C3N4, and the ZnS/g-C3N4 composites, (b) the band gaps (Eg) of pristine ZnS and g-C3N4. | |
3.3 Photocatalytic H2-evolution activity and photostability
The H2 evolution experiments were performed by taking 50 mL of a 25 vol% methanol solution and 0.05 g of powdered photocatalyst. Before the actual photocatalytic experiment was performed, control experiment was carried out by taking a pure methanol solution in the absence of either photocatalyst or irradiation. As there was no appreciable H2 production was detected, it was confirmed that H2 was produced by photocatalytic reaction. The influence of different ZnS contents in the ZnS/g-C3N4 composites on the H2 evolution is presented in Fig. 7a. It can be found that both the pure g-C3N4 and ZnS show a negligible visible light H2 evolution activity. As shown in Fig. 7a, the rate of H2 evolution increases with increasing of ZnS amount in the composites, and can reach to about 194 μmol h−1 g−1 when the mass ratio of ZnS to g-C3N4 is increased to 0.5. With the further increasing of the content of ZnS, the photocatalytic activity of the nanocomposite samples decreases. But they still show enhanced photocatalytic activities than that of pure g-C3N4 and ZnS. As a control, we also test the H2-production activity of the m-50% ZnS/g-C3N4 synthesized by the physically mixed method. Notably, we found that the photocatalytic H2 efficiency of the m-50% ZnS/g-C3N4 is significantly lower than that of ZnS/g-C3N4 sample, which prepared by the precipitation method. This result conforms that the formation of intact interface structure between the ZnS and g-C3N4 is the crucial parameter to the high photocatalytic activity of the resulting nanocomposite samples. The stability and reusability of 50% ZnS/g-C3N4 nanocomposite were evaluated by the cycling H2 evolution experiment, and the results are shown in Fig. 7b. The results show that the 50% ZnS/g-C3N4 nanocomposite photocatalysts do not display obvious decrease of photocatalytic H2 production activity under visible light. Only insignificant loss in photocatalytic activity is observed, which might be partly caused by incomplete collection of the photocatalyst during each step. This indicates that the ZnS/g-C3N4 nanocomposite photocatalysts is sufficiently stable for photocatalytic H2 production.
 |
| Fig. 7 (a) Photocatalytic activities of Pt-dispersed pure g-C3N4 (A), 20% ZnS/g-C3N4 (B), 30% ZnS/g-C3N4 (C), 40% ZnS/g-C3N4 (D), 50% ZnS/g-C3N4 (E), 60% ZnS/g-C3N4 (F), 70% ZnS/g-C3N4 (G), pure ZnS (H), and m-50% ZnS/g-C3N4 (I) for H2 production under visible-light irradiation. (b) Cycle runs for the photocatalytic H2 production over 50% ZnS/g-C3N4. | |
3.4 Possible mechanism of enhanced photocatalytic activity
It is generally accepted that photocatalytic activity is mainly governed by light-absorption ability, surface properties (including adsorption property, specific surface area, etc.), and photogenerated charge-separation efficiency. The H2-production activity of the 50% ZnS/g-C3N4 was greatly enhanced relative to the pure ZnS and g-C3N4. The BET surface area of the pure g-C3N4, ZnS and 50% ZnS/g-C3N4 is 32.35, 3.30 and 13.40 m2 g−1 (Fig. S4†), indicating that the BET surface area was greatly reduced by the introduction of ZnS. Therefore, the enhanced photocatalytic activity of the samples has no direct relation with the BET surface areas. As discussed above, the ZnS loading did not significantly change the light-absorption property of g-C3N4, indicating that the light-adsorption property factor is also not crucial to the photocatalytic activity for ZnS/g-C3N4 nanocomposites. Therefore, the enhanced photocatalytic activities of ZnS/g-C3N4 nanocomposites might be mainly attributed to the synergic effect between ZnS and g-C3N4, which can greatly accelerate the separation of photogenerated carriers. The mechanism for the superior photocatalytic hydrogen evolution over the as-prepared ZnS/g-C3N4 nanocomposite under visible light irradiation was proposed in Fig. 8. Under visible light irradiation, the VB electrons of g-C3N4 are excited to the CB, creating holes in the VB. Normally, these photogenerated electrons and holes are easy to recombine, resulting in a low photocatalytic activity of g-C3N4 itself. However, in the ZnS/g-C3N4 system, due to the intimate contact of ZnS and g-C3N4, the CB electrons of g-C3N4 can be injected into the ZnS because the CB level of g-C3N4 is higher than that of ZnS. Therefore, the recombination process of the electron–hole is inhibited, leading to an enhanced photoactivity of the ZnS/g-C3N4 nanocomposites.
 |
| Fig. 8 Schematic diagram of electron–hole pairs separation and the possible reaction mechanism over ZnS/g-C3N4 nanocomposite photocatalyst under visible light irradiation. | |
To investigate the migration, transfer and recombination processes of photogenerated electron–hole pairs, PL spectra of the samples were recorded. Fig. 9a shows the photoluminescence spectra of the as-prepared samples activated at room temperatures and excitation wavelength λex = 370 nm. The main emission peak centers at about 450 nm for the pure g-C3N4 sample, which can be ascribed to the band gap recombination of electron–hole pairs. Obviously, after the formation of ZnS/g-C3N4 composite, the PL intensity of the composite decreased rapidly, demonstrating the efficient charge separation of the composite. Therefore, it is reasonable to conclude that the enhanced photocatalytic activity could be attributed to the effective interfacial charge transfer between ZnS and g-C3N4. For the comparison purpose, we also examine the PL spectra of the m-50% ZnS/g-C3N4 sample. As expected, we found that the PL intensity of the m-50% ZnS/g-C3N4 sample is higher than that of 50% ZnS/g-C3N4 sample. This result indicates that the formation of the intact interface structure between the ZnS and CN is beneficial to the interfacial charge transfer. To further study the charge transfer properties, we conducted electrochemical impedance spectroscopy (EIS) measurements over the pristine g-C3N4, ZnS and 50% ZnS/g-C3N4 photocatalysts, and the result was shown in Fig. 9b. In the Nyquist plot, the radius of each arc is characteristic of the charge-transfer process at the corresponding electrode/electrolyte interface with a smaller radius corresponding with a lower charge-transfer resistance. Notable, the 50% ZnS/g-C3N4 exhibits the smallest charge transfer resistance among all tested electrodes under irradiation, suggesting that effective shuttling of charges between the electrode and the electrolyte, and faster interfacial charge transfer occurred at the nanocomposite interface owing to the formation of the heterojunctions between g-C3N4 and ZnS.
 |
| Fig. 9 (a) Photoluminescence (PL) spectra of the as-prepared samples at room temperature. Excitation wavelength λex = 370 nm; (b) Nyquist impedance plots of pristine g-C3N4, ZnS and 50% ZnS/g-C3N4 photocatalysts under visible-light irradiation in 0.2 M Na2SO4 aqueous solution (pH 6.8). | |
4. Conclusions
In summary, this work shows a simple and feasible method to enhance the photocatalytic H2-production activity of g-C3N4 by the combination of wide band gap semiconductor. ZnS microsphere/g-C3N4 composite photocatalyst were successfully prepared via a simple precipitation method. All the prepared ZnS microsphere/g-C3N4 nanocomposites exhibit higher photocatalytic activity for H2 evolution from ethanol aqueous solution than ZnS or g-C3N4. The content of ZnS in the composite influences the photocatalytic performance, and the optimum proportion of ZnS is 50 wt%. Importantly, it is believed that the positive synergetic effect between ZnS microsphere and g-C3N4 can greatly suppress the charge recombination, thus markedly improving the photocatalytic H2-production activity. On the basis of the results of this study, the ZnS/g-C3N4 nanocomposite sample is expected to be an effective visible-light H2-production photocatalyst for practical applications.
Acknowledgements
This work was supported by the financial supports of National Nature Science Foundation of China (no. 21406091), Natural Science Foundation of Jiangsu Province (BK20140530), and College Natural Science Research Program of Jiangsu Province (13KJB610003).
Notes and references
- A. Fujishima and K. Honda, Nature, 1972, 238, 37 CrossRef CAS.
- X. B. Chen, S. H. Shen, L. J. Guo and S. S. Mao, Chem. Rev., 2010, 110, 6503 CrossRef CAS PubMed.
- H. Tong, S. X. Ouyang, Y. P. Bi, N. Umezawa, M. Oshikiri and J. H. Ye, Adv. Mater., 2012, 24, 229 CrossRef CAS PubMed.
- M. Cargnello and B. T. Diroll, Nanoscale, 2014, 6, 97 RSC.
- Z. G. Zou and H. Arakawa, J. Photochem. Photobiol., A, 2003, 158, 145 CrossRef CAS.
- Y. X. Li, G. Chen, Q. Wang, X. Wang, A. K. Zhou and Z. Y. Shen, Adv. Funct. Mater., 2010, 20, 3390 CrossRef CAS.
- X. C. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domen and M. Antonietti, Nat. Mater., 2009, 8, 76 CrossRef CAS PubMed.
- R. M. Navarro Yerga, M. C. Alvarez Galvan, F. Del Valle, J. Villoria De La Mano and J. L. G. Fierro, ChemSusChem, 2009, 2, 471 CrossRef CAS PubMed.
- Q. J. Xiang, J. G. Yu and M. Jaroniec, J. Am. Chem. Soc., 2012, 134, 6575 CrossRef CAS PubMed.
- C. S. Xing, Y. Zhang, Z. D. Wu, D. L. Jiang and M. Chen, Dalton Trans., 2014, 43, 2772 RSC.
- X. C. Wang, S. Blechert and M. Antonietti, ACS Catal., 2012, 2, 1596 CrossRef CAS.
- S. W. Cao and J. G. Yu, J. Phys. Chem. Lett., 2014, 5, 2101 CrossRef CAS.
- D. M. Chen, K. W. Wang, D. G. Xiang, R. L. Zong, W. Q. Yao and Y. F. Zhu, Appl. Catal., B, 2014, 147, 554 CrossRef CAS PubMed.
- X. X. Xu, G. Liu, C. Randorn and J. T. S. Irvine, Int. J. Hydrogen Energy, 2011, 36, 13501 CrossRef CAS PubMed.
- J. X. Wang, J. Huang, H. L. Xie and A. L. Qu, Int. J. Hydrogen Energy, 2014, 39, 6354 CrossRef CAS PubMed.
- J. X. Sun, Y. P. Yuan, L. G. Qiu, X. Jiang, A. J. Xie, Y. H. Shen and J. F. Zhu, Dalton Trans., 2012, 41, 6756 RSC.
- L. A. Gu, J. Y. Wang, Z. J. Zou and X. J. Han, J. Hazard. Mater., 2014, 268, 216 CrossRef CAS PubMed.
- Y. P. Zang, L. P. Li, X. G. Li, R. Lin and G. S. Li, Chem. Eng. J., 2014, 246, 277 CrossRef CAS PubMed.
- W. Liu, M. L. Wang, C. X. Xu, S. F. Chen and X. L. Fu, J. Mol. Catal. A: Chem., 2013, 368–369, 9 CrossRef CAS PubMed.
- C. J. Li, S. P. Wang, T. Wang, Y. J. Wei, P. Zhang and J. L. Gong, Small, 2014, 10, 2783 CrossRef CAS PubMed.
- L. Ge, C. C. Han and J. Liu, Appl. Catal., B, 2011, 108–109, 100 CrossRef CAS PubMed.
- Y. L. Tian, B. B. Chang, J. L. Lu, J. Fu, F. N. Xi and X. P. Dong, ACS Appl. Mater. Interfaces, 2013, 5, 7079 CAS.
- Z. Y. Jin, N. Murakami, T. Tsubota and T. Ohno, Appl. Catal., B, 2014, 150–151, 479 CrossRef CAS PubMed.
- D. L. Jiang, L. L. Chen, J. M. Xie and M. Chen, Dalton Trans., 2014, 43, 4878 RSC.
- D. L. Jiang, J. J. Zhu, M. Chen and J. M. Xie, J. Colloid Interface Sci., 2014, 417, 115 CrossRef CAS PubMed.
- D. L. Jiang, L. L. Chen, J. J. Zhu, M. Chen, W. D. Shi and J. M. Xie, Dalton Trans., 2013, 42, 15726 RSC.
- J. Wang, P. Guo, Q. S. Guo, P. G. Jönsson and Z. Zhao, CrystEngComm, 2014, 16, 4485 RSC.
- C. S. Xing, Z. D. Wu, D. L. Jiang and M. Chen, J. Colloid Interface Sci., 2014, 433, 9 CrossRef CAS PubMed.
- T. T. Li, L. H. Zhao, Y. M. He, J. Cai, M. F. Luo and J. J. Lin, Appl. Catal., B, 2013, 129, 255 CrossRef CAS PubMed.
- Y. M. He, J. Cai, T. T. Li, Y. Wu, Y. M. Yi, M. F. Luo and L. H. Zhao, Ind. Eng. Chem. Res., 2012, 51, 14729 CrossRef CAS.
- Y. M. He, L. H. Zhang, X. X. Wang, Y. Wu, H. J. Lin, L. H. Zhao, W. Z. Weng, H. L. Wan and M. H. Fan, RSC Adv., 2014, 4, 13610 RSC.
- J. Cai, Y. M. He, X. X. Wang, L. H. Zhang, L. Z. Dong, H. J. Lin, L. H. Zhao, X. D. Yi, W. Z. Weng and H. L. Wan, RSC Adv., 2013, 3, 20862 RSC.
- Y. J. Wang, Q. S. Wang, X. Y. Zhan, F. M. Wang, M. Safdar and J. He, Nanoscale, 2013, 5, 8326 RSC.
- R. Marschall, Adv. Funct. Mater., 2014, 24, 2421 CrossRef CAS.
- J. S. Hu, L. L. Ren, Y. G. Guo, H. P. Liang, A. M. Cao, L. J. Wan and C. L. Bai, Angew. Chem., Int. Ed., 2005, 44, 1269 CrossRef CAS PubMed.
- X. X. Yu, J. G. Yu, B. Cheng and B. B. Huang, Chem.–Eur. J., 2009, 15, 6731 CrossRef CAS PubMed.
- Y. P. Hong, J. Zhang, X. Wang, Y. J. Wang, Z. Lin, J. G. Yu and F. Huang, Nanoscale, 2012, 4, 2859 RSC.
- Y. Feng, N. N. Feng, G. Y. Zhang and G. X. Du, CrystEngComm, 2014, 16, 214 RSC.
- G. J. Lee, A. Manivel, V. Batalova, G. Mokrousov, S. Masten and J. Wu, Ind. Eng. Chem. Res., 2013, 52, 11904 CrossRef CAS.
- H. X. Sang, X. T. Wang, C. C. Fan and F. Wang, Int. J. Hydrogen Energy, 2012, 37, 1348 CrossRef CAS PubMed.
- H. Labiadh, T. B. Chaabane, L. Balan, N. Becheik, S. Corbel, G. Medjahdi and R. Schneider, Appl. Catal., B, 2014, 144, 29 CrossRef CAS PubMed.
- X. F. Li, J. Zhang, L. H. Shen, Y. M. Ma, W. W. Lei, Q. L. Cui and G. T. Zou, Appl. Phys. A: Mater. Sci. Process., 2009, 94, 387 CrossRef CAS.
- S. C. Yan, Z. S. Li and Z. G. Zou, Langmuir, 2009, 25, 10397 CrossRef CAS PubMed.
- M. J. Bojdys, J. O. Müller, M. Antonietti and A. Thomas, Chem.–Eur. J., 2008, 14, 8177 CrossRef CAS PubMed.
- G. H. Dong and L. Z. Zhang, J. Mater. Chem., 2012, 22, 1160 RSC.
- B. Chai, T. Y. Peng, J. Mao, K. Li and L. Zan, Phys. Chem. Chem. Phys., 2012, 14, 16745 RSC.
- Y. J. Cui, J. S. Zhang, G. G. Zhang, J. H. Huang, P. Liu, M. Antonietti and X. C. Wang, J. Mater. Chem., 2011, 21, 13032 RSC.
- M. A. Butler, J. Appl. Phys., 1977, 48, 1914 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available: SEM and TEM images of the pure g-C3N4, ZnS and the samples with different weight ratios of ZnS to g-C3N4, XPS spectra of g-C3N4, ZnS, and 50% ZnS/g-C3N4 samples, and N2 adsorption/desorption isotherms of g-C3N4, ZnS and 50% ZnS/g-C3N4. See DOI: 10.1039/c4ra11740a |
|
This journal is © The Royal Society of Chemistry 2014 |
Click here to see how this site uses Cookies. View our privacy policy here.