Highly efficient visible-light photocatalysts: reduced graphene oxide and C3N4 nanosheets loaded with Ag nanoparticles

Xiaomeng Lü*, Jiayu Shen, Jiaxi Wang, Zhengshan Cui and Jimin Xie
School of Chemistry and Chemical Engineering, Jiangsu University, Zhenjiang 212013, P. R. China. E-mail: lvxm@mail.ujs.edu.cn; Fax: +86-511-88791800; Tel: +86-511-88791708

Received 15th October 2014 , Accepted 19th January 2015

First published on 19th January 2015


Abstract

In the current study, the degradation of methyl orange under visible-light irradiation was performed to gain an insight into the performance of the novel Ag/C3N4/reduced graphene oxide photocatalyst (Ag/C3N4/RGO). Using N,N-dimethylformamide (DMF) as an effective reducing agent as well as solvent, Ag nanoparticles were well anchored on C3N4/RGO nanosheets, which were prepared by a facile hydrothermal reaction. Inexpensive, stable C3N4 was coupled with well-conductive RGO and the plasmon resonance effect of Ag exhibited higher activities compared with pure C3N4, C3N4/RGO, Ag/RGO, and Ag/C3N4, respectively. The presence of Ag and RGO was confirmed by FTIR, XRD, and TEM, and their influence on the activity of C3N4 was demonstrated with the photocatalytic results, detection of reactive species and mechanism analysis. The photoluminescence spectra (PL) and the transient photocurrent response were also tested to determine the enhanced separation of photogenerated charge carriers. The facile synthesis of Ag/C3N4/RGO, together with their superior catalytic performance, successfully provide a promising route for the rational design of comparatively inexpensive and highly active C3N4-based ternary nanostructured composites.


1. Introduction

Semiconductor photocatalysis has attracted intense attention as a potential solution to the worldwide energy shortage and in environmental purification, especially in the photodegradation of organic pollutants.1–5 Over the past years, various semiconductor photocatalysts, such as TiO2, SnO2, ZnO, WO3, Fe2O3, Bi2WO6, and In2S3, have been studied and employed in environmental remediation.6 However, the application of these pure semiconductor catalysts is limited due to various factors, including the high recombination rate of photogenerated electron–hole pairs, low absorption coefficients, or mismatch with the solar spectrum. Constructing efficient sunlight or visible-light-driven photocatalysts with an appropriate bandgap, strong oxidative/reductive ability, and high stability in water solution systems is thus a promising but challenging task.

Recently, a polymeric semiconductor material graphitic carbon nitride (C3N4), only made up of carbon and nitrogen, was used as a metal-free photocatalyst for solar energy conversion, hydrogen production, and environment purification, due to its cheap, abundant, medium bandgap (2.7 eV) and stable merits.7–14 Nevertheless, pure g-C3N4 suffers from a high recombination rate of photogenerated electron–hole pairs, low specific surface area, and low visible-light utilization efficiency, which limit its photocatalytic performance. To solve these problems, many methods, such as chemical doping and modification, have been proposed. Silver nanoparticles15 (AgNPs) are often used in combination with a semiconductor as an efficient photocatalyst, because of their specific optical, electrical, magnetic, catalytic, and antibacterial features, and particularly due to their cheap price compared with platinum or gold. Some Ag-based composite have been demonstrated to be efficient photocatalysts under visible light by our group.16,17 We have recently designed and prepared Ag/C3N4 composites with different AgNPs contents by a facile one-pot hydrothermal method and demonstrated their significantly enhanced photocatalytic efficiency, which was attributed to the more efficient separation of photoinduced electrons from holes and from more reactive species.18 In addition, graphene is known to be an exciting material in material science19 and has recently attracted much attention in photocatalysis,20–25 due to its optical transmittance, large surface area, and high electrical conductivity.26 It has been reported that graphene is also an excellent electron-donating modifier for g-C3N4, due to their similar layered structure and suitable electronic, mechanical, and chemical properties.27 The binary photocatalyst graphene/g-C3N4, synthesized from the combination of graphene and g-C3N4, has also been used for hydrogen production, and showed that the H2-production rate of the composite exceeded that of pure g-C3N4 by more than 3.07 times.28 The combination of graphene sheets and AgNPs to C3N4 may further increase the photocatalytic activity, but this has never been studied.

Herein, we first report a ternary photocatalyst by anchoring AgNPs onto C3N4 and reduced graphene oxide (Ag/C3N4/RGO) sheets, as depicted in Fig. 1, for the degradation of methyl orange (MO) under visible light. We found that the prepared Ag/C3N4/RGO composite exhibited a significantly more rapid photodegradation activity compared with pure C3N4 and the binary photocatalysts Ag/C3N4, Ag/RGO, and C3N4/RGO. We also discuss possible photodegradation mechanisms of the composites in detail based on the experimental results, the PL spectra, their detection of reactive species, and the transient photocurrent response.


image file: c4ra12395f-f1.tif
Fig. 1 Schematic of the formation process of Ag/C3N4/RGO composite.

2. Experimental

2.1. Preparation of graphene oxide (GO)

GO was synthesized using the modified Hummers method from natural graphite powder (natural, −325 mesh, Alfa Aesar).29 Concentrated H2SO4 (69 mL) was added to a mixture of graphite powder (3.0 g) and NaNO3 (1.5 g) in a 100 mL three-neck boiling flask, and the mixture was cooled to 0 °C using an ice bath. KMnO4 (9 g) was very slowly added to the mixture under continuous stirring, and the temperature was kept below 25 °C. The mixture was then warmed to 35 °C and stirred for 7 h. Additional KMnO4 (9 g) was added and the mixture was continuously stirred for 12 h at 35 °C. After cooling down to room temperature, the mixture was then diluted with 400 mL of deionized water, followed by the addition of 30% H2O2 (3 mL), causing the color of the mixture to change to bright yellow. Afterwards, the mixture was centrifuged and washed successively with 1 M HCl solution (at least three centrifugation cycles, 8000 rpm for 5 min, TG16-WS, Hunan Xiangyi Laboratory Instrument Development Co., Ltd. China) and deionized water (at least five centrifugation cycles, 10[thin space (1/6-em)]000 rpm for 10 min) until the filtrate became neutral. The product was separated by centrifugation, washed with ethanol, and dried under vacuum at 60 °C overnight. The resulting product was dispersed in deionized water, followed by ultrasonic treatment for 2 h. The final product was separated by centrifugation, washed with ethanol, and dried under vacuum at 60 °C.

2.2. Preparation of C3N4/RGO

C3N4 was prepared according to our previous report.18 In a typical case, 0.21 g of as-prepared C3N4 was dispersed into 20 mL of deionized water by stirring for 10 min, followed by ultrasonication for 30 min. Afterwards, 0.09 g of GO was dispersed into 20 mL of deionized water, and sonicated for 1 h. Then, it was added to the above solution dropwise. The mixture was stirred for 30 min, and then transferred into a 50 mL Teflon-lined stainless steel autoclave. The temperature was increased to 140 °C and maintained for 6 h.30 Afterwards, the mixture was cooled down to room temperature and washed with deionized water and ethanol several times, then dried under vacuum at 60 °C.

2.3. Preparation of Ag/C3N4/RGO, Ag/C3N4 and Ag/RGO

The typical procedures for the fabrication of the Ag/C3N4/RGO ternary hybrid were as follows: first, 0.17 g of dried C3N4/RGO powder was sonicated in 30 mL N,N-dimethylformamide (DMF) for 1 h in a beaker. In a separate beaker, 0.047 g of AgNO3 and 0.05 g of PVP were stirred in 10 mL DMF for 30 min. Then, the two mixtures were poured into a screw-capped test tube and heated over a silicon oil bath for 20 h. A schematic of the preparation of Ag/C3N4/RGO nanocomposite is shown in Fig. 1. After being centrifuged repeatedly at 4500 rpm with deionized water to remove the adsorbed Ag+ ions and the unbound Ag NPs, the final product was obtained after washing with ethanol and drying at 50 °C under vacuum for 12 h. The final product was then stored for further characterization and photocatalytic study. Ag/RGO and Ag/C3N4 photocatalysts were prepared using the same method without the presence of C3N4 and GO, respectively.

2.4. Characterization

The crystal structure and phase purity of the Ag/C3N4 were analyzed by X-ray diffraction (XRD) using a D8 Advance X-ray diffraction (Bruker axs company, Germany) equipped with Cu-Kα radiation (λ = 1.5406 Å), employing a scanning rate of 7° min−1 in the 2θ range from 5° to 80°. Fourier transform infrared spectroscopy (FT-IR) analysis was carried out on Nicolet Model Nexus 470 IR equipment to determine the specific functional groups present on the surface. The morphology of the as-prepared catalysts were also examined by transmission electron microscopy (TEM), which was recorded on a JEOL-JEM-2010 (JEOL, Japan) operating at 200 kV. UV-vis diffuse reflectance spectra (UV-vis/DRS) were recorded on a Shimadzu UV2450 spectrophotometer. The PL of the samples was obtained using a Varian Cary Eclipse spectrometer.

2.5. Photocatalytic experiments

The photocatalytic activity of the as-prepared Ag/C3N4/RGO composite was evaluated by the degradation of MO. All the experiments were performed in a self-made quartz photoreactor fitted with a circulation water system to maintain a constant temperature and under visible-light irradiation of a 300 W xenon lamp (PLS-SXE300, Trusttech Co. Ltd., Beijing) with a 400 nm cutoff filter to remove the UV irradiation. In all the photocatalytic degradation experiments, the catalyst (50 mg) was suspended in 100 mL of 10 mg L−1 MO aqueous solution. Before the illumination, the suspension was magnetically stirred for 30 min in the dark to reach the adsorption–desorption equilibrium between the dye and the catalyst. At given irradiation time intervals, 3 mL of the suspension was collected and subsequently centrifuged to remove the catalyst particles. The concentration was analyzed by measuring the maximum absorbance at 464 nm for MO using a Shimadzu UV-2450 spectrophotometer. The degradation efficiency of methyl orange was calculated by the following equation:
 
image file: c4ra12395f-t1.tif(1)
where C0 and A0 are the initial concentration and absorbance of methyl orange solution at 464 nm, corresponding to maximum absorption wavelength; Ct and A are the concentration and absorbance of methyl orange solution at 464 nm after UV light irradiation at any time.

2.6. Photoelectrochemical measurements

The photocurrents were measured using an electrochemical analyzer (CHI660B, ChenHua Instruments, Shanghai, China) to investigate the transition of the photogenerated electrons in C3N4, C3N4/RGO, Ag/C3N4, and Ag/C3N4/RGO, respectively. A standard three-electrode cell was used equipped with a platinum wire as the counter electrode, a saturated calomel electrode (SCE) as the reference electrode, and an indium-tin oxide glass (ITO) as the working electrode, respectively. A 300 W Xe arc lamp with λ ≥ 400 nm was utilized as the light source. 0.1 M Na2SO4 aqueous solution was used as the electrolyte solution. For the working electrodes, 20 μL 5.0 mg mL−1 C3N4, C3N4/RGO, Ag/C3N4, and Ag/C3N4/RGO solution (dispersed in ethanol) were dropped onto the pretreated ITO with a 1 cm × 0.5 cm area and dried to form the modified ITO.

3. Results and discussion

3.1. Characterization of photocatalysts

As revealed by the XRD pattern (Fig. 2a), two distinct peaks at 13.0° and 27.4° (corresponding to 0.681 nm and 0.326 nm, respectively), in the pure C3N4 sample can be indexed as the (100) and (002) diffractions for the graphitic materials, respectively, corresponding to the in-planar structural packing motif and the interlayer stacking of the aromatic system.7 In the XRD pattern of GO, the intense and sharp diffraction peak at 2θ = 9.8° is attributed to the (001) lattice planes. The C3N4/RGO exhibits a similar XRD pattern as pure C3N4, suggesting that the incorporation of GO sheets has little influence on the crystallite size and phase structure of C3N4. The four diffraction peaks of (111) at 38.2°, (200) at 44.6°, (220) at 64.8°, and (311) at 77.9° in the patterns of Ag/C3N4 and Ag/C3N4/RGO match well with the face-centered cubic structure Ag, indicating the formation of metallic Ag in these nanocomposites. We further analyzed the formation of GO, C3N4, C3N4/RGO, Ag/C3N4, and Ag/C3N4/RGO composites by using FTIR. The broad absorption at 3000–3700 cm−1 and 1625 cm−1 of GO correspond to the stretching mode of hydroxyl, and the stretching vibration at 1735 cm−1 and 1061 cm−1 are assigned to the C[double bond, length as m-dash]O and C–O–C groups, respectively (Fig. 1b). The weak bending vibration and stretching vibration at 1424 cm−1 and 1218 cm−1 are assigned to the O–H deformations of the C–OH groups. However, the FTIR spectra of C3N4/RGO and Ag/C3N4/RGO composites show no characteristic peaks of GO, indicating the reduction of GO. The result is consistent with the result of the XRD pattern. The FTIR spectra of all the doped C3N4 samples show the typical stretching modes of CN heterocycles in the 1200–1650 cm−1 range, and the breathing mode of the tri-s-triazine units at 808 cm−1 indicating that modification with GO or/and AgNPs does not influence the structure of the composites. Therefore, the FTIR analysis, as well as the results of the XRD, supports the formation of the Ag/C3N4/RGO composite.
image file: c4ra12395f-f2.tif
Fig. 2 XRD patterns (a) and FTIR spectra (b) of the as-prepared GO, C3N4, C3N4/RGO, Ag/C3N4, and Ag/C3N4/RGO composites.

The morphology and microstructure of the samples are examined by TEM and high-resolution TEM (HR-TEM). The TEM image of GO reveals that the surface morphology of GO is smooth and has a wrinkle-like structure (Fig. 3a), and the pure C3N4 structure is made up of a few nanometer-sized sheets and pores (Fig. 3b). As seen from the TEM image of Ag/C3N4, spherical-shaped Ag nanoparticles are observed uniformly on the surface of C3N4 (Fig. 3c), since adding PVP during the synthesis affects the molecular motion of the reduced silver and subsequently limits the aggregation of colloids.18 Though they are tightly connected with each other, we can also observe the intrinsic structure of C3N4 and reduced GO (Fig. 3d). From the TEM image of Ag/C3N4/RGO (Fig. 3e), it can be seen that the AgNPs and C3N4 sheets completely cover the RGO sheets to form the Ag/C3N4/RGO composites. The average diameter of AgNPs is about 10 nm. The existence of AgNPs is also confirmed by the HRTEM image in Fig. 3f, the lattice distance of 0.236 nm corresponding to the d-spacing of the (111) crystallographic planes of metallic Ag.


image file: c4ra12395f-f3.tif
Fig. 3 TEM of the as-prepared GO (a), C3N4 (b), Ag/C3N4 (c), C3N4/RGO (d), Ag/C3N4/RGO (e), and HRTEM of Ag (f).

3.2. Photocatalytic activity and mechanisms

Fig. 4a shows the photocatalytic activity of all the samples under visible light. It clearly shows that the degradation efficiency of MO over Ag/C3N4/RGO was higher than that of other photocatalysts. MO degradation follows a pseudo-first-order kinetics (Fig. 4b), and the kinetic constant with C3N4, Ag/RGO, C3N4/RGO, Ag/C3N4 and Ag/C3N4/RGO are 0.035, 0.060, 0.085, 0.117, and 0.432 h−1, respectively. The obtained degradation rate on Ag/C3N4/RGO was 12.3, 7.2, 5.1, and 3.7 times larger than that on C3N4, Ag/RGO, C3N4/RGO and Ag/C3N4, respectively.
image file: c4ra12395f-f4.tif
Fig. 4 (a) Degradation rate of methyl orange under visible-light irradiation in the presence of catalysts: pure C3N4; Ag/RGO; C3N4/RGO; Ag/C3N4 and Ag/C3N4/RGO, and (b) MO degradation curves of ln(C/C0) versus time for different catalysts.

Since the absorption property of a catalyst is very important, UV-vis diffuse reflectance spectra (DRS) was used to test the optical absorption properties of the photocatalysts (Fig. 5). Coupling with RGO, the absorption intensity in the visible-light region of C3N4 was clearly improved, albeit, modified with AgNPs, leading to the absorption edge being shifted towards a higher wavelength, indicating a decreased bandgap.17 Both Ag/C3N4 and Ag/C3N4/RGO show intensive absorption in the visible region, which is consistent with the surface plasmon resonance (SPR) effect of Ag, further confirming the formation of AgNPs. The bandgaps of C3N4, C3N4/RGO, Ag/C3N4, and Ag/C3N4/RGO are estimated, from their UV-vis spectra, to be 2.61, 2.37, 2.02, and 1.78 eV, respectively, showing an intrinsic semiconductor-like absorption in the blue region of the visible spectrum.


image file: c4ra12395f-f5.tif
Fig. 5 UV-vis diffuse reflectance spectra (DRS) of bulk C3N4, C3N4/RGO, Ag/C3N4, and Ag/C3N4/RGO composites.

The improved photocatalytic performance of Ag/C3N4/RGO can be further explained by the PL spectra in Fig. 6. The pure C3N4 sample was found to have a higher PL intensity than the composites. This indicates that C3N4 had the highest optical recombination rate, which could deteriorate the photodegradation efficiency. However, PL quenching was observed in the composites, and was most significant in Ag/C3N4/RGO, which indicates that this composite had a lower rate of recombination of photoelectrons compared with pure C3N4 and the other composites, thus greatly improving the photocatalytic activity. To provide additional evidence for the separation of the photoelectrons and holes, the transient photocurrent responses of the C3N4, C3N4/RGO, Ag/C3N4, and Ag/C3N4/RGO samples were recorded over several cycles of on-off visible-light irradiation, as shown in Fig. 7. The on–off cycles of the photocurrent are highly reproducible, and all the samples show an apparently boosted photocurrent response under visible-light illumination, which is attributed to the separation of the photoelectrons and holes within the photo-electrode. The photoelectron may transport to the cathode, via C3N4 or RGO sheets, and photoinduced holes are trapped or captured by a reduced species in the electrolyte. Compared to the C3N4, C3N4/RGO, and Ag/C3N4 samples, the Ag/C3N4/RGO sample shows the highest transient photocurrent. This is presumably due to the following two reasons: (1) in the presence of AgNPs, the interface of Ag/C3N4 can form the Schottky junction, which results in the separation of the photoelectrons and holes; (2) RGO sheets, due to their two-dimensional π-conjugation structure, can act as an electron collector and transporter in the Ag/C3N4/RGO sample, which effectively suppresses the charge recombination, leaving more charge carriers to form reactive species.


image file: c4ra12395f-f6.tif
Fig. 6 Room-temperature PL emission spectra of different photocatalysts excited by wavelength at 393 nm.

image file: c4ra12395f-f7.tif
Fig. 7 The transient photocurrent response of C3N4, C3N4/RGO, Ag/C3N4, and Ag/C3N4/RGO composites.

To further explore the mechanism, reactive oxidative species trapping experiments were performed to investigate the main reactive oxidative species involved in the photocatalytic process by using three different scavengers: disodium ethylenediaminetetraacetate (EDTA, holes scavenger), tert-butanol (t-BuOH, OH˙ radicals scavenger), and p-benzoquinone (BQ, O2˙ radicals scavenger) (Fig. 8). The photodegradation of MO was significantly suppressed by the introduction of 1 mmol BQ, indicating that the O2˙ radical species are the main reactive species involved in the photocatalysis.31 In contrast, the photodegradation of MO undergoes no conspicuous changes by the introduction of 1 mmol EDTA anions, implying that the holes make no contribution to the degradation process.32 When 1 mmol of t-BuOH was added, the degradation rate of MO was reduced from 93.4% to 76%, indicating that OH˙ radical species are minor reactive species involved in the photocatalytic oxidation process.33


image file: c4ra12395f-f8.tif
Fig. 8 Degradation ratio of Ag/C3N4/RGO composites with different scavengers.

On the basis of the above experimental results and discussion, the improved photocatalytic performance arises from the enhanced light harvesting and more efficient separation of photogenerated electron–hole pairs due to the Ag/C3N4/RGO composite. The possible mechanism for the photocatalytic degradation of MO by Ag/C3N4/RGO is thus proposed and schematically illustrated in Fig. 9.


image file: c4ra12395f-f9.tif
Fig. 9 Schematic of methyl orange degradation over Ag/C3N4/RGO composite photocatalyst under visible light.

The relevant reactions are listed in eqn (2)–(8). Under visible-light irradiation, owing to the visible-light wavelength bandgap energy of pure C3N4, it may absorb visible light and generate electron–hole pairs when it is irradiated under the light with photon energy equal to or greater than approximately 2.61 eV (eqn (2)). For the present Ag/C3N4/RGO composite, the process becomes quicker, because it has a smaller bandgap than pure C3N4. The photogenerated electrons from C3N4 can then transfer easily to the RGO sheet due to their suitable energy levels (eqn (3)),34 leaving positively charged holes behind. The electrons were captured by the RGO sheet and utilized for the reduction of oxygen, thus forming O2˙ radical species (eqn (4)). Meanwhile, under light irradiation, the plasmonic Ag nanocrystals also generate electron–hole pairs through surface plasmon resonance (SPR), in which the active photo-induced electrons can reduce O2 molecules to form O2˙ radicals,35 leaving behind holes (Ag+) that can be filled by electrons from C3N4/RGO (eqn (5)–(7)). The O2˙ radicals together with the photo-generated holes in the VB of C3N4 can oxidize organic dye directly (eqn (8)).

 
C3N4 + hv → C3N4(e + h+) (2)
 
C3N4(e + h+) + RGO → C3N4(h+) + RGO(e) (3)
 
RGO(e) + O2 → RGO + O2˙ (4)
 
Ag + hv → Ag* (5)
 
Ag* + O2 → O2˙ + Ag+ (6)
 
Ag+ + RGO(e) → Ag + RGO (7)
 
O2˙/C3N4(h+) + MO → CO2 + H2O (8)

Therefore, the RGO sheet and Ag nanocrystals play important roles in the efficient separation of photo-generated electron–hole pairs and the generation of reactive oxidation species of O2˙, and thereby enhance the photodegradation activity of ternary composites. In addition, the relative narrow bandgap, as well as the increased absorbance in the visible-light region, also contributes to the improved photocatalytic activity of the Ag/C3N4/RGO composite.

4. Conclusion

We synthesized a ternary photocatalyst Ag/C3N4/RGO by anchoring Ag nanoparticles onto C3N4 and RGO sheets, which exhibited superior photocatalytic activity over pure C3N4 and binary C3N4/RGO, Ag/RGO, and Ag/C3N4 photocatalysts under visible light. The enhanced photocatalytic activity could be attributed to the enhanced light harvesting and more efficient separation of photogenerated electron–hole pairs due to the addition of AgNPs and the RGO sheets. The Ag/C3N4/RGO photocatalyst may have potential applications in the fabrication of industrial photocatalytic devices and in environmental conservation.

Acknowledgements

This work was financially supported by the National Natural Science Foundation of China (no. 21003065). This work was partly financially supported by the Scientific Innovation Research of College Graduate in Jangsu Province (SJLX_0480) and Jiangsu Province Research Joint Innovation Fund-Prospective Joint Research Project (BY2014123-08). We also thank support of China Council Scholarship.

References

  1. H. Tong, S. X. Ouyang, Y. P. Bi, N. Umezawa, M. Oshikiri and J. H. Ye, Adv. Mater., 2012, 24, 229–251 CrossRef CAS PubMed.
  2. A. Kubacka, M. Fernández-García and G. Colón, Chem. Rev., 2012, 112, 1555–1614 CrossRef CAS PubMed.
  3. X. M. Lü, D. J. Mao, X. J. Wei, H. Zhang, J. M. Xie and W. Wei, J. Mater. Res., 2013, 28, 400–404 CrossRef.
  4. C. Cui, Y. P. Wang, D. Y. Liang, W. Cui, H. H. Hu, B. Q. Lu, S. Xu, X. Y. Li, C. Wang and Y. Yang, Appl. Catal., B, 2014, 158, 150–160 CrossRef PubMed.
  5. J. Y. Xiong, G. Cheng, G. F. Li, F. Qin and R. Chen, RSC Adv., 2011, 1, 1542–1553 RSC.
  6. M. Mecklenburg, A. Schuchardt, Y. K. Mishra, S. Kaps, R. Adelung, A. Lotnyk, L. Kienle and K. Schulte, Adv. Mater., 2012, 24, 3486–3490 CrossRef CAS PubMed.
  7. X. Wang, K. Maeda, A. Thomas, K. Takanabe, G. Xin, J. M. Carlsson, K. Domen and M. Antonietti, Nat. Mater., 2009, 8, 76–80 CrossRef CAS PubMed.
  8. Y. Wang, X. Wang and M. Antonietti, Angew. Chem., Int. Ed., 2012, 51, 68–89 CrossRef CAS PubMed.
  9. S. Zhou, Y. Liu, J. M. Li, Y. J. Wang, G. Y. Jiang, Z. Zhao, D. X. Wang, A. J. Duan, J. Liu and Y. C. Wei, Appl. Catal., B, 2014, 158, 20–29 CrossRef PubMed.
  10. Z. F. Jiang, D. Liu, D. L. Jiang, W. Wei, K. Qian, M. Chen and J. M. Xie, Dalton Trans., 2014, 13792–13802 RSC.
  11. S. Y. Yang, W. Y. Zhou, C. Y. Ge, X. T. Liu, Y. P. Fang and Z. S. Li, RSC Adv., 2013, 3, 5631–5638 RSC.
  12. Y. P. Zang, L. P. Li, Y. Zuo, H. F. Lin, G. S. Li and X. F. Guan, RSC Adv., 2013, 3, 13646–13650 RSC.
  13. S. Samanta, S. Martha and K. Parida, ChemCatChem, 2014, 6, 1453–1462 CAS.
  14. S. Martha, A. Nashim and K. M. Parida, J. Mater. Chem. A, 2013, 1, 7816–7824 CAS.
  15. M. H. Fulekar, A. Singh, D. P. Dutta, M. Roy, A. Ballald and A. K. Tyagic, RSC Adv., 2014, 4, 10097–10107 RSC.
  16. Z. F. Jiang, X. M. Lv, D. L. Jiang, J. M. Xie and D. J. Mao, J. Mater. Chem. A, 2013, 1, 14963–14972 CAS.
  17. Z. F. Jiang, J. J. Zhu, D. Liu, W. Wei, J. M. Xie and M. Chen, CrystEngComm, 2014, 16, 2384–2394 RSC.
  18. X. M. Lü, J. Y. Shen, Z. W. Wu, J. X. Wang and J. M. Xie, J. Mater. Res., 2014, 29, 2170–2178 CrossRef.
  19. D. R. Dreyer, R. S. Ruoff and C. W. Bielawski, Angew. Chem., Int. Ed., 2010, 49, 9336–9344 CrossRef CAS PubMed.
  20. Q. J. Xiang and J. G. Yu, Chem. Soc. Rev., 2012, 41, 782–796 RSC.
  21. H. Liu, Z. Chen, Z. T. Jin, Y. Su and Y. Wang, Dalton Trans., 2014, 7491–7498 RSC.
  22. J. Q. Xu, M. D. Chen and Z. M. Wang, Dalton Trans., 2014, 3537–3544 RSC.
  23. H. F. Li, H. T. Yu, S. Chen, H. M. Zhao, Y. B. Zhang and X. Quan, Dalton Trans., 2014, 2888–2894 RSC.
  24. Y. Feng, N. N. Feng, Y. Z. Wei and G. Y. Zhang, RSC Adv., 2014, 4, 7933–7943 RSC.
  25. J. J. Guo, Y. Li, S. M. Zhu, Z. X. Chen, Q. L. Liu, D. Zhang, W. J. Moon and D. M. Song, RSC Adv., 2012, 2, 1356–1363 RSC.
  26. A. K. Geim and K. S. Novoselov, Nat. Mater., 2007, 6, 183–191 CrossRef CAS PubMed.
  27. X. H. Li, J. S. Chen, X. C. Wang, J. H. Sun and M. Antonietti, J. Am. Chem. Soc., 2011, 133, 8074–8077 CrossRef CAS PubMed.
  28. Q. J. Xiang, J. G. Yu and M. Jaroniec, J. Phys. Chem. C, 2011, 115, 7355–7363 CAS.
  29. D. C. Marcano, D. V. Kosynkin, J. M. Berlin, A. Sinitskii, Z. Sun, A. Slesarev, L. B. Alemany, W. Lu and J. M. Tour, ACS Nano, 2010, 4, 4806–4814 CrossRef CAS PubMed.
  30. Y. Zhou, Q. L. Bao, L. A. L. Tang, Y. L. Zhong and K. P. Loh, Chem. Mater., 2009, 21, 2950–2956 CrossRef CAS.
  31. J. P. Zou, S. L. Luo, L. Z. Zhang, J. Ma, S. L. Lei, L. S. Zhang, X. B. Luo, Y. Luo, G. S. Zeng and C. T. Au, Appl. Catal., B, 2013, 140, 608–618 CrossRef PubMed.
  32. J. P. Zou, J. Ma, Q. Huang, S. L. Luo, J. Yu, X. B. Luo, W. L. Dai, J. Sun, G. C. Guo, C. T. Au and S. L. Sui, Appl. Catal., B, 2014, 156, 447–455 CrossRef PubMed.
  33. Y. S. Xu and W. D. Zhang, Dalton Trans., 2014, 1094–1101 Search PubMed.
  34. R. C. Pawar, V. Khare and C. S. Lee, Dalton Trans., 2014, 12514–12527 RSC.
  35. M. Rycenga, C. M. Cobley, J. Zeng, W. Li, C. H. Moran, Q. Zhang, D. Qin and Y. N. Xia, Chem. Rev., 2011, 111, 3669–3712 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.