DOI:
10.1039/C4RA09546D
(Review Article)
RSC Adv., 2015,
5, 6543-6552
Graphene-based photocatalysts for oxygen evolution from water
Received
31st August 2014
, Accepted 8th December 2014
First published on 8th December 2014
Abstract
Graphene (GR) has triggered new research fields in material science, due to its unique monolayer structure, remarkably high conductivity, superior electron mobility, extremely high specific surface area and chemical stability. It is considered to be an ideal matrix and electron mediator of semiconductor nanoparticles for environmental and energy applications. In particular, GR-based nanocomposites have attracted significant attention when used as photocatalysts. This review will focus on oxygen evolution from water using GR-semiconductor photocatalytic systems. Recent achievements of effective strategies in the fabrication of GR-based semiconductor photocatalysts for oxygen evolution from water are summarized. Furthermore, the morphology control and composition design of semiconductors on GR sheets are also reviewed in relation to their photocatalytic oxygen generation properties. This review ends with a summary and some perspectives on the major challenges and opportunities in the future research.
 H. Pan | H. Pan received his Master's degree from Donghua University at the State Key Laboratory for Modification of Chemical Fibers and Polymer Materials, College of Materials Science and Engineering in 2014. Presently he is a PhD student under the supervision of Professor S. Zhu. His current work focuses on graphene-based functional materials. |
 S. Zhu | S. Zhu received her PhD degree from Shanghai Jiaotong University in 2001. She is presently a professor at the School of Materials Science and Engineering, State Key Lab of Metal Matrix Composites, Shanghai Jiao Tong University. Her current fields of interest are graphene-based functional materials, porous carbon, and biomimetic materials based on biological templates. |
1. Introduction
Photocatalytic water splitting is of great importance, due to its efficient conversion of solar energy to chemical energy, and it represents a promising technology to solve the global energy and environmental challenges.1–5 Since the pioneering work reported by Honda,6 the concept of sunlight-induced H2 or O2 production has stimulated considerable research efforts, leading to the development of numerous photocatalytic catalysts.7 The semiconductor's band gap and electronic band edge positions with respect to water oxidation/reduction potential levels are very crucial in determining the feasibility of solar hydrogen/oxygen production. The ideal band gap of the semiconductors should be around 2.0 eV for the effective utilization of solar energy. Moreover, the bottom of the conduction band (CB) must be located at a more negative potential than the reduction potential (H+/H2), whereas the top of the valence band (VB) must be located at a more positive potential than the oxidation potential (H2O/O2).8–10 Unfortunately, rare semiconductor photocatalysts possess suitable redox potentials and band gaps for both simultaneous water reduction and oxidation. As a result, many studies have focused on half of the reaction by using sacrificial reagents as electron donors or acceptors.11 Generally, hydrogen production from water reduction requires two electrons, whereas oxygen evolution from water oxidation is more challenging, because it requires four holes to generate two oxygen–oxygen bonds for the production of one oxygen molecule.12–18 For this reason, reported active photocatalysts for oxygen evolution from water, such as WO3 and BiVO4,19–21 are rare. Recently, some new semiconductors, such as BiCu2VO6, BiZn2VO6, TiNxOyFz, nitrogen-doped CsCa2Ta3O10, layered double hydroxides (LDH) and Ag3PO4 crystals, have also been investigated.22–27 Evidently, the oxygen evolution reaction is crucial for renewable energy technologies, including water splitting and fuel cells.28 Unfortunately, oxygen-evolving photocatalysts represent a bottleneck for the development of energy-conversion schemes based on sunlight.29
The principle and main process of photocatalytic oxygen generation from water is schematically shown in Fig. 1. Two key requirements should be considered in the fabrication of photocatalysts for efficient solar oxygen generation from water: one is excellent photo-adsorption ability within the solar spectrum, and the other is a low recombination rate of photogenerated electron–hole pairs.1,30–33 In terms of the first issue, some semiconductors have been found to possess efficient absorption in the visible-light solar spectrum, which occupies ∼43% of solar radiation energy; these include Bi2WO6 and Bi2MoO6 with an aurivillius structure, and BiVO4 with a monoclinic scheelite structure.34–36 However, most of these semiconductors are not ideal photocatalytic materials, because of their low quantum efficiencies. A great deal of effort has been made toward improving their photocatalytic performance. Coupling with other substances to fabricate semiconductor-based nanocomposites, such as the utilization of co-catalysts, Z-scheme photocatalysis, carbon quantum dots (CQD), and graphene (GR), is often viewed as a feasible route for efficient photocatalysis.37–40 Among these, semiconductors incorporated with GR have attracted considerable interest due to their unique properties.41 With a flat monolayer of carbon atoms tightly packed into a two-dimensional honeycomb lattice, GR possesses remarkably high conductivity, superior electron mobility and an extremely high specific surface area (2630 m2 g−1), and can be produced on a large scale at low cost.42–44 Given the excellent electronic conductivity endowed by its two-dimensional planar p-conjugation structure, GR in composites acts not only as a superior supporting matrix for bonding with functional components, but also as an excellent electron mediator to adjust electron transfer. Thus, it restrains the recombination of photoexcited charges and enhances the efficiency of electron–hole separation.45
 |
| Fig. 1 The principle and main process of photocatalytic oxygen generation from water. | |
As is known, semiconductors possessing valence band positions that are more positive than the oxidation potential (H2O/O2) are potential photocatalysts for producing oxygen from water (Fig. 2); examples include WO3, Fe2O3, TiO2, and ZnO.46,47 When integrated with graphene, the unique characteristics of GR make it more attractive for use in oxygen evolution. Photocatalytic systems, such as BiWO6/GR, hematite/reduced graphene oxide (RGO), NiTi-LDH/RGO, WO3/GR, BiVO4/graphene oxide (GO) and Ag3PO4/Ag/AgBr/RGO, have been reported to have greatly improved catalytic properties for oxygen evolution.48–53 The combination of GR in photocatalysts has two evident advantages: one is to widen light absorption range; the other one is to promote a stable catalytic system. For example, it is known that TiO2 and ZnO can only be excited by ultraviolet (UV) or near UV radiation, owing to their wide band gaps. Inspiringly, the band gap energy values for these GR-based composites (GR nanosheets–TiO2 and GR nanosheet–ZnO), were measured to be 1.39 and 1.26 eV, respectively. The narrowed band gap originating from GR makes this type of composite a promising candidate for water splitting under sunlight.54 Bai et al. found that the incorporation of GR greatly enhanced the stability of Ag3PO4, which is otherwise photocorrosive when used for photocatalytic O2 production.55 In addition, the introduction of GR to WO3 results in an enhanced surface area and an efficient separation of charges. It is worth mentioning that the excess addition of GR leads to a decrease in catalytic efficiency, because of the shielding effect of black GR on the active sites.56
 |
| Fig. 2 The conduction band and valence band positions of selected semiconductors (V vs. NHE, pH = 7). Reproduced from ref. 8. | |
Thus, GR-based semiconductor photocatalysts have attracted extensive attention because of their usefulness in environmental and energy applications. This critical review summarizes the recent progress in the fabrication of GR-based semiconductor photocatalysts for oxygen evolution from water under visible-light irradiation. The morphology control and composition design of semiconductors on GR sheets are reviewed in relation to their photocatalytic oxygen generation properties. The importance of the interface between the semiconductors and GR is highlighted. The simultaneous evolution of H2 and O2 from water using GR-based materials is also discussed. This review ends with a summary and some perspectives on the challenges and new directions in this emerging area of research.
2. Synthesis of graphene-based photocatalysts
2.1. Solution mixing and in situ growth
Solution mixing is one of the most widely used methods to fabricate GR-based photocatalysts, with easy operation and satisfactory results.41,57 To improve the photocatalytic properties of WO3, Ng et al. fabricated WO3/GR composites by mixing WO3 powder with GR oxide (GO).58 WO3 was added to the GO suspension and ultrasonicated for 30 min to produce a WO3/GO dispersion. The dispersion was then exposed to either UV or visible light irradiation for 3 h to obtain WO3/RGO composites. Although the morphological features of the WO3/RGO composites were essentially identical, the WO3 particle sizes were as large as 100 nm and tended to aggregate. Using the same strategy, WO3/RGO composites were prepared using polyvinylpyrrolidone (PVP) as an intermediate to combine tungsten with RGO. The mixture of GO and ammonium metatungstate hydrate was ultrasonicated and stirred to obtain the WO3/GO precursor solution. The resultant WO3/RGO composite was obtained by calcining the precursors at 450 °C for 5 h in air. As a result, WO3 nanocrystallites ranging from 20–40 nm were uniformly distributed on the RGO.59
In 2014, an in situ growth method was developed for the fabrication of visible-light-driven NiTi-LDH/RGO catalysts by anchoring NiTi-LDH nanosheets onto the surface of RGO, which displayed excellent photocatalytic behaviour for the splitting of water into oxygen.50 The monolayer RGO suspension was obtained by sonication in deionized water. Titanium, nickel sources (Ni(NO3)2·6H2O, TiCl4) and urea were dissolved in the RGO suspension. After stirring at 90 °C, the final precipitate was dried in an oven at 60 °C for 24 h. The NiTi-LDH nanosheets on the surface of RGO were highly dispersed and possessed a plate-like morphology with a lateral diameter of 100–200 nm. Elemental mapping images displayed the uniform and homogeneous distribution of both Ni and Ti.
2.2. Hydrothermal and solvothermal approaches
Apart from the approaches mentioned above, the hydrothermal method is another efficient method for the synthesis of GR-based photocatalysts.60–62 As compared with the solution mixing method, hydrothermal and solvothermal approaches are more attractive, owing to the controllable morphology of the semiconductor particles. More importantly, the nanoparticles bond well to GR even without the use of intermediates. Meng et al. used a hydrothermal process to prepare α-Fe2O3/RGO composites, in which hematite nanoparticles are supported on RGO nanosheets. A proper amount of FeCl3·6H2O was mixed with GO, and then the mixture was dissolved in deionized water and sonicated. Then, the prepared solution was mixed with ethanol and placed in a boiling aqueous bath for thermal hydrolysis. The sample collected by centrifuging was heated in air at 350 °C for 2 h and then in pure nitrogen at 800 °C for 15 min. The crystalline size of monolithic α-Fe2O3 (63.8 nm) was larger than that of α-Fe2O3 (41.4 nm) grown on the RGO sheets. The TEM image showed single-crystalline structure features of the α-Fe2O3 particles on the RGO sheets. The Fe2O3/RGO composite showed an enhanced photocatalytic activity toward water oxidation compared with pristine α-Fe2O3 nanoparticles.49 Similarly, Niu et al. prepared RGO-cuprous oxide (Cu2O/RGO) photocatalysts using the same method. Well monodispersed cube-like Cu2O particles (300–500 nm) were observed to precipitate on the RGO layers.63 Moreover, RGO/TiO2 microspheres were produced by the non-hydrolytic sol–gel reaction of tetrabutyl orthotitanate (TBOT) and acetone followed by hydrothermal treatment.64 The pretreated TiO2 was prepared by adding TBOT to excess acetone. TiO2/RGO microspheres were prepared by hydrothermal treatment with temperatures varying from 120 to 180 °C. During this process, ammonia was used as a solution medium. It was found that the pre-treated TiO2 was smooth and spherical in shape with particle diameters in the range between 1 and 2 μm. After treatment at 120 °C, microspheres with aggregates of nanoflakes on their surface were produced. When the hydrothermal treatment temperature was increased from 150 to 160 °C, rough microspheres with aggregates of nanorods were obtained. More recently, a solvothermal method was developed for the fabrication of Bi2WO6/RGO (BWO/RGO) composites. A mixture of GO, Bi(NO3)3·5H2O and Na2WO4·2H2O was sealed into a Teflon-lined autoclave and then maintained at 180 °C. Finally, Bi2WO6 nanoparticles deposited on RGO sheets were produced, which showed some wrinkles. Evidently, the solvothermal approach is widely used in the fabrication of these GR-based photocatalysts. Unfortunately, most BWO particles were as large as a few hundred nanometers, and their dispersion on GO sheets were not uniform, owing to the difficulty in controlling the composition and phase of the complex ternary compounds.65
2.3. Sonochemical methods
As mentioned above, particles more than several hundred nanometers in size are obtained on GR sheets by using the in situ growth method. Two key issues must be considered for the design and fabrication of GR-based photocatalysts: (i) particle size control and (ii) the interface between the catalyst and GR. Controlling the particle size and improving the interaction between the semiconductor and GR is a challenge. In our previous work, ultrasonic waves proved to be effective in solving this problem.66–68 Using ultrasonication, we succeeded in a controlled incorporation of TiO2 nanoparticles onto GR layers homogeneously in a few hours. The average size of the nanoparticles was controlled at around 4–5 nm on the sheets without using any surfactant, which is attributed to pyrolysis and the condensation of the dissolved TiCl4 into TiO2 by the ultrasonic waves. The uniform dispersion of TiO2 nanoparticles on both the GR surface and the interlayers was confirmed by SEM and TEM images. The results suggest that ultrasound is very effective in dispersing TiO2 nanoparticles on GR layers.68
Inspired by the effective ultrasonic waves, we reported the synthesis of a composite (WO3/GR) consisting of WO3 nanoparticles and GR sheets using a sonochemical method. The average particle size of the WO3 was controlled at around 12 nm on the GR sheets without using any surfactant. The composite consisted of nano-WO3 particles, and the two-dimensional GR sheets are a promising photocatalyst for oxygen production. When used as photocatalyst for water splitting, the amount of evolved O2 from WO3/GR with 40 wt% GR inside was considerably higher than that from pure WO3 and mixed-WO3/GR, 1.8 times and 2 times as much as that from mixed-WO3/GR (ca. 214 mmol L−1) and pure WO3 (ca.186 mmol L−1), respectively. The improved performance is due to the synergistic effects of chemically bonded WO3 and GR. The sensitization of WO3 by GR enhanced the visible light absorption properties of WO3/GR. Moreover, the chemical bonding between WO3 and GR reduced the recombination of the photo-generated electron–hole pairs, leading to improved photoconversion efficiency. This simple strategy opens up a new method to design more optimized systems for photodissociating water under visible light. The same process has been extended to the fabrication of BWO on the surface of GR sheets. As is expected, the combination of the functionality of BWO with the unique properties of GR results in improved performance in the production of O2 from water.51
In addition, some other approaches have been developed to construct GR-based photocatalysts, such as layer-by-layer assembly, template-assisted approaches, and photoassisted methods.69–73 Usually, the GR-based photocatalysts are prepared by a combination of the abovementioned methods, rather than by an individual method.
3. Morphology control
As is well known, the morphology of semiconductors on GO plays a key role in their photocatalytic performance. In our previous work, BWO nanoneedles were fabricated on the surface of GR sheets by a facile sonochemical method, followed by calcination.74 The reduction of GO plays an important role in the fabrication of nanoneedles (Fig. 3a and b) instead of nanoparticles on GR sheets. The oxygen-containing groups show a strong influence on the morphology control of BWO on the surface of GR sheets. As compared with GO, RGO sheets have lower quantities of oxygen-containing groups, providing fewer sites for both the physisorption of Bi ions and the deposition of BWO on the sheets, and this leads to the formation of BWO nanoneedles on the sheets after calcination (Fig. 3c). The nanoneedles have a cross-section area of ∼450 nm at the bottom and a length of 2.5 μm, and they form on the surface of the GR sheets and disperse homogeneously. When used as a photocatalyst for oxygen production, the oriented BWO nanoneedles grown on GR sheets produce 188.9 μmol L−1 oxygen, which is higher than that of particulate BWO on GR (164.8 μmol L−1). The high photocatalytic property is mainly attributed to the improved contact area of oriented BWO on GR sheets, resulting in increased active sites.75 Similar results were obtained in photocatalytic systems of TiO2 nanowires/GR and TiO2 particles/GR.
 |
| Fig. 3 (a and b) The morphology of GO-BWO. (c) Proposed mechanism of the formation of BWO-T nanoneedles on GR sheets. (d and e) The photocatalytic O2 (d) and H2 (e) creation activities of Bi2WO6 powder (BWO-T), the physically mixed Bi2WO6 and the reduced GO (BWO-T/GR). Reproduced from ref. 66. | |
It is widely recognized that a photocatalyst with high photocatalytic activity requires both high crystallinity and a large surface area to reduce recombination of the photogenerated electron–hole pairs and to increase the density of active surface sites, as well as to enhance the light harvesting. On the basis of the abovementioned consideration, exploring mesoporous materials as photocatalysts might be a significant subject because they not only possess the merits of both high crystallinity and large surface area, but also contain continuous pore channels to provide a short distance for photogenerated charge carrier transfer within the mesoporous framework. In 2013, Huang et al. reported a novel photocatalytic system by hybridizing GR with a mesoporous semiconductor for solar energy conversion.47 The composite is composed of highly-ordered mesoporous WO3 (m-WO3) and RGO. It was found that the obtained m-WO3/RGO composite had a greater surface area and pore volume than m-WO3. Furthermore, the specific surface area and pore volume increased with increasing amounts of RGO. Under visible light irradiation, the amount of oxygen evolving from the optimized photocatalyst (ca. 6 wt% RGO) reached 437.3 μmol g−1 in 5 h when KIO3 was used as an electron acceptor. In addition to the contribution of RGO, this superior photocatalytic activity could be ascribed to the mesoporous structure of m-WO3, which provides a large surface area and ordered meso-channels, thus more active sites could be achieved and charges could be efficiently transferred. A similar phenomenon was observed in C3N4/GR composites.76
4. Composition design
4.1. Nitrogen doping
GO is a p-doped material, because oxygen atoms are more electronegative than carbon atoms.77 However, GO itself shows very small catalytic activity toward oxygen evolution from water. Doping with heteroatoms, such as nitrogen (N), has been reported to alter the electrical properties, which can induce a charge rearrangement on the GR sheets and enhance their catalytic activity.78 Li et al. reported the synthesis of N-doped GR (GN) by nitrogen plasma treatment of GR, and found that GN exhibited high electrocatalytic activity for the reduction of hydrogen peroxide and fast direct electron transfer kinetics for glucose oxidase.79
More recently, Jing et al. prepared GR doped with different amounts of N through a one-pot ammonia-modified hydrothermal process, and then successfully coupled the doped GR with nanocrystalline α-Fe2O3 by a common wet-chemical method.80 The increased amount of doped quaternary-type N was quite favourable for photogenerated charge transfer and transportation. As a result, the photogenerated charge separation of the resulting GN–Fe2O3 nanocomposite was greatly promoted. This was responsible for the visibly improved activities of α-Fe2O3 in photoelectrochemical water oxidation to produce O2. Interestingly, it was suggested for the first time that an increased amount of doped quaternary-type N would be very favourable for photogenerated charge transfer and transportation and for O2 adsorption, further leading to greatly increased charge separation in the resulting GN–Fe2O3 nanocomposite. On the basis of the abovementioned results, it is reasonable to conclude that the photogenerated charge separation of α-Fe2O3 could be enhanced after coupling with a certain ratio of GR, particularly with that doped with an appropriate amount of N species.81
A similar result was found in the synthesis of GR modified with Fe and N (Fe–N–GR) by a rapid heat treatment process.82 N atoms were doped into the graphene planes and GO was completely reduced during the heat treatment. Fe atoms in the composite are supported by the doped N atoms via coordination bonds. The oxygen reduction reaction for the Fe–N–GR catalyst had an onset-potential of 850 mV vs. RHE (pH = 0.33).
4.2. Multicomponent synergism
A new visible light photocatalyst for O2 evolution from water, Ag3PO4, is unstable upon photo-illumination, which is easily corroded by the photogenerated electrons.26 Compared with bare Ag3PO4, incorporating Ag3PO4 with GO not only improves its photocatalytic activity, but also enhances its stability during the photocatalytic process.83 As is known, plasmon-mediated photocatalysis has recently become a rising research star in the harvesting and conversion of solar energy, because photocatalysts containing semiconductors and plasmonic nanostructures of noble metals can enhance photocatalytic activity primarily by extending the optical absorption region and enhancing the concentration of charge carriers through an excitation of surface plasmon resonance (SPR).84 In 2012, an Ag3PO4/Ag/AgBr/RGO hybrid composite with high visible light photocatalytic O2-production activity was successfully prepared by a photoassisted deposition–precipitation strategy, followed by facile hydrothermal treatment.53 Under irradiation with >420 nm light, the effectiveness of these processes manifests itself in a 1.3 times higher O2 evolution yield of Ag3PO4/Ag/AgBr/RGO as compared with that of Ag3PO4/Ag/AgBr and a 2 times higher yield compared with that of bare Ag3PO4 (Fig. 4a). The photocurrent traces show a rapid response both at the start and at the end of illumination, and an improved photocurrent density of the Ag3PO4/Ag/AgBr/RGO hybrid over all other composites (Fig. 4b). It is considered that the addition of Ag/AgBr caused conduction band depletion and lowered the valence band of Ag3PO4; RGO supports this effect in a synergistic manner through delocalization of the transferred charge. Moreover, RGO provides significant parasitic absorption that partially counters the observed increase in efficiency. The proposed mechanism is shown in Fig. 4c. The as-synthesized N-doped Ag3PO4 nanoparticles are denuded of the majority of charge carriers by transfer to the Ag nanoparticles, eliminating the availability of the extra conduction-band electrons for recombination with the photogenerated holes (resulting in increased hole availability for water oxidation). This leads to the pinning of the Ag3PO4 conduction band at the silver Fermi level, shifting the Ag3PO4 valence band edge downward and rendering the photogenerated holes more active in water oxidation. Charge transfer to the silver creates a substantial negative charge on the very small metal nanoparticles, limiting their beneficial effect on the photocatalyst. Charging of the nanoparticles can be reduced by distribution of the charge onto RGO sheets, further lowering the Ag3PO4 valence band position.
 |
| Fig. 4 The photocatalytic performance of oxygen evolution and the photocatalytic mechanism model. (a) Photocatalytic O2 production under visible light irradiation (λ > 420 nm) from a 0.05 M aqueous AgNO3 solution over bare Ag/AgBr, Ag3PO4, Ag3PO4/RGO, Ag3PO4/Ag/AgBr, and Ag3PO4/Ag/AgBr/RGO (values 0, 38, 43, 48, 76 from bottom to top, unit μmol h−1). (b) Transient photocurrent responses of electrodes functionalized with the Ag3PO4-based materials in the same order (bottom to top) as in panel (a). Measurements proceeded in a 0.01 M Na2SO4 aqueous solution under visible light irradiation (λ > 420 nm, I0 = 64 mW cm−2) at 0.5 V bias vs. SCE. (c) Model of the synergistic increase of photocatalytic activity of Ag3PO4 upon functionalization with Ag/AgBr and RGO. Reprinted with permission from ref. 77. Copyright 2012 American Chemical Society. | |
Further study was reported by Tian for the fabrication of CoPi nanoplates integrated with GO by photochemical deposition from an aqueous solution under visible-light illumination. Compared with GO, the CoPi/GO composites exhibited a 3.6-fold enhancement in the photocurrent. The photocurrent of the CoPi/GO-modified electrode after 2 h of illumination remained similar without any significant sign of catalyst dissolution or degradation. The GO not only serves as a substrate for CoPi growth, but also increases charge transfer in the composites.73
In addition, some metal oxides have been employed as co-catalysts to build novel semiconductor composites for photocatalysis, such as hausmannite (Mn3O4), a mixed valence manganese oxide. In 2014, Yin et al. reported the fabrication of a nanocomposite consisting of α-Fe2O3, Mn3O4 and RGO for photocatalytic water oxidation to produce oxygen.85 The α-Fe2O3/Mn3O4 hybrid prepared by modifying α-Fe2O3 with Mn3O4 nanoparticles was proved to anchor well on the RGO sheets of the nanocomposite, resulting in superior interfacial contacts between the hybrid and RGO. The typical p-type semiconductor Mn3O4 in the nanocomposite forms a heterojunction with α-Fe2O3, enhancing the charge transfer. In addition, Mn3O4 also acts as a noble-metal free co-catalyst, reducing the oxygen evolution overpotential of hematite. RGO in the nanocomposite serves not only as a superior supporting matrix for anchoring the semiconductor nanoparticles, but also as an excellent electron mediator to adjust the electron transfer. As a result, the nanocomposite exhibited remarkably enhanced photocatalytic activity toward water oxidation compared with bare hematite or α-Fe2O3/Mn3O4 under UV-vis light irradiation. The quantum efficiency of the optimized photocatalyst reached up to 4.35% at 365 nm.
4.3. Core/shell heterojunction structure
In addition to N-doping and multicomponent synergism systems, core/shell geometry is another intriguing architecture that can improve the efficiency of energy conversion. With α-Fe2O3 nanorod cores, GR interlayers, and BiV1−xMoxO4 shells, a core/shell heterojunction array was fabricated by Hou et al. for photoelectrochemical water splitting,86 as shown in Fig. 5a. The heterojunction yielded a pronounced photocurrent density of ∼1.97 mA cm−2 at 1.0 V vs. Ag/AgCl, and a high photoconversion efficiency of ∼0.53% at −0.04 V vs. Ag/AgCl under the irradiation of a Xe lamp (Fig. 5b). The unique core/shell architecture enhances the light absorption due to the “window effect”87 behaviour between the α-Fe2O3 core and BiV1−xMoxO4 shell, and improves the separation of photogenerated carriers at the α-Fe2O3/GR/BiV1−xMoxO4 interface.
 |
| Fig. 5 (a) Proposed mechanism of photoelectrochemical water splitting; (b) variation of photocurrent density vs. bias potential (left), and photoconversion efficiency as a function of applied potential (right). Reprinted with permission from ref. 68. Copyright 2012 American Chemical Society. | |
5. Interface between semiconductors and graphene
It has been reported that the transfer of photoexcited electrons from TiO2 to nanocarbons, such as carbon nanotubes or GR, hinders the recombination process, thereby enhancing the oxidative reactivity.88,89 For this reason, GO–TiO2 composites exhibit excellent photochemical responses under visible light (>400 nm) irradiation. It was found that the unpaired p electrons on GO can bond with the surface Ti atoms of TiO2 to form Ti–O–C bonds and extend the light absorption range of TiO2.90–92 Further studies indicate that the interaction between GR and TiO2 can significantly affect the interfacial electron transfer properties, which is a key issue for photocatalytic activity.
A similar phenomenon has been observed in a Co3O4/N-doped RGO (Co3O4/N-RGO) composite synthesized with chemical bonding via a facile hydrothermal method.93 X-ray absorption near edge structure (XANES) measurements were conducted to explore the interaction between GO and Co3O4. It was found that the Co3O4/N-RGO hybrid exhibited an increase of carbon K-edge peak intensity at ∼288 eV compared to pure N-RGO, corresponding to carbon atoms in GR attached to oxygen or other species. This result indicated the possible existence of Co–O–C or Co–N–C in the hybrid. Oxygen K-edge XANES and Co L-edge XANES measurements also suggest the same chemical bonding between the two components. Liang's report demonstrated that chemical bonding between GR and a semiconductor could be achieved through a hydrothermal synthesis, which is relatively simple. Chemical bonding in GR-based photocatalysts could also be achieved through a simple sonochemical method, demonstrated by our previous fabrication of WO3/GR used for photocatalytic oxygen evolution from water.51 The bonding between WO3 and GR minimizes the interface defects, reducing the recombination of photo-generated charges, responsible for the enhanced O2 evolution.
Because the bridge facilitates charge transfer and can be engineered through relatively simple methods, including hydrothermal or sonochemical synthesis, this kind of chemical bonding or other intimate interface is generally being considered for GR-based photocatalyst architectures.
6. Simultaneous H2 and O2 evolution from water using graphene-based materials
As known, semiconductors with large band gaps or improper band location cannot produce O2 or H2 effectively. In fact, semiconductor photocatalysts possessing suitable redox potentials and band gaps for both simultaneous water reduction and oxidation are rare. There are many reports about O2 production from half reaction systems containing sacrificial reagents, which are generally considered to not accurately reflect their photocatalytic ability. In this part, GR-based photocatalytic systems that simultaneously produce O2 and H2 are introduced in detail.
6.1. Graphene oxide photocatalytic systems
The electronic properties of GO are related to the composition of oxygen bonding on the GR sheets.94 The high electronegativity of the oxygen atoms on carbon sheets causes the charge flow that exerts p-type semiconductivity to GO.95,96 As oxygen bonds to GR, the valence band changes from the π-orbital of GR to the O 2p orbital, leading to a larger band gap for a higher oxidation level of GO. Introducing more oxygen enlarges the band gap, and the valence band maximum (VBM) gradually changes from the p orbital of GR to the 2p orbital of oxygen; the p* orbital remains as the conduction band minimum (CBM). Yeh et al. researched the electronic band energy levels of GO specimens with various oxidation levels using electrochemical methods.97 The results reflect that with sufficient oxidation, the electronic structure of GO is suitable for both the reduction and oxidation of water under illumination, and the production of H2 and O2 gases in the presence of sacrificial reagents. They found that the downward shift in the valence band edge was predominantly responsible for the enlargement of the band gap in the GO sheets. During the photocatalytic reaction, the mutual reduction between GO sheets narrows the band gap, leading to an activity decay of GO in catalysing O2 evolution from an AgNO3 solution because of the upward shift of the valence band edge, whereas the activity for H2 evolution from a methanol solution remains unchanged. Strong evolution of O2 from a NaIO3 solution under illumination was observed, probably due to a more effective GO dispersion to suppress mutual reduction.
6.2. Z-Scheme system
Further studies have found that the combination of GO with a Z-scheme system makes it an effective mediator to produce H2 and O2 via overall water splitting.52 GO was used as a solid-state redox mediator, and BiVO4 was used as an O2-generating photocatalyst. The GO–BiVO4 composite was mixed with a H2-generating Ru–SrTiO3–Rh photocatalyst in an aqueous solution with a pH value of 3.5 to enable inter-particulate contact. Under visible light irradiation, the photogenerated electrons transferred from the conduction band of BiVO4 to GO for accumulation. When Ru–SrTiO3–Rh came in contact with GO–BiVO4, the accumulated electrons on GO transferred to Ru–SrTiO3–Rh and combined with the photogenerated holes. The remaining electrons on Ru–SrTiO3–Rh and the holes on BiVO4 subsequently reacted with water, forming H2 and O2, respectively. The evolution of H2 and O2 in a stoichiometric ratio without decay indicated the occurrence of Z-scheme overall water splitting. The minimum turnover number (TON), which was calculated as the number of moles of reactive electrons per mole of GO, was 3.2 over 24 h (Fig. 6a). The result suggested that photoreduced GO (PRGO) is stable as a solid electron mediator, and the Z-scheme system splits water photocatalytically. The mechanism of water splitting in a Z-scheme photocatalysis system consisting of Ru–SrTiO3–Rh and PRGO/BiVO4 is shown in Fig. 6b.
 |
| Fig. 6 (a) Overall water splitting under visible-light irradiation by the (Ru–SrTiO3–Rh)–(PRGO/BiVO4) system. (b) Schematic image of a suspension of Ru–SrTiO3 and PRGO/BiVO4 in water (top); mechanism of water splitting in a Z-scheme photocatalysis system consisting of Ru–SrTiO3–Rh and PRGO/BiVO4 under visible-light irradiation (down). Reprinted with permission from ref. 89. Copyright 2011 American Chemical Society. | |
6.3. Graphene oxide quantum dots
Compared with 2D GR, graphene quantum dots (GQDs), a type of C-dot, exhibit new phenomena due to quantum confinement and edge effects.98 GQDs are advantageous because their band gaps can be tuned from 0 eV to that of benzene by varying their sizes.99 The incorporation of zero dimension GQDs can extend the photo-response of a photocatalyst to the visible-light range.100 The combination of GQDs with photocatalysts, such as CdS-modified TiO2 nanotube arrays (TNAs) and ZnO nanowires, has been reported for photoelectrochemical water splitting.101,102 Very recently, Yeh et al. reported N-doped GO-quantum dots (NGO-QDs) as photocatalysts for overall water-splitting under visible light illumination.103 The NGO-QDs exhibited both p- and n-type conductivities. Visible light (>420 nm) illumination of the NGO-QDs resulted in simultaneous H2 and O2 production from pure water at an H2
:
O2 molar ratio of 2
:
1 (Fig. 7a). Fig. 7b shows the time courses of H2 production over the GO-QD photocatalysts, and no O2 evolution was observed. Moreover, the evolution of O2 over NH3–NGO-QDs also cannot observe H2 evolution (Fig. 7c). The results shown in Fig. 7 (b and c) proved that the p- and n-domains were responsible for the evolution of H2 and O2, respectively. The authors suggested a p–n type photochemical diode configuration (Fig. 7d), which resulted in an energetic band bending existing at the interface between the semiconductor and the solution. The p–n type photochemical diode configuration provided a favourable situation to achieve vectorial charge displacement for overall water-splitting.
 |
| Fig. 7 (a) Time courses of H2 and O2 evolution over 1.2 g NGO-QDs suspended in 200 mL of pure water under visible-light (420 nm < λ < 800 nm) irradiation. (b)Time courses of H2 evolution over 1.2 g GO-QDs under the same conditions as (a). (c) Time courses of O2 evolution over 1.2 g NH3–NGO-QDs under the same conditions as (a). (d) The configuration and energy diagram for the NGO-QD photochemical diode. Reprinted with permission from ref. 95. Copyright 2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. | |
7. Summary and perspective
In summary, GR-based semiconductor photocatalysts for oxygen evolution from water are attracting considerable attention, as they are crucial for renewable energy technologies. Considerable approaches have been developed for the fabrication of these photocatalysts. The morphologies of the semiconductors on GO greatly influence their photocatalytic performance. Composition design is an effective method to enhance the photocatalytic properties, including nitrogen doping, multicomponent synergism systems and core/shell heterojunction structure. With appropriate tuning of the CBM and VBM positions, these composites are suitable for H2 and O2 generation.
There are several challenges facing research on GO-based photocatalytic O2 evolution from water. First, theoretical electronic-structure calculations and experimental identification efforts are required. The dependence of the photocatalytic properties on the particle size and layer number of GO should be explored, as they determine the charge transport mechanism in the GR-based composites.
Second, exploration and identification of the interfacial contact and bonding between RGO and the semiconductors in the composites is needed.104 The interaction of GR and the semiconductor affects the morphology of the loaded particles, which exerts a strong influence on the photocatalytic performance of the resultant materials. On the other hand, the interface determines the efficiency of the electron–hole separation. To date, only a few techniques have succeeded in directly characterizing the interaction of GR and nanoparticles. Atomic resolution scanning transmission electron microscopy (STEM) and surface-enhanced Raman scattering (SERS) spectra may be the most potential techniques for determining the interaction of GR with nanoparticles, although these techniques only have been applied in metal–GR systems.105,106
Third, studies on the preparation of a ternary hybrid as a photocatalyst for visible-light-driven water oxidation have been seldom reported,107,108 particularly, for the design of ternary hybrids with non-metal oxide, such as polymers. It has been reported that PANI–GR–TiO2, a ternary hybrid, shows remarkable photocatalytic activity and photostability for visible-light photoelectrocatalytic water oxidation.108 This rationally designed ternary hybrid possesses unique advantages over traditional photocatalysts towards water oxidation.
The final challenge is to design a structure for overall water splitting with simultaneous evolution of H2 and O2. Achieving this goal may require further exploiting GO by chemically modifying it; for example, the development of new nanostructured GO composites that efficiently transport charge inside the composites and inject charges for reactions at the water–composite interface. Ultimately, a cheap, reliable and highly efficient photocatalyst for oxygen evolution still needs to be developed, considering its critical role in the development of high performance energy conversion and storage devices.
Acknowledgements
The authors gratefully acknowledge the financial support of the National Science Foundation of China (no. 51072117) and Shanghai Science and Technology Committee (no. 13JC1403300). We also thank the Shanghai Jiao Tong University (SJTU) Instrument Analysis Center for the measurements.
Notes and references
- Z. G. Zou, J. H. Ye, K. Sayama and H. Arakawa, Nature, 2001, 414, 625–627 CrossRef CAS PubMed.
- L. Z. Wu, B. Chen, Z. J. Li and C. H. Tung, Acc. Chem. Res., 2014, 47, 2177–2185 CrossRef CAS PubMed.
- F. Y. Song, Y. Ding and C. C. Zhao, Acta Chim. Sin., 2014, 72, 133–144 CrossRef CAS.
- P. Zhang, J. J. Zhang and J. L. Gong, Chem. Soc. Rev., 2014, 43, 4395–4422 RSC.
- M. Matsuoka, M. Kitano, M. Takeuchi, K. Tsujimaru, M. Anpo and J. M. Thomas, Catal. Today, 2007, 122, 51–61 CrossRef CAS PubMed.
- K. H. A. Fujishima, Nature, 1972, 238, 37–38 CrossRef.
- A. Kudo and Y. Miseki, Chem. Soc. Rev., 2009, 38, 253–278 RSC.
- J. Zhao, X. Wang, Z. C. Xu and J. S. C. Loo, J. Mater. Chem. A, 2014, 2, 15228–15233 CAS.
- X. Q. An and J. C. Yu, RSC Adv., 2011, 1, 1426–1434 RSC.
- J. Yu, Y. Hai and B. Cheng, J. Phys. Chem. C, 2011, 115, 4953–4958 CAS.
- Q. J. Xiang, J. G. Yu and M. Jaroniec, J. Am. Chem. Soc., 2012, 134, 6575–6578 CrossRef CAS PubMed.
- K. Maeda, K. Teramura, D. L. Lu, T. Takata, N. Saito, Y. Inoue and K. Domen, Nature, 2006, 440, 295 CrossRef CAS PubMed.
- K. Maeda, M. Higashi, B. Siritanaratkul, R. Abe and K. Domen, J. Am. Chem. Soc., 2011, 133, 12334–12337 CrossRef CAS PubMed.
- K. Maeda, A. K. Xiong, T. Yoshinaga, T. Ikeda, N. Sakamoto, T. Hisatomi, M. Takashima, D. L. Lu, M. Kanehara, T. Setoyama, T. Teranishi and K. Domen, Angew. Chem., Int. Ed., 2010, 49, 4096–4099 CrossRef CAS PubMed.
- X. B. Chen, S. H. Shen, L. J. Guo and S. S. Mao, Chem. Rev., 2010, 110, 6503–6570 CrossRef CAS PubMed.
- K. Maeda, M. Higashi, D. L. Lu, R. Abe and K. Domen, J. Am. Chem. Soc., 2010, 132, 5858–5868 CrossRef CAS PubMed.
- M. W. Kanan and D. G. Nocera, Science, 2008, 321, 1072–1075 CrossRef CAS PubMed.
- M. H. V. Huynh and T. J. Meyer, Chem. Rev., 2007, 107, 5004–5064 CrossRef CAS PubMed.
- J. X. Yang, D. G. Wang, X. Zhou and C. Li, Chem.–Eur. J., 2013, 19, 1320–1326 CrossRef CAS PubMed.
- J. A. Seabold and K. S. Choi, Chem. Mater., 2011, 23, 1105–1112 CrossRef CAS.
- J. A. Seabold and K. S. Choi, J. Am. Chem. Soc., 2012, 134, 2186–2192 CrossRef CAS PubMed.
- H. M. Liu, R. Nakamura and Y. Nakato, ChemPhysChem, 2005, 6, 2499–2502 CrossRef CAS PubMed.
- H. M. Liu, R. Nakamura and Y. Nakato, Electrochem. Solid-State Lett., 2006, 9, G187–G190 CrossRef CAS PubMed.
- K. Maeda, B. J. Lee, D. L. Lu and K. Domen, Chem. Mater., 2009, 21, 2286–2291 CrossRef CAS.
- X. Zong, C. H. Sun, Z. G. Chen, A. Mukherji, H. Wu, J. Zou, S. C. Smith, G. Q. Lu and L. Z. Wang, Chem. Commun., 2011, 47, 6293–6295 RSC.
- Z. G. Yi, J. H. Ye, N. Kikugawa, T. Kako, S. X. Ouyang, H. Stuart-Williams, H. Yang, J. Y. Cao, W. J. Luo, Z. S. Li, Y. Liu and R. L. Withers, Nat. Mater., 2010, 9, 559–564 CrossRef CAS PubMed.
- K. Maeda, Y. Shimodaira, B. Lee, K. Teramura, D. Lu, H. Kobayashi and K. Domen, J. Phys. Chem. C, 2007, 111, 18264–18270 CAS.
- A. A. Gewirth and M. S. Thorum, Inorg. Chem., 2010, 49, 3557–3566 CrossRef CAS PubMed.
- J. L. Fillol, Z. Codola, I. Garcia-Bosch, L. Gomez, J. J. Pla and M. Costas, Nat. Chem., 2011, 3, 807–813 CrossRef CAS PubMed.
- K. Chang, Z. W. Mei, T. Wang, Q. Kang, S. X. Ouyang and J. H. Ye, ACS Nano, 2014, 8, 7078–7087 CrossRef CAS PubMed.
- R. Marschall, Adv. Funct. Mater., 2014, 24, 2421–2440 CrossRef CAS.
- D. M. Schultz and T. P. Yoon, Science, 2014, 343, 1239176 CrossRef PubMed.
- R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki and Y. Taga, Science, 2001, 293, 269–271 CrossRef CAS PubMed.
- K. R. Lai, Y. T. Zhu, J. B. Lu, Y. Dai and B. B. Huang, Comput. Mater. Sci., 2013, 67, 88–92 CrossRef CAS PubMed.
- J. Q. Yu and A. Kudo, Chem. Lett., 2005, 34, 1528–1529 CrossRef CAS.
- H. M. Liu, R. Nakamura and Y. Nakato, J. Electrochem. Soc., 2005, 152, G856–G861 CrossRef PubMed.
- G. W. Busser, B. Mei, A. Pougin, J. Strunk, R. Gutkowski, W. Schuhmann, M. G. Willinger, R. Schlogl and M. Muhler, ChemSusChem, 2014, 7, 1030–1034 CrossRef CAS PubMed.
- Y. X. Yang, W. Guo, Y. N. Guo, Y. H. Zhao, X. Yuan and Y. H. Guo, J. Hazard. Mater., 2014, 271, 150–159 CrossRef CAS PubMed.
- M. Miyauchi, Y. Nukui, D. Atarashi and E. Sakai, ACS Appl. Mater. Interfaces, 2013, 5, 9770–9776 CAS.
- H. J. Yu, Y. F. Zhao, C. Zhou, L. Shang, Y. Peng, Y. H. Cao, L. Z. Wu, C. H. Tung and T. R. Zhang, J. Mater. Chem. A, 2014, 2, 3344–3351 CAS.
- Q. J. Xiang, J. G. Yu and M. Jaroniec, Chem. Soc. Rev., 2012, 41, 782–796 RSC.
- C. Lee, X. D. Wei, J. W. Kysar and J. Hone, Science, 2008, 321, 385–388 CrossRef CAS PubMed.
- A. A. Balandin, S. Ghosh, W. Z. Bao, I. Calizo, D. Teweldebrhan, F. Miao and C. N. Lau, Nano Lett., 2008, 8, 902–907 CrossRef CAS PubMed.
- Y. W. Zhu, S. Murali, W. W. Cai, X. S. Li, J. W. Suk, J. R. Potts and R. S. Ruoff, Adv. Mater., 2010, 22, 3906–3924 CrossRef CAS PubMed.
- L. Mao, S. M. Zhu, J. Ma, D. A. Shi, Y. X. Chen, Z. X. Chen, C. Yin, Y. Li and D. Zhang, Nanotechnology, 2014, 25, 215401 CrossRef PubMed.
- M. J. Zhou, N. Zhang and Z. H. Hou, Int. J. Photoenergy, 2014, 943537 Search PubMed.
- H. Huang, Z. K. Yue, G. Li, X. M. Wang, J. Huang, Y. K. Du and P. Yang, J. Mater. Chem. A, 2013, 1, 15110–15116 CAS.
- Z. H. Sun, J. J. Guo, S. M. Zhu, L. Mao, J. Ma and D. Zhang, Nanoscale, 2014, 6, 2186–2193 RSC.
- F. K. Meng, J. T. Li, S. K. Cushing, J. Bright, M. J. Zhi, J. D. Rowley, Z. L. Hong, A. Manivannan, A. D. Bristow and N. Q. Wu, ACS Catal., 2013, 3, 746–751 CrossRef CAS.
- B. Li, Y. F. Zhao, S. T. Zhang, W. Gao and M. Wei, ACS Appl. Mater. Interfaces, 2013, 5, 10233–10239 CAS.
- 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.
- A. Iwase, Y. H. Ng, Y. Ishiguro, A. Kudo and R. Amal, J. Am. Chem. Soc., 2011, 133, 11054–11057 CrossRef CAS PubMed.
- Y. Hou, F. Zuo, Q. Ma, C. Wang, L. Bartels and P. Y. Feng, J. Phys. Chem. C, 2012, 116, 20132–20139 CAS.
- R. A. Rakkesh, D. Durgalakshmi and S. Balakumar, J. Mater. Chem. C, 2014, 2, 6827–6834 RSC.
- S. Bai, X. P. Shen, H. W. Lv, G. X. Zhu, C. L. L. Bao and Y. X. Shan, J. Colloid Interface Sci., 2013, 405, 1–9 CrossRef CAS PubMed.
- Q. Li, B. D. Guo, J. G. Yu, J. R. Ran, B. H. Zhang, H. J. Yan and J. R. Gong, J. Am. Chem. Soc., 2011, 133, 10878–10884 CrossRef CAS PubMed.
- J. F. Dai, T. Xian, L. J. Di and H. Yang, J. Nanomater., 2013, 642897 Search PubMed.
- Y. H. Ng, A. Iwase, N. J. Bell, A. Kudo and R. Amal, Catal. Today, 2011, 164, 353–357 CrossRef CAS PubMed.
- J. D. Lin, P. Hu, Y. Zhang, M. T. Fan, Z. M. He, C. K. Ngaw, J. S. C. Loo, D. W. Liao and T. T. Y. Tan, RSC Adv., 2013, 3, 9330–9336 RSC.
- X. F. Yang, H. Y. Cui, Y. Li, J. L. Qin, R. X. Zhang and H. Tang, ACS Catal., 2013, 3, 363–369 CrossRef CAS.
- S. Y. Dong, Y. R. Cui, Y. F. Wang, Y. K. Li, L. M. Hu, J. Y. Sun and J. H. Sun, Chem. Eng. J., 2014, 249, 102–110 CrossRef CAS PubMed.
- Y. Hou, X. Y. Li, Q. D. Zhao and G. H. Chen, Appl. Catal., B, 2013, 142, 80–88 CrossRef PubMed.
- Z. G. Niu, Mater. Sci. Semicond. Process., 2014, 23, 78–84 CrossRef CAS PubMed.
- P. T. N. Nguyen, C. Salim, W. Kurniawan and H. Hinode, Catal. Today, 2014, 230, 166–173 CrossRef CAS PubMed.
- J. J. Xu, Y. H. Ao and M. D. Chen, Mater. Lett., 2013, 92, 126–128 CrossRef CAS PubMed.
- S. M. Zhu, J. J. Guo, J. P. Dong, Z. W. Cui, T. Lu, C. L. Zhu, D. Zhang and J. Ma, Ultrason. Sonochem., 2013, 20, 872–880 CrossRef CAS PubMed.
- S. M. Zhu, C. L. Zhu, J. Ma, Q. Meng, Z. P. Guo, Z. Y. Yu, T. Lu, Y. Li, D. Zhang and W. M. Lau, RSC Adv., 2013, 3, 6141–6146 RSC.
- J. J. Guo, S. M. Zhu, Z. X. Chen, Y. Li, Z. Y. Yu, Q. L. Liu, J. B. Li, C. L. Feng and D. Zhang, Ultrason. Sonochem., 2011, 18, 1082–1090 CrossRef CAS PubMed.
- B. H. R. Suryanto, X. Y. Lu and C. Zhao, J. Mater. Chem. A, 2013, 1, 12726–12731 CAS.
- S. Y. Dong, Y. K. Li, J. Y. Sun, C. F. Yu, Y. H. Li and J. H. Sun, Mater. Chem. Phys., 2014, 145, 357–365 CrossRef CAS PubMed.
- C. Lavorato, A. Primo, R. Molinari and H. Garcia, ACS Catal., 2014, 4, 497–504 CrossRef CAS.
- W. Yan, F. He, S. L. Gai, P. Gao, Y. J. Chen and P. P. Yang, J. Mater. Chem. A, 2014, 2, 3605–3612 CAS.
- J. Q. Tian, H. Y. Li, A. M. Asiri, A. O. Al-Youbi and X. P. Sun, Small, 2013, 9, 2709–2714 CrossRef CAS PubMed.
- Z. H. Sun, J. J. Guo, S. M. Zhu, J. Ma, Y. L. Liao and D. Zhang, RSC Adv., 2014, 4, 27963–27970 RSC.
- X. Pan, Y. Zhao, S. Liu, C. L. Korzeniewski, S. Wang and Z. Y. Fan, ACS Appl. Mater. Interfaces, 2012, 4, 3944–3950 CAS.
- Q. J. Xiang, J. G. Yu and M. Jaroniec, J. Phys. Chem. C, 2011, 115, 7355–7363 CAS.
- X. R. Wang, X. L. Li, L. Zhang, Y. Yoon, P. K. Weber, H. L. Wang, J. Guo and H. J. Dai, Science, 2009, 324, 768–771 CrossRef CAS PubMed.
- L. Ci, L. Song, C. H. Jin, D. Jariwala, D. X. Wu, Y. J. Li, A. Srivastava, Z. F. Wang, K. Storr, L. Balicas, F. Liu and P. M. Ajayan, Nat. Mater., 2010, 9, 430–435 CrossRef CAS PubMed.
- Y. Wang, Y. Y. Shao, D. W. Matson, J. H. Li and Y. H. Lin, ACS Nano, 2010, 4, 1790–1798 CrossRef CAS PubMed.
- L. M. He, L. Q. Jing, Y. B. Luan, L. Wang and H. G. Fu, ACS Catal., 2014, 4, 990–998 CrossRef CAS.
- Y. J. Wang, R. Shi, J. Lin and Y. F. Zhu, Appl. Catal., B, 2010, 100, 179–183 CrossRef CAS PubMed.
- K. Kamiya, K. Hashimoto and S. Nakanishi, Chem. Commun., 2012, 48, 10213–10215 RSC.
- L. Liu, J. C. Liu and D. D. Sun, Catal. Sci. Technol., 2012, 2, 2525–2532 CAS.
- S. Linic, P. Christopher and D. B. Ingram, Nat. Mater., 2011, 10, 911–921 CrossRef CAS PubMed.
- S. L. Yin, X. M. Wang, Z. G. Mou, Y. J. Wu, H. Huang, M. S. Zhu, Y. K. Du and P. Yang, Phys. Chem. Chem. Phys., 2014, 16, 11289–11296 RSC.
- Y. Hou, F. Zuo, A. Dagg and P. Y. Feng, Nano Lett., 2012, 12, 6464–6473 CrossRef CAS PubMed.
- H. T. Yu, S. Chen, X. Quan, H. M. Zhao and Y. B. Zhang, Appl. Catal., B, 2009, 90, 242–248 CrossRef CAS PubMed.
- G. X. Wang, B. Wang, J. Park, J. Yang, X. P. Shen and J. Yao, Carbon, 2009, 47, 68–72 CrossRef CAS PubMed.
- G. Williams, B. Seger and P. V. Kamat, ACS Nano, 2008, 2, 1487–1491 CrossRef CAS PubMed.
- H. Zhang, X. J. Lv, Y. M. Li, Y. Wang and J. H. Li, ACS Nano, 2010, 4, 380–386 CrossRef CAS PubMed.
- I. Y. Kim, J. M. Lee, T. W. Kim, H. N. Kim, H. I. Kim, W. Choi and S. J. Hwang, Small, 2012, 8, 1038–1048 CrossRef CAS PubMed.
- J. S. Lee, K. H. You and C. B. Park, Adv. Mater., 2012, 24, 1084–1088 CrossRef CAS PubMed.
- Y. Y. Liang, Y. G. Li, H. L. Wang, J. G. Zhou, J. Wang, T. Regier and H. J. Dai, Nat. Mater., 2011, 10, 780–786 CrossRef CAS PubMed.
- R. J. W. E. Lahaye, H. K. Jeong, C. Y. Park and Y. H. Lee, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79, 125435 CrossRef.
- J. Ito, J. Nakamura and A. Natori, J. Appl. Phys., 2008, 103, 113712 CrossRef PubMed.
- G. H. Lu, L. E. Ocola and J. H. Chen, Nanotechnology, 2009, 20, 445502 CrossRef PubMed.
- T. F. Yeh, F. F. Chan, C. T. Hsieh and H. S. Teng, J. Phys. Chem. C, 2011, 115, 22587–22597 CAS.
- J. H. Shen, Y. H. Zhu, X. L. Yang and C. Z. Li, Chem. Commun., 2012, 48, 3686–3699 RSC.
- Y. W. Son, M. L. Cohen and S. G. Louie, Phys. Rev. Lett., 2006, 97, 216803 CrossRef.
- D. K. L. Chan, P. L. Cheung and J. C. Yu, Beilstein J. Nanotechnol., 2014, 5, 689–695 CrossRef PubMed.
- Y. F. Yu, J. L. Ren and M. Meng, Int. J. Hydrogen Energy, 2013, 38, 12266–12272 CrossRef CAS PubMed.
- C. X. Guo, Y. Q. Dong, H. B. Yang and C. M. Li, Adv. Energy Mater., 2013, 3, 997–1003 CrossRef CAS.
- T. F. Yeh, C. Y. Teng, S. J. Chen and H. S. Teng, Adv. Mater., 2014, 26, 3297–3303 CrossRef CAS PubMed.
- E. Gracia-Espino, G. Z. Hu, A. Shchukarev and T. Wagberg, J. Am. Chem. Soc., 2014, 136, 6626–6633 CrossRef CAS PubMed.
- J. Lee, K. S. Novoselov and H. S. Shin, ACS Nano, 2011, 5, 608–612 CrossRef CAS PubMed.
- R. Zan, U. Bangert, Q. Ramasse and K. S. Novoselov, Nano Lett., 2011, 11, 1087–1092 CrossRef CAS PubMed.
- W. J. Han, L. Ren, X. Qi, Y. D. Liu, X. L. Wei, Z. Y. Huang and J. X. Zhong, Appl. Surf. Sci., 2014, 299, 12–18 CrossRef CAS PubMed.
- L. Jing, Z. Y. Yang, Y. F. Zhao, Y. X. Zhang, X. Guo, Y. M. Yan and K. N. Sun, J. Mater. Chem. A, 2014, 2, 1068–1075 CAS.
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.