Photoactivity and electronic properties of graphene-like materials and TiO2 composites using first-principles calculations

Hongtao Gao*a, Baichuan Lua, Dongyang Lia, Fengjuan Guoa, Dongmei Daia, Chongdian Sib, Guangjun Liu*b and Xian Zhaoc
aState Key Laboratory Base of Eco-chemical Engineering, College of Chemistry and Molecular Engineering, Qingdao University of Science & Technology, Qingdao, 266042, P. R. China. E-mail: gaohtao@126.com; Tel: +86-0532-84022681
bDepartment of Chemistry and Chemical Engineering, Jining University, Qufu 273155, China. E-mail: liugjun@126.com
cState Key Laboratory of Crystal Materials, Shandong University, Jinan, 250100, China

Received 19th May 2016 , Accepted 25th June 2016

First published on 28th June 2016


Abstract

Graphene-like materials (GLM) have attracted intense attention in the research on photocatalytic technology due to their superior properties. First-principles calculations based on DFT were used to explore the photoactivity and electronic properties of graphene-like materials and TiO2 composites in this work. It was found that TiO2/GLM composites might be demonstrated for high thermodynamic stability. The electronic properties of TiO2/GLM, such as geometric structure, density of states, charge density, charge density difference and optical properties, were characterized. There were interactions between GLM sheets and TiO2, which caused charge accumulation on the GLM surface and charge depletion on the other side of TiO2 in the heterojunction. Electrons in the highest occupied molecular orbital (HOMO) consisted of the O2p orbital from TiO2 could be directly excited or dispersed to the lowest unoccupied molecular orbital (LUMO) composed of the hybridized orbital from GLM and TiO2 under irradiation. The produced well-separated electron–hole pairs induced an enhanced photocatalytic performance of TiO2/GLM. The theoretical results pointed out the electron migration and transfer path at the interface, which might illustrate the mechanism of enhanced photocatalytic activity of TiO2/GLM.


1. Introduction

The efficient and cost-effective direct conversion of solar energy to chemical energy and solar fuels has been considered as one of the most perceptive strategies to solve the environmental and energy problems in the future.1 Photocatalytical technology has been recognized as an effective means to environmental contaminant control and remediation.2,3 Due to its chemical stability, low cost and strong oxidizing characteristics, TiO2 has been widely used as a photocatalyst over the past several decades.4,5 But unfortunately, TiO2 can utilize no more than 5% of the total solar energy due to its wide band gap (3.0–3.2 eV), which limits its application in the industrial application of photocatalytic technology. Many strategies, such as doping with elements, surface modification and coupling with other semiconductors, have been made to improve its photocatalytic performance.

Owing to excellent conductivity, superior chemical stability and high specific surface area, graphene has attracted a great deal of attention in recent years,6,7 which is valuable as a hot substrate that supports the heterogeneous growth of desired active guest materials.8 It seems that graphene can serve as a suitable candidate to be combined with TiO2 and shows enhanced photocatalytic performance compared to bare TiO2.9 Several reports have been devoted to the preparation of TiO2 modified with graphene for the photodegradation of organic dyes.10,11 For example, TiO2/grapheme (hereinafter referred to TiO2/G) composite produced by direct growth of (TiO2)3 nanocluster (hereinafter referred to TiO2 cluster or TiO2) on graphene oxide sheets showed a strong photocatalytic activity for the degradation of rhodamine B under UV irradiation.12 The addition of graphene nanoplatelets to TiO2 has been recently considered as a method to improve the photocatalytic efficiency of TiO2 by favoring charge carrier separation.13 The conjugation of TiO2 with graphene has been proved to be an effective approach to improve the photocatalytic properties of TiO2.7,9,14

There are various two-dimensional (2D) layered compounds with similar properties like graphene, which can be called graphene-like materials (hereinafter referred to GLM, including graphene), such as, graphitic carbon nitride, black phosphorus and transition metal dichalcogenides, might be used to combine with TiO2 to enhance photocatalytic activity.14 Two-dimensional (2D) graphitic carbon nitride (g-C3N4) has attracted immense attention due to its unique electronic band structure, visible-light-responsiveness, high thermal and chemical stability, making it potentially suitable for solar energy conversion and environmental remediation.15 Despite the aforementioned benefits of g-C3N4 material and photocatalytic process, the photocatalytic efficiency is still low to make the system viable for practical applications owing to the high recombination rate of photogenerated charge carriers in the individual semiconductor.16 Sheet-like TiO2/g-C3N4 heterojunctions exhibited excellent stability and good visible-light photocatalytic performance for photodegradation of rhodamine B.17 The enhanced photocatalytic activities of TiO2/g-C3N4 photocatalysts could be primarily attributed to the formation of a hybrid structure in the contact interface between TiO2 and g-C3N4, which greatly promoted the separation of photogenerated electron–hole charge carriers.18 The in situ growth of TiO2 on the surface of g-C3N4 not only formed a TiO2/g-C3N4 nanojunction with a more efficient interparticle electron transfer but also promoted further protonation, which increased the photocatalytic activities of TiO2/g-C3N4.19 Molybdenum disulfide (MoS2) has been commonly studied as a noble metal free low-cost and abundant cocatalyst which is comparative and even superior to noble metal cocatalyst for enhancing the photocatalytic H2 evolution efficiency of the semiconductor.20 Q. Xiang et al. reported a new composite material consisting of TiO2 cluster grown in the presence of a layered MoS2/graphene hybrid as a high-performance photocatalyst for H2 evolution.21 Both MoS2 and g-C3N4 played positive role for enhanced photocatalytic H2 production activity of TiO2.22 Moreover, black phosphorus (BP) is a layered semiconductor and has great potential applications in optical, electronic and nano-materials.23,24

As described above, there have been some experimental investigations on GLM layered nanomaterials, which improved the photocatalytic activity of TiO2 during the past decades. However, the mechanism on enhanced photocatalytic activity of TiO2/GLM has not been understood clearly. Meanwhile, there are few comprehensive theoretical studies reported on structural and electronic properties of the composites, which might resolve the photocatalytic mechanism of TiO2/GLM.

Herein, first-principles calculations based on density functional theory (DFT) were used to explore the effect of GLM hybridization on the geometry structure and electronic properties of TiO2 in this work. The hybridization of TiO2 with GLM was proved to be susceptible thermodynamically. The theoretical investigation illustrated charge accumulated on the GLM surface and depleted on the side of TiO2 cluster, indicating a probable charge migration or transference across the interface from GLM to TiO2. The charge migration path and the well-separated electron–hole pairs could illustrate the mechanism of enhanced photocatalytic activity of TiO2/GLM.

2. Model and computational details

In our work, all of the calculations were performed using the well tested Cambridge Serial Total Energy Package (CASTEP) code,25,26 which employs planewave basis sets to treat valence electrons and norm-conserving pseudopotentials to approximate the potential field of ionic cores. We employed a generalized gradient approximation (GGA) method, the Perdew, Burke, Ernzerhof (PBE) exchange–correlation functional, and the ultrasoft pseudopotential in the calculations. The basis set cut-off energy was chosen at 340 eV, and we used a 5 × 5 × 1 Monkhorst–Pack k-point mesh for geometry optimization of the TiO2 cluster and a 9 × 9 × 1 mesh to calculate its electronic properties. For the TiO2 cluster hybridization with graphene or GLM, such as MoS2, g-C3N4 and BP, Monkhorst–Pack special k-point was set to 4 × 4 × 1 and the k-point was set to 9 × 9 × 1 on the density of state for geometry optimization. Further increasing cutoff energy and k-points shows little difference in the calculation results. The convergence threshold for self-consistent iteration was set at 2 × 10−6 eV per atom.

S. A. Shevlin and S. M. Woodley have made a systematic study on (TiO2)n cluster by DFT calculation.27 It indicated that (TiO2)3 nanocluster was a suitable structural model to explore its structure and electronic properties, which has been proved to be in good agreement with the experimental data. To reduce time-consuming calculations, a small stoichiometric TiO2 cluster was employed in our study based on the above consideration, which was presented in Fig. 1. The optimized monolayer geometric structures of graphene, g-C3N4, MoS2, and BP were presented in Fig. 2a–d, respectively. The g-C3N4 monolayer studied in this work was built on triazine rings with a hexagonal unit cell, in which all C atoms have three nearest neighbor N atoms, as shown in Fig. 2b.28 The monolayer of MoS2 is made up of hexagons with the Mo and S atoms located at alternating corners (Fig. 2c). Inside the monolayer of BP, each phosphorus atom is covalently bonded with three adjacent phosphorus atoms to form a puckered honeycomb structure,29 as shown in Fig. 2d.


image file: c6ra12980c-f1.tif
Fig. 1 Crystal structure of (TiO2)3 cluster. Gray and red balls represent Ti and O atom, respectively.

image file: c6ra12980c-f2.tif
Fig. 2 Crystal structures of monolayer graphene (a), g-C3N4 (b), MoS2 (c) and black phosphorus (d). Dark gray, blue, yellow, aquamarine, pink balls represent C, N, and S, Mo and P atoms, respectively.

The geometry structures of TiO2 cluster combined with graphene (a bare 4 × 4 monolayer sheet), g-C3N4 (a 3 × 3 monolayer sheet), MoS2 (a 3 × 3 sandwich-like monolayer sheet) and BP (a bare 4 × 4 monolayer sheet) lattice parameters were (9.85 Å × 9.85 Å × 20.00 Å), (14.23 Å × 14.23 Å × 20.00 Å), (9.51 Å × 9.51 Å × 20.00 Å) and (12.72 Å × 9.72 Å × 20.00 Å), respectively. The thickness of vacuum layer was set as 20 Å to counteract stress. The distance from the adjacent Ti atom to the surface of GLM was set as 3 Å.

3. Results and discussion

3.1. Theoretical model of TiO2 cluster combined with graphene

Graphene was taken as an example to combine with TiO2 cluster to study the theoretical hybridization model of TiO2/GLM. The atomic arrangement of TiO2 cluster and the relative distance from the cluster to the graphene sheet were investigated to find out the suitable spatial position of TiO2 cluster, which hybridized with graphene. In order to investigate the stability of the composite, the binding energy of TiO2/G composite was calculated according to the eqn (1).
 
Eb = ETiO2/GETiO2EG (1)
where Eb indicate the binding energy of TiO2/G composite, ETiO2 is the energy of TiO2 cluster, EG is the energy of graphene, and ETiO2/G is the total energy of TiO2/G. The negative value means that the hybridization process release heat, which implies that the hybridization process is energetically favourable on thermodynamics stability. The smaller the binding energy is, the easier the hybridization process can be achieved.

There are four types of TiO2/G theoretical structure with different geometric location of TiO2 cluster (Fig. 3) and the binding energies were listed blow each model. It could be seen that model A (Fig. 3a) was the most stable geometry as its binding energy was lowest. Since C and Ti atoms have a greater electronegativity difference than C and O atoms have, electrons transfer more easier from Ti to graphene, which contribute to a better thermodynamic stability of model A. Comparing with Ti2 and Ti3 atoms, the Ti1 atom has a lower saturation of chemical bonds, which made model A stand out. As a result, a theoretical basis for the model of TiO2/GLM has been done and details of electrons transfer in TiO2/GLM will be discussed in the following chapters.


image file: c6ra12980c-f3.tif
Fig. 3 The different theoretical structure of TiO2/G with its binding energy.

3.2. Binding energies and geometric structures

Similarly, the binding energies of TiO2/GLM composites were also calculated according to the eqn (2).
 
Eb = ETiO2/GLMETiO2EGLM (2)
where Eb is the binding energy of the composite, ETiO2 is the energy of TiO2 cluster, EGLM is the energy of GLM, and ETiO2/GLM is the total energy of TiO2 cluster hybridized with GLM. From stability perspective of TiO2/GLM composites were in the order of TiO2/g-C3N4(−7.67 eV) > TiO2/BP(−1.33 eV) > TiO2/MoS2(−0.78 eV) = TiO2/G(−0.78 eV). The theoretical result implied that the combination of TiO2 cluster with GLM were all favourable, especially TiO2/g-C3N4 composite.

Interaction occurred between TiO2 cluster and GLM when they hybridized, which would make geometric structure of TiO2 cluster change. Both bond lengths and bond angles were measured to characterize the effect of GLM hybridization on the structural variety of TiO2 cluster. The bond lengths and the bond angles of TiO2 cluster were respectively listed in Tables 1 and 2, which could be seen in the ESI.

For TiO2/G composite, the interaction between C and Ti were the electrostatic attraction, while C and O were the electrostatic repulsion. For this reason, it produced a series of changes. Both the bond numbered 1 (from 1.745 to 1.754 Å) and the bond numbered 2 (from 1.744 to 1.754 Å) have grown. The ∠O5Ti1O6 increased from 108.49° to 110.58°, and ∠Ti2O1Ti3 increased from 86.37° to 88.39°. The bond numbered 5 shorted from 2.076 to 2.037, and the bond numbered 6 shorted from 2.084 to 2.040. The ∠O1Ti2O5 increased from 77.63° to 78.29°, and ∠O1Ti3O6 increased from 77.43° to 78.42°, while ∠O4Ti3O6 decreased from 124.86° to 122.64°. Above-mentioned bond length and angle changes of TiO2 cluster in TiO2/G system described the structural changes as a whole. Additionally, the TiO2/MoS2 system was very similar to the case of TiO2/G and we also considered as one of the factors which caused the same binding energies between them. By contrast, geometric variations caused by the interaction of TiO2 and g-C3N4 were very obvious. TiO2 cluster overall structure has a certain extent variation which is a chain reaction in order to protect structural integrality. Meanwhile, the N atom which contains two C–N bonds of g-C3N4 monolayer shift upward clearly and g-C3N4 becomes irregularity. In the case of TiO2/BP system, the TiO2 cluster has C2 symmetry and the interaction is mainly on one side of TiO2 cluster: for bond length, bond numbered 2 > bond numbered 1, bond numbered 10 > bond numbered 9. For angles, the ∠O5Ti1O6 decreased from 108.49° to 106.79°, ∠Ti2O1Ti3 increased from 86.37° to 89.03°, ∠O4Ti3O6 decreased from 124.86° to 118.68° and ∠O1Ti3O6 increased from 77.43° to 82.31°.

In summary, there was electrostatic interaction between GLM sheet and TiO2 cluster. The interaction between GLM and Ti was electrostatic attraction, while the interaction between GLM and O were electrostatic repulsion mainly. The electrostatic forces of TiO2 cluster from graphene and MoS2 sheet were relatively weak, while those from g-C3N4 and BP were much stronger. The reason might come from the electronegativity difference of various elements. The electronegativity of C (2.55) approached to Mo (2.16) and S (2.58), but it was between Ti (1.54) and O (3.44). In the same way, the electronegativity of N (3.04) approached to O (3.44) and the electronegativity of P (2.19) was closed to Ti (1.54). Therefore, the electrostatic interaction might not bring change to the geometry structure of graphene, MoS2 and TiO2 cluster. In contrast, larger electrostatic interaction brought change to the geometry structure of g-C3N4, BP and TiO2 cluster, which might be correlated to the change of electronic properties.

3.3. Charge density and charge density difference analysis

In order to further explore how GLM affect the electronic properties and bonding character of the TiO2/GLM composites, we calculated the electronic total charge density and difference charge density. The Fig. 4 exhibited the three-dimensional electron cloud and its potential distribution in TiO2/GLM composites. The interface charge-redistribution could adjust the electrostatic potential distribution in those systems as a whole respectively. In the case of TiO2/G composite (Fig. 4a), it was observed that the electron overlap between TiO2 cluster and graphene was very weak. Similarly, the electron overlap of three other TiO2/GLM composites were in the order of TiO2/MoS2 < TiO2/BP < TiO2/g-C3N4. All composites transient potential overlap was transition potential, which meant the reaction was spontaneous. That charge density of overlap increased signifies stronger covalent interaction and those calculation results agreed with the binding energy of TiO2/GLM.
image file: c6ra12980c-f4.tif
Fig. 4 The optimized geometric structures of TiO2/GLM with charge density (set as the isovalue of 0.2 e Å−3) and potential. (a) TiO2/G, (b) TiO2/g-C3N4, (c) TiO2/MoS2, (d) TiO2/BP.

The interaction between the TiO2 and GLM sheets implied a substantial charge transfer in the involved constituents. In order to explore the effect of GLM hybridization on the electron distribution of TiO2 cluster and the interaction between TiO2 and GLM, charge density difference was calculated. This could be visualized by three-dimensional charge density difference:

 
Δρ = ρTiO2/GLM − (ρTiO2 + ρGLM) (3)
where ρTiO2/GLM, ρTiO2 and ρGLM are the charge densities of the composite, TiO2 cluster and GLM sheets, respectively. For comparison, the charge density difference of TiO2/G, TiO2/g-C3N4, TiO2/MoS2 and TiO2/BP were presented in Fig. 5a–d, respectively. The orange color represented charge accumulation, while green color represented charge depletion. In the planar-averaged and 3D electron density difference plots of TiO2/G (Fig. 5a), we found that charge accumulation occurred on the graphene surface and charge depletion happened on the other side of TiO2 cluster. It was conclusive that the electron moved from the TiO2 cluster to graphene. By analysing the principle and rule of TiO2/G system's charge density difference, it would be suited to TiO2/GLM, as well as explaining the charge distribution between the interfaces. Remarkably, the O4 atom of TiO2/BP system had some interaction with BP, it also explained the effects of BP to TiO2 cluster on one side. As has been mentioned above, we concluded that the electron migrated from the TiO2 cluster to GLM. In photocatalytic process, GLM sheets would work as electron acceptors and provide a conduction platform to transport electrons participating in the oxidation–reduction reactions during the photodegradation of contaminant. This theoretical study pointed out the electron migration and transfer path at the interface, which might demonstrate the mechanism of enhanced photocatalytic activity of TiO2/GLM.


image file: c6ra12980c-f5.tif
Fig. 5 The planar-averaged electron density difference as a function of position in the z-direction and the 3D charge density difference with an isovalue of 0.02 e Å−3. (a) TiO2/G, (b) TiO2/g-C3N4, (c) TiO2/MoS2, (d) TiO2/BP.

3.4. Density of states, frontier molecular orbital and optical properties

To explore the electron distribution in different energy levels and the charge carrier migration path at the interface between TiO2 cluster and GLM, the density of states of TiO2 and TiO2/GLM were calculated.

The plot of total density of states (TDOS) for TiO2 cluster and the partial density of states (PDOS) on Ti, O atoms was presented in Fig. 6, in which both the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) of TiO2 cluster were presented. It could be seen that the upper of valence band (VB) was predominantly composed of O2p state, while the bottom of conduction band (CB) was mainly composed of Ti3d state. The calculated TiO2 cluster energy gap with a value of 2.04 eV, which was similar to that of rutile TiO2 we calculated, was lower than the experimental data (3.0–3.2 eV).30 It might attribute to two main reasons: on the one hand, as the size effect the difference existed between a small cluster and the bulk matter. On the other hand, the well-known shortage that DFT underestimated the band gap due to the self-correlation error of electrons and inherent lack of derivative discontinuity.31,32 HOMO was consisted of O2py state, whereas the LUMO was consisted by Ti3dz2 state, indicating that the O2py and Ti3dz2 states were the key-states of photogenerated electron transition in TiO2 cluster. The electrons transit from O2py and Ti3dz2, leaving hole in the O2py orbit under light irradiation.


image file: c6ra12980c-f6.tif
Fig. 6 The calculated TDOS and PDOS, and the side view of HOMO and LUMO with an isovalue of 0.02 e Å−3 of TiO2 cluster.

For the TiO2/G system (Fig. 7a), it was observed that the TDOS of TiO2/G and PDOS of Ti, O and C atoms, in which HOMO and LUMO of TiO2/G were also presented. From the perspective of DOS, the general shape of TDOS moved to lower energy state. Due to the hybridization of graphene, C2p orbit participated in both VB and CB. The upper of VB was predominantly composed of O2p, while the bottom of CB was mainly composed of C2p. And the electron from both VB and CB moved to the lower energy region. The above analysis showed that TiO2 cluster tended to stabilize after combination with graphene. Interestingly, the HOMO was consisted of O2py state, whereas LUMO were dominated by 2pz states of C and 3dyz of Ti, indicating that the O2py, C2pz and Ti3dyz were key-states to the photo-generated carrier of TiO2/G composite. Observably, the C2pz and Ti3dyz states have been hybridized and the CB shifted to lower energy region. So the electrons transit from O2py to LUMO leaving hole in the O2py orbit, which dispersed all over HOMO. The electron–hole pairs were easy to separate and difficult to recover so that graphene could enhance the photocatalytic activity of TiO2.


image file: c6ra12980c-f7.tif
Fig. 7 The calculated TDOSs and PDOS, and the side view of HOMO and LUMO with an isovalue of 0.02 e Å−3. (a) TiO2/G, (b) TiO2/g-C3N4, (c) TiO2/MoS2, (d) TiO2/BP. Orange and green isosurfaces represent LUMO and HOMO in the space with respect to isolated TiO2/GLM.

Because of the interaction and redistribution, TDOS shape of TiO2/g-C3N4 (Fig. 7b), TiO2/MoS2 (Fig. 7c) and TiO2/BP (Fig. 7d) showed the characteristic features of significant changes. The O2p and Ti3d states of the three systems shifted slightly to lower energy levels. It was found that GLM could affect CB ingredient but BP could affect VB ingredient as well. Intuitively, we analyzed the HOMO–LUMO orbit of TiO2/g-C3N4, TiO2/MoS2 and TiO2/BP, respectively. For the TiO2/g-C3N4 system, HOMO was consisted of O2px state and LUMO was hybridized from 2pz state of C and 2pz state of N. There was no doubt that 2p orbit had the lower energy than the 3d orbit so that photo-generated carriers transfer easier. In the case of TiO2/MoS2 system, O2py orbit constituted HOMO and Mo4dz2 orbit constituted LUMO. It was conceivable that photo-excited electrons were difficult to recover and those hole-rich O atoms were active sites for oxidation. The P3pz state forms HOMO, LUMO was formed by P3py state and O2py state at TiO2/BP system. Because the HOMO and LUMO diffused all around TiO2/BP system, the mobility of electrons increased substantially. To further verify the photocatalytic of TiO2/GLM systems, optical properties should be investigated.

The optical property of semiconductor photocatalyst is one of the key factors in determining photocatalytic properties. The UV-visible absorption spectra of TiO2 cluster and TiO2/GLM were calculated to characterize optical properties, as shown in Fig. 8. The photoelectric response and absorption intensity of TiO2 cluster have been changed after incorporation with GLM. Not only the TiO2/GLM systems gave rise to obvious red-shift of absorption edge but also the absorption intensity in the visible region improved greatly. Thus, it was concluded that coupling TiO2 cluster on the GLM would enhance the absorption in the UV-vis region and modify the absorption edge. Meanwhile, it improved the photocatalytic activity of TiO2 and likely to demonstrate excellent photocatalytic performance under vis-light irradiation.


image file: c6ra12980c-f8.tif
Fig. 8 Calculated absorption spectra of TiO2 cluster and TiO2/GLM.

In summary, interaction between GLM and TiO2 cluster caused electrons diffuse to lower energy region, which could explain why the TiO2/GLM composites became stable. HOMO and LUMO of TiO2 cluster could be changed by GLM to the benefit of photoactivity. This character could yield well-separated electron–hole pairs, and red-shift of absorption edge, absorption range extending and absorption intensity of TiO2/GLM increasing, which induced the enhanced photocatalytic performance of TiO2/GLM.

4. Conclusions

First-principles calculation based on DFT was used to explore the photoactivity and electronic properties of TiO2/GLM composites in this work. The electronic properties and geometric structures of TiO2/GLM systems were calculated to expound the interface interaction and account for the photocatalytic mechanism. It was found that TiO2/GLM composites might be demonstrated for high thermodynamical stability. The interaction of GLM and TiO2 caused charge accumulation on the GLM surface and charge depletion on the other side of TiO2. Therefore, not only was the charge transport path indicated, but also the photo-electron transition tracks were changed. We discussed the cause of photocatalytic performance enhancements from two aspects: on the one hand, the hybridization of GLM brought HOMO and LUMO of TiO2 cluster to the benefit changing of photoactivity. On the other hand, the absorption edge of TiO2/GLM systems occur red-shift and its absorption intensity aggrandize. This work can provide theoretical basis for the feasibility of experiment and the TiO2/GLM systems could be promising for photocatalytic applications.

Acknowledgements

This work has been supported by The National Natural Science Foundation of China (Grant No. 41573103, 41340037) and the scientific research program of Shandong Province (2013G0021701, 2014GGH217001) P. R. China.

Notes and references

  1. N. Zhang, M. Q. Yang, S. Liu, Y. Sun and Y. J. Xu, Chem. Rev., 2015, 115, 10307–10377 CrossRef CAS PubMed.
  2. A. Fujishima and K. Honda, Nature, 1972, 238, 37–38 CrossRef CAS PubMed.
  3. M. R. Hoffmann, S. T. Martin, W. Choi and D. W. Bahnemann, Chem. Rev., 1995, 95, 69–96 CrossRef CAS.
  4. H. G. Yang, C. H. Sun, S. Z. Qiao, J. Zou, G. Liu, S. C. Smith, H. M. Cheng and G. Q. Lu, Nature, 2008, 453, 638–641 CrossRef CAS PubMed.
  5. H. Gao, X. Sui, D. Dai, C. Si and G. Liu, J. Nanosci. Nanotechnol., 2015, 15, 5193–5197 CrossRef CAS.
  6. X. An and J. C. Yu, RSC Adv., 2011, 1, 1426–1434 RSC.
  7. Q. Xiang, J. Yu and M. Jaroniec, J. Am. Chem. Soc., 2012, 134, 6575–6578 CrossRef CAS PubMed.
  8. W. Geng, H. Liu and X. Yao, Phys. Chem. Chem. Phys., 2013, 15, 6025–6033 RSC.
  9. F. Perreault, A. Fonseca de Faria and M. Elimelech, Chem. Soc. Rev., 2015, 44, 5861–5896 RSC.
  10. J. Liu, H. Bai, Y. Wang, Z. Liu, X. Zhang and D. D. Sun, Adv. Funct. Mater., 2010, 20, 4175–4181 CrossRef CAS.
  11. H. Zhang, P. Xu, G. Du, Z. Chen, K. Oh, D. Pan and Z. Jiao, Nano Res., 2011, 4, 274–283 CrossRef CAS.
  12. Y. Liang, H. Wang, H. Sanchez Casalongue, Z. Chen and H. Dai, Nano Res., 2010, 3, 701–705 CrossRef CAS.
  13. F. Sordello, E. Odorici, K. Hu, C. Minero, M. Cerruti and P. Calza, Nanoscale, 2016, 8, 3407–3415 RSC.
  14. X. Chen and B. Chen, Environ. Sci. Technol., 2015, 49, 6181–6189 CrossRef CAS PubMed.
  15. W. J. Ong, L. L. Tan, S. P. Chai and S. T. Yong, Dalton Trans., 2015, 44, 1249–1257 RSC.
  16. W.-J. Ong, L. K. Putri, L.-L. Tan, S.-P. Chai and S.-T. Yong, Appl. Catal., B, 2016, 180, 530–543 CrossRef CAS.
  17. W. Li, C. Li, B. Chen, X. Jiao and D. Chen, RSC Adv., 2015, 5, 34281–34291 RSC.
  18. J. Zhou, M. Zhang and Y. Zhu, Phys. Chem. Chem. Phys., 2015, 17, 3647–3652 RSC.
  19. J. Shen, H. Yang, Q. Shen, Y. Feng and Q. Cai, CrystEngComm, 2014, 16, 1868 RSC.
  20. M.-Q. Yang, C. Han and Y. J. Xu, J. Phys. Chem. C, 2015, 119, 27234–27246 CAS.
  21. Q. Xiang, J. Yu and M. Jaroniec, J. Am. Chem. Soc., 2012, 134, 6575–6578 CrossRef CAS PubMed.
  22. X. Yang, H. Huang, M. Kubota, Z. He, N. Kobayashi, X. Zhou, B. Jin and J. Luo, Mater. Res. Bull., 2016, 76, 79–84 CrossRef CAS.
  23. H. Liu, Y. Du, Y. Deng and P. D. Ye, Chem. Soc. Rev., 2015, 44, 2732–2743 RSC.
  24. H. Du, X. Lin, Z. Xu and D. Chu, J. Mater. Chem. C, 2015, 3, 8760–8775 RSC.
  25. M. D. Segall, P. J. D. Lindan, M. J. Probert, C. J. Pickard, P. J. Hasnip, S. J. Clark and M. C. Payne, J. Phys.: Condens. Matter, 2002, 14, 2717–2744 CrossRef CAS.
  26. H. Gao, X. Li, J. Lv and G. Liu, J. Phys. Chem. C, 2013, 117, 16022–16027 CAS.
  27. S. A. Shevlin and S. M. Woodley, J. Phys. Chem. C, 2010, 114, 17333–17343 CAS.
  28. L. Xu, W.-Q. Huang, L.-L. Wang, Z.-A. Tian, W. Hu, Y. Ma, X. Wang, A. Pan and G.-F. Huang, Chem. Mater., 2015, 27, 1612–1621 CrossRef CAS.
  29. L. Li, Y. Yu, G. J. Ye, Q. Ge, X. Ou, H. Wu, D. Feng, X. H. Chen and Y. Zhang, Nat. Nanotechnol., 2014, 9, 372–377 CrossRef CAS PubMed.
  30. A. J. Bard and M. A. Fox, Acc. Chem. Res., 1995, 28, 141–145 CrossRef CAS.
  31. P. Mori-Sanchez, A. J. Cohen and W. T. Yang, Phys. Rev. Lett., 2008, 100, 146401 CrossRef PubMed.
  32. L. J. Sham and M. Schluter, Phys. Rev. Lett., 1983, 51, 1888–1891 CrossRef.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra12980c

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.