DOI:
10.1039/C6RA13947G
(Paper)
RSC Adv., 2016,
6, 67488-67494
Sol–gel fabrication of WO3/RGO nanocomposite film with enhanced electrochromic performance
Received
29th May 2016
, Accepted 4th July 2016
First published on 4th July 2016
Abstract
Tungsten trioxide/reduced graphene oxide (WO3/RGO) nanocomposite film was fabricated by the sol–gel spin-coating technique, using a mixed colloidal dispersion of WO3 precursor and GO. The structure, surface morphology and optical properties of WO3 and WO3/RGO films were investigated by means of X-ray diffraction, Raman spectroscopy, X-ray photoelectron spectroscopy, scanning electron microscopy, high-resolution transmission electron microscopy and UV-visible spectrophotometry. The electrochemical and electrochromic performances of the WO3 and WO3/RGO films were studied by chronoamperometry, in situ optical transmittance and cyclic voltammogram analysis. In the WO3/RGO nanocomposite film, monoclinic WO3 nanoparticles were grafted onto the surface of RGO sheets via C–W and C–O–W chemical bonds. The results showed that the WO3/RGO nanocomposite film exhibited shorter coloration–bleaching times (tc = 9.5 s and tb = 7.6 s), higher coloration efficiency (75.3 cm2 C−1 at 633 nm), larger optical modulatory range (59.6% at 633 nm) and better cyclic stability compared with WO3 films, which are attributed to faster Li+ ion diffusion and electron transfer rate. The methodology provides a facile and industrially scalable approach to obtain fast switching electrochromic materials.
1. Introduction
Electrochromism describes a reversible phenomenon of a material's optical properties (transmittance, absorbance and reflectance) produced by oxidation–reduction reactions under an applied electric field.1 Since it was discovered by Deb in 1969,2 considerable progress has been achieved in the study of electrochromic materials and their applications, such as energy-saving smart windows,3 intelligent optical displays,4 antiglare rearview mirrors,5 and active camouflages.6 Recently, electrochromic smart windows have attracted tremendous attention because the energy crisis requires a substantial energy-saving solution to combat conventional energy source consumption.7,8 As an important kind of inorganic electrochromic material, tungsten oxide (WO3) has been extensively studied, due to its relatively fast response time and high coloration efficiency as compared with other electrochromic materials.9–11 In order to improve the electrochemical stability and switching rate of crystalline structure WO3 films, many kinds of WO3 nanocomposites, such as WO3/carbon nanotubes,12 WO3/Ti,13 WO3/V2O5,14 and WO3/TiO2,15 have been studied, and exhibited enhanced electrochromic properties.
The excellent characteristics of graphene, such as high electrical conductivity, large surface area, high optical transmittance and high chemical stability, endow its composites with many enhanced properties.16,17 There have been several reports of metal oxide/graphene composites for electrochromic application. For example, Fu et al.18 prepared WO3/RGO composites by a one-step facile electrochemical deposition method which showed faster switching time, longer cycle life and larger coloration efficiency due to the synergistic effect of incorporation of RGO into WO3. A WO3 nanowires/RGO composite was prepared by the Chang group, and the composite exhibited high-quality electrochromic performance with fast color-switching speed, good cyclic stability, and high coloration efficiency.19 Tu et al.20 also prepared a porous nickel oxide (NiO)/RGO hybrid film with improved electrochromic properties by the combination of electrophoretic deposition and chemical-bath deposition methods, and the porous hybrid film exhibited high coloration efficiency, fast switching speed and better cycling performance compared with a porous NiO thin film. Zhang et al.21 synthesized vanadium pentoxide (V2O5)/graphene nanocomposite films by a direct intercalation method using V2O5 sol and graphene. The intercalation of graphene improves the stability, reversibility, response rate and optical modulatory range the V2O5 xerogel films. However, there have been very few studies on the effect of structure and morphology on electrochromic performance of metal oxide electrochromic materials via the incorporation of RGO sheets.
Many methods have been adopted to fabricate WO3 electrochromic films, including electrochemical deposition,22 spray deposition,23 electrophoretic deposition,24 chemical vapor deposition,25 sputter deposition,26 and the sol–gel technique.27 Among them, the sol–gel technique is very attractive because of its cost effectiveness and ease of depositing uniform large-area films for window applications. Meanwhile, it is efficient for tailoring the microstructure of films by the introduction of chemical dopants to the reactant sols.28 Herein, we introduce a facile and industrially scalable sol–gel approach to fabricate WO3/RGO nanocomposite film. The surface morphology, microstructure and electrochromic properties of WO3 and WO3/RGO films are studied. We envisage that the unique morphology and high electrical conductivity of RGO sheets could enhance the electrochromic performance by overcoming the intrinsic drawback of the crystalline structure of WO3.
2. Experimental
2.1 Materials
Natural graphite powder (particle sizes ≤ 500 mesh; purity ≥ 99.9%) was purchased from Qingdao Xingyuan Graphite Co. Ltd, China. Potassium permanganate, sulfuric acid, hydrogen peroxide, sodium nitrate and anhydrous ethanol were purchased from Chengdu Kelong Chemical Reagent Factory, China. Tungsten hexachloride (WCl6, 99.5%) was purchased from Aladdin Chemistry Co. Ltd. All reagents were of analytical reagent grade and used without further purification.
2.2 WO3/RGO nanocomposite film fabrication
GO powder was prepared by the modified Hummers method using graphite powder,29 washed with deionized water, and dried in a vacuum freeze drier. The GO powder was dispersed in anhydrous ethanol by sonication to fabricate a stable GO dispersion. 10 g of WCl6 was dissolved in 100 mL of anhydrous ethanol and stirred for 20 min to produce tungsten alkoxide precursor under nitrogen atmosphere. The tungsten alkoxide precursor solution was aged by heating at 50 °C for 24 hours and kept for another 3 days at room temperature, resulting in the formation of a homogeneous WO3 precursor solution. GO suspension was added slowly to the WO3 precursor solution and dispersed by sonication for 2 hours to obtain the mixed colloidal dispersions. The ratio of GO to WO3 was 3 wt%. Indium tin oxide (ITO, Rs = 5–15 Ω, Zhuhai Kaivo Optoelectronic Technology Co. Ltd, China) glass substrates were sequentially rinsed with 1 mol L−1 NaOH/H2O2 mixed solution, acetone, anhydrous ethanol and deionized water. The mixed colloidal dispersions were deposited onto the ITO glass substrates by the spin-coating procedure to fabricate the WO3/RGO nanocomposite film at a controlled speed of 2000 rpm. Subsequently, the film was annealed in air at 100 °C for 1 h and heated to 500 °C at a rate of 5 °C min−1 and kept for 2 hours. The WO3 film was prepared using the same conditions but without the addition of GO.
2.3 Characterization techniques
The structure, surface morphology and optical properties of the WO3 and WO3/RGO films were investigated by X-ray diffraction (XRD, X'Pert Pro MPD, Cu radiation, λ = 1.5418 Å), Raman spectroscopy (LabRAM HR, laser excitation wavelength of 632.8 nm), X-ray photoelectron spectroscopy (XPS, Kratos XSAM800), scanning electron microscopy (SEM, JEOL JSM-7500F), high-resolution transmission electron microscopy (HRTEM, Philips Tecnai G2 F20) and UV-visible spectrophotometry (Analytik Jena AG, Specord 200). Electrochemical and electrochromic properties were determined in a conventional three-electrode electrochemical cell with an electrochemical workstation (Shanghai CH Instruments Co., CHI660E), where platinum sheet electrode and saturated calomel electrode were used as counter and reference electrodes, respectively. Chronoamperometry (CA) and cyclic voltammetry (CV) measurements were performed in a 1 mol L−1 lithium perchlorate/propylene carbonate (LiClO4/PC) electrolyte. The optical transmittance of the WO3 and WO3/RGO films were measured in situ by a UV-visible spectrophotometer (Analytik Jena AG, Specord 200). Coloration and bleaching times were calculated as the time required for achieving 90% of the total transmission change. The potential range of CV measurements for the films was from 1.5 V to −1.5 V at a scan rate of 50 mV s−1.
3. Results and discussion
3.1 Structural analysis
Fig. 1 shows the XRD spectra of the spin-coated WO3 and WO3/RGO films. Four main peaks corresponding to the ITO glass substrate were identified. The XRD patterns of the WO3 and WO3/RGO films show WO3 monoclinic structure (JCPDS no. 43-1035). Susanti et al.30 also found similar results for their WO3 nanomaterial which was synthesized by a sol–gel method using WCl6 and C2H5OH as precursors followed by annealing at 500 °C. The XRD pattern for the WO3 film shows three dominant crystalline plane orientations: (002) at 2θ = 23.2°, (020) at 2θ = 23.7° and (200) at 2θ = 24.3°. However, the diffraction pattern of the WO3/RGO nanocomposite film was found to be shifted towards lower diffraction angles mainly due to the interaction between the oxygen functional groups in RGO and the outermost oxygen in WO3.31 The XRD pattern for the WO3/RGO nanocomposite film also shows three dominant crystalline plane orientations: (002) at 2θ = 23.1°, (020) at 2θ = 23.5° and (200) at 2θ = 24.1°. The estimate of average crystallite sizes (D, nm) in the materials was calculated by the Scherrer formula (1):32 |
 | (1) |
where λ is the wavelength of the X-ray radiation (nm), B is the full width at half maximum (radian), and θ is Bragg's angle (degree). The mean WO3 crystallite sizes estimated from (002) peaks of the WO3 and WO3/RGO films are 21 nm and 17 nm respectively, in agreement with the following SEM results. Meanwhile, the result shows that the incorporation of RGO sheets has a certain restraining effect on grain growth of WO3. The diffraction peak of RGO was not observed in the pattern of the WO3/RGO film probably due to too low a content, but the existence of RGO was confirmed by Raman spectra.
 |
| Fig. 1 XRD spectra of ITO glass, WO3 and WO3/RGO films. | |
The Raman spectra of GO powder and WO3 and WO3/RGO films are shown in Fig. 2. Raman analysis on WO3 film shows two peaks centered at 713 cm−1 and 807 cm−1, characteristic features of monoclinic WO3, that can be ascribed to O–W–O stretching vibrations of the bridging oxygen of WO6 octahedra.33 The Raman spectrum of GO powder shows D-band (1331 cm−1) and G-band (1589 cm−1) peaks. The G-band represents the in-plane bond-stretching motion of the pairs of C sp2 atoms (the E2g phonons), whereas the D-band corresponds to the breathing modes of rings or κ-point phonons of A1g symmetry.34 In addition, the observation of D-band (1326 cm−1) and G-band (1590 cm−1) peaks confirms the existence of RGO in WO3/RGO nanocomposite film. As for GO and RGO of the WO3/RGO composite film, the D/G ratios were calculated as 1.23 and 1.38, respectively. The D/G intensity ratio of RGO is larger than that of GO, indicating a decrease in the average size of the sp2 domains and formation of a high quantity of defects upon heat reduction of GO, which is in agreement with the literature.35,36 Compared with that of the WO3 film, the band at 713 cm−1 was broadened and shifted to 701 cm−1 in the spectrum of the WO3/RGO composite film, probably because the formation of C–O–W bonds makes the initial W
O bond weaker and a similar phenomenon has been reported.37,38 It means that WO3 was grafted onto the surface of RGO sheets via C–O–W bonds rather than physically adsorbed on the surfaces of RGO sheets. This kind of structure is desirable for electron transfer in the WO3/RGO nanocomposite film.39
 |
| Fig. 2 Raman spectra of GO powder, and WO3 and WO3/RGO films. | |
XPS was performed to determine the surface element composition of samples and the valence states of various species. From the full spectrum of WO3/RGO nanocomposite film in Fig. 3(a), it can be seen that the sample consists of W, O, and C elements without impurities. Fig. 3(b) depicts the W 4f XPS spectra of the WO3 and WO3/RGO films. The main peaks of W 4f7/2 and W 4f5/2 are located at 35.5 eV and 37.5 eV for the WO3 film. While the main peaks of W 4f7/2 and W 4f5/2 are slightly shifted to 35.6 eV and 37.7 eV for the WO3/RGO nanocomposite film, associated with interaction between WO3 and RGO. Furthermore, these are consistent with the typical binding energies of W6+ oxidation states in stoichiometric WO3 materials.40 Fig. 3(c) shows the C 1s XPS spectrum of the WO3/RGO nanocomposite film. Peaks at 284.6, 286.5, 287.6 and 288.8 eV are observed, which can be assigned to the C–C, C–O, C
O and O–C
O species, respectively.41 Based on other research reports about GO and RGO,34,41,42 it is concluded that most oxygen functional groups are successfully removed during thermal treatment. Moreover, the peaks at about 280.6 eV indicated the formation of C–W chemical bond.
 |
| Fig. 3 (a) XPS full spectrum of WO3/RGO nanocomposite film, (b) W 4f XPS spectra of the WO3 and WO3/RGO films, and (c) C 1s XPS spectrum of WO3/RGO nanocomposite film. | |
3.2 Morphological analysis
The morphology of the WO3 and WO3/RGO films were characterized using SEM. Fig. 4 shows the SEM images with low and high magnifications of WO3 and WO3/RGO films. The WO3 film in Fig. 4(a) demonstrates a smooth and compact surface with a few cracks which are probably attributed to the release of internal stresses during the solvent evaporation.43 Fig. 4(b) shows that the WO3 film consists of uniform spherical nanoparticles with an average size of around 20 nm. However, the WO3/RGO nanocomposite film displays a coarse and loose surface with more microcracks (Fig. 4(c)). From the high magnification image of the WO3/RGO nanocomposite film (Fig. 4(d)), it can be observed that the surfaces and edges of the RGO sheets are covered by WO3 nanoparticles with smaller size (around 15 nm). Therefore, the WO3/RGO nanocomposite film has more space to facilitate more ions of the electrolyte to penetrate through the film with shorter diffusion paths.
 |
| Fig. 4 SEM images of the WO3 film (a and b) and WO3/RGO nanocomposite film (c and d). | |
The morphology and structural characteristics of the WO3/RGO nanocomposite were also analyzed by HRTEM. The low magnification image shown in Fig. 5(a) demonstrates that there are good interfacial contacts formed between WO3 nanoparticles and wrinkled RGO sheets. The lattice distance of WO3 nanoparticles measured in the HRTEM image (Fig. 5(b)) equals 0.387 nm, which can be indexed to the (002) crystal plane of monoclinic WO3, and it can be concluded that the WO3 nanoparticles are attached to the RGO sheets from the HRTEM image.
 |
| Fig. 5 (a) TEM and (b) HRTEM images of WO3/RGO nanocomposite film. | |
3.3 Optical properties and bandgap calculations
Fig. 6 compares the absorption spectra and optical bandgap (Eg) of WO3 and WO3/RGO films prepared by the spin-coating technique. The optical absorption edge of the WO3/RGO nanocomposite film shows a red shift compared with the WO3 film owing to the incorporation of RGO sheets. The Eg of a semiconductor material may be estimated from absorption spectra using the following eqn (2):44,45where α, hν, A and Eg are the absorption coefficient, the photon energy, a constant of proportionality, and optical bandgap energy. The exponent n depends on the type of optical transition of a semiconductor material (n = 1/2 for allowed direct transition and n = 2 for allowed indirect transition). WO3 has an indirect band transition between the 2p electrons from the valence band of the oxygen and the 5d conduction band of tungsten.46 WO3 is an indirect bandgap semiconductor,47 and the optical bandgap energy of WO3 varies from about 2.6 to 3.0 eV,48 so the value of n is taken as 2 for WO3. Thus, the Eg of the WO3 and WO3/RGO films could be estimated from Tauc plots of (αhν)1/2 versus (hν), and the values of Eg were calculated by extrapolating the linear part to zero absorption coefficient α = 0. The Eg for the WO3 film is calculated as about 3.01 eV, greater than that of bulk WO3 (2.62 eV) due to quantum size effects of the small crystallite size.49 And Eg of WO3/RGO nanocomposite film is decreased slightly to 2.87 eV. The incorporation of RGO could decrease the bond length of W
O or O–W–O, which is possibly ascribed to the interaction between WO3 and GO, such as C–O–W and C–W bonds, and as a result the Eg decreases.
 |
| Fig. 6 UV-visible optical absorption spectra of WO3 and WO3/RGO films. Inset: Tauc plots of (αhν)1/2 versus (hν). | |
3.4 Electrochemical and electrochromic performance
The assessment of the coloration–bleaching kinetics is of great importance to determine the practical applications of electrochromic films. Therefore, the switching characteristics of the WO3 and WO3/RGO films were measured by CA and the corresponding in situ optical transmittance at an optical wavelength of 633 nm, as shown in Fig. 7(a) and (b). The switching times for the WO3 film are calculated to be 14.3 s for coloration and 10.8 s for bleaching. As expected, for the WO3/RGO nanocomposite film, the switching times were calculated to be 9.5 s for coloration and 7.6 s for bleaching. Coloration efficiency (CE) is a crucial characteristic parameter for evaluating the electrochromic performance of materials, which is defined as the change in optical density (ΔOD) per unit of charge density (Q) inserted into (or extracted from) the electrochromic films.50 It can be calculated according to the following formulas (3) and (4):51 |
 | (3) |
|
 | (4) |
where Tb and Tc denote the transmittances of the film in bleached and colored states, respectively. Fig. 7(c) shows the plots of ΔOD at a wavelength of 633 nm versus the inserted charge density at an applied potential of −1.5 V. The CE value is calculated to be 46.7 and 75.3 cm2 C−1 at 633 nm for the WO3 and WO3/RGO films, respectively. Combining the results of coloration–bleaching switching times and coloration efficiency of the WO3 and WO3/RGO films, it can be concluded that the WO3/RGO nanocomposite film has shorter coloration–bleaching times and greater coloration efficiency than the WO3 film, and can be explained from two facets. Firstly, there are plentiful microcracks and holes in the WO3/RGO nanocomposite film, facilitating intercalation and de-intercalation of Li+ ions at the interface between the WO3/RGO film and electrolyte. Secondly, it is known that RGO sheets have a high electrical conductivity and specific surface area.15,16 Massive WO3 nanoparticles are grafted onto the surface of RGO sheets via chemical bonds, such as C–O–W and C–W bonds. The chemical bonds between WO3 and RGO sheets can effectively reduce the interface defects and contact resistance, and therefore the RGO sheets promote the electron transfer rate in the WO3/RGO nanocomposite film by providing low-resistance conduction pathways.37
 |
| Fig. 7 (a) Chronoamperometry measurements and (b) corresponding in situ optical transmittance curves for WO3 and WO3/RGO films at a fixed wavelength of 633 nm under potentials of −1.5 V (30 s) and 1.5 V (30 s), and (c) variation of the in situ optical density (ΔOD) versus charge density (Q) for WO3 and WO3/RGO films at a fixed wavelength of 633 nm under a coloration potential of −1.5 V. | |
Fig. 8 shows the in situ optical transmittance spectra of both the WO3 and WO3/RGO films measured at 1.5 V (bleached state, 30 s) and −1.5 V (colored state, 30 s). As can be seen, both films display low transmittance at −1.5 V and high transmittance at 1.5 V. It is observed that the optical modulation of the WO3/RGO nanocomposite film is around 59.6% at a wavelength of 633 nm, slightly higher than that of the WO3 film (53.8%). This proves that the incorporation of RGO has increased the optical modulatory range, in agreement with the literature.17 Photographs of the WO3 and WO3/RGO films in the colored and bleached states are shown in the insets in Fig. 8. It can be seen that the color changes are reversible and homogeneous with varying potential. Briefly, when the applied potential was −1.5 V, the color of the WO3 and WO3/RGO films was dark blue (colored states). As the potential increased from −1.5 to 1.5 V, the color of the WO3 and WO3/RGO films gradually turned powder blue (bleached states).
 |
| Fig. 8 In situ optical transmittance spectra of both WO3 and WO3/RGO films measured at 1.5 V (bleached state, 30 s) and −1.5 V (colored state, 30 s). The inset shows photographs of the colored and the bleached states for the WO3 and WO3/RGO films. | |
Fig. 9(a) and (b) shows the cyclic voltammograms of the WO3 and WO3/RGO films. During Li+ ion intercalation/de-intercalation, the maximum cathodic and anodic peak current densities for WO3 and WO3/RGO films are −2.18/1.65 mA cm−2 and −1.27/0.98 mA cm−2, respectively. The maximum cathodic and anodic peak currents for the WO3/RGO nanocomposite film at the 150th cycle are −2.16/1.36 mA cm−2. But, for the WO3 film, the maximum cathodic and anodic peak current densities at the 150th cycle drastically decreased to −0.78/0.59 mA cm−2. The WO3/RGO nanocomposite film demonstrates much higher current density than the WO3 film, meaning that more ions and electrons are involved at the interface between the WO3/RGO nanocomposite film and the electrolyte. The result indicates that the incorporation of RGO can enhance the cyclic stability and the intercalation properties of Li+ ions.
 |
| Fig. 9 Cyclic voltammograms of WO3 film (a) and WO3/RGO nanocomposite film (b). | |
4. Conclusions
In summary, WO3/RGO nanocomposite film has been fabricated by a facile and industrially scalable sol–gel spin-coating approach. In the WO3/RGO nanocomposite film, the incorporation of RGO brings about more microcracks and holes, and massive WO3 nanoparticles grafted onto the surface of RGO sheets via chemical bonds, facilitating Li+ ion diffusion and accelerating the electron transfer rate. As a result, the synergistic effect between the WO3 nanoparticles and the RGO sheets delivers shorter coloration–bleaching times, higher coloration efficiency and better cyclic stability in comparison with the WO3 film. The results suggest that the sol–gel spin-coating approach is an effective way to prepare WO3/RGO nanocomposite electrochromic film for electrochromic devices and energy-saving smart windows.
Acknowledgements
The authors are grateful for the financial support from the National Natural Science Foundation of China (no. 51473104 and 61271075).
References
- P. R. Somani and S. Radhakrishnan, Mater. Chem. Phys., 2003, 77, 117 CrossRef CAS.
- S. K. Deb, Appl. Opt., Suppl., 1969, 3, 192 CrossRef.
- H. Geizerová, Z. Hejl and M. Santrůcek, J. Electrochem. Soc., 1993, 140, 3560 CrossRef.
- C. G. Granqvist, E. Avendaño and A. Azens, Thin Solid Films, 2003, 442, 201 CrossRef CAS.
- F. G. K. Baucke, K. Bange and T. Gambke, Displays, 1988, 9, 179 CrossRef CAS.
- S. Beaupre, A. C. Breton, J. Dumas and M. Leclerc, Chem. Mater., 2009, 21, 1504 CrossRef CAS.
- J. M. Wang, X. W. Sun and Z. H. Jiao, Materials, 2010, 3, 5029 CrossRef CAS.
- M. Pittaluga, Energ. Build., 2013, 66, 49 CrossRef.
- G. A. Niklasson and C. G. Granqvist, J. Mater. Chem., 2006, 17, 127 RSC.
- V. V. Kondalkar, S. S. Mali, R. R. Kharade, R. M. Mane, P. S. Patil, C. K. Hong, J. H. Kim, S. Choudhury and P. N. Bhosale, RSC Adv., 2015, 5, 26923 RSC.
- I. Karakurt, J. Boneberg and P. Leiderer, Appl. Phys. A, 2006, 83, 1 CrossRef CAS.
- P. M. Kadam, N. L. Tarwal, S. S. Mali, H. P. Deshmukh and P. S. Patil, Electrochim. Acta, 2011, 58, 556 CrossRef CAS.
- G. F. Cai, X. L. Wang, D. Zhou, J. H. Zhang, Q. Q. Xiong, C. D. Gu and J. P. Tu, RSC Adv., 2013, 3, 6896 RSC.
- N. Ozer and C. M. Lampert, Thin Solid Films, 1999, 349, 205 CrossRef CAS.
- L. Y. Wang, L. Yuan, X. F. Wu, J. Wu, C. M. Hou and S. H. Feng, RSC Adv., 2014, 4, 47670 RSC.
- L. L. Xu, S. W. Bian and K. L. Song, J. Mater. Sci., 2014, 49, 6217 CrossRef CAS.
- K. S. Novoselov, V. I. Fal'ko, L. Colombo, P. R. Gellert, M. G. Schwab and K. Kim, Nature, 2012, 490, 192 CrossRef CAS PubMed.
- C. P. Fu, C. Foo and P. S. Lee, Electrochim. Acta, 2014, 117, 139 CrossRef CAS.
- X. Chang, S. Sun, L. Dong, X. Hu and Y. Yin, Electrochim. Acta, 2014, 129, 40 CrossRef CAS.
- G. F. Cai, J. P. Tu, J. Zhang, Y. J. Mai, Y. Lu, C. D. Gu and X. L. Wang, Nanoscale, 2012, 4, 5724 RSC.
- X. Y. Zhang, H. J. Sun, Z. L. Li, J. Xu, S. S. Jiang, Q. Y. Zhu, A. P. Jin and G. S. Zakharova, J. Electrochem. Soc., 2013, 160, H587 CrossRef CAS.
- S. I. C. D. Torresi, A. Gorenstein, R. M. Torresi and M. V. Vázquez, J. Electroanal. Chem., 1991, 318, 131 CrossRef.
- R. Mukherjee and P. P. Sahay, J. Mater. Sci.: Mater. Electron., 2015, 26, 1 Search PubMed.
- E. Khoo, P. S. Lee and J. Ma, J. Eur. Ceram. Soc., 2010, 30, 1139 CrossRef CAS.
- W. B. Henley and G. J. Sacks, J. Electrochem. Soc., 1997, 144, 1045 CrossRef CAS.
- E. Washizu, A. Yamamoto, Y. Abe, M. Kawamura and K. Sasaki, Solid State Ionics, 2003, 165, 175 CrossRef CAS.
- M. Deepa, N. Sharma, P. Varshney, S. P. Varma and S. A. Agnihotry, J. Mater. Sci., 2000, 35, 5313 CrossRef CAS.
- S. A. Agnihotry, R. Sharma, M. Kar and T. K. Saxena, Sol. Energy Mater. Sol. Cells, 2006, 90, 15 CrossRef CAS.
- W. S. Hummers and R. E. Offeman, J. Am. Chem. Soc., 1958, 80, 1339 CrossRef CAS.
- D. Susanti, N. S. Haryo, H. Nisfu, E. P. Nugroho, H. Purwaningsih, G. E. Kusuma and S. J. Shih, Front. Chem. Sci. Eng., 2012, 6, 371 CrossRef CAS.
- S. Thangavel, M. Elayaperumal and G. Venugopal, Mater. Express, 2012, 2, 327 CrossRef CAS.
- B. D. Cullity and S. R. Stock, Elements of X-ray Diffraction, Prentice Hall, Upper Saddle River, NJ, 3rd edn, 2001 Search PubMed.
- Y. Djaoued, S. Balaji and N. Beaudoin, J. Sol-Gel Sci. Technol., 2013, 65, 374 CrossRef CAS.
- D. Chen, L. Li and L. Guo, Nanotechnology, 2011, 22, 325601 CrossRef PubMed.
- R. S. Dey and C. R. Raj, RSC Adv., 2013, 3, 25858 RSC.
- H. L. Guo, X. F. Wang, Q. Y. Qian, F. B. Wang and X. H. Xia, ACS Nano, 2009, 3, 2653 CrossRef CAS PubMed.
- 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 RSC.
- S. M. Zhu, X. Y. Liu, Z. X. Chen, C. J. Liu, C. L. Feng, J. J. Gu, Q. L. Liu and D. Zhang, J. Mater. Chem., 2010, 20, 9126 RSC.
- Y. W. Zhu, S. Murali, M. D. Stoller, K. J. Ganesh, W. W. Cai, P. J. Ferreira, A. Pirkle, R. M. Wallace, K. A. Cychosz, M. Thommes, D. Su, E. A. Stach and R. S. Ruoff, Science, 2011, 332, 1537 CrossRef CAS PubMed.
- J. C. Dupin, D. Gonbeau, P. Vinatier and A. Levasseur, Phys. Chem. Chem. Phys., 2000, 2, 1319 RSC.
- J. Gao, F. Liu, Y. L. Liu, N. Ma, Z. Q. Wang and X. Zhang, Chem. Mater., 2010, 22, 2213 CrossRef CAS.
- N. Song, H. Q. Fan and H. L. Tian, J. Mater. Sci., 2015, 50, 2229 CrossRef CAS.
- J. Li, Q. L. Zhao, G. Y. Zhang, J. Z. Chen, L. Zhong, L. Li, J. Huang and Z. Ma, Solid State Sci., 2010, 12, 1393 CrossRef CAS.
- F. P. Koffyberg, K. Dwight and A. Wold, Solid State Commun., 1979, 30, 433 CrossRef CAS.
- J. Tauc, Amorphous and liquid semiconductor, Plenum Press, New York, 1974 Search PubMed.
- C. G. Granqvist, A. Azens, A. Hjelm, L. Kullman, G. A. Niklasson, D. Rönnow, M. Strømme Mattsson, M. Veszelei and G. Vaivars, Sol. Energy, 1998, 63, 199 CrossRef CAS.
- P. P. González-Borrero, F. Sato, A. N. Medina, M. L. Baesso, A. C. Bento, G. Baldissera, C. Persson, G. A. Niklasson, C. G. Granqvist and A. Ferreira da Silva, Appl. Phys. Lett., 2010, 96, 061909 CrossRef.
- C. G. Granqvist, Sol. Energy Mater. Sol. Cells, 2000, 60, 201 CrossRef CAS.
- T. Kubo and Y. Nishikitani, J. Electrochem. Soc., 1998, 145, 1729 CrossRef CAS.
- H. Z. Li, J. M. Wang, G. Y. Shi, H. Z. Wang, Q. H. Zhang and Y. G. Li, RSC Adv., 2015, 5, 196 RSC.
- C. G. L. C. Chen and K. C. Ho, Electrochim. Acta, 2001, 46, 2151 CrossRef.
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.