DOI:
10.1039/C5RA18613G
(Communication)
RSC Adv., 2015,
5, 103567-103572
Sputtering and sulfurization-combined synthesis of a transparent WS2 counter electrode and its application to dye-sensitized solar cells†
Received
11th September 2015
, Accepted 24th November 2015
First published on 25th November 2015
Abstract
In this work, continuous and large-area tungsten sulfide (WS2) films, deposited by radio frequency sputtering followed by a sulfurization process, were applied as a low-cost platinum (Pt)-free counter electrode (CE) for dye-sensitized solar cells (DSSCs). The composition and structure of WS2 films were confirmed using X-ray diffraction, field-emission scanning electron microscopy, Raman spectroscopy and X-ray photoemission spectroscopy techniques. The WS2 CE was phase pure and considerably transparent. The cyclic voltammetry, electrochemical impedance spectroscopy and Tafel curve showed that the WS2 CE possesses high electrocatalytic activity and fast reaction kinetics for the reduction of tri-iodide to iodide, which can be attributed to its inherent catalytic property. Finally, TiO2-based DSSC with an optimized WS2 CE (sputtered for 10 min) showed as high as 6.3% power conversion efficiency, which was comparable to the performance of DSSC with a Pt-based CE (6.64%). Our study demonstrated the feasibility to develop low-cost, transparent, catalytically active, stable and abundant metal chalcogenide catalysts by an RF sputtering method to replace Pt CE for photovoltaic application.
Introduction
The dye-sensitized solar cells (DSSCs) is a promising candidate for replacing traditional silicon-based solar cells due to its simple fabrication process, environmental friendliness, high power conversion efficiency (PCE), availability of easy and low-cost sensitizers, simple production, high stability and wider range of design possibilities.1,2 To date, a myriad of efforts have been made to improve the PCE of DSSCs by using platinum (Pt) counter electrode (CE), cobalt(II/III) electrolytes and zinc porphyrin.3 Among them, conventional Pt as a CE is an expensive noble metal, which greatly restricts the commercial production of DSSCs. Pt can also be decomposed to PtI4 by
redox couple in electrolyte.4,5 Many attempts have been made to replace the conventional Pt CE with low-cost alternatives such as graphite,6 carbon black,7 carbon nanotubes,8 conducting polymers,9 metal chalcogenides,10,11 nitrides,12 carbides13 and phosphides.14 On the other hand, environmentally benign, abundant and non-expensive transition metal di-chalcogenides (MX2, M = Mo, W and X = S, Se) have been the subject of extensive investigation due to their strong catalytic activities and potential electrochemical applications.15–17 Among the varieties of layered transition metal sulfides, tungsten sulfide (WS2) is composed of three atomic layers: a tungsten layer is sandwiched between two sulfur layers, and the triple layers are stacked and held together by weak van der Waals interactions. It is an indirect band gap semiconductor with an energy gap of ∼1.2 eV in bulk form.16,18 Due to its analogous structure to graphene and potential electrocatalytic activity, WS2 is considered as versatile for electrochemical applications. For example, WS2 is extensively used for variety of applications such as solid state lubricants, catalysts for hydride sulfurization and hydrogen evolution reaction and anode materials for lithium ion batteries.19–21 However, only a few reports have so far focused on the application of WS2 as CE in DSSCs. Wu et al.11 synthesized molybdenum sulphide (MoS2) and WS2 by a hydrothermal method. They employed them as CE materials for DSSCs and achieved PCEs of 7.59% and 7.73%, respectively. Recently, Yue et al.22 reported DSCCs with a PCE of 6.41% under a simulated solar illumination of 100 mW cm−2 (AM 1.5) using multi-wall carbon nanotubes decorated with WS2 (MWCNTs-WS2) CE, which were obtained using a hydrothermal method. However, to avoid the complex solution-based preparation processes which are not eco-friendly, development of a pure WS2 thin film directly grown onto a conductive substrate deserves further exploration. In this study, we present a scalable and reproducible RF sputtering method for the fabrication of a large area, continuous WS2 film onto fluorine-tin-oxide (FTO)/glass substrate. The TiO2-based DSSCs with an optimized WS2 CE exhibited an impressive 6.3% PCE, which is comparable to DSSCs result with a Pt CE (6.64%).
Experimental details and device fabrication
Pieces of FTO/glass substrates (∼2 × 2 cm2) were ultrasonically degreased in acetone, methanol, isopropyl alcohol and deionized water and then baked at 120 °C for 5 min. After loading the pieces in a sputter chamber, the chamber was evacuated by a rotary pump and a turbomolecular pump combination to achieve the pressure at ∼1 × 10−7 torr. Next, WO3 thin films were deposited onto FTO/glass substrates using a W target by RF magnetron sputtering. During the film deposition, the Ar and O2 flow ratio was maintained at 5 and 15 sccm, respectively and the RF power was fixed at 180 W. The substrate temperature was fixed at room temperature. The sputtering time was varied (5, 10, 15 min) to control the thickness of WS2 film. After taking out the samples from the sputter chamber, the as-sputtered films were placed downstream of a chemical vapor deposition (CVD) chamber and heated. The as-sputtered films were subjected to sulfurization process at 650 °C for 30 min to form WS2 film and also to improve the crystalline quality of the films. A pure sulfur powder (99.99%) was placed upstream of the CVD chamber, and a heating filament for the sulfur boat was fixed at 120 °C. The sulfur powder was evaporated at 120 °C using a carrier gas, Ar (100 sccm), and the pressure of the CVD chamber was kept at 2 × 10−2 mTorr. The distance between the sulfur boat and the substrate and the flow of the carrier gas were fixed for all the experiments. The amount of the sulfur powder was optimized at 0.5–0.8 g for the controlled growth of WS2 film.
2.1 Fabrication of DSSCs using Pt CEs
A Pt CE was prepared by using a drop-cast method, as reported previously.23,24 In brief, a thin TiO2 blocking layer with 100 nm thickness was deposited onto a FTO/glass substrate. A 0.2 M TiCl4 (20–30 wt% in HCl) was prepared using H2O2 at 50–55 °C and crystallized at 350 °C for 60 min. After preparation of the solution, few drops were placed onto the FTO/glass substrate and put on a hot plate at 70 °C for 45 min. A TiO2 layer (thickness ∼4 μm) with an average particle size of ∼20 nm was coated by the doctor blade method. A light scattering TiO2 layer (thickness ∼3 μm) was applied over the previously deposited TiO2 and sintered again at 450 °C for 30 min. The prepared TiO2 photoanode was dipped in 0.3 mM N719 (ethanol + acetonitrile) dye for 24 h so as to form working electrode for DSSCs. The DSSCs were fabricated by injecting a liquid electrolyte (0.005 M I2, 0.1 M Lil, 0.6 M tetrabutylammonium iodide, and 0.5 M 4-tert-butylpyridine in acetonitrile) through an aperture between the dye-sensitized TiO2 electrode and the CEs (MoS2 & Pt in the present case).
2.2 Characterization techniques
The phase purity analysis and structure confirmation of WS2 CEs were studied using a Raman microscope at 514 nm wavelength. The structural formation and crystal orientation were verified by using X-ray diffraction (XRD, Rigaku) with Cu-Kα radiation operated at 50 kV and 300 mA. The morphological evolution of WS2 films as a function of RF sputtering time was confirmed from the field-emission scanning electron microscopy (FE-SEM, Nova nano SEM200-100 FEI) images. Chemical configurations were determined by X-ray photoelectron spectroscopy (XPS), model PHI 5000 Versa Probe (Ulvac-PHI) using a monochromatic Al Kα X-ray source (1486.6 eV). The data were collected from a spot-size of 100 × 100 μm2. The carbon 1s peak (284.6 eV) was used as a reference for internal calibration.
2.3 Photo-electrochemical measurements
A solar simulator (150 W Xe lamp, Sun 2000 solar simulator, ABET 5 Technologies, USA) equipped with an AM 1.5G filter was used to generate simulated sunlight with a corrected intensity of 1 sun (100 mW cm−2). The photocurrent density–applied voltage (J–V) spectra were obtained using a Keithley 2400 source meter. Incident photon to current conversion efficiency (IPCE) spectra were obtained without bias potential under illumination with respect to a calibrated Melles-Griot silicon diode, and by changing the excitation wavelength (McScience K3100 spectral IPCE measurement system and Polaronix® K102 signal amplifier). Electrochemical impedance spectroscopy (EIS) analysis was performed by IviumStat Technologies, Netherlands. The frequency of applied sinusoidal AC voltage signal varied from 150 kHz to 0.1 MHz.
Results and discussion
The schematic diagram of the WS2 RF sputtering preparation process and DSSCs module is illustrated in Fig. 1.
 |
| Fig. 1 Schematic diagram for the preparation of WS2 CE and its role in DSSCs. | |
Fig. 2a–c shows optical microscopy images of WS2 films obtained at different sputtering times. The observed films seem to be uniform and continuous. Additionally, the films become darker (thicker) with increasing the sputtering time. Also, 3D images of prepared WS2 films for different sputtering time are presented in Fig. (S1).†
 |
| Fig. 2 (a–c) Optical images of WS2 films obtained at different sputtering times; (a) 5, (b) 10 and (c) 15 min and (d) the Raman spectra of the corresponding WS2 films (sputtered for 5, 10 and 15 min). | |
Raman analysis was performed to perceive the crystal quality of WS2. Raman spectra of the annealed WS2 thin films are shown in Fig. 2d. From Fig. 2d, two characteristic WS2 Raman peaks related to the in-plane vibration of W and sulfur atoms, E2g1 mode at ∼354.01 cm−1 and out-of-plane vibration of sulfur atoms, A1g mode at ∼418.4 cm−1, were obtained. The frequency difference (Δk) between two Raman peaks is closely related with the layer number and can be used to determine the thickness of atomically thin WS2.25,26 The Δk increased with increasing sputtering time. The obtained Δk were ∼64.3 cm−1, ∼65 cm−1 and ∼65.5 cm−1 for 5 min, 10 min and 15 min sputter time, respectively, corresponding to few-layer WS2 films [Fig. 2d].
The crystallinity and purity of the WS2 thin films were analysed by XRD analysis. From the Fig. 3, all the reflections were corresponding solely to WS2 and substrate. The diffraction pattern revealed intense peaks at the positions of 37.2°, 42.9°, 51.1°, 54.6°, 61.6°, 65.7°, 70.9°, 78.3°, 80.5°, 83.5° and 87.5°, which were assigned to (103), (006), (105), (106), (107), (114), (202), (109), (205), (206) and (118) lattice planes of WS2 crystal structure, respectively, evidencing that there was no trace of WO3 after sulfurization process. The substrate-related FTO peaks were clearly noticed in the XRD patterns. Presence of intense XRD peaks was an indication of the polycrystalline nature and the resultant diffraction peaks corroborated well with the standard patterns for the hexagonal crystal structure of WS2 (JCPDS # 87-2417 file). The XRD results demonstrated no other phase transformation or other impurities in the crystalline film. The observed XRD reflections were highly sharp, indicating the high crystallinity of the WS2 film surface. No other significant variations were observed in the XRD patterns for WS2 films prepared at different sputtering time periods.
 |
| Fig. 3 XRD patterns of WS2 films sputtered at different sputtering times. | |
Surface morphology acts as a key role in the development of thin film solar cells.27 FE-SEM was used to see the surface morphology of the WS2 material over FTO/glass substrate [Fig. 4a–c]. Some crystallites larger than the normal due to the agglomeration process were detected. The valleys and hillocks were noticed over the sample prepared at 5 min-sputtering time [Fig. 4a]. The various sizes and shapes of crystallites were noticed for WS2 sputtered for 10 min [Fig. 4b]. The highest crystallite sizes were found for WS2 film sputtered for 15 min (Fig. 4c). The elemental mapping graph of WS2 (Fig. 4d) confirmed the presence of Sn (from substrate), tungsten, and sulfur; consistent to XPS results presented below.
 |
| Fig. 4 FE-SEM micrographs of WS2 CE sputtered for; (a) 5, (b) 10, and (c) 15 min. (d) Elemental mapping of WS2 (sputter time 10 min). | |
To further study the chemical composition and surface electronic states of WS2, XPS analysis was carried out (sample sputtered for 10 min). The survey XPS spectra (Fig. 5a) indicate the presence of sulfur, carbon, and oxygen etc., elements in the WS2 film. The C1s peak, related to sp3 hybridized carbon, was observed at 284.8 eV [Fig. 5b]. Fig. (5c and d) shows the XPS spectrum of WS2 film. The W 4f core-level spectrum showed three peaks corresponding to the W 4f5/2 and W 4f7/2 levels for W4+ and W 5p3/2 electronic states at 34.8, 32.7, and 38.0 eV, respectively. The S 2p core-level spectrum exhibited two peaks at 163.5 and 162.3 eV, corresponding to the S 2p1/2 and S 2p3/2 states, respectively. These results are well-consistent with the reported values for WS2 single crystals.
 |
| Fig. 5 XPS spectra of WS2 film sputtered at 10 min; (a) survey spectra, (b) carbon related C1s peak, (c) tungsten doublet peak (W4+), and (d) sulphur (S2−). | |
In this study, the fabricated transparent WS2 CEs were applied for DSSCs (Fig. 1). Fig. S2† shows J–V characteristic curves of DSSCs with WS2 CEs with different thicknesses. The obtained photovoltaic performance parameters are summarized in Table S1.† The observed PCEs (η) of TiO2-based DSSCs with WS2 CEs sputtered for 5, 10 and 15 min were 5.4%, 6.3% and 5.8%, respectively. The highest efficiency was obtained from WS2 CE sputtered at 10 min. J. Wu et al.28 demonstrated DSSCs with WS2 CE, but the PCE of DSSCs based on pristine WS2 CE was limited to 5.32%. The PCE could be enhanced by inclusion of high-conductive MWCNT along with WS2 in the presence of glucose so called as G-A WS2/MWCNT hybrid CE. The PCEs of our WS2-based electrode were also superior to that of the previously reported plastic and carbon-coated WS2 CEs.29
Fig. 6a presents J–V characteristic curves of DSSCs with Pt and WS2 CEs under illumination of 100 mW cm−2. J–V curves of two DSSCs are nearly rectangular and identical, demonstrating an excellent catalytic behaviour of WS2 as like Pt. DSSCs employing WS2 CE exhibits PCE of 6.3%, short-circuit current (Jsc) of 13.43 mA cm−2, open-circuit voltage (Voc) of 0.71 V, and fill factor (FF) of 0.66 [Table 1]. The IPCE curves of TiO2-based DSSCs with Pt and WS2 CEs (Fig. 6b) also present a similar behavior, suggesting that almost similar (to Pt) catalytic activity is obtained by the transparent WS2 CE.
 |
| Fig. 6 (a) J–V characteristic curves, (b) IPCE measurements of TiO2-based DSSCs with WS2 and Pt CEs. | |
Table 1 DSSCs parameters of WS2 (sputtered at 10 min) and Pt-CEs
|
DSSCs type |
Jsc (mA cm−2) |
Voc (V) |
FF (%) |
η (%) |
1 |
TiO2–Pt CE |
16.50 |
0.66 |
0.61 |
6.64 |
2 |
TiO2–WS2 CE |
13.43 |
0.71 |
0.66 |
6.3 |
The PCE value for WS2 CE is close to that obtained for Pt CE (6.64%). The slight difference in the Voc and Jsc values for the DSSCs with Pt and WS2 CE can be attributed to several factors i.e., (a) surface roughness difference, (b) conductivity, (c) catalytic activity, and (d) chemical stability etc. Electrocatalytic activity is highly dependent on catalyst morphology, grain size, surface texture, crystalline structure and electrical conductivity of CE.30,31 The PCE dependence in DSSCs with different WS2 film thicknesses (sputter time periods) could be attributed to the different morphologies of WS2 films obtained for different time periods (Fig. 4). The catalytic activity of WS2 originates from the sulfur edges as reported previously. It is believed that with increasing the sputtering time more active sulfur edges are formed which ultimately enhances the electro-catalytic activity.30,31
In this work, transparent (few-layers) WS2 electrode was used for the first time as the CE material in DSSCs. To estimate the electrocatalytic performance of the WS2 towards oxidation and reduction, cyclic voltammetry (CV) studies were performed. Fig. 7a shows the CVs of the system for Pt and WS2 in the potential interval between −0.8 and 0.8 V. In this process, electrons are injected into photo-oxidized dye from I− ions in the electrolyte [eqn (1)], and the produced I3− ions are reduced on the CE [eqn (2)]. In this interval, the anodic peak is contributed by the reaction (1) in an anodic sweep and the cathodic peak is contributed by the reaction (2) in a cathodic sweep, respectively.
 |
| Fig. 7 Cyclic voltammetry curves of; (a) I−/I3− redox system using Pt and WS2 CEs, (b and c) WS2 CE at different scan rates and durability test for 100 consecutive cycles at a 10 mV s−1 scan rate. (d) Tafel polarization curves of symmetrical cells obtained with Pt and WS2 CEs. | |
More CV tests of WS2 CE for the
redox reaction were carried out by changing scan rates (25–150 mV s−1) as shown in Fig. 7b. It was observed that the peak current densities were gradually increased with the increasing the scan rate. Simultaneously, the diffusion coefficient D of Pt and WS2 CEs can be estimated from the Randles–Sevcik equation:
|
Ip = Kn3/2AC(D)1/2V1/2
| (3) |
where
Ip represents the peak current,
K is the constant of 2.69 × 10
5,
n is the number of electrons transferred in the redox event,
A is the electrode area,
C represents the bulk concentration of I
3− species,
D is the diffusion coefficient and
V means the scan rate. The diffusion coefficient of I
3− for the Pt (∼7.56 × 10
−5) is larger than that for the WS
2 CE (∼4.98 × 10
−5), presumably arising from the high active surface area and surface roughness difference of both electrodes.
To investigate the electrochemical stability of the WS2 CE, 100 consecutive CV cycles were performed in the potential region from −0.8 to 0.8 V vs. Ag/AgCl [Fig. 7c]. After 100 consecutive scans, the CV shape of WS2 was changed slightly and the current densities of its redox peaks remained stable, suggesting that the WS2 CE possessed the characteristics of reversible activity, excellent electrochemical, chemical stability and strong adhesiveness on the FTO glass substrate. TiO2 was used as a dye material and plays an important role due to many advantages such as having the most efficient photoactivity, the highest stability and the lowest cost.32,33 Two types of photochemical reactions occur on the TiO2 surface when irradiated with ultraviolet light. One includes the photo-induced redox reactions of adsorbed substances, and the other is the photo-induced hydrophilic conversion of TiO2 itself.32,34
Tafel polarization measurements were performed on the symmetrical cells using WS2 (10 min sputter time) and Pt CEs, and the corresponding curves are shown in Fig. 7d. Also, Tafel polarization measurements of WS2 obtained for different deposition time periods are given in Fig. (S4).† The tangent slope of Tafel curves provides the information about the exchange current density (Jo). The commercial Pt electrode has larger Jo compared with sputtered WS2 electrode, suggesting its higher electrocatalytic activity and lower charge-transfer resistance at the electrolyte–electrode interface.
Furthermore, EIS spectrums for symmetrical cells were carried out to examine the overall cell resistance. The Nyquist plots (Fig. 8a) for the symmetric cells with the Pt and WS2 CEs illustrate impedance characteristics. The semicircle in the high frequency region (left) is due the charge-transfer process of the electrolyte–counter electrode interface, which changes inversely with the catalytic activity of the CE on the reduction of tri-iodide, and the corresponding constant phase element (CPE). The one in the low frequency region (right) is assigned to the Nernst diffusion process of tri-iodine ions. The impedance values (calculated using Z-View curve fitting software) are presented in Table 2. The Rct value of the WS2 and Pt CEs is 7.2 and 5.0 Ω cm2, respectively. In addition, Rs is another important factor affecting the fill factor and performance of DSSCs.35 Due to semiconducting behaviour, the WS2 CE showed relatively higher Rs value than Pt CE.36,37 Higher Jsc value in Pt CE DSSCs could be due to an availability of its larger active surface area for
redox reaction.38
 |
| Fig. 8 Nyquist plots for (a) Pt and (b) WS2 CEs with inset of (b) as an equivalent circuit. | |
Table 2 EIS fitting parameters used for WS2 and Pt CEs
CE |
Pt |
WS2 |
Rs (Ω cm2) |
3.1 |
19.5 |
R1 (Ω cm2) |
8.11 |
13.39 |
CPE1−t × 10−3 |
2.45 |
7.29 |
CPE1−p |
0.88 |
0.90 |
R2 (Ω cm2) |
1.96 |
2.11 |
CPE2−t × 10−3 |
0.98 |
0.50 |
CPE2−p |
0.60 |
0.096 |
Conclusion
In summary, continuous and large-area transparent WS2 films deposited onto FTO substrate by RF magnetron sputtering were used as a low-cost and platinum-free counter electrode materials for dye-sensitized solar cells. X-ray diffraction and XPS results have verified the phase purity and surface composition of WS2 deposited onto FTO substrate. CV, EIS and Tafel curve measurements indicated that the WS2-based CE hold good electrocatalytic activity and fast reaction kinetics for the
redox reaction. Finally, the DSSC with an optimized WS2 CE (sputtered for 10 min) showed as high as 6.3% PCE, which is comparable to the performance of DSCC with Pt CE (6.64%). The present work demonstrated the feasibility to develop low-cost, transparent, efficient and abundant metal chalcogenide electrocatalysts (WS2) with high catalytic activity and stability by RF sputtering method to replace Pt electrocatalyst for photovoltaic applications.
Acknowledgements
This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Education (2010-0020207, 2012R1A1A2007211), by the Human Resources Development of the Korea Institute of Energy Technology Evaluation and Planning (KETEP) grant funded by the Korea Government Ministry of Trade, Industry & Energy (No. 20154030200630), and by the MOTIE (Ministry of Trade, Industry & Energy) (10052928) and KSRC (Korea Semiconductor Research Consortium) support program for the development of the future semiconductor device. Authors, RSM and MN extend their gratitude to the Visiting Professor Program (VPP) Unit of King Saud University (KSU), Saudi Arabia for the financial support.
References
- B. O'regan and M. Grätzel, Nature, 1991, 353, 737–740 CrossRef.
- E. Stathatos, P. Lianos, U. Lavrencic-Stangar and B. Orel, Adv. Mater., 2002, 14, 354 CrossRef CAS.
- A. Yella, H.-W. Lee, H. N. Tsao, C. Yi, A. K. Chandiran, M. K. Nazeeruddin, E. W.-G. Diau, C.-Y. Yeh, S. M. Zakeeruddin and M. Grätzel, Science, 2011, 334, 629–634 CrossRef CAS PubMed.
- E. Olsen, G. Hagen and S. E. Lindquist, Sol. Energy Mater. Sol. Cells, 2000, 63, 267–273 CrossRef CAS.
- A. Kay and M. Grätzel, Sol. Energy Mater. Sol. Cells, 1996, 44, 99–117 CrossRef CAS.
- M. J. Ju, J. C. Kim, H.-J. Choi, I. T. Choi, S. G. Kim, K. Lim, J. Ko, J.-J. Lee, I.-Y. Jeon and J.-B. Baek, ACS Nano, 2013, 7, 5243–5250 CrossRef CAS PubMed.
- H. Choi, H. Kim, S. Hwang, W. Choi and M. Jeon, Sol. Energy Mater. Sol. Cells, 2011, 95, 323–325 CrossRef CAS.
- T. Chen, L. Qiu, Z. Cai, F. Gong, Z. Yang, Z. Wang and H. Peng, Nano Lett., 2012, 12, 2568–2572 CrossRef CAS PubMed.
- L. Kavan, J.-H. Yum and M. Grätzel, Nano Lett., 2011, 11, 5501–5506 CrossRef CAS PubMed.
- C.-T. Li, C.-P. Lee, Y.-Y. Li, M.-H. Yeh and K.-C. Ho, J. Mater. Chem. A, 2013, 1, 14888–14896 CAS.
- M. Wu, Y. Wang, X. Lin, N. Yu, L. Wang, L. Wang, A. Hagfeldt and T. Ma, Phys. Chem. Chem. Phys., 2011, 13, 19298–19301 RSC.
- G. R. Li, F. Wang, Q. W. Jiang, X. P. Gao and P. W. Shen, Angew. Chem., Int. Ed., 2010, 49, 3653–3656 CrossRef CAS PubMed.
- A. Hauch and A. Georg, Electrochim. Acta, 2001, 46, 3457–3466 CrossRef CAS.
- M. Wu, J. Bai, Y. Wang, A. Wang, X. Lin, L. Wang, Y. Shen, Z. Wang, A. Hagfeldt and T. Ma, J. Mater. Chem., 2012, 22, 11121–11127 RSC.
- J. N. Coleman, M. Lotya, A. O'Neill, S. D. Bergin, P. J. King, U. Khan, K. Young, A. Gaucher, S. de and R. J. Smith, Science, 2011, 331, 568–571 CrossRef CAS PubMed.
- Y. Zhang, Y. Zhang, Q. Ji, J. Ju, H. Yuan, J. Shi, T. Gao, D. Ma, M. Liu and Y. Chen, ACS Nano, 2013, 7, 8963–8971 CrossRef CAS PubMed.
- D. Voiry, H. Yamaguchi, J. Li, R. Silva, D. C. Alves, T. Fujita, M. Chen, T. Asefa, V. B. Shenoy and G. Eda, Nat. Mater., 2013, 12, 850–855 CrossRef CAS PubMed.
- A. L. Elías, N. Perea-López, A. Castro-Beltrán, A. Berkdemir, R. Lv, S. Feng, A. D. Long, T. Hayashi, Y. A. Kim and M. Endo, ACS Nano, 2013, 7, 5235–5242 CrossRef PubMed.
- R. Bhandavat, L. David and G. Singh, J. Phys. Chem., 2012, 3, 1523–1530 CAS.
- A. Sobczynski, A. Yildiz, A. J. Bard, A. Campion, M. A. Fox, T. Mallouk, S. E. Webber and J. M. White, J. Phys. Chem., 1988, 92, 2311–2315 CrossRef CAS.
- L. Rapoport, Y. Bilik, Y. Feldman, M. Homyonfer, S. Cohen and R. Tenne, Nature, 1997, 387, 791–793 CrossRef CAS.
- G. Yue, J. Wu, J.-Y. Lin, Y. Xiao, S.-Y. Tai, J. Lin, M. Huang and Z. Lan, Carbon, 2013, 55, 1–9 CrossRef CAS.
- S. B. Ambade, R. B. Ambade, R. S. Mane, G.-W. Lee, S. F. Shaikh, S. A. Patil, O.-S. Joo, S.-H. Han and S.-H. Lee, Chem. Commun., 2013, 49, 2921–2923 RSC.
- X. Xin, M. He, W. Han, J. Jung and Z. Lin, Angew. Chem., Int. Ed., 2011, 50, 11739–11742 CrossRef CAS PubMed.
- H. Li, Q. Zhang, C. C. R. Yap, B. K. Tay, T. H. T. Edwin, A. Olivier and D. Baillargeat, Adv. Funct. Mater., 2012, 22, 1385–1390 CrossRef CAS.
- C. Lee, H. Yan, L. E. Brus, T. F. Heinz, J. Hone and S. Ryu, ACS Nano, 2010, 4, 2695–2700 CrossRef CAS PubMed.
- G. Yue, J.-Y. Lin, S.-Y. Tai, Y. Xiao and J. Wu, Electrochim. Acta, 2012, 85, 162–168 CrossRef CAS.
- J. Wu, G. Yue, Y. Xiao, M. Huang, J. Lin, L. Fan, Z. Lan and J.-Y. Lin, ACS Appl. Mater. Interfaces, 2012, 4, 6530–6536 CAS.
- Y. Wang, S. Li, Y. Bai, Z. Chen, Q. Jiang, T. Li and W. Zhang, Electrochim. Acta, 2013, 114, 30–34 CrossRef CAS.
- B. Hinnemann, P. G. Moses, J. Bonde, K. P. Jørgensen, J. H. Nielsen, S. Horch, I. Chorkendorff and J. K. Nørskov, J. Am. Chem. Soc., 2005, 127, 5308–5309 CrossRef CAS PubMed.
- T. F. Jaramillo, K. P. Jørgensen, J. Bonde, J. H. Nielsen, S. Horch and I. Chorkendorff, Science, 2007, 317, 100–102 CrossRef CAS PubMed.
- M. Thomalla and H. Tributsch, J. Phys. Chem. B, 2006, 110, 12167–12171 CrossRef CAS PubMed.
- M. Shanmugam, T. Bansal, C. A. Durcan and B. Yu, Appl. Phys. Lett., 2012, 100, 153901 CrossRef.
- N. G. Park and K. Kim, Phys. Status Solidi A, 2008, 205, 1895–1904 CrossRef CAS.
- P. Joshi, L. Zhang, Q. Chen, D. Galipeau, H. Fong and Q. Qiao, ACS Appl. Mater. Interfaces, 2010, 2, 3572–3577 CAS.
- L. Han, N. Koide, Y. Chiba, A. Islam, R. Komiya, N. Fuke, A. Fukui and R. Yamanaka, Appl. Phys. Lett., 2005, 86, 213501 CrossRef.
- N. Koide, A. Islam, Y. Chiba and L. Han, J. Photochem. Photobiol., A, 2006, 182, 296–305 CrossRef CAS.
- C. H. Yoon, R. Vittal, J. Lee, W.-S. Chae and K.-J. Kim, Electrochim. Acta, 2008, 53, 2890–2896 CrossRef CAS.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra18613g |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.