Preparation of CoS2 nanoflake arrays through ion exchange reaction of Co(OH)2 and their application as counter electrodes for dye-sensitized solar cells

Jung-Che Tsaia, Min-Hsiung Honab and Ing-Chi Leu*c
aDepartment of Materials Science and Engineering, National Cheng Kung University, Tainan, 701 Taiwan, Republic of China
bResearch Center for Energy Technology and Strategy, National Cheng Kung University, Tainan 701, Taiwan, Republic of China
cDepartment of Materials Science, National University of Tainan, Tainan, 701 Taiwan, Republic of China. E-mail: icleu@mail.mse.ncku.edu.tw

Received 16th September 2014 , Accepted 8th December 2014

First published on 8th December 2014


Abstract

The cobalt sulfide nanoflake arrays prepared by the transformation of Co(OH)2 nanoflake arrays using the ion exchange reaction method were incorporated into Pt-free dye-sensitized solar cells (DSSCs). Morphologies and crystal structures of the cobalt sulfide and cobalt hydroxide nanoflakes arrays were characterized by SEM, TEM and XRD analyses. The electrochemical properties were determined by cyclic voltammetry (CV) measurement. The cobalt sulfide nanoflakes which were composed of CoS2 single crystals and their aggregates dispersing in the amorphous cobalt sulfide matrix were completely transferred by substitution of S2− for O2− in the ion exchange reaction. The DSSC assembled with cobalt sulfide nanoflake arrays as the counter electrode showed a photovoltaic conversion efficiency of 5.20%, which was close to that of DSSC with sputtered Pt as the counter electrode (5.34%). Therefore, the cobalt sulfide nanoflake array film can be considered as a promising alternative counter electrode for use in DSSCs due to its large surface area and high electrocatalytic performance.


Introduction

Dye-sensitized solar cells (DSSCs) have received a lot of attention due to their low cost, relatively high performance and easy fabrication process.1 Generally, a typical DSSC consists primarily of a photoanode, an electrolyte and a counter electrode. A conventional photoanode used is a dye-sensitized porous titania (TiO2) on FTO (SnO2:F) coated glass. During the photovoltaic process, photoelectrons generated from the excited state dyes are injected into the conduction band of TiO2 and further transferred across the external circuit to the counter electrode. Meanwhile, the remained holes in oxidized dyes are regenerated by I in the electrolyte. Then the resultant I3 ions diffuse to the surface of counter electrode and reduce to I by receiving the external circuit electron for completing the continuous loops of dye regeneration and photoelectron injection. The most common material for counter electrodes is platinum (Pt) deposited on FTO coated glass due to its superior electrocatalytic activity for the I3/I redox couple, the electrical conductivity and reliability.2 However, the high price noble metal Pt as the counter electrode restricts DSSCs large-scale manufacture and their widespread application. So far, many reports have been carried out to search for cheaper and effective substitutes for Pt in counter electrodes.

Carbonaceous materials (carbon black,3 graphene,4 carbon nanotubes,5 etc.) and conductive polymers (PEDOT,6 polypyrrole,7 etc.) were investigated as counter electrode materials of DSSCs for their good catalytic capability and conductivity. Some catalytically active transition metal compounds such as carbides,8 nitrides,9 oxides10 and various hybrids3 have been shown to have a great potential to replace Pt due to the intrinsic catalytic activity for I3 reduction and the various morphologies to offer high specific surface. Compared to carbonaceous or polymer materials, inorganic materials have unique properties such as low cost, material diversity, abundance, high catalytic activity and ease of modification. Recent reports showed that some catalytic metal sulfides such as MoS2,11 WS2,11 NiS,12 and CoS13 also can serve as efficient catalytic counter electrodes of DSSCs owing to their outstanding catalytic capability and abundant feedstock. In 2009, CoS electrodeposited on the flexible substrate showed comparable catalytic activity to Pt as the counter electrodes, which opened up the way to exploring more transition meal sulfides.13 It is well known that a high specific surface area is beneficial for providing more active sites of counter electrodes. The hierarchical CoS spindles in the counter electrodes in DSSCs have been reported to have a high catalytic activity.14 However, it is known that the electrochemical properties are determined by the kinetic features controlled by the transport of electrons and ions into the active materials. It is an effective approach to enhance the kinetic features by directly constructing the nanoarray electrodes on the collector (FTO substrate). The deposited architecture provides an effective diffusion path for ions, and a lower internal resistance to realize the subsequent high-power performance. A large variety of approaches have been reported for synthesis of cobalt sulfides, such as electrochemical deposition,13 thermal decomposition method,15 hydrothermal method,16 and solvothermal method.17 These methods are suitable for powder materials, but difficult to directly synthesize oriented cobalt sulfide nanostructures on the FTO. Therefore, it is of great interest to researchers to develop a facile synthesis method for cobalt sulfide nanostructures for energy applications.

The solution-based ion exchange reaction (IER) has been proven to be a low-cost and effective method for chemical transformation of nanomaterials into dedicated composition and morphology.18 The solution-based IER avoids the disadvantages of high temperatures, high pressure and toxic gas sources such as H2S. At the same time, it reduces the danger and cost of the fabrication process. Conventionally, the different nanostructures of metal sulfides such as nickel sulfide or cobalt sulfide have been derived from nanostructures of metal oxides or hydroxides.19 During the IER process, cobalt sulfide nanostructures were transformed from Co3O4 nanostructures and maintained the morphologies of metal oxides.19 Therefore, IER is a promising technique to synthesize the metal sulfides for energy applications. To date, there are only a limited number of reports on nanoflakes of cobalt sulfide grown directly on FTO substrate because their cobalt hydroxide precursor is usually synthesized in solution or on the nickel foam. In the present work, cobalt hydroxide is used as the skeleton to be transformed into cobalt sulfide via the IER method. As a result of shape-preserved reaction, the CoS2 nanoflake arrays were fabricated from Co(OH)2 nanoflake arrays and retained strong contact with the FTO substrate. It is found that the CoS2 nanoflake arrays exhibit outstanding I/I3 redox catalytic activity and a power conversion efficiency of 5.2% is achieved for the DSSC with CoS2 nanoflake arrays as the counter electrode, which is nearly as that of the DSSC with sputtered Pt as the counter electrode (5.34%).

Experimental

Chemicals

Cobalt acetate 9-hydrate (Co(CH3COO)2·9H2O, 98%), cobalt sulfate (CoSO4·9H2O, 99%), sodium sulfide 9-hydrate (Na2S·9H2O, 98%), and hexamethylenetetramine (C6H12N4, HMTA, 99%) were obtained from Sigma-Aldrich. Titania nanoparticles (P25, size: ∼21 nm, 20% rutile and 80% anatase) and di-tetrabutylammonium cis-bis(isothiocyanato)bis(2,2′-bipyridyl-4,4′-dicarboxylato)ruthenium(II) (N719, C58H86N8O8RuS2, 99.9%) were purchased from UniRegion Bio-Tech, Inc. Anhydrous lithium iodide (LiI, >98%), lithium perchlorate, (LiClO4, >99.8%), iodine (I2, >98%), tert-butanol (C4H10O, 99.9%) and acetonitrile (CH3CN, ACN, 99.9%) were obtained from Alfa Aesar. 1.3-dimethylimidazoliumiodine (C5H9IN2, 99.8%), 4-tert-butylpyridine (C9H13N, 99.8%), guanidine thiocyanate (CH6ClN3, 99.8%) were purchased from Merck. Ethanol (99.5%) was purchased from J. T. Baker. Deionized water was used throughout the work.

Preparation of CoS2 nanoflake array films

The experimental details are as follows. Clean FTO glasses with 1 cm × 2 cm in size were used as substrates. To facilitate the nucleation of Co(OH)2 nanoflake arrays, a 5 nm-thick Co3O4 thin film (seed layer) was first deposited on the FTO substrate by spin coating a 5 mM ethanolic solution of cobalt acetate, followed by conversing to Co3O4 at 350 °C for 30 min in air.20 In a hydrothermal synthesis of Co(OH)2 nanoflake array film, 0.1 M cobalt acetate and 0.3 M hexamethylenetetramine were dissolved in 20 mL of deionized water to form a homogeneous solution with pink color, which was transferred into the teflon-lined stainless steel autoclave. After that, the pieces of FTO glass substrate were immersed into the reaction solution with an angle against the wall. Then, the autoclave was sealed and maintained at 120 °C for 0.5–5 h in an electric oven. After cooling down to room temperature naturally, the samples were fetched out and washed with deionized water. To transform the hydroxide into cobalt sulfide, the ion exchange reaction was performed on the samples by rinsing in 0.1 M sodium sulfide aqueous solution at 75 °C for 8 h.

Fabrication of the DSSCs

The TiO2 paste prepared by mixing P25, ethanol, DI-water and X100 was coated on a cleaned FTO substrate by the doctor-blade method. The TiO2 electrode (∼20 μm) was gradually heated to 450 °C (rate = 5 °C min−1) and subsequently sintered at that temperature for 30 min. After cooling it to 80 °C, the electrode was dipped in a solution containing 0.3 mM N719 dye in a mixed solvent of acetonitrile and tert-butanol in a volume ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]1 for a day. The dye-sensitized TiO2 electrode was coupled with the various cobalt sulfide electrodes (or sputtered Pt electrode for comparison) to fabricate the DSSCs. These two electrodes were separated by a 60 μm thick thermal plastic Surlyn film and sealed by heating. A mixture of 1 M 1.3-dimethylimidazoliumiodine, 0.5 M 4-tert-butylpyridine, 0.15 M iodine, and 0.1 M guanidine thiocyanate in ACN was used as the electrolyte and injected into the gap between the two electrodes by capillarity. An area of 0.237 cm2, defined by a metal mask, was selected as the active region for the photovoltaic measurement. The symmetrical cell was composed of two identical electrodes and its electrochemistry measurement had the same conditions and arrangement as the DSSCs.

Characterization

The surface morphology and composition of cobalt hydroxide and cobalt sulfide were characterized by field-emission scanning electron microscopy (FESEM, Auriga 39-50, Zeiss) equipped with an energy dispersive X-ray spectrometer (EDX, XFlash 5030, Bruker). X-ray diffraction experiments were carried out with an automatic X-ray powder diffractometer (D/MAX2500, Rigaku). The transmission electron microscopy (TEM) images were obtained from a high-resolution analysis electron microscope (HRAEM, JEM-2100F CS STEM, JEOL) with an accelerating voltage of 200 kV. Cyclic voltammetry (CV) measurement was carried out in a three-electrode system in an acetonitrile solution contained 0.1 M LiClO4, 10 mM LiI and 1 mM I2 at a scanning rate of 50 mV s−1 using an electrochemical analyzer (263A, EG&G). Pt worked as the counter electrode and Ag/Ag+ worked as the reference electrode. Tafel-polarization curves of the symmetrical cells were measured by the same electrochemical analyzer. The photovoltaic performance of the DSSCs was studied under simulated AM 1.5 illumination (100 mW cm−2, Sol2A Class ABA Solar Simulators, Oriel) with a digital source meter (Model 2400, Keithley). In addition, the incident light illuminated from the photoanode (working electrode) side and its intensity was calibrated with a standard Si-based solar cell (9115DV, Oriel).

Results and discussion

The cobalt sulfide nanoflake arrays are transformed from the Co(OH)2 nanoflake arrays first prepared on the FTO glass, therefore it is important to know the growth characteristics of the Co(OH)2 nanoflake arrays. To investigate the growth of Co(OH)2 nanoflake arrays, the hydrothermal synthesis method was kept at 120 °C for different growth time. The reactions involved in the hydrothermal synthesis of Co(OH)2 nanoflake arrays can be illustrated as follows.21
 
C6H12N4 + 6H2O → 6HCHO + 4NH3 (1)
 
NH3 + H2O ↔ NH4+ + OH (2)
 
Co2+ + 2OH → Co(OH)2 (3)

At the growth time of 0.5 h, the random nucleated cobalt hydroxide reveals dark spots on the FTO surface, which is shown in Fig. 1(b) SEM image. When HMTA is continuously thermally decomposed to generate more OH, the tiny flake structures are grown from the nucleate sites. After prolonging the growth time, neighboring flakes become larger and are bound together by weak van der Waal force to form the nanoflake arrays. Finally, the surface of FTO was covered by the Co(OH)2 nanoflake arrays. The as-deposited film is uniform in appearance and exhibits bluish-green in color. It is reported in the literature that cobalt hydroxides prefer to grow into nanoflakes which align vertically to the substrate due to their intrinsic lamellar structure.22 Because Co(OH)2 has a layered brucite crystal structure of the CdI2 type, which shows a weak interaction between layers and strong binding within the layered planes.16 The two-dimensional monolayer consists of edge-shared CoO6 octahedrons.23 Therefore, (001) plane of Co(OH)2 is stable with the lowest surface energy and Co(OH)2 will preferentially grow along the layered plane to form two-dimensional nanoflake array films.24


image file: c4ra10584b-f1.tif
Fig. 1 SEM images of Co(OH)2 nanoflake arrays on FTO substrate via a hydrothermal synthesis at 120 °C for (a) 0, (b) 0.5, (c) 1, and (d) 5 h.

The SEM top-view and cross-section images of the hydrothermally synthesized Co(OH)2 nanoflake arrays for 5 h are shown in Fig. 2(a) and (b). The Co(OH)2 films exhibit a highly porous structure with pore diameter ranging from 20 nm–1 μm and its height is about 4.4 μm. The synthesized Co(OH)2 film has a network of interconnected nanoflake arrays with a thickness of 20 nm. It is believed that this porous feature is helpful to enhance the follow-up IER process and the resulting performance of devices. During the cross-section preparation of samples, the nanoflake arrays did not detach from the substrate and basically maintained the original structure at the action of the strong shearing stress. Impressively, it is also indicated that Co(OH)2 nanoflake array film has a good adhesion with the FTO substrate in the high magnification SEM image, which is shown as inset of Fig. 2(b). Fig. 2(c) and (d) are TEM analyses of scraped Co(OH)2 nanoflake from FTO substrate. TEM image shows that the individual nanoflake has a smooth feature. Additionally, 6-fold symmetric diffraction spots in the nano beam electron diffraction pattern (NBDP) (inset of Fig. 2(c)) for the nanoflake can be indexed with the α-Co(OH)2 phase, and a high resolution TEM image shows that the α-Co(OH)2 nanoflake is polycrystalline in nature.


image file: c4ra10584b-f2.tif
Fig. 2 (a) and (b) SEM images of Co(OH)2 nanoflake arrays on FTO substrate. (c) TEM image and (d) high resolution TEM image of a scraped Co(OH)2 nanoflake (NBDP in inset).

The XRD pattern indicates that highly crystalline α-Co(OH)2 is formed without other crystal phases, as shown in Fig. 3(b). It reveals a predominant brucite-like phase in which the 2θ scan has peaks at 12.4°, 21.3°, 32.9°, 35.3°, and 58.8°, corresponding to the (003), (006), (101), (012), and (110). The other peaks belong to the peaks of FTO substrate, which match the diffraction peaks in Fig. 3(a). The result of XRD conforms the results obtained in a previous study.25 The cobalt hydroxide has two phases that α-phase is in bluish-green color and β-phase is in red-pink color. The XRD results also match the as-prepared Co(OH)2 nanoflake arrays with bluish-green color in the present work. After IER process, the peaks of α-Co(OH)2 XRD pattern disappear but those of FTO pattern remain in Fig. 3(c). It is a clear evidence of transformation from Co(OH)2 nanoflake arrays to cobalt sulfide nanoflake arrays. The details of cobalt sulfide phase will be discussed in Fig. 4.


image file: c4ra10584b-f3.tif
Fig. 3 XRD patterns of (a) FTO substrate, (b) as-prepared Co(OH)2 nanoflake array film and (c) Co(OH)2 nanoflake array film after the IER method at 75 °C for 8 h.

image file: c4ra10584b-f4.tif
Fig. 4 (a) The SEM image and (b) EDS element mapping of CoS2 nanoflake arrays on FTO substrate. (c)–(e) TEM analyses of scrapped CoS2 nanoflakes (NBDP in inset).

The SEM image of the cobalt sulfide nanoflake array films is shown in Fig. 4(a). Compared to the Co(OH)2 nanoflake arrays with a relatively smooth surface (Fig. 2(a)), it can be seen that the cobalt sulfide nanoflake arrays remain the same nanoflake morphology but surface of nanoflakes becomes rougher after the IER process. This observation provides the evidence that the IER method does not damage the basic morphology of the nanoflake arrays. During the IER method, the substitution of S2− for O2− occurs when the S2− source contacts with the Co(OH)2 surface. It can be thought that the IER method is a diffusion controlled process, so the nanoflake structure will be rarely damaged. Therefore, nanoflake array structure is suitable for IER method because of short distance diffusion from surface to center in a piece of nanoflake. Using the EDS mapping technique, it reveals that Co (red dots) and S (green dots) are homogeneously spread over the whole nanoflake structure. Fig. 4(c)–(e) show the TEM analyses of the nanoflakes scrapped from the film of cobalt sulfide nanoflake arrays. The TEM result indicates that the transformed cobalt sulfide nanoflakes exhibit CoS2 single crystals and aggregates dispersing in the amorphous cobalt sulfide matrix. From element ratio of EDS and NBDP analyses, the tiny grains are CoS2 phase. Interestingly, the high resolution TEM image (Fig. 4(e)) shows that cobalt sulfide nanoflakes are composed of CoS2 crystals and amorphous cobalt sulfide. The EDS analyses of specific areas on the single CoS2 nanoflake are shown in Fig. S1.This result shows that there are no obvious peaks of cobalt sulfide in the XRD pattern because the intensity of CoS2 peaks is too weak to be detected or hidden under the background noise, i.e., the amount of CoS2 is below the detection limit of XRD, while the major component in the matrix is the amorphous sulfide phase.

The synthesis mechanism of cobalt sulfides using the IER method is proposed based on the difference of the solubility product constant (Ksp) of the materials. A material with lower Ksp value is more thermodynamically stable than that with higher Ksp value.18 For example, the Ksp values of Co(OH)2, Co3O4 and CoS are about 5.9 × 10−15, 3.1 × 10−18 and 3 × 10−26, respectively.26 It is believed that Ksp values of cobalt sulfides are much smaller than those of cobalt hydroxides and cobalt oxides. It is to say that the latter is tendentious to convert into the former by providing the excess sulfur source (like S2−). In our studies, when the Co(OH)2 samples are soaked into the Na2S solution, the samples change the color from bluish-green to dark gray rapidly, which means that the ion exchange reaction occurs immediately. Therefore, the CoS2 nanoflake arrays can be obtained by the ion exchange reactions from Co(OH)2 nanoflake arrays in a solution containing S2− anions. Additionally, previous studies have been reported that both thermodynamics and kinetics of IER method depended on the interface reaction between metal oxide (or hydroxide) and S2− anions.27 The reaction rate of nanostructured materials is much faster than that of bulky materials. Consequently, the nanoflake structure is beneficial for the IER method to exchange the anions due to their high surface areas and short distance of diffusion.

To characterize the catalytic activity of the cobalt sulfide nanoflake arrays electrode toward the reduction of I3, cyclic voltammetry (CV) experiments were carried out. For comparison, the CV of the sputtered Pt electrode was measured under the same conditions. Fig. 5 shows the CVs of the CoS2 nanoflake arrays and Pt electrodes. Two typical pairs of redox peaks are observed, implying a high catalytic activity for the reduction of I3. The relative negative pair is assigned to the oxidation and reduction of I/I3 [eqn (4)], whereas the relative positive pair was assigned to the oxidation and reduction of I2/I3 [eqn (5)] according to the literature.28

 
I3 + 2e ↔ 3I (4)
 
3I2 + 2e ↔ 2I3 (5)


image file: c4ra10584b-f5.tif
Fig. 5 Cyclic voltammograms of I3/I for CoS2 nanoflake arrays and sputtered Pt electrodes.

The profile and location of the two pairs of redox peaks for CoS2 are almost similar to those of the Pt electrode, indicating that the former can behave a similar catalytic activity to the latter as a counter electrode in DSSCs. But the cathodic peak potential of the CoS2 electrode is more positive than that of Pt electrode. It indicates that the overpotential for reduction of I3 to I of CoS2 electrode is smaller than that of Pt electrode. And CoS2 electrode exhibits larger current densities, which implies that surface area of CoS2 electrode was larger than that of sputtered Pt electrode. For further investigation to verify the electrocatalytic capabilities of the counter electrodes, Tafel-polarization is measured using the symmetrical cells. Fig. 6 demonstrates the Tafel curves of the cells based on the CoS2 and sputtered Pt electrodes with the logarithmic current density (log[thin space (1/6-em)]J) as a function of the voltage. The slope of a tangent to the CoS2 curve is similar to that of Pt, indicating that the two electrodes have the similar electrocatalytic activity.


image file: c4ra10584b-f6.tif
Fig. 6 Tafel polarization curves of the symmetrical cells based on CoS2 nanoflake arrays and sputtered Pt electrodes.

The current density–voltage (JV) behavior of DSSCs under one Sun illumination (AM 1.5, 100 mW cm−2) using sputtered Pt or CoS2 nanoflake arrays as the counter electrodes is shown in Fig. 7, and the photovoltaic parameter of device performance is summarized in Table 1. Both the DSSCs with CoS2 nanoflake arrays and sputtered Pt used as the counter electrodes produce power conversion efficiencies of 5.20% and 5.34%, and yield a high fill factor of 0.688 and 0.694, respectively, indicating the high catalytic activity for triiodide reduction. The counter electrode with CoS2 nanoflake arrays is compared with other counter electrodes with cobalt sulfide reported in the literature. Different kinds of cobalt sulfide have been prepared and used as the catalyst on the counter electrode of a DSSC in the literature. The CoS films were prepared by the potentiostatic method13 and honeycomb-like CoS films were obtained by the potentiodynamic method.29 The power conversion efficiencies of the DSSC with CoS as counter electrodes were 6.5% and 6.01%, respectively. For the comparison, efficiencies of the DSSCs with Pt counter electrode were 6.5% and 5.71%, respectively. And it was also shown that a conversion efficiency of 7.67% was achieved by using the CoS acicular nanorod arrays as the counter electrode of DSSCs, which was near the efficiency of a DSSC with Pt counter electrode (7.70%).30 It is noticed that the catalyst, electrolytes, and spacers in DSSCs are different in these cases and there is no perfectly comparable DSSC in the literature with DSSC in this study. The difference in efficiencies measured is within ∼3% compared to the maximum efficiency obtained for each counter electrode in this study. The JV performance of CoS2 nanoflake arrays indicates that the cobalt sulfide nanostructure is an impressive counter electrode catalyst for I3/I based DSSCs.


image file: c4ra10584b-f7.tif
Fig. 7 Photocurrent density–voltage (JV) curves of the DSSCs using sputtered Pt and CoS2 nanoflake arrays as the counter electrodes.
Table 1 Photovoltaic parameters of the DSSCs using sputtered Pt and CoS2 nanoflake arrays as the counter electrodes
C.E. Voc (V) Jsc (mA cm−2) F.F. η (%)
Sputtered Pt 0.767 10.04 0.694 5.34
CoS2 nanoflake 0.747 10.13 0.688 5.20


Conclusions

In summary, the hydrothermal growth mechanism of Co(OH)2 nanoflake arrays on FTO substrate is demonstrated in this work. Due to the difference of the solubility product constant, the polycrystalline CoS2 nanoflake arrays were prepared via the solution-based ion exchange reaction without any damage to the basic morphology of the nanoflake arrays. The cobalt sulfide nanoflakes are composed of CoS2 single crystals and their aggregates dispersing in the amorphous cobalt sulfide matrix. The introduction of nanoflake array structure as the counter electrodes leads to a large interfacial contact between the electrolyte and CoS2 counter electrode, which is beneficial for the catalytic reaction of the I/I3. The synthesized film of CoS2 nanoflake arrays was characterized using various electrochemical techniques and demonstrated the comparable catalytic activity as the counter electrode to sputtered Pt. The DSSC using CoS2 nanoflake arrays as the counter electrode achieves a power conversion efficiency of 5.20%, which is close to the photovoltaic performance of DSSC using sputtered Pt as the counter electrode (5.34%). The low-cost, highly catalytic nanomaterial and facile fabrication method should have enormous potential to substitute the Pt counter electrode in the DSSCs.

Acknowledgements

We acknowledge the financial support of the Ministry of Science and Technology (MOST) grant funded by Taiwan government through MOST 103-2221-E-006-089 and NSC 102-2221-E-024-003-MY3.

Notes and references

  1. B. Oregan and M. Gratzel, Nature, 1991, 353, 737–740 CrossRef CAS.
  2. A. Hauch and A. Georg, Electrochim. Acta, 2001, 46, 3457–3466 CrossRef CAS.
  3. T. N. Murakami, S. Ito, Q. Wang, M. K. Nazeeruddin, T. Bessho, I. Cesar, P. Liska, R. Humphry-Baker, P. Comte, P. Pechy and M. Gratzel, J. Electrochem. Soc., 2006, 153, A2255–A2261 CrossRef CAS PubMed.
  4. J. D. Roy-Mayhew, D. J. Bozym, C. Punckt and I. A. Aksay, ACS Nano, 2010, 4, 6203–6211 CrossRef CAS PubMed.
  5. S. H. Seo, S. Y. Kim, B. K. Koo, S. I. Cha and D. Y. Lee, Langmuir, 2010, 26, 10341–10346 CrossRef CAS PubMed.
  6. J. B. Xia, N. Masaki, K. J. Jiang and S. Yanagida, J. Mater. Chem., 2007, 17, 2845–2850 RSC.
  7. J. H. Wu, Q. H. Li, L. Q. Fan, Z. Lan, P. J. Li, J. M. Lin and S. C. Hao, J. Power Sources, 2008, 181, 172–176 CrossRef CAS PubMed.
  8. M. X. Wu, X. A. Lin, A. Hagfeldt and T. L. Ma, Angew. Chem., Int. Ed., 2011, 50, 3520–3524 CrossRef CAS PubMed.
  9. Q. W. Jiang, G. R. Li and X. P. Gao, Chem. Commun., 2009, 7603 CAS.
  10. X. Lin, M. X. Wu, Y. D. Wang, A. Hagfeldt and T. L. Ma, Chem. Commun., 2011, 47, 11489–11491 RSC.
  11. M. X. Wu, Y. D. Wang, X. Lin, N. S. Yu, L. Wang, L. L. Wang, A. Hagfeldt and T. L. Ma, Phys. Chem. Chem. Phys., 2011, 13, 19298–19301 RSC.
  12. J. Yang, C. X. Bao, K. Zhu, T. Yu, F. M. Li, J. G. Liu, Z. S. Li and Z. G. Zou, Chem. Commun., 2014, 50, 4824–4826 RSC.
  13. M. K. Wang, A. M. Anghel, B. Marsan, N. L. C. Ha, N. Pootrakulchote, S. M. Zakeeruddin and M. Gratzel, J. Am. Chem. Soc., 2009, 131, 15976–15977 CrossRef CAS PubMed.
  14. G. Q. Wang and S. P. Zhuo, Phys. Chem. Chem. Phys., 2013, 15, 13801–13804 RSC.
  15. H. K. Mulmudi, S. K. Batabyal, M. Rao, R. R. Prabhakar, N. Mathews, Y. M. Lam and S. G. Mhaisalkar, Phys. Chem. Chem. Phys., 2011, 13, 19307–19309 RSC.
  16. W. J. Dong, X. B. Wang, B. J. Li, L. N. Wang, B. Y. Chen, C. R. Li, X. A. Li, T. R. Zhang and Z. Shi, Dalton Trans., 2011, 40, 243–248 RSC.
  17. Q. R. Hu, S. L. Wang, Y. Zhang and W. H. Tang, J. Alloys Compd., 2010, 491, 707–711 CrossRef CAS PubMed.
  18. L. Dloczik and R. Konenkamp, Nano Lett., 2003, 3, 651–653 CrossRef CAS.
  19. X. H. Xia, C. R. Zhu, J. S. Luo, Z. Y. Zeng, C. Guan, C. F. Ng, H. Zhang and H. J. Fan, Small, 2014, 10, 766–773 CrossRef CAS PubMed.
  20. X. H. Xia, J. P. Tu, Y. J. Mai, X. L. Wang, C. D. Gu and X. B. Zhao, J. Mater. Chem., 2011, 21, 9319–9325 RSC.
  21. Z. Chen, Y. Chen, C. Zuo, S. Zhou, A. G. Xiao and A. X. Pan, Bull. Mater. Sci., 2013, 36, 239–244 CrossRef CAS PubMed.
  22. Y. L. Hou, H. Kondoh, M. Shimojo, T. Kogure and T. Ohta, J. Phys. Chem. B, 2005, 109, 19094–19098 CrossRef CAS PubMed.
  23. Y. Oaki, S. Kajiyama, T. Nishimura and T. Kato, J. Mater. Chem., 2008, 18, 4140–4142 RSC.
  24. X. H. Xia, J. P. Tu, J. Y. Xiang, X. H. Huang, X. L. Wang and X. B. Zhao, J. Power Sources, 2010, 195, 2014–2022 CrossRef CAS PubMed.
  25. R. Z. Ma, Z. P. Liu, K. Takada, K. Fukuda, Y. Ebina, Y. Bando and T. Sasaki, Inorg. Chem., 2006, 45, 3964–3969 CrossRef CAS PubMed.
  26. A. P. Miller, Am. J. Public Health, 1941, 31, 1324 CrossRef PubMed.
  27. D. H. Son, S. M. Hughes, Y. D. Yin and A. P. Alivisatos, Science, 2004, 306, 1009–1012 CrossRef CAS PubMed.
  28. C. H. Yoon, R. Vittal, J. Lee, W. S. Chae and K. J. Kim, Electrochim. Acta, 2008, 53, 2890–2896 CrossRef CAS PubMed.
  29. J. Y. Lin, J. H. Liao and T. C. Wei, Electrochem. Solid-State Lett., 2011, 14, D41–D44 CrossRef CAS PubMed.
  30. C. W. Kung, H. W. Chen, C. Y. Lin, K. C. Huang, R. Vittal and K. C. Ho, ACS Nano, 2012, 6, 7016–7025 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra10584b

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