Optimized CdS quantum dot-sensitized solar cell performance through atomic layer deposition of ultrathin TiO2 coating

Kehan Yu , Xiu Lin , Ganhua Lu , Zhenhai Wen , Chris Yuan * and Junhong Chen *
Department of Mechanical Engineering, University of Wisconsin-Milwaukee, 3200 North Cramer Street, Milwaukee, Wisconsin 53211, U.S.A. E-mail: jhchen@uwm.edu; cyuan@uwm.edu; Tel: +1-414-229-2615

Received 17th May 2012 , Accepted 17th June 2012

First published on 18th June 2012


Abstract

Here we demonstrate a CdS quantum dot (QD) sensitized solar cell with significantly enhanced stability and depressed recombination in I/I3 electrolyte. The CdS QDs were deposited in a mesoporous TiO2 film using chemical bath deposition. Following the coating of an ultrathin TiO2 protection layer using atomic layer deposition (ALD), the performance and stability of CdS QD sensitized solar cells were pronouncedly enhanced. Current–voltage measurements and electrochemical impedance spectroscopy analysis show that the ALD-TiO2 coating slows down the charge recombination and protects the QDs from the photocorrosion of the I/I3 electrolyte. A systematic study shows that a ∼2 nm ALD-TiO2 layer can best promote solar cell efficiency by balancing protection with charge transfer. We have obtained a photoconversion efficiency of 1.41% and stable operation (no degradation in photocurrent) for at least 20 min under full sun illumination on CdS QD solar cells. The photocurrent gradually decreased down to ∼70% of its original value after 1 h of operation.


Introduction

Quantum dot sensitized solar cells (QDSSCs) have been studied extensively because they are potential low-cost alternatives to existing crystalline silicon cells. Semiconductor QDs such as CdS,1 CdSe,2 PbS,3 PbSe,4 InAs,5 and InP6 have been used to sensitize metal oxides, like nanocrystalline (nc) TiO2 or SnO2, by absorbing the most intense regions of the solar spectrum.7 In a recent breakthrough, a photoconversion efficiency (PCE) of 6% was achieved using PbS QDs.8 Unlike their dye counterparts, most QDs suffer from photodegradation when used with I/I3 in a photoelectrochemical cell.9 In the QDSSC, the optimum redox couple was found to be sulfide/polysulfide (S2−/Sn2−);10,11 however, sulfides react very strongly with platinum, the catalyst of the counter electrode (CE), and can poison the surface, thereby reducing the catalytic activity.12 Alternative CE materials for polysulfide solution, such as CoS, Cu2S, NiS, carbon, and PbS, have been investigated.13 To overcome the attack of iodine, Zaban and co-workers have designed a strategy to employ a TiO2 barrier layer over CdS QDs using electrophoretic deposition.14 With the amorphous TiO2 passivation shell, they obtained a CdS QDSSC that was stable for 120 s in I/I3 electrolyte. However, control over the TiO2 layer thickness, which is much needed for further optimization of the process, was quite limited by the means of electrophoretic deposition used in the study.

Atomic layer deposition (ALD) has been used to introduce extremely thin and conformal coating due to its unique excellent step coverage, atomic-scale thickness control, exceptional composition control for nanostructures, and excellent uniformity over large areas. Researchers have grown metal oxide materials like TiO2 and Al2O3 by ALD to improve the performance of dye-sensitized solar cells;15,16 for example, the Cao group introduced an ALD-TiO2 film that suppressed surface charge recombination.15 An ALD-Al2O3 film employed by the Hupp group was found to passivate surface states.16 For a QDSSC, the passivation layer can be neither too thick, which would impede the charge transfer, nor too thin, which may allow the electrolyte to penetrate through the layer. ALD is especially suitable for depositing a TiO2 thin film as the passivation layer for a stable QDSSC, due to its ultrafine thickness control.

In this study, different thicknesses of ultrathin TiO2 layers were deposited on the surface of CdS QDSSCs using ALD. In addition, once the intended thickness was reached, the ultrathin TiO2 coating was subsequently annealed to form a crystalline phase, which has better stability and electronic properties than amorphous TiO2. Photoanodes with different TiO2 coating thicknesses were systematically examined using electrochemical impedance spectroscopy (EIS) and photocurrent–voltage measurements. An intermediate coating thickness was found to be optimal, by balancing the protection effect and the charge transfer efficiency. A power efficiency of 1.4% and stability of 20 min were achieved as a result of ultrafine thickness control of TiO2 coating by ALD.

Experimental

Equipment

A field-emission scanning electron microscope (SEM) (Hitachi S 4800) equipped with a Bruker Quantax energy-dispersive X-ray spectrometry (EDS) system was used to observe the morphology of TiO2–CdS structures and element distribution; the SEM renders a resolution of 1.4 nm at 1 kV acceleration voltage. High-resolution transmission electron microscopy (HRTEM) characterization of TiO2–CdS was carried out using a Hitachi H 9000 NAR TEM, which has a point resolution of 0.18 nm at 300 kV in the phase contrast HRTEM imaging mode. The EIS measurements were conducted using an electrochemical workstation (CH Instruments 600D).

Preparation of QDSSCs

Mesoporous TiO2 films were prepared by doctor-blade casting of a titania paste (1.0 g Degussa P25 TiO2 powder, 50 μl titanium(IV) butoxide, 2 ml isopropanol) onto fluorine-doped tin oxide (FTO, Solaronix) coated glass substrates with 12 Ω □−1 sheet resistance. A TiO2 blocking/protection layer was grown on all FTO glass for 1,000 ALD cycles before applying mesoporous TiO2. The detailed ALD process is described below. After drying in air, the TiO2–FTO glass was sintered at 500 °C for 30 min in air.

The CdS QDs were assembled on TiO2 films using chemical bath deposition (CBD).17 After cooling down, the TiO2–FTO glass was transferred immediately to a Teflon-lined stainless steel autoclave (100 ml) containing Cd(CH3COO)2·2H2O (0.110 g, 98%, Sigma Aldrich) and dimethyl sulfoxide (DMSO, 40 ml). The autoclave was tightly sealed and reacted at 180 °C for 12 h. The obtained anodes were then rinsed extensively with DI water and isopropanol and dried in air. The resulting anodes were then annealed in Ar flow (1.0 liter min−1, lpm) at 500 °C for 30 min.

The ultrathin TiO2 protection layers were coated on the TiO2–CdS photoanodes at 200 °C using the Cambridge NanoTech Savannah S100 ALD System. Titanium isopropoxide (Ti(OiPr)4) and H2O were used as precursors, with high-purity nitrogen as a carrier and purging gas. The reactants were alternatively pulsed and purged in the following sequence: pulsing H2O for 0.3 s, purging for 6 s, pulsing Ti(OiPr)4 for 0.015 s, purging for 4 s. The thicknesses of the TiO2 layers were measured with a Horiba spectroscopic ellipsometer to be 1.5 ± 0.08 nm, 2.2 ± 0.09 nm, and 3.2 ± 0.12 nm for coatings of 50 cycles, 100 cycles, and 150 cycles, respectively. The measurements of thickness were carried out on silicon wafers coated with a TiO2 thin film under the same conditions used for the mesoporous photoanodes. The same procedure was used for the deposition of the 1000-cycle blocking layer; the TiO2-coated films were annealed at 500 °C in Ar (1.0 lpm) for 30 min to increase crystallinity.

The Pt/FTO counter electrodes (CEs) were prepared by drop casting a 0.005 M H2PtCl6 aqueous solution on FTO glass. After drying in air at room temperature, the CEs were elevated to 500 °C in air and maintained for 30 min.

A hot-melt sealing foil (25 microns, Meltonix 1170-25, Solaronix) was sandwiched between the photoanode and CE. Light pressure was applied on the anode–spacer–cathode assembly on a hot plate at 150 °C to seal the cell. An I/I3 electrolyte (Iodolyte AN-50, Solaronix) was introduced into the cells by capillary force. No further sealing process was adopted.

Photovoltaic characterization

The photocurrent–voltage (JV) characteristics were obtained under simulated solar illumination (solar simulator, Newport, 94021A) at one sun (AM 1.5G, 100 mW cm−2) using a Keithley 2420 source meter equipped with a calibrated Si-reference cell (Oriel, P/N 91150V). The incident-photon-to-electron conversion efficiency (IPCE) of solar cells was analyzed using a system consisting of a 300 W Xenon lamp simulated light source (Newport 66902), a monochromator (Newport cornerstone 74125), and a radiometer (Newport Merlin 70104). The photon flux of light incident on the samples was calibrated using a silicon photodiode (Newport 70356). Measurements were typically made at 10 nm wavelength intervals between 350 and 1100 nm.

Electrochemical characterization

The EIS measurements were conducted under AM 1.5G simulated solar illumination at 100 mW cm−2. The same solar cells were measured for EIS immediately after JV characterization. EIS spectra were recorded over a frequency range of 100 kHz–10 mHz. The applied bias voltage and ac amplitude were set at open circuit voltage (VOC) of the cells and at 5 mV, respectively.

Results and discussion

Structure of photoanodes

The photoanodes of QDSSCs were fabricated by chemical bath deposition of CdS QDs on mesoporous TiO2 nanocrystal films. Conformal ultrathin TiO2 films were coated on the CdS–TiO2 nanocomposite using ALD. Typically, the photoanodes were deposited for 50 cycles, 100 cycles, and 150 cycles, denoted as 50c, 100c, and 150c, respectively. The thicknesses of the TiO2 coating layers were measured with an ellipsometer as 1.5 ± 0.08 nm, 2.2 ± 0.09 nm, and 3.2 ± 0.12 nm for the 50c, 100c, and 150c cases, respectively. Comparisons were made among three differently coated solar cells and a non-coated solar cell, and the schematic diagrams of coated and non-coated QDSSCs are shown in Fig. 1a and b. The thickness of the CdS–TiO2 photoanode film is 15–20 μm, confirmed by a cross-sectional SEM image (Fig. 1c). EDS reveals a homogeneous CdS distribution across the mesoporous TiO2 film (Fig. 1d). The ultrathin ALD-TiO2 coating can be clearly observed in an HRTEM image, as shown in Fig. 2, in which a thin layer of TiO2 (∼2 nm) can be discerned covering the CdS–TiO2 nanostructure. The observed thickness of the 100c ALD-TiO2 agrees well with the ellipsometer measurement of 2.2 ± 0.09 nm.
Schematics of the CdS QD-sensitized nanocrystalline TiO2 electrode (a) without and (b) with an ALD-TiO2 protection layer. (c) Cross-sectional SEM image of the CdS QD-sensitized mesoporous TiO2 film with 100c ALD-TiO2 coating supported on FTO glass. (d) EDS analysis of Ti and Cd along the white arrow in (c).
Fig. 1 Schematics of the CdS QD-sensitized nanocrystalline TiO2 electrode (a) without and (b) with an ALD-TiO2 protection layer. (c) Cross-sectional SEM image of the CdS QD-sensitized mesoporous TiO2 film with 100c ALD-TiO2 coating supported on FTO glass. (d) EDS analysis of Ti and Cd along the white arrow in (c).

HRTEM image of 100c ALD-TiO2-coated CdS–TiO2 nanocomposite. TiO2 and CdS particles are outlined by dashed lines.
Fig. 2 HRTEM image of 100c ALD-TiO2-coated CdS–TiO2 nanocomposite. TiO2 and CdS particles are outlined by dashed lines.

Electrochemical properties

EIS analysis revealed noticeable differences in charge transfer resistance at the CdS–ALD-TiO2–electrolyte interfaces with different TiO2 coating thicknesses. Fig. 3a shows EIS Nyquist plots of coated QDSSCs; they all exhibit Gerischer impedances, indicating low charge collection efficiencies.18 Apart from the components corresponding to the electrolyte and the counter electrode, the EIS Nyquist plot ranging from 100 kHz to 1 Hz is of particular interest because it is directly related to the charge transfer at the CdS–ALD-TiO2–electrolyte interface.15,19 The diameter of the Gerischer arc represents the active layer resistance of the photoanode, RAL = (RrecRtr)1/2, where Rtr is the total transport resistance in the mesoporous TiO2 film and at the CdS–ALD-TiO2–electrolyte interface, and Rrec is the recombination resistance across the surface of the nanoporous solid.18 The Gerischer impedance happens when the electron diffusion length is much smaller than the thickness of the active layer, that is, RtrRrec. It is obvious that RAL for the 100c case is appreciably lower than that of the 150c. A better charge transfer can be achieved at the CdS–100c-ALD-TiO2–electrolyte interface since the ALD-TiO2 layer is simultaneously serving as a protection layer and a thinner charge transfer barrier. For the similar interface of the 150c case, although the ALD-TiO2 layer is thicker and better for protection, it hinders the hole tunneling and displays slower recombination and poor charge transfer; however, the ALD-TiO2 coating thickness cannot be further reduced such that the effective protection of the CdS against the electrolyte is not sacrificed. A rapid bleaching was observed in the 50c cell for both the QDs and the electrolyte, that is the redox I/I3 couple concentration decreased and the QDs were dissolved, thus resulting in a slow charge transfer (50c, Nyquist plot).
Electrochemical impedance spectra of the three differently coated cells (-Δ- 50c, -□- 100c, and -○- 150c) under illumination of one sun (AM 1.5G, 100 mW cm−2) at open circuit voltage. (a) Nyquist plots, frequencies of peak points are indicated; (b) Bode phase plots.
Fig. 3 Electrochemical impedance spectra of the three differently coated cells (-Δ- 50c, -□- 100c, and -○- 150c) under illumination of one sun (AM 1.5G, 100 mW cm−2) at open circuit voltage. (a) Nyquist plots, frequencies of peak points are indicated; (b) Bode phase plots.

The charge recombination in the solar cells can be evaluated by EIS. As proposed by Kern et al.20 and Bisquert,21 the “recombination” reflected by the EIS represents reaction of the electrons either on the conduction band or in the trap states of TiO2 with oxidized redox species I3 in a dye-sensitized solar cell. For QDSSCs, the QDs introduce additional surface states, which make the recombination more complicated. However, the EIS analysis is still effective since the recombination is generally defined as the back reaction of electrons. The effective rate constant for recombination keff can be estimated with

 
ugraphic, filename = c2ra20979a-t1.gif(1)
where ωmax is the frequency at the peak point of the EIS Nyquist plots (as indicated in Fig. 3a).22 Effective lifetime of electrons is thus defined as
 
ugraphic, filename = c2ra20979a-t2.gif(2)

As shown in Fig. 3a, the peak frequencies ωmax decrease from 50c to 150c. The recombination rate is lower for the anode with the thicker coating, which is calculated using eqn (1). This result clearly shows that the ALD coating passivates the surface states or trap states. A TiO2 shell structure is often employed to suppress recombination, which is usually caused by surface traps (defect states) and back electron transfer from mesoporous TiO2 to the electrolyte.23 Also, the passivation effect is presented directly by the EIS Bode phase plots. The characteristic peak shifts to a lower frequency with increasing ALD coating thickness, suggesting an increase in electron lifetime (Fig. 3b). It should be noted that frequencies at characteristic peaks in the Bode phase plots (Fig. 3b) are slightly different from peak frequencies in the Nyquist plots (Fig. 3a). The lifetimes of electrons calculated with eqn (2), as well as the rate constants of recombination, are listed in Table 1. We observed the same trend on additional batches of solar cells, which is shown in the ESI, Fig. S1.

Table 1 Solar cell parameters of the CdS quantum dot sensitized mesoporous TiO2 electrode with and without an ALD-TiO2 protection layer
  J SC (mA cm−2) V OC (mV) FF η (%) k eff (s−1) τ eff (ms)
Non-coated 1st 1.47 572 0.71 0.6
2nd 1.36 449 0.13 0.08
3rd 1.32 468 0.13 0.08
50c 1st 1.69 563 0.5 0.47 84.9 11.8
2nd 1.40 543 0.53 0.41
3rd 1.68 554 0.48 0.45
100c 1st 3.16 668 0.56 1.19 39.3 25.4
2nd 3.02 667 0.57 1.15
3rd 2.97 667 0.57 1.14
Al-back 3.55 673 0.59 1.41
150c 1st 1.39 668 0.62 0.58 22.1 45.2
2nd 1.52 660 0.60 0.60
3rd 1.54 660 0.59 0.60


Photovoltaic performances of QDSSCs

A comparison of the performance of CdS QDSSCs with various ALD-TiO2 coating thicknesses is shown in Fig. 4. Fig. 4a and b show the JV characteristics of the 50c, 100c, and 150c cells, under illumination and in the dark, respectively, using I/I3 as a redox electrolyte. Each cell was tested three times at 1 min intervals, and then measured in the dark. Under illumination (Fig. 4a), the performance of the 100c cell is significantly better than that of the other cells. Insignificant degradation was observed during the JV test, indicating a relatively good photo-stability. Table 1 summarizes the parameters of three measurements of all the cells, and a power conversion efficiency (η) of 1.19% was achieved by the 100-cycle coating. Based on the three measurements, the open circuit voltages VOC,100cVOC,150c > VOC,50c > VOC,non were observed. Carrier recombination is considered as one of the factors for the reduction in open circuit voltage.11,15,18 Although a thicker coating can favor the VOC due to less recombination, the 150-cycle coating hinders charge separation and leads to hole accumulation. Based on our observation, the 100-cycle coating promotes the VOC most by balancing the surface passivation and charge separation.
Characterization of the CdS-QD sensitized TiO2 solar cells with and without the ALD-TiO2 protection layer. J–V measurements of three differently coated cells (50c, 100c, and 150c) (a) under illumination of one sun (AM 1.5G, 100 mW cm−2) and (b) under dark conditions. For the curves measured under illumination, each cell was measured three times: solid line (first time), dashed line (second time), and dash-dotted line (third time). (c) J–V measurement of a non-coated cell under illumination of one sun (AM 1.5G, 100 mW cm−2) three times. (d) IPCE of the three differently coated cells.
Fig. 4 Characterization of the CdS-QD sensitized TiO2 solar cells with and without the ALD-TiO2 protection layer. JV measurements of three differently coated cells (50c, 100c, and 150c) (a) under illumination of one sun (AM 1.5G, 100 mW cm−2) and (b) under dark conditions. For the curves measured under illumination, each cell was measured three times: solid line (first time), dashed line (second time), and dash-dotted line (third time). (c) JV measurement of a non-coated cell under illumination of one sun (AM 1.5G, 100 mW cm−2) three times. (d) IPCE of the three differently coated cells.

The fill factor FF150c > FF100c > FF50c > FFnon is a direct indication of the slower recombination for thicker coatings. Although a larger series resistance (resistance of the TiO2 coating) usually leads to a smaller FF, it is clear that the FFs here are mainly affected by recombination control.18Fig. 4b shows that electron transfer (or dark current) from the electrode with a thicker coating into the electrolyte at forward bias is significantly reduced compared to the electrode with a thinner coating. This result is directly correlated with the reduced recombination of electrons from the mesoporous electrode into the redox electrolyte caused by the ALD-TiO2 shell, which is consistent with the EIS analysis. In contrast, a non-coated cell degraded quickly during the three consecutive measurements as the QDs dissolved quickly in the I/I3 electrolyte (Fig. 4c). Considering all factors, an intermediate coating thickness is optimal to achieve slow recombination, stability against corrosion, and fast charge transfer in the nc-TiO2–CdS–ALD-TiO2 structure; as a result, the highest short circuit current density JSC and η were obtained for the 100c cell. This conclusion is based on observations of multiple batches of solar cells. The JV characteristics of additional solar cells are shown in the ESI, Fig. S2. The incident photon-to-current efficiency (IPCE) results show the same trend as the JV characteristics (Fig. 4d)—the better the cell was protected, the smaller the loss of charge during the photoconversion.

We also noticed that the TiO2–QD film did not absorb 100% incident light. Increasing the TiO2–QDs film thickness will not lead to improved light absorption because a thicker film will lower the charge collection efficiency, as suggested by the EIS result. A reflecting aluminum foil was placed at the back of the 100c cell (100c Al-back) for a further test, and a power efficiency of η = 1.41% was achieved. Other solar cell parameters are listed in Table 1. We note that the Cd2+ ion concentration (Cd[thin space (1/6-em)]:[thin space (1/6-em)]Ti = 1[thin space (1/6-em)]:[thin space (1/6-em)]4) in our anodes produced using CBD is lower than that reported by Zaban et al. (Cd[thin space (1/6-em)]:[thin space (1/6-em)]Ti ∼ 1[thin space (1/6-em)]:[thin space (1/6-em)]2).14 Lower CdS loading will lead to a lower photoelectron concentration, which is a direct reason for a large Rtr. It also contributes to a low IPCE as many incident photons cannot be captured by CdS. Additionally, only a fraction of the visible light was utilized by CdS QDs because of their relatively large band gap (2.4 eV). This issue can be overcome by replacing CdS QDs with other low band gap QDs, such as CdSe, PbS, PbSe, or a combination of them.

The photo-stability of the ALD-TiO2 protected electrode was tested under periodic on–off one-sun illumination in the presence of the I/I3 electrolyte. The short circuit current of a 100c cell was measured as a function of time with shuttered light at 12 s intervals. Fig. 5a shows the reproducible photo-current of the coated photoanode. A long-term stability test of a 100c cell was carried out under the illumination of AM 1.5G, 100 mW cm−2. The JSC was recorded for over 1 h (Fig. 5b); the current was very stable during the first 20 min, and then gradually decreased down to ∼70% of its original value in the next 40 min. After the measurements, the photoanode (originally yellow) was found slightly bleached (light yellow). Nevertheless, such a stability is much better than what has been reported by Zaban et al.14


Photo-stability test of a 100c solar cell. (a) Short circuit current density as a function of time using periodic illumination intervals (12 s interval length). (b) Short circuit current versus time of a 100c solar cell under illumination of one sun (AM 1.5G, 100 mW cm−2) over 1 h.
Fig. 5 Photo-stability test of a 100c solar cell. (a) Short circuit current density as a function of time using periodic illumination intervals (12 s interval length). (b) Short circuit current versus time of a 100c solar cell under illumination of one sun (AM 1.5G, 100 mW cm−2) over 1 h.

An Al2O3 coating was introduced to improve the performance of dye-sensitized solar cells,16 and as a comparison we also made a similar Al2O3 coating to protect CdS QDs using ALD. The JSC was as low as 0.33, 0.31, and 0.24 mA cm−2 when the thickness of the Al2O3 layer was 1.3, 1.9, and 2.6 nm, respectively. Obviously, this non-conductive coating limited the photocurrent to a very low level. A thinner Al2O3 coating could reduce the series resistance and increased the JSC to 2.8 and 1.28 mA cm−2 for coating thicknesses of 2.6 and 3.9 Å, respectively. However, QDs with an extremely thin coating (e.g. 1.3 Å) were even less stable. For example, the photoanode with the 1.3 Å ALD-Al2O3 coating degraded quickly during the test and the resulting JSC was as low as 0.29 mA cm−2. Detailed JV curves are shown in the ESI (Fig. S3 and S4).

Conclusions

ALD can introduce an ultrathin protection layer of TiO2 to passivate and enhance the stability of CdS QD sensitized solar cells. The ALD-TiO2 layer slows down charge recombination and increases VOC and FF; however, ALD-TiO2 coating that is too thick may hinder the hole tunneling, resulting in a low charge separation efficiency. And a protection layer that is too thin cannot effectively provide the protection effect. A systematic study shows that a ∼2 nm ALD-TiO2 layer, corresponding to 100 ALD cycles, can best promote solar cell efficiency by balancing the protection and charge transfer. An ALD-Al2O3 coating was employed to compare with the TiO2 coating. But the ALD-Al2O3 must be thinner to obtain a sufficiently high photocurrent because of its insulating nature. As a result, the protection effect of ALD-Al2O3 is worse than that of ALD-TiO2. For future work, a deposition process with both chemical stability and a no hole-transfer barrier is desired for photoanode coating in QDSSCs. The ALD coating method is applicable not only to the current solar cell, but also to other QDSSC systems, such as solar cells sensitized with CdSe, PbS, and PbSe QDs. Therefore, this study can potentially guide the design of QDSSCs with enhanced performance.

Acknowledgements

This work was financially supported by the NSF (CMMI-1001039) and DOE (DE-EE0003208). SEM/EDS and TEM were performed in the UWM Electron Microscope Laboratory and UWM HRTEM Laboratory, respectively. We thank M. Gajdardziska-Josifovska for providing TEM access, D. Robertson for technical support with TEM, H. A. Owen for technical support with SEM.

References

  1. (a) C.-H. Chang and Y.-L. Lee, Appl. Phys. Lett., 2007, 91, 053503 CrossRef; (b) S. Hotchandani and P. V. Kamat, J. Phys. Chem., 1992, 96, 6834 CrossRef CAS; (c) S. Lin, Appl. Phys. Lett., 2007, 90, 143517 CrossRef; (d) P. Ardalan, T. P. Brennan, H. B. R. Lee, J. R. Bakke, I. K. Ding, M. D. McGehee and S. F. Bent, ACS Nano, 2011, 5, 1495 CrossRef CAS; (e) W. T. Sun, Y. Yu, H. Y. Pan, X. F. Gao, Q. Chen and L. M. Peng, J. Am. Chem. Soc., 2008, 130, 1124 CrossRef CAS; (f) L. Sheeney-Haj-Khia, B. Basnar and I. Willner, Angew. Chem., Int. Ed., 2005, 44, 78 CrossRef; (g) I. Robel, B. A. Bunker and P. V. Kamat, Adv. Mater., 2005, 17, 2458 CrossRef CAS.
  2. (a) D. Liu and P. V. Kamat, J. Phys. Chem., 1993, 97, 10769 CrossRef CAS; (b) S. Q. Fan, B. Fang, J. H. Kim, J. J. Kim, J. S. Yu and J. Ko, Appl. Phys. Lett., 2010, 96, 063501 CrossRef; (c) S. R. Sun, L. A. Gao, Y. Q. Liu and J. Sun, Appl. Phys. Lett., 2011, 98, 093112 CrossRef; (d) B. Farrow and P. V. Kamat, J. Am. Chem. Soc., 2009, 131, 11124 CrossRef CAS; (e) A. Kongkanand, K. Tvrdy, K. Takechi, M. Kuno and P. V. Kamat, J. Am. Chem. Soc., 2008, 130, 4007 CrossRef CAS; (f) N. Fuke, L. B. Hoch, A. Y. Koposov, V. W. Manner, D. J. Werder, A. Fukui, N. Koide, H. Katayama and M. Sykora, ACS Nano, 2010, 4, 6377 CrossRef CAS; (g) H. Lee, M. Wang, P. Chen, D. R. Gamelin, S. M. Zakeeruddin, M. Grätzel and M. K. Nazeeruddin, Nano Lett., 2009, 9, 4221 CrossRef CAS; (h) R. Loef, A. J. Houtepen, E. Talgorn, J. Schoonman and A. Goossens, Nano Lett., 2009, 9, 856 CrossRef CAS; (i) P. Brown and P. V. Kamat, J. Am. Chem. Soc., 2008, 130, 8890 CrossRef CAS; (j) K. S. Leschkies, R. Divakar, J. Basu, E. Enache-Pommer, J. E. Boercker, C. B. Carter, U. R. Kortshagen, D. J. Norris and E. S. Aydil, Nano Lett., 2007, 7, 1793 CrossRef CAS; (k) I. Robel, V. Subramanian, M. Kuno and P. V. Kamat, J. Am. Chem. Soc., 2006, 128, 2385 CrossRef CAS.
  3. (a) R. Plass, S. Pelet, J. Krueger, M. Gratzel and U. Bach, J. Phys. Chem. B, 2002, 106, 7578 CrossRef CAS; (b) J. B. Gao, J. M. Luther, O. E. Semonin, R. J. Ellingson, A. J. Nozik and M. C. Beard, Nano Lett., 2011, 11, 1002 CrossRef CAS.
  4. (a) K. S. Leschkies, T. J. Beatty, M. S. Kang, D. J. Norris and E. S. Aydil, ACS Nano, 2009, 3, 3638 CrossRef CAS; (b) J. J. Choi, Y. F. Lim, M. B. Santiago-Berrios, M. Oh, B. R. Hyun, L. F. Sung, A. C. Bartnik, A. Goedhart, G. G. Malliaras, H. D. Abruna, F. W. Wise and T. Hanrath, Nano Lett., 2009, 9, 3749 CrossRef CAS; (c) M. C. Beard, A. G. Midgett, M. Law, O. E. Semonin, R. J. Ellingson and A. J. Nozik, Nano Lett., 2009, 9, 836 CrossRef CAS; (d) M. Law, M. C. Beard, S. Choi, J. M. Luther, M. C. Hanna and A. J. Nozik, Nano Lett., 2008, 8, 3904 CrossRef CAS.
  5. (a) P. R. Yu, K. Zhu, A. G. Norman, S. Ferrere, A. J. Frank and A. J. Nozik, J. Phys. Chem. B, 2006, 110, 25451 CrossRef CAS; (b) W. Wei, X. Y. Bao, C. Soci, Y. Ding, Z. L. Wang and D. Wang, Nano Lett., 2009, 9, 2926 CrossRef CAS.
  6. A. Zaban, O. I. Micic, B. A. Gregg and A. J. Nozik, Langmuir, 1998, 14, 3153 CrossRef CAS.
  7. P. V. Kamat, K. Tvrdy, D. R. Baker and J. G. Radich, Chem. Rev., 2010, 110, 6664 CrossRef CAS.
  8. J. Tang, K. W. Kemp, S. Hoogland, K. S. Jeong, H. Liu, L. Levina, M. Furukawa, X. Wang, R. Debnath, D. Cha, K. W. Chou, A. Fischer, A. Amassian, J. B. Asbury and E. H. Sargent, Nat. Mater., 2011, 10, 765 CrossRef CAS.
  9. C. Liu, P. C. Wu, T. Sun, L. Dai, Y. Ye, R. M. Ma and G. G. Qin, J. Phys. Chem. C, 2009, 113, 14478 CAS.
  10. P. K. Mahapatra and A. R. Dubey, Sol. Energy Mater. Sol. Cells, 1994, 32, 29 CrossRef CAS.
  11. G. Hodes, J. Phys. Chem. C, 2008, 112, 17778 CAS.
  12. (a) R. Mohtadi, W. K. Lee, S. Cowan, J. W. Van Zee and M. Murthy, Electrochem. Solid-State Lett., 2003, 6, A272 CrossRef CAS; (b) Z. Paal, K. Matusek and M. Muhler, Appl. Catal., A, 1997, 149, 113 CrossRef CAS.
  13. (a) G. Hodes, J. Manassen and D. Cahen, J. Appl. Electrochem., 1977, 7, 181 CrossRef CAS; (b) G. Hodes, J. Manassen and D. Cahen, J. Electrochem. Soc., 1980, 127, 544 CrossRef CAS; (c) Z. R. Yu, J. H. Du, S. H. Guo, H. Y. Zhang and Y. Matsumoto, Thin Solid Films, 2002, 415, 173 CrossRef CAS; (d) Z. S. Yang, C. Y. Chen, C. W. Liu and H. T. Chang, Chem. Commun., 2010, 46, 5485 RSC; (e) Q. X. Zhang, Y. D. Zhang, S. Q. Huang, X. M. Huang, Y. H. Luo, Q. B. Meng and D. M. Li, Electrochem. Commun., 2010, 12, 327 CrossRef CAS; (f) Z. Tachan, M. Shalom, I. Hod, S. Ruhle, S. Tirosh and A. Zaban, J. Phys. Chem. C, 2011, 115, 6162 CrossRef CAS.
  14. M. Shalom, S. Dor, S. Rühle, L. Grinis and A. Zaban, J. Phys. Chem. C, 2009, 113, 3895 CAS.
  15. K. Park, Q. Zhang, B. B. Garcia, X. Zhou, Y.-H. Jeong and G. Cao, Adv. Mater., 2010, 22, 2329 CrossRef CAS.
  16. T. W. Hamann, O. K. Farha and J. T. Hupp, J. Phys. Chem. C, 2008, 112, 19756 CAS.
  17. (a) A. Cao, Z. Liu, S. Chu, M. Wu, Z. Ye, Z. Cai, Y. Chang, S. Wang, Q. Gong and Y. Liu, Adv. Mater., 2010, 22, 103 CrossRef CAS; (b) R. Elbaum, S. Vega and G. Hodes, Chem. Mater., 2001, 13, 2272 CrossRef CAS.
  18. F. Fabregat-Santiago, G. Garcia-Belmonte, I. Mora-Sero and J. Bisquert, Phys. Chem. Chem. Phys., 2011, 13, 9083 RSC.
  19. (a) Q. Wang, J. E. Moser and M. Gratzel, J. Phys. Chem. B, 2005, 109, 14945 CrossRef CAS; (b) F. Fabregat-Santiago, J. Bisquert, G. Garcia-Belmonte, G. Boschloo and A. Hagfeldt, Sol. Energy Mater. Sol. Cells, 2005, 87, 117 CrossRef CAS.
  20. R. Kern, R. Sastrawan, J. Ferber, R. Stangl and J. Luther, Electrochim. Acta, 2002, 47, 4213 CrossRef CAS.
  21. J. Bisquert, J. Phys. Chem. B, 2001, 106, 325 CrossRef.
  22. M. Adachi, M. Sakamoto, J. Jiu, Y. Ogata and S. Isoda, J. Phys. Chem. B, 2006, 110, 13872 CrossRef CAS.
  23. (a) F. Fabregat-Santiago, J. Garcia-Canadas, E. Palomares, J. N. Clifford, S. A. Haque, J. R. Durrant, G. Garcia-Belmonte and J. Bisquert, J. Appl. Phys., 2004, 96, 6903 CrossRef CAS; (b) J. Nelson, Phys. Rev. B: Condens. Matter, 1999, 59, 15374 CrossRef CAS; (c) J.-Y. Hwang, S.-A. Lee, Y. H. Lee and S.-I. Seok, ACS Appl. Mater. Interfaces, 2010, 2, 1343 CrossRef CAS.

Footnote

Electronic Supplementary Information (ESI) available. See DOI: 10.1039/c2ra20979a

This journal is © The Royal Society of Chemistry 2012
Click here to see how this site uses Cookies. View our privacy policy here.