You are on page 1of 5

Journal of Physics and Chemistry of Solids 73 (2012) 535539

Contents lists available at SciVerse ScienceDirect

Journal of Physics and Chemistry of Solids


journal homepage: www.elsevier.com/locate/jpcs

Successive ionic layer adsorption and reaction of ZnSe shells for ZnO
nanowire-based dye-sensitized solar cells
Jooyoung Chung, Jihyun Myoung, Jisook Oh, Sangwoo Lim n
Department of Chemical and Biomolecular Engineering, Yonsei University, 262 Seongsanno, Seodaemun-gu, Seoul 120749, Korea

a r t i c l e i n f o

abstract

Article history:
Received 6 April 2011
Received in revised form
1 August 2011
Accepted 1 December 2011
Available online 17 December 2011

Thin ZnSe layers were deposited on ZnO nanowires by a novel successive ionic layer adsorption and
reaction technique in order to solve recombination problems in ZnO nanowire-based dye-sensitized
solar cells (DSSCs). Cell efciency increased from 0.1 to 1.31.4% with the deposition of a 9- to13-nmthick ZnSe shell on ZnO nanowires due to a large increase in JSC. The dramatic increase in JSC and cell
efciency is due to the facilitation of electron transfer related to ambipolar diffusion by the formation of
a type II band alignment and the suppression of recombination in the presence of the ZnSe shell.
& 2011 Elsevier Ltd. All rights reserved.

Keywords:
A. Nanostructures
B. Solgel growth
C. Electron microscopy
D. Optical properties

1. Introduction
ZnO nanowires are used as anode materials in dye-sensitized
solar cells (DSSCs) because they are relatively easy to synthesize
and modify and exhibit outstanding electric 1-D II-VI semiconductor characteristics [1,2]. In addition, they provide a direct
electric pathway along the direction of the c-axis to ensure the
rapid collection of electrons, and have internal electric elds that
can assist electron collection by preventing recombination with
electrolytes [1]. Further, this type of nanostructure exhibits a
large electron diffusion coefcient of 1.8  10  3 cm2/s [2]. However, many factors such as the kind of dye molecule, the amount
of dye loading relative to the surface area of the ZnO nanowires,
and the electrical and optical properties of the ZnO nanowires
affect the efciency of ZnO nanowire-based DSSCs [35]. The
surface conditions of the ZnO nanowires are particularly important to DSSC performance [6,7]. For example, improvement in
electronic coupling between the surfaces of ZnO nanowires and
the dye increases light harvesting efciency [6]. In addition, it has
been reported that an increased number of hydroxide functional
groups on the surface of ZnO nanowires leads to facilitation of dye
loading, which contributes to the improvement of DSSC performance [7]. Efforts to increase the surface area of the ZnO
nanostructure have also been made by preparing hierarchical
ZnO NW growth through nanoforests [8] and nanoowers [9] to
improve performance.

Corresponding author. Tel.: 82 2 2123 5754; fax: 82 2 312 6401.


E-mail address: swlim@yonsei.ac.kr (S. Lim).

0022-3697/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jpcs.2011.12.001

However, one of the major problems in operating ZnO nanowire-based DSSCs is the recombination of injected electrons due
to defects in ZnO nanostructures [10]. Several methods have been
developed to resolve such recombination problems, such as the
formation of core/shell structures [1114] and Zn(OAc)2 treatment [15]. The formation of shells on ZnO nanowires is a
particularly good candidate approach for solving recombination
problems because core/shell nanowires reduce radiative recombination losses [12], while dark exciton formation minimizes the
exciton recombination rate and improves the carrier collection in
the solar cell [13]. In addition, core/shell structured nanowires
adjust the band gap alignment in solar cells [14]. For example,
Greene et al. reported TiO2 core shell formation on a ZnO
nanowire, though the cell efciency rose only 0.29%. In that
report, the TiO2 shell was deposited by an atomic layer deposition
method followed by annealing at 220 1C, which made the process
somewhat complicated [16].
ZnSe is a prospective material for use in optoelectronic devices
due to its direct band gap and large exciton binding energy
(22 meV) [17,18]. In addition, ZnO/ZnSe structures form type II
band alignments that facilitate electron transfer [19,20]. Type II
band alignments of heterostructures, which are formed by coating
nanowires, induce charge separation to different regions [21].
Charge separation is advantageous for photovoltaic devices
because injected electrons in such devices are conned within
the core material [21]. In particular, when a ZnO/ZnSe core/shell
nanostructure was prepared using Na2SeSO3 in an autoclave at
160 1C on a Zn sheet, the structure exhibited superior photoelectrochemical properties [22]. However, the actual effect of the
ZnSe shell on the solar cell was not elucidated. Wu et al. observed

536

J. Chung et al. / Journal of Physics and Chemistry of Solids 73 (2012) 535539

a strong photoresponse from well-aligned ZnO/ZnSe core/shell


nanowire arrays with type-II energy alignment, which were
prepared by chemical vapor deposition at 620 1C [23]. However,
in that study, the effect of the ZnSe shell on DSSC performance
was also not determined. Therefore, in this study, we prepared
ZnSe shells on ZnO nanowires using a relatively simple successive
ionic layer adsorption and reaction (SILAR) method and investigated their effects on ZnO nanowire-based DSSCs. This strategy
may offer opportunities for the generation of Zn-based type II
band alignment structures and for their application to solar cells.

Abet-Technologies). The morphologies of the nanowires were


observed using a eld-emission scanning electron microscope (FESEM, Hitachi S-4200). The formation of ZnSe shells on the ZnO
nanowires was examined by X-ray photoelectron spectroscopy (XPS,
Thermo Electron Corporation, K-Alpha) and high-resolution transmission electron microscopy (HR-TEM, JEOL, JEM-3010). The crystallinities of the ZnO and ZnO/ZnSe core/shell structures were
investigated by X-ray diffraction (XRD, Bruker D8 Discover). Photoluminescence spectroscopy (PL, SPEX 1403) was used to investigate
the optical properties and surface defects of ZnO nanostructures.

3. Results and discussion

2. Experiments
A ZnO seed layer was deposited on uorine-doped tin oxide (FTO)
glass by radio frequency (RF) sputtering and was annealed at 300 1C
in an O2 ambient atmosphere for 1 h. For the growth of ZnO
nanowires, a precursor was prepared by dissolving 30 mM of zinc
nitrate hexahydrate (Zn(NO3)2  6H2O, 98%, Aldrich) in deionized
water. Then, ammonium hydroxide (NH4OH, 28% NH3 in water; J.T.
Baker) was added to set the pH of the solution to 10.3. ZnO
nanowires were grown on the seed layer by a hydrothermal method
at 60 1C for 24 h. The precursor solution was refreshed every 6 h. The
FTO glass on which the ZnO nanowires had grown was taken out of
the precursor solution, rinsed in deionized water and dried.
The SILAR method was used to deposit ZnSe layers on the ZnO
nanowires. A zinc nitrate precursor was prepared by dissolving
2 mM of zinc nitrate hexahydrate in deionized water. A sodium
selenosulfate solution was prepared by stirring 0.02 M selenium
(Se, 99.99%, Aldrich) and 0.05 M sodium sulte (Na2SO3, 98100%,
Aldrich) at 70 1C for 24 h. Then, a 2 mM sodium selenosulfate
solution was prepared by dilution. ZnO nanowires grown on the
FTO glass were sequentially immersed in 2 mM zinc nitrate
hexahydrate and 2 mM sodium selenosulfate solution at room
temperature for 20 s each. These SILAR steps were repeated 30 or
60 times, and the resulting samples were labeled as S-30 and
S-60, respectively. The samples were rinsed in deionized water
between each step.
The ZnO nanowire-grown FTO glass with or without a ZnSe
shell was immersed in 0.5 mM N719 dye (cis-bis(isothiocyanato)
bis(2,20 -bipyridyl-4,40 -dicarboxylato)ruthenium(II)bistetra-butylammonium) in ethanol for 2 h. In order to form a cathode, a Pt
layer was deposited on another piece of glass using a DC
sputtering process. Sputtering was performed using a Pt (99.99%
purity) target in an Ar ambient atmosphere at 100 W. The ZnO
nanowire-grown FTO glass was covered with the Pt-deposited
glass separated by 50-mm-thick hot-melt spacers (Type 1702,
Himilan of Mitsui-DuPont Polychemical). Then, the space
between the anode and the cathode was lled with a liquid
electrolyte (0.5 M LiI, 50 mM I2, 0.5 M 4-tertbutylpyridine in
3-methoxypropionitrile).
The performances of the DSSCs were assessed by current density
voltage measurements using a potentiostat (CompactStat) under AM
1.5 G of simulated sunlight (Sun 2000 Series Solar Simulators, 11000,

500 nm

500 nm

The morphologies of the ZnO nanowires and ZnO/ZnSe core/


shell nanostructures observed by FE-SEM are shown in Fig. 1. In
order to observe the effects of the addition of ZnSe layers, sample
S-60 (subjected to 60 cycles of ZnSe SILAR) was compared with
as-grown ZnO nanowires in Fig. 1. The diameters and lengths of
the ZnO nanowires were 50100 nm and 5.3 mm, respectively,
with or without the ZnSe SILAR process.
HR-TEM images of ZnO nanowires with and without ZnSe
layers are shown in Fig. 2. While the as-grown ZnO nanowires had
atomically smooth surfaces without additional surface layers
(Fig. 2(a)), thin ZnSe layers were observed on the ZnO nanowires
that underwent the SILAR process (Fig. 2(b and c)). The average
thickness of the ZnSe shells was 9 nm and 13 nm after 30 and 60
SILAR cycles, respectively, and a nearly linear increase in shell
thickness with increasing SILAR cycles was observed. The thickness of the ZnSe shell was uniform over the ZnO nanowires (data
not shown). The change in crystallinity was investigated by
measuring selected area electron diffraction (SAED) patterns of
the ZnO nanowires. The SAED pattern of the ZnO nanowires
without ZnSe layers shown in Fig. 2(d) indicates that ZnO
nanowires were grown with a preferential [0002] orientation.
Rings corresponding to the interplanar (111) orientation of the
ZnO/ZnSe core/shell nanostructure, as seen in Figs. 2(e and f),
represent the zinc blend structure of ZnSe [24]. Consequently, the
HR-TEM and SAED patterns of the ZnO/ZnSe core/shell nanostructures reveal that thin ZnSe layers were successfully synthesized on single crystalline ZnO nanowires.
We used a SILAR deposition mechanism to prepare a ZnSe shell
layer. In the rst SILAR step, ZnO nanowires were immersed in the
zinc nitrate solution, and Zn2 ions were adsorbed onto the surfaces
of the ZnO nanowires. In the next step, Se2 anions produced in the
sodium selenosulfate solution [25] reacted with Zn2 ions on the
surfaces of the ZnO nanowires to deposit ZnSe. Thus, the reactions
required for the formation of the ZnSe layer are as follows:
Zn(NO3)2-Zn2 2NO3 ,

(1)

Na2SeSO3 OH -Na2SO4 HSe ,

(2)

HSe  OH  -H2OSe2  ,

(3)

Zn2 Se2  -ZnSe.

(4)

1 m

1 m

Fig. 1. FE-SEM images of ZnO nanowires with and without ZnSe layers: (a) top view of as-grown ZnO nanowires; (b) top view of sample S-60; (c) cross-sectional view of
as-grown ZnO nanowires; and (d) cross-sectional view of sample S-60.

J. Chung et al. / Journal of Physics and Chemistry of Solids 73 (2012) 535539

537

Zn

Se

Zn

Zn

ZnS

ZnO

5 nm

5 nm

5 nm

0001

0110
111

111

Fig. 2. HR-TEM images of: (a) as-grown sample; (b) sample S-30; and (c) sample S-60. SAED patterns of: (d) as-grown sample, (e) sample S-30, and (f) sample S-60.

S-60

Intensity (a. u.)

Intensity (a. u.)

S-60

S-30

S-30

as-grown
as-grown

53

54
55
56
57
Binding energy (eV)

Current density (A

Intensity (a. u.)

/m2)

as-grown
S-30
S-60

300

20

58

400

500
600
700
Wavelength (nm)

800

30

80

40

50
60
2 (deg.)

70

80

as-grown
S-30
S-60

60
40
20
0

0.1

0.2
0.3
0.4
Voltage (V)

0.5

Fig. 3. (a) XPS Se3d spectra of ZnO nanowires with and without ZnSe shells; (b) XRD patterns of ZnO nanowires with and without ZnSe shells; (c) PL spectra
of ZnO nanowires with and without ZnSe shells; (d) photocurrent density vs. voltage curves of DSSCs fabricated using ZnO nanowires with and without ZnSe
shells.

538

J. Chung et al. / Journal of Physics and Chemistry of Solids 73 (2012) 535539

core/shell nanostructures resulted from these large increases in


the JSC and ll factor. On the other hand, samples S-30 and S-60
exhibited degraded VOC, which is thought to have resulted from
the positive shift of the conduction band edge of the anode [32] or
back electron transfer [33]. The positive shift in the conduction
band edge may increase the injection rate of the dyes due to an
increased driving force for injection and a higher density of
conduction band states accessible to the dye excited state [34],
which will result in an enhanced photocurrent. However, further
study is needed to fully understand the current observation.
Increases in the JSC of the DSSCs can be attributed to the
increased amount of dye on the nanowires and/or suppression of
electron recombination. However, because no signicant changes
in nanowire surface area, e.g., nanowire density and/or length,
were observed after ZnSe layer deposition (Fig. 1), the amount of
dye on the nanowire surface was not the primary factor inuencing JSC and cell efciency. As illustrated in Fig. 4(a), ZnO
nanowires with ZnSe shells and N719 dye form a type II band
alignment [18,35,36], which is known to facilitate electron
transfer [20]. The advantage of the ZnSe shell may be conferred
only when the LUMO-HOMO structures of the ZnSe shell and dye
form a type II band alignment, as shown in the energy band
diagram in Fig. 4(a). In addition, type-II alignment produces
longer recombination lifetimes [37], which is favorable in DSSCs
because the increase in lifetime slows interfacial recombination
[38]. In addition, the photocurrent is determined by ambipolar
diffusion of electrons coupled to counterion charges [39]. Thus,
small-radius Li in the liquid electrolyte may penetrate into the
dye-coated anode surface and form an ambipolar Li e  with the
electrons in the conduction band of the ZnO/ZnSe nanostructure. As a
result, electron transfer in the anode material accelerates, leading to
an increase in the JSC of the DSSC [40,41]. Therefore, the behavior
illustrated in Fig. 3(d) can also be explained by a change in the
ambipolar diffusion constant for electrons due to the ZnSe coating. It
has been reported that electrons need to diffuse to a near-substrate
region in the case of near-substrate recombination [42].
Conversely, without ZnSe shells the dye is adsorbed directly
onto the ZnO nanowire surfaces (Fig. 4(a)). In such cases, defects
near the ZnO surface provide recombination sites and degrade JSC
in the DSSC [43]. The quantity of defects was decreased by
depositing ZnSe layers onto the ZnO nanowires (Fig. 3(a)). The
relationship between the JSC of the DSSCs and the PL visible
emission intensity of ZnO nanowires with and without ZnSe
layers obtained from Fig. 3(c) is plotted in Fig. 4(b). Here, the PL
visible emission intensity was normalized to as-grown ZnO
nanowires without ZnSe layers. In the gure, JSC increases with
decreasing visible emission intensity, which is proportional to the
amount of defects on the ZnO. Therefore, deposition of ZnSe

In order to conrm the formation of the ZnSe layer on the ZnO


nanowires, XPS Se3d peaks were observed for the samples, as
shown in Fig. 3(a). Se3d peaks at 55.455.7 correspond to the
binding energy of ZnSe [26], and therefore, the thin layers
observed on the ZnO nanowires were concluded to be ZnSe. The
crystallinities of the nanowires were investigated by XRD, and the
results are shown in Fig. 3(b). All nanowires with and without
ZnSe layers had hexagonal wurtzite structures with preferential
(002) c-axis orientation. Although the formation of ZnSe layers
was conrmed by XPS analysis, no additional XRD peaks representative of ZnSe were observed, possibly due to the fact that the
ZnSe layers were extremely thin [27].
In order to determine the effect of the ZnSe shell on the optical
properties of the ZnO nanowire structure, room temperature PL
spectra of ZnO nanowires with and without ZnSe layers were
measured. As Fig. 3(c) shows the appearance of UV and visible
emission peaks in all PL spectra, additionally, the visible emissions of the ZnO nanowires at 560 nm decreased after ZnSe layer
formation. The visible emissions can be attributed to defects in
the ZnO crystals, such as singly ionized oxygen vacancies [28] and
interstitial oxygen defects [29], while the UV emission represents
near band edge emission relative to free exciton recombination
[30,31]. Therefore, the results in Fig. 3(c) suggest that the
formation of ZnSe layers on ZnO nanowires minimizes ZnO
surface defects.
DSSCs were fabricated using ZnO nanowires in order to
examine the improved surface properties of ZnO nanowires after
ZnSe shell formation. Surprisingly, as shown in Table 1, the cell
efciency of the DSSCs was dramatically increased after ZnSe
layer formation from 0.11% for as grown ZnO nanowires to 1.37%
and 1.29% for samples S-30 and S-60, respectively. The photocurrent density vs. voltage curves in Fig. 3(d) also show that the
JSC and ll factor of the DSSCs dramatically increased after
formation of the ZnSe layer. The S-30 and S-60 samples showed
JSC measurements of 79.7 and 66.9 A/m2 and ll factors of 0.39
and 0.40, respectively, while for the non-ZnSe-layered sample, the
measurements were 11.4 A/m2 and 0.26. The signicant improvements in cell efciency observed for the DSSCs with ZnO/ZnSe

Table 1
Performances of the DSSCs prepared using ZnO nanowires and ZnO/ZnSe
core/shell nanostructures.
Sample

JSC [A/m2]

VOC [V]

FF

Efciency [%]

as-grown
S-30
S-60

11.4
79.7
66.9

0.54
0.46
0.51

0.26
0.39
0.40

0.11
1.37
1.29

-3.89 eV

80

-4.4 eV

-5
-5.05 eV
-6
-6.28 eV

-7
-8

ZnSe

60
40
20

-7.6 eV
ZnO

JSC (A /m2)

E vs. vaccuum (eV)

-4

100

-3.25 eV

-3

N719

0
100
98
96
94
92
90
Normalized PL visible intensity (%)

Fig. 4. (a) Schematic illustration of energy band diagrams of ZnO, ZnSe, and N719 dye. (b) Changes in JSC of the ZnO nanowire-based DSSC with the normalized PL visible
emission intensity.

J. Chung et al. / Journal of Physics and Chemistry of Solids 73 (2012) 535539

layers on the ZnO nanowires likely diminished the recombination


of injected electrons. Finally, the suppression of the recombination of injected electrons and the facilitation of electron transfer
due to the formation of ZnSe layers on the ZnO nanowires were
the main factors leading to increases in JSC and efciency in ZnO
nanowire-based DSSCs.

4. Conclusions
In summary, thin ZnSe layers were synthesized on ZnO
nanowires using a SILAR method. TEM, SAED, and XPS measurements indicated that uniform ZnSe layers were deposited on the
ZnO nanowires. ZnSe shell thickness increased almost linearly
with the number of SILAR cycles. Using the PL spectra, we
conrmed that the defect-related visible emission was decreased
by depositing ZnSe layers on the ZnO nanowires. DSSCs were
fabricated using ZnO nanowires with and without ZnSe layers and
compared. DSSCs with ZnSe shells exhibited signicant increases
in cell performance compared to unmodied ZnO nanowires due
to a large increase in JSC. Finally, we concluded that the increased
JSC of the DSSCs with ZnSe shells on ZnO nanowires may be
attributed to the successful formation of type II band alignments
of ZnO nanowires, ZnSe shells, and an N719 dye structure, which
reduced the number of defect sites on the ZnO nanowire surfaces
and facilitated electron transfer.

Acknowledgments
This research was supported by the Basic Science Research
Program through the National Research Foundation of Korea
(NRF) funded by the Ministry of Education, Science and Technology (2010-0010573). This work was also supported by the Priority
Research Centers Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science
and Technology (2009-0093823).
References
[1] M. Law, L.E. Greene, J.C. Johnson, R. Saykally, P. Yang, Nat. Mater. 4 (2005)
455.
[2] C.-H. Ku, J.-J. Wu, Appl. Phys. Lett. 91 (2007) 093117.
[3] G. Hua, Y. Zhang, J. Zhang, X. Cao, W. Xu, L. Zhang, Mater. Lett. 62 (2008)
4109.
[4] Y.F. Hsu, Y.Y. Xi, A.B. Djurisic, W.K. Chan, Appl. Phys. Lett. 92 (2008) 133507.
[5] J.-J. Wu, G.-R. Chen, H.-H. Yang, C.-H. Ku, J.-Y. L ai, Appl. Phys. Lett. 90 (2007)
213109.

539

[6] I. Bedja, P.V. Kamat, X. Hau, A.G. Lappin, S. Hotchandani, Langmuir 13 (1997)
2398.
[7] J. Chung, J. Lee, S. Lim, Physica B 405 (2010) 2593.
[8] S.H. Ko, D. Lee, H.W. Kang, K.H. Nam, J.Y. Yeo, S.J. Hong, C.P. Grigoropoulos,
H.J. Sung, Nano Lett. 11 (2011) 666.
[9] C.Y. Jiang, X.W. Sun, G.Q. Lo, D.L. Kwong, J.X. Wang, Appl. Phys. Lett. 90 (2007)
263501.
[10] W. Yang, F. Wan, S. Chen, C. Jiang, Nanoscale. Res. Lett. 4 (2009) 1486.
[11] M. Law, L.E. Greene, A. Radenovic, T. Kuykendall, J. Liphardt, P. Yang, J. Phys.
Chem. B 110 (2006) 22652.
[12] Y. Zhang, L.-W. Wang, A. Mascarenhas, Nano. Lett. 7 (2007) 1264.
[13] J. Schrier, D.O. Demchenko, L.-W. Wang, Nano. Lett. 7 (2007) 2377.
[14] Y. Tak, S.J. Hong, J.S. Lee, K. Yong, J. Mater. Chem. 19 (2009) 5945.
[15] J. Han, F. Fan, C. Xu, S. Lin, M. Wei, X. Duan, Z.L. Wang, Nanotechnology 21
(2010) 405203.
[16] L.E. Greene, M. Law, B.D. Yuhas, P. Yang, J. Phys. Chem. C 111 (2007) 18451.

[17] H. Wenisch, K. Schull,


D. Hommel, G. Landwehr, D. Siche, H. Hartmann,
Semicond. Sci. Technol. 11 (1996) 107.

[18] S. Lee, F. Michl, U. Rossler,


M. Dobrowolska, J.K. Furdyna, Phys. Rev. B 57
(1998) 9695.
[19] J. Zheng, Z. Wu, W. Yang, S. Li, J. Kang, J. Mater. Res. 25 (2010) 1272.
[20] S. Gubbala, V. Chakrapani, V. Kumar, M.K. Sunkara, Adv. Funct. Mater. 18
(2008) 2411.
[21] H. Zhong, Y. Zhou, Y. Yang, C. Yang, Y. Li, J. Phys. Chem. C 111 (2007) 6538.
[22] P. Chen, L. Gu, X. Cao, Cryst. Eng. Comm. 12 (2010) 3950.
[23] Z. Wu, Y. Zhang, J. Zheng, X. Lin, X. Chen, B. Huang, H. Wang, K. Huang, S. Li,
J. Kang, J. Mater. Chem. 21 (2011) 6020.
[24] W.C.H. Choy, S. Xiong, Y. Sun, J. Phys. D Appl. Phys. 42 (2009) 125410.
[25] R.B. Kale, S.D. Sartale, B.K. Chougule, C.D. Lokhande, Semicond. Sci. Technol.
19 (2004) 980.
[26] N. Xu, B.H. Boo, J.K. Lee, J.H. Kim, J. Phys. D Appl. Phys. 33 (2000) 180.
[27] K. Wang, J. Chen, W. Zhou, Y. Zhang, Y. Yan, J. Pern, A. Mascarenhas, Adv.
Mater. 20 (2008) 3248.
[28] L.E. Greene, M. Law, J. Goldberger, F. Kim, J.C. Johnson, Y. Zhang, R.J. Saykally,
P. Yang, Angew. Chem. Int. Ed. 42 (2003) 3031.
[29] D. Li, Y.H. Leung, A.B. Djurisic, Z.T. Liu, M.H. Xie, S.L. Shi, S.J. Xu, W.K. Chen,
Appl. Phys. Lett. 85 (2004) 1601.
[30] L.H. Quang, S.J. Chua, K.P. Loh, E. Fitzgerald, J. Cryst. Growth 287 (2006) 157.
[31] Y.J. Xing, Z.H. Xi, Z.Q. Xue, X.D. Zhang, J.H. Song, R.M. Wang, J. Xu, Y. Song,
S.L. Zhang, D.P. Yu, Appl. Phys. Lett. 83 (2003) 1689.

[32] D. Cahen, G. Hodes, M. Gratzel,


J.F. Guillemoles, I.J. Riess, Phys. Chem. B 104
(2000) 2053.
[33] Y. Diamant, S.G. Chen, O. Melamed, A. Zaban, J. Phys. Chem. B 107 (2003)
1977.
[34] S. Ferrere, B.A. Gregg, J. Phys. Chem. B 105 (2001) 7602.
[35] R. Zhu, C.-Y. Jiang, B. Liu, S. Ramakrishna, Adv. Mater. 21 (2009) 994.
[36] H.-L. Yip, S.K. Hau, N.S. Baek, A.K.-Y. Jen, Appl. Phys. Lett. 92 (2008) 193313.
[37] K. Ohdaira, H. Murata, S. Koh, M. Baba, H. Akiyama, R. Ito, Y. Shiraki, J. Phys.
Soc. Jpn. 72 (2003) 3271.
[38] X.-T. Zhang, H.-W. Liu, T. Taguchi, Q.-B. Meng, O. Sato, A. Fujishima, Sol.
Energy Mater. Sol. Cells 81 (2004) 197.
[39] N. Kopidakis, E.A. Schiff, N.-G. Park, J. van de Lagemaat, A.J. Frank, J. Phys.
Chem. B 104 (2000) 3930.
[40] C.L. Olson., J. Phys. Chem. B 110 (2006) 9619.
[41] Y. Liu, A. Hagfeldt, X.R. Xiao, S.E. Lindquist., Sol. Energy Mater. Sol. Cells 55
(1998) 267.
[42] K. Zhu, E.A. Schiff, N.-G. Park, J. van de Lagemaat, A.J. Frank, Appl. Phys. Lett.
80 (2000) 685.
[43] J. Weidmann, Th. Dittrich, E. Konstantinova, I. Lauermann, I. Uhlendorf,
F. Koch, Sol. Energ. Mat. Sol. Cells 56 (1999) 153.

You might also like