You are on page 1of 25

ARTICLE IN PRESS

Solar Energy Materials & Solar Cells 90 (2006) 549–573


www.elsevier.com/locate/solmat

Review

Review of recent progress in solid-state


dye-sensitized solar cells
Bin Li, Liduo Wang, Bonan Kang, Peng Wang, Yong Qiu
Key Lab of Organic Optoelectronics & Molecular Engineering of Ministry of Education,
Department of Chemistry, Tsinghua University, Beijing 100084, China
Received 26 January 2005; received in revised form 2 April 2005; accepted 29 April 2005
Available online 28 June 2005

Abstract

The dye-sensitized nanocrystalline TiO2 solar cells (DSSCs) provide a promising alternative
concept to conventional p–n junction photovoltaic devices. However, liquid-state DSSCs
possess the problem of low stability since a volatile liquid electrolyte is utilized. An effective
approach to solve such a problem is by replacing the volatile liquid electrolyte with solid-state
or quasi solid-state hole conductor, such as p-type semiconductors, ionic liquid electrolyte and
polymer electrolyte. In this paper, the recent progress on the selection and utilization of these
hole conductors are mainly discussed. Research on mechanisms of solid-state DSSCs was also
summarized here including the hole transfer process at dye/hole conductor interface, ionic
transportation inside hole conductor media and the factors which depress the efficiency of
solid-state cells. With a thorough analysis of the problems of solid-state DSSCs, several ways
towards higher efficiency and lower cost are suggested.
r 2005 Elsevier B.V. All rights reserved.

Keywords: Solid state; Solar cell; TiO2; Hole conductor; Mechanism

Corresponding authors. Key Lab of Organic Optoelectronics & Molecular Engineering of Ministry of
Education, Beijing 100084, China. Tel.: +86 1062782197; fax: +86 1062795137.
E-mail addresses: chldwang@mail.tsinghua.edu.cn (L. Wang), qiuy@mail.tsinghua.edu.cn (Y. Qiu).

0927-0248/$ - see front matter r 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.solmat.2005.04.039
ARTICLE IN PRESS
550 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 550
2. Structure and operation principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 551
3. Development of hole conductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 552
3.1. P-type semiconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 552
3.2. Ionic liquid electrolytes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 556
3.3. Polymer electrolytes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 558
4. Mechanisms in solid-state DSSCs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 564
5. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 568
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 569

1. Introduction

It is assumed that the world energy demand will increase by about 70% between
2000 and 2030. Fossil fuels, supplying 80% of all energy consumed worldwide, are
facing rapid resource depletion. The resource reserves of fossil fuels throughout the
whole world in 2002 were projected to last 40 years for oil, 60 years for natural gas
and 200 years for coal. Because of a growing demand for energy, combined with the
depletion of fossil resources, global warming and its associated climate change, there
is an urgent need for environmentally sustainable energy technologies. Among all the
renewable energy technologies, such as wind turbines, hydropower, wave and tidal
power, solar cells, solar thermal, biomass-derived liquid fuels and biomass-fired
electricity generation, photovoltaic technology utilizing solar energy is considered as
the most promising one. Fortunately the supply of energy from the sun to the earth is
gigantic: 3  1024 J a year, or about 10,000 times more than that the global
population currently consumes. In other words, covering 0.1% of the earth’s surface
with solar cells with an efficiency of 10% would satisfy our present needs [1].
A photovoltaic system consists of solar cells and ancillary components. Solar cells
utilize the energy from the sun by converting solar radiation directly into electricity.
In 1954, researchers at the Bell Telephone Laboratories demonstrated the first
practical conversion of solar radiation into electric energy by use of a p–n junction
type solar cell with 6% efficiency [2]. With the advent of the space program,
photovoltaic cells made from semiconductor-grade silicon quickly became the power
source of choice for use on satellites. The common solar power conversion
efficiencies are between 15 and 20% [3]. However, the relatively high cost of
manufacturing these silicon cells has prevented their widespread use. Another
disadvantage of silicon cells is the use of toxic chemicals in their manufacture. These
aspects prompted the search for environmentally friendly and low cost solar cell
alternatives.
In 1991, dye-sensitized nanocrystalline TiO2 solar cells (DSSCs) based on the
mechanism of a fast regenerative photoelectrochemical process were first reported by
Grätzel et al. [4]. The overall efficiency of this new type of solar cell was 7.1–7.9%
(under simulated solar light), which is comparable to that of amorphous silicon solar
cells [1]. The main difference between this type of solar cell and conventional cells is
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 551

that in the new cells the functional element, which is responsible for light absorption
(the dye), is separated from the charge carrier transport. Such a feature makes it
possible for DSSCs to use low- to medium-purity materials through low-cost
processes, and exhibits a commercially realistic energy-conversion efficiency. In
addition, it is of great importance that the materials used in DSSCs are environment-
benign.
Although DSSCs based on liquid electrolytes have reached efficiency as high as
10% under AM 1.5 (1000 W/m2) [5], the main problem is that the liquid electrolytes
may limit device stability because the liquid may evaporate when the cell is
imperfectly sealed, and more generally, permeation of water or oxygen molecules
and their reaction with the electrolytes may worsen cell performance. Liquid
electrolytes also makes the manufacture of multi-cell modules difficult because cells
must be connected electrically yet separated chemically, preferably on a single
substrate [6–8]. Recently, many attempts have been made to solve the above
problems by the replacement of liquid electrolyte with solid or quasi solid-state hole
conductors [7–46].
In this review, we focus our attention on the recent development of solid-state
DSSCs, including solid-state hole conductors and previous research on the
mechanism of solid-sate DSSCs. The problems researchers have encountered and
the prospects of solid-state DSSCs are also provided.

2. Structure and operation principle

A schematic presentation of the structure of solid-state DSSCs is given in Fig. 1


[47]. At the heart of the system is a mesoporous TiO2 film, which is placed in contact

Fig. 1. Structure of solid-state dye-sensitized nanocrystalline TiO2 solar cells.


ARTICLE IN PRESS
552 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

with a solid-state hole conductor. Attached to the surface of the nanocrystalline TiO2
film is a monolayer of a charge transfer dye. Photo-excitation of the dye results in the
injection of an electron into the conduction band of the TiO2. The original state of
the dye is subsequently restored by electron donation from the hole conductor. The
regeneration of the sensitizer by the hole conductor intercepts the recapture of the
conduction band electron by the oxidized dye. The hole conductor is regenerated in
turn at the counter-electrode, and the circuit is completed via electron migration
through the external load [48].

3. Development of hole conductors

The hole conductors employed in solid-state DSSCs can be classified as p-type


semiconductors, ionic liquid electrolytes and polymer electrolytes.

3.1. P-type semiconductors

The most common approach to fabricate solid-state DSSCs is by using p-type


semiconductors. Several aspects are essential for any p-type semiconductor in a
DSSC [10,47]:

(1) It must be able to transfer holes from the sensitizing dye after the dye has injected
electrons into the TiO2; that is, the upper edge of the valence band of p-type
semiconductors must be located above the ground state level of the dye.
(2) It must be able to be deposited within the porous nanocrystalline layer.
(3) A method must be available for depositing the p-type semiconductors without
dissolving or degrading the monolayer of dye on TiO2 nanocrystallites.
(4) It must be transparent in the visible spectrum, or, if it absorbs light, it must be as
efficient in electron injection as the dye.

Many inorganic p-type semiconductors satisfy several of the above requirements;


however, the familiar large-band gap p-type semiconductors such as SiC and GaN
are not suitable for use in DSSCs since the high-temperature deposition techniques
for these materials will certainly degrade the dye. After extensive experimentation, a
type of inorganic p-type semiconductor based on copper compounds such as CuI,
CuBr, or CuSCN [10,24,49,50] was found to meet all of these requirements.
These copper-based materials can be cast from solution or vacuum deposition to
form a complete hole-transporting layer, and CuI and CuSCN share good
conductivity in excess of 102 Scm1, which facilitates their hole conducting ability
[47].
A solid-state DSSC based on CuI was first demonstrated by Tennakone et al. [10]
in 1995. The short-circuit current density was about 1.5–2.0 mA cm2 in sunlight
(about 800 Wm2), which was probably the highest record for solid-state DSSCs up
to then. By replacing cyanidin with a Ru-bipyridyl complex [7,51], Tennakone et al.
[7] reported an efficiency of 2.4% in solid-state DSSCs made with CuI.
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 553

However, DSSCs utilizing CuI are not very stable. In 2003, Sirimanne et al. [52]
found that the deterioration of CuI-based solid-state DSSCs was very rapid, even
much faster than that of liquid-state DSSCs. One of the reasons responsible for the
degradation is that the excessive iodine in the CuI film strongly decreased the
photocurrent of the cell [38]. Also, CuI tends to be oxidized under continuous
illumination [53]. Taguchi et al. [42] and Kumara et al. [54] fabricated solid-state
DSSCs with a MgO-coated TiO2 nanoporous film and a CuI layer. The solar cell as
fabricated showed both improved conversion efficiency and stability. They
attributed the improved lifetime to the suppression of the photooxidation capability
of TiO2 by blocking the transfer of photogenerated holes to the CuI layer.
Another important factor for the degradation of CuI-based DSSCs is the
loosening of the contact between the dye-sensitized TiO2 surface and CuI crystallites.
CuI deposited from the acetonitrile solution produces large (10 mm) crystallites that
do not penetrate into the pores of the nanocrystalline matrix and form loose contacts
with TiO2 [30]. Tennakone et al. [30] found that the stability of the CuI-based DSSC
can be greatly improved by incorporation of a small quantity (103 M) of 1-
methyl-3-ethylimidazolium thiocyanate (MEISCN) in the coating solution (i.e., CuI
in acetonitrile). In this case, MEISCN acted as a CuI crystal growth inhibitor and at
the same time the thin film of this compound remaining at the CuI grain boundaries
or the interface between CuI and the dyed TiO2 seemed to allow hole conduction.
The same MEISCN was utilized by Meng et al. [55] to improve the cells’ efficiency to
3.8% with improved stability under continuous illumination for about 2 weeks.
However, MEISCN is fairly expensive as its purification involves a chromatographic
separation. In the subsequent work of Tennakone et al. [53], they found that the
simpler substance triethylamine hydrothiocyanate (THT) was equally or more
effective as a CuI crystal growth inhibitor and could be adopted for fabrication of
solid-state DSSCs.
One alternative to using CuI is employing CuSCN, which does not decompose to
SCN and there is no indication that stoichiometrically excessive SCN creates
surface traps in CuSCN. This is consistent with the observation that solid-state
DSSCs based on CuSCN have more stable performance [38]. The main constraint for
experimentation with CuSCN was the need for a satisfactory method of deposition.
In 2001, Kumara et al. [24] found that CuSCN can be deposited on Ru-dye coated
nanocrystalline TiO2 lms from a solution in n-propylsulphide ([C2H7]2S). The cell
efficiency was about 1.25%, higher than that of cells that make use of CuSCN
deposited electrolytically [56] or from acetonitrile solution containing HCl [16]. In
2002, O’Regan et al. [32] reported another solid-state DSSC with CuSCN as the hole
conductor, where CuSCN was also deposited into the pores of the dye-sensitized
nanocrystalline TiO2 film from a dilute solution in propylsulphide. The cells showed
an efficiency of 2% at 1 sun. It is noteworthy that the degree of pore filling achieved
was near 100% for TiO2 films less than 2 mm thick and fell to 65% for films about
6 mm thick. Such a degree of pore filling is especially high in solid-state DSSCs.
Although the performance of cells using electrodeposited CuSCN reported by
O’Regan et al. [56] in 1995 was relatively low, they developed electrodeposition
technology and achieved an overall energy conversion efficiency of 1.5% and
ARTICLE IN PRESS
554 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

efficiency of charge separation and collection of near 100% for absorbed photons in
2000 [19]. Nevertheless, the performance of cells made by CuSCN is still lower than
that of cells utilizing CuI, probably due to the relatively lower hole conductance.
Inorganic semiconductors other than Cu complexes have seldom been reported up
to now. In 2005, an inorganic oxide semiconductor (NiO) was first used in a DSSC
by Bandara et al. [57]. NiO was prepared by dipping TiO2 electrodes in NiSO4
solution for 3 h, and then sintered at 400 1C for 30 min. In this type of solid-state
DSSC, dye was absorbed on the surface of the NiO-coated TiO2 film and charge
carriers were transported to different layers by tunneling. The current of this kind of
DSSC was 0.15 mA and the voltage was 480 mV, not as high as DSSCs utilizing Cu
complexes. However, by the method described in the above work and with a suitable
p-type oxide semiconductor, one could fabricate a very stable DSSC.
Compared with inorganic p-type semiconductors, organic p-type semiconductors
(i.e. organic hole-transport materials) possess the advantages of having plentiful
sources, easy film formation and low cost. Such organic materials have been widely
used in organic solar cells [58–61], organic thin film transistors [62–64], and organic
light-emitting diodes [65–67].
In 1998, Grätzel et al. [14] reported the first efficient solid-state DSSC
employing amorphous organic hole-transport material (HTM), 2,20 ,7,70 -tetrakis
(N,N-di-p-methoxyphenyl-amine)9,90 -spirobifluorene (OMeTAD) (Fig. 2). By utiliz-
ing N(PhBr)3SbCl6 and Li[(CF3SO2)2N] as the dopants, a maximum value of the
incident monochromatic photo-to-current conversion efficiency (IPCE) of 33% and
overall efficiency of 0.74% were obtained. In 2001, the performance of solid-state
DSSCs based on spiro-OMeTAD was considerably improved by Krüger et al. [22]

Fig. 2. Several organic p-type semiconductors utilized in solid-state DSSCs.


ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 555

through controlling the charge recombination across the interface of the hetero-
junction. This was achieved by blending the hole conductor matrix with a
combination of 4-tert-butylpyridine (tBP) and Li[CF3SO2]2N. An open-circuit
voltage Voc of over 900 mV and a short-circuit current Isc of 5.1 mA were obtained,
yielding an overall efficiency of 2.56% at AM1.5 illumination. We can know from
this report that the Voc increases with the tBP concentration in the hole conductor
film. But when the tBP concentration is kept constant, an increase in lithium ion
concentration results in an increase of the Isc. A further work by Krüger et al. [29]
reported that by performing the dye adsorption in the presence of silver ions in the
dye solution, the efficiency of solid-state DSSCs employing spiro-OMeTAD was
improved to 3.2%, which is the highest efficiency of solid-state DSSCs obtained by
utilizing an organic p-type semiconductor up to now. It was confirmed that the silver
ions could not only increase the dye absorption but also cause a spectra blueshift by
coordinating the ions with to the thiocyanate ligands of dye. The hole conductor
matrix used in these cells was also formed by spin coating of a solution of spiro-
OMeTAD in chlorobenzene, containing Li[CF3SO2]2N, tBP and N(PhBr)3SbCl6.
As we can see from the above report by Grätzel et al. [14] that the performance of
solid-state DSSCs utilizing organic HTMs is always poor if no ionic salts (lithium
salts for common use) exist. However, ionic salts are not soluble in most organic
solvents as HTMs, and therefore require the addition of cosolvents such as
polyethylene oxide or tert-butylpyridine (tBP). In 2003, a new HTM based on an
arylamine TPD (TDP ¼ N,N0 -diphenyl-N,N0 -(m-tolyl)-benzidine) structure compris-
ing a tetraethylene glycol (TEG) group on the fluorene unit (referred to as HTM-
TEG1 in Fig. 2) was reported by Park et al. [68]. The side TEG group can help to
dissolve lithium ions without co-solvents such as tert-butylpyridine. With further
optimization, Park et al. [69] reported another supramolecule with the TEG group
(referred to as HTM-TEG2 in Fig. 2) in 2004. By utilizing such a lithium ion-bonded
supramolecule, they achieved an efficiency of 0.8%, which is comparable to that of
the solid-state DSSCs using OMeTAD [14].
Several other organic HTMs used in solid-state DSSCs and their performance are
listed in Table 1.
However, the conversion efficiencies of most of the solid-state DSSCs employing
organic p-type semiconductors are relatively low particularly under high light
irradiation. This is due to the low intrinsic conductivities of organic HTMs, the high
frequencies of charge recombination from TiO2 to HTMs, the poor electronic
contact between dye molecules and HTMs caused by incomplete penetration of solid
HTMs in the pores of the mesoporous TiO2 electrodes.
Among all the disadvantages of organic HTMs, the poor pore filling is considered
as the most important factor. In order to improve the pore filling of organic p-type
semiconductors, Yanagida et al. [13,70,71] proposed in situ photo-electrochemically
polymerized pyrrole as the hole transport phase in the pores of the dye-sensitized
mesoporous TiO2 films. However, the performance of these cells was poor in that the
polypyrrole itself absorbs visible light [13,71]. In 2004, they reported the fabrication
of solid-state DSSCs using in situ photo-electrochemically polymerized poly(3,4-
ethylenedioxythiophene) (PEDOT) as the hole transport phase [72,73]. PEDOT has
ARTICLE IN PRESS
556 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

Table 1
Performance of solid-state DSSCs utilizing different organic HTMs

Organic HTM Performance Reference

Pentacene 0.8%(Z) [35]


TPD 0.2%(IPCE) [11]
600 mV (Voc), 1.65 mA/cm2 (Isc) [129,130]
Polythiophene 570 mV (Voc), 27 mA/cm2 (Isc) [20]
0.16%(Z) [21,28]
650 mV (Voc), 60 mA/cm2 (Isc) [131]
440 mV (Voc), 68 mA/cm2 (Isc) [72]
Polyaniline 565 mV (Voc), 450 mA/cm2 (Isc), 0.12%(Z) [132]
1.15%(Z) [133]
Polypyrrole 0.1%(Z) [13]
630 mV(Voc), 70 mA/cm2 (Isc) [70]
770 mV(Voc), 94.7 mA/cm2 (Isc), 0.2%(Z) [71]

Voc: open-circuit voltage, Isc: short-circuit current, Z: overall conversion efficiency, IPCE: incident
monochromatic photo-to-current conversion efficiency.

high transparency in the visible range, high conductivity and remarkable stability at
room temperature. By this means, good pore filling and an overall efficiency of
0.53% at AM 1.5 illumination (100 mW cm2) was achieved [73].

3.2. Ionic liquid electrolytes

Room-temperature ionic liquids (RTILs) have good chemical and thermal


stability, negligible vapor pressure, nonflammability, and high ionic conductivity
[74,75]. When incorporated into DSSCs, they can be both the source of iodide and
the solvent themselves.
Papageorgiou et al. [76] reported in 1996 that these unique properties of RTILs
were effective for long-term operation of electrochemical devices such as DSSCs.
They employed methyl-hexyl-imidazolium iodide (MHImI) as an involatile
electrolyte, which melts at lower than room temperature. Cell performance showed
outstanding stability, with an estimated sensitizer redox turnover number in excess
of 50 million. However, the short-circuit photocurrent (Jsc) of the system was smaller
by one order of magnitude than that usually reported in DSSC systems using organic
solvents such as acetonitrile. Since the slow diffusion of iodide redox species in
highly viscous melted MHImI (ca. 1800 mPa s at 25 1C) causes a low Jsc, a low-
viscosity RTIL would be expected to improve the conversion efficiency of the DSSC.
In 1999, Hagiwara et al. [77] reported a low-viscosity RTIL EMIm-F?2.3HF
obtained by the reaction of 1-ethyl-3-methyl imidazolium chloride (EMIC) and
hydrogen fluoride. The viscosity of EMIm-F?2.3HF was 4.8 mPa s at 25 1C [74],
much lower than that of MHImI, and the specific conductivity was about 12 S m1 at
298 K. The application of such a low-viscosity RTIL to a DSSC system was
investigated by Matsumoto et al. [74] in 2001. By utilizing EMIm-F?2.3HF as an
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 557

involatile solvent containing 0.9 M of 1,2-dimethyl-3-hexylimidazolium iodide


(DMHImI) and 30 mM of I2, a cell efficiency of 2.1% was obtained. What was
particularly mentioned is that the triiodide that formed after the regeneration of
the oxidative form of dye with iodide might be transferred to the Pt counter
electrode by a Grotthus-like hopping mechanism rather than by the diffusion of
triiodide. This hopping mechanism for iodide redox will be important to improving
the performance of DSSCs using relatively high-viscosity RTILs.
Wang et al. [75] reported an ionic liquid electrolyte composed of 1-methyl-3-
propylimidazolium iodide (PMII), 1-methyl-3-ethylimidazolium dicyanamide
(EMIDCN), and lithium iodide. Combined with an amphiphilic polypyridyl
ruthenium sensitizer called Z-907 ((Ru(H2dcbpy)(dnbpy)(NCS)2, where H2dcbpy
is 4,40 -dicarboxylic acid-2,20 -bipyridine and dnbpy is 4,40 -dinonyl-2,20 -bipyridine)), a
quasi solid-state DSSC employing this ionic liquid electrolyte demonstrated an
efficiency of 6.6% at an irradiance of air mass 1.5. The addition of lithium iodide to a
pure ionic liquid electrolyte significantly enhanced the efficiency because of an
increase in the electron injection yield and dye regeneration rate.
In order to improve the conductivity of RTIL-based electrolytes, Yamanaka et al.
[78] reported in 2005 the utilization of an ionic liquid crystal (ILC) as a constituent of
the electrolyte, which forms a self-assembled structure and promotes the exchange
reaction by the locally increased concentrations of I2 and I 3 . The ionic liquid crystal
used here was 1-dodecyl-3-methylimidazolium iodide (C12MImI) and the current
along with the efficiency of the DSSC were improved.
Due to their nonvolatility, RTIL-based electrolytes have been applied to a
100 mm  100 mm-sized DSSC [79]. For the purpose of fabricating such a large-sized
DSSC, the influence of electrode distance, TiO2 nano-particle size, and then
thickness of the TiO2 nano-porous layer was investigated by Matsui et al. in ionic
liquid with a 1-ethyl-3-methylimidazolium bis(trifluoromethanesulfonyl)amide
(EMIm-TFSA) electrolyte system. In the optimal conditions, an energy conversion
efficiency of 4.5% was obtained using a 9 mm  5 mm small sized cell. The energy
conversion efficiency of a large sized cell was 2.7% in the ionic-liquid system.
Imidazolium compounds have been the most popular materials employed in
RTIL-based DSSCs. Besides the above imidazolium salts, some other materials
have also been investigated, such as 1-ethyl-3-methylimidazolium tetrafluoroborate
(EMImBF4) [80,81], and 1-butyl-3-methylimidazolium hexafluorophosphate
(BMImPF6) [81]. Recently, trialkylsulfonium polyiodides have been found to
exhibit good electrical conductivity, 103–104 S/cm, which increases with the
content of iodine. This can be explained by a mechanism of electrical conduction in
polyiodide chains via a Grotthus mechanism, a relay mechanism, in which the net
transport of charge is achieved without any net transport of mass [37]. Paulsson et al.
[37,82] synthesized a series of trialkylsulfonium iodides, such as (Et2MeS)I,
(Bu2MeS)I, and (Bu2EtS)I. With low concentrations of iodine, they exhibited good
conducting ability. DSSCs using iodine-doped (Bu2MeS)I as electrolyte achieved an
overall light-to-electricity conversion efficiency of 3.7% in simulated AM 1.5 solar
light at a light intensity of 0.1 Sun. Such efficiency is comparable to that of cells using
imidazolium RTILs.
ARTICLE IN PRESS
558 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

Although most of the efficient RTILs used in DSSCs are iodide, in 2004, a new
solvent-free, SeCN/(SeCN) 3 -based ionic-liquid electrolyte was reported by Wang
et al. [83]. The pseudohalides SCN/(SCN)2 and SeCN/(SeCN)2 redox couples in
acetonitrile have been investigated in liquid-state DSSCs by Oskam et al. [84]. Their
performance was found to be disappointing; the incident photon-to-current
conversion efficiency (IPCE) was 20% for the (SeCN)2/SeCN- couple and 4% for
the (SCN)2/SCN- couple, compared with 480% for the iodide/triiodide-based
system. This was attributed to the inefficient dye regeneration by the SeCN or
SCN electron donor. However, Wang et al. synthesized an SeCN/(SeCN) 3 -based
RTIL and got an exciting result. The viscosity of this ionic liquid, EMISeCN (1-
ethyl-3-methylimidazolium selenocyanate), was determined to be 25 cP at 21 1C,
which is extraordinarily low for an ionic liquid. For comparison, the most fluid
imidazolium iodide-based analogue up to now, i.e., 1-propyl-3-methylimidazolium
iodide (PMII), has a 35 times higher viscosity, i.e. around 880 cP at room
temperature. The specific conductivity of EMISeCN is 14.1 mS/cm compared with
0.5 mS/cm for PMII, and even at 30 1C, it is still 1 mS/cm, which is negligibly small
for PMII. Electron donation from the SeCN to the oxidized Z-907 dye, which was
used in the quasi solid-state DSSCs they fabricated, is even faster than from iodide
ions to the oxidized dye in the PMII ionic liquid [75]. Thus an unprecedented
7.5–8.3% power conversion efficiency under AM 1.5 sunlight has been achieved for
DSSCs with EMISeCN.
It is of great interest that the fill factors of ionic liquid electrolytes-based DSSCs
easily reach extremely high values of over 0.75 even in full sunlight. Dilution of the same
ionic liquid with a low-viscosity organic solvent often decreases the fill factor. This
unexpected behavior is likely to arise from the effective screening of the electric charges
that are produced under light illumination in the mesoporous films. The very high
density of ions present in these RTILs appears to facilitate charge separation although
the mechanism by which the screening is achieved remains obscure to date [48].

3.3. Polymer electrolytes

Since the efficiencies of solid-state DSSCs utilizing p-type semiconductors were


found to be unsatisfactory compared with those using organic liquid-phase redox
electrolytes and the RTIL-based DSSCs still have the disadvantage of being fluid, the
polymer electrolytes have been taken into account recently. It is of great importance
that polymer electrolytes have the advantages of relatively high ionic conductivity
and easy solidification. The use of conducting polymers as electrolytes or electrode
materials for the development of plastic-like electrochemical devices is an appealing
concept that dates back to the late 1960s. In the late 1970s Shirakawa et al. [85]
demonstrated that after halogenation, electronically conducting polymers, com-
monly called polymer electrolytes, such as polyacetylene (PAC) could be cycled in
lithium batteries as well as the common inorganic oxide intercalation cathodes.
From then on, polymer electrolytes have been widely used in lithium batteries
[86–88], redox type laminated supercapacitors [89], and multichromatic optical
windows [90].
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 559

Using a gel network polymer electrolyte is the usual way to improve the contact
between solid electrolytes and nanoporous TiO2 layers. Actually, quasi solid-state
DSSCs containing physical gels have been reported to have almost the same
performance as those containing liquid gel precursors, which remain in a liquid state
before becoming gelated to a solid-state gel. The process for the fabrication of quasi
solid-state DSSCs is given in Fig. 3 [91]. Gel electrolyte precursors containing liquid
electrolytes and gelators are injected into cells that are already set. Gelation is
usually carried out in the cell by heating the cell. Gel electrolyte precursors are liquid
at first, which makes it possible for the precursors to be impregnated into nanoscale
pores of TiO2 films. By heating, gelators propagate along a three-dimensional
polymer network in the cell. Several conditions are necessary for gel electrolytes [91]:

(1) Polymerization must occur in the presence of iodine.


(2) Polymerization must occur at a temperature below which the dye would not
decompose.
(3) Polymerization should be able to be initiated and completed even in the presence
of some impurities such as oxygen, water, ions and so on.
(4) Polymerization has to be completed without generating byproducts that could
decrease the photovoltaic performance.
(5) Polymerization can proceed without an initiator because the resultant decom-
position products of the initiator may decrease the photovoltaic performance.

In the past years, many gel polymer systems have been employed. In 1998, some
amino acid derivatives with phenyl and alkyl groups were reported to work as
gelators for liquid-phase organic electrolytes at very low concentrations [15]. The
structure of such gelators is given in Fig. 4a. This was the first time that the
gelatinized hole-transport layer had comparable characteristics to the liquid-phase
hole-transport layer. Further studies on quasi solid-state DSSCs containing amino

Fig. 3. Process for fabrication of quasi solid-state DSSCs based on gel electrolyte.
ARTICLE IN PRESS
560 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

O H
N N
O
H O
O H O
N
O N N
H O H
O

O
H H
N N O
O N HN
H O O

(a) Amino acid gelators

O CH3
H2C [( OC2H4 )l (OC3H8) m ]n O C C CH2

O CH3

HC [( OC2H4 )l (OC3H8) m ]n O C C CH2

O CH3

H2C [( OC 2H4 )l (OC3H8 ) m ]n O C C CH2

(b) Poly(ethylene oxide-co-propylene oxide) trimethacrylate oligomer)

OEt O O OEt
H
OEt Si(CH2)3 NH C NH C CH2 [ OCH2CH] n NH C NH (CH2)3Si OEt
OEt
OEt CH3 CH3

(c) Ureasil PPG 230 (n=3)


Fig. 4. Structures of some of the gel polymer precursors.

acid gelators and the charge transport property of gel-state I/I 3 redox electrolytes
were reported by Kubo et al. [23] in 2001. An optimized quasi solid-state DSSC
showed the values of 0.67 V for open-circuit voltage, 12.8 mA cm2 for short-circuit
photocurrent density, and 5.91% for overall conversion efficiency under AM 1.5
irradiation with higher durability.
Poly(ethylene glycol) (PEO) based polymer electrolytes have been widely
investigated for use in lithium batteries [86–88]. In 2000, network polymer
electrolytes containing PEO chains were synthesized by a cross-linking reaction
[92] and in 2002, quasi solid-state DSSCs containing these PEO gel network polymer
electrolytes were fabricated by Ren et al. [33] and showed an efficiency as high as
3.6%. It was concluded that the molecular weight of PEO segments was important
for the ionic conductivity, and thus for the overall efficiency of solar cells. Recently,
such poly(ethylene glycol) segments have also been utilized in fabricating co-polymer
electrolytes. By cross-linking reactions among the reactive groups of poly(ethylene
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 561

oxide-co-propylene oxide) trimethacrylate oligomers (as presented in Fig. 4b), a


three-dimensional network polymer was formed [93]. Solid-state DSSCs with ion
conducting polymer electrolytes formed by such a three-dimensional network
polymer showed an extremely high efficiency of 8.1% under AM 1.5 irradiation
(100 mW/cm2). It was also indicated that the conversion efficiency could be
improved further by optimizing the thickness of TiO2 nanoporous electrodes and
also by improving the composition of the viscosity and dielectric constant of the
solvent. Another way to enhance the ionic conductivity of PEO-based polymer
electrolytes and to enlarge the interfacial contact area between dye-sensitized TiO2
particles and polymer electrolytes is to add an amorphous oligomer to the polymer
electrolyte. By using this method, Kang et al. [94] got an efficiency of 3.84% at
100 mW/cm2.
The most effective I 3 /I

redox couple introduced in gel electrolytes is usually
obtained by co-dissolving an iodide salt and iodine. However, no solvent can be
equally good for both substances, and for this reason crystallization, for example, of
KI, has been a major source of deterioration of the cell. A unique solvent, sulfolane,
was employed by Stathatos et al. [46] to solve the problem. Their gel electrolyte was
based on a ureasil precursor oligomer (as shown in Fig. 4c) and sulfolane, and it
incorporated the I 3 /I

redox couple. Encouraging results showed that the
combination of these two reagents prevents crystallization of KI, thus ensuring a
long life for the cell and a satisfactory overall efficiency surpassing 5%.
Mixing liquid electrolytes directly with polymers without in situ polymerization is
another important approach for the fabrication of solid-state DSSCs utilizing
polymer electrolytes. The first attempt was made by Cao et al. [9] in 1995. Their
polymer electrolyte was prepared by refluxing a mixture of polyacrylonitrile,
ethylene carbonate, propylene carbonate, acetonitrile and a desired concentration of
NaI at about 90 1C under an N2 atmosphere for a few hours. A cell using this
polymer electrolyte showed an overall conversion efficiency of 3–5%. Encouraged by
such a result, many other approaches have been investigated [17,95,96]. In 2003,
Kaneko et al. [97] reported solid-state DSSCs incorporating an inexpensive
polysaccharide solid involving redox electrolytes and an organic medium with an
overall efficiency of over 7%.
It is regarded that higher ionic mobilities occur in the amorphous rather than in
the crystalline phases of polymers in polymer electrolytes. As many polymers tend to
crystallize at room temperature, Nogueira et al. [17,18,25] employed one copolymer,
poly(epichlorohydrin-co-ethylene oxide), Epichlomer-16 for the application of solid-
state DSSCs. In their work, the Epichlomer-16 electrolyte was used without the
addition of any polymeric agent or plasticizer. Although the efficiency was only
0.22% at the beginning [18], it was improved to 2.6% with further optimization [25].
What is to be emphasized is that such a polymer has been utilized to make flexible
DSSCs. In 2003, Longo et al. [98] fabricated a soft solid-state DSSC containing the
Epichlomer-16 electrolyte with an efficiency of 0.32%, while Haque et al. [96]
reported a flexible solid-state DSSC based upon dye-sensitized nanocrystalline Al2O3
coated TiO2 films and an I2/NaI-doped solid-state Epichlomer-16 electrolyte with an
efficiency as high as 5.3%.
ARTICLE IN PRESS
562 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

As mentioned above, many polymer electrolytes exhibit very low ambient ionic
conductivity because of the severe crystallinity of polymers. The most common
approach for hindering the crystallinity, thus lowering the operational temperature
of polymer electrolytes, is to add liquid plasticizers, but this promotes deterioration
of the electrolyte’s mechanical properties and increases its reactivity with the back
electrodes. Croce et al. [86,88,99] once reported nanocomposite polymer electrolytes
for lithium batteries. They demonstrated conductivities as high as 104 S cm1 at
50 1C and 105 S cm1 at 30 1C in a PEO-LiClO4 mixture containing powders of
TiO2 and Al2O3 with particle sizes of 5.8–13 nm, while that of ceramic-free
electrolytes was only 104–108 S cm1 in the temperature range of 80–30 1C [86].
The increase in conductivity with respect to the corresponding unfilled electrolytes
that contain no inorganic nano-particles was attributed to the enlargement of the
amorphous phase in the polymer matrix.
To improve the mechanical, interfacial, and conductivity properties of polymer
electrolytes, Falaras et al. [100] presented a new type of solvent-free composite
polymer electrolyte based on high-molecular mass PEO filled with titania (TiO2), LiI
and I2. Efficiencies of 0.96% with a TiO2 electrode and 0.14% with an SnO2 [101]
electrode were obtained. They also reported that the structure modification of the
polymer electrolyte might be made not only by a combination of steric hindrance
effects but also by acid–base interactions between the filler surface groups and the
oxygen of the PEO. Via optimization of the corresponding solar cells in terms of a
polymer’s molecular weight and the nature and particle size distribution of the
organic particle filler, the efficiency of devices containing TiO2 electrodes was
improved to about 4.2% under direct sunlight illumination.
In 2001, Stathatos et al. [27] reported a new nanocomposite SiO2/poly(ethylene
glycol)-200 (PEG-200) electrolyte. In their work, emphasis was given to the nature
and the applicability of the SiO2/PEG-200 film. Compared with the binary PEO/
TiO2 composite electrolyte, the SiO2/PEG-200 electrolyte was prepared by
hydrolyzing tetramethoxysilane (TMOS) in PEG solution rather than by directly
mixing SiO2 particles with PEG. Such a sol–gel process facilitates the formation of
well cross-linked composite electrolytes. The inorganic sub-phase here provides the
gelling agent and simultaneously works as a gluing material holding the working and
the counter electrode together, without additional aids [41]. The efficiency of solid-
state DSSCs employing the above SiO2/PEG-200 electrolyte was 1.2%. This rather
low efficiency comes from the fact that the slow evaporation of volatile compo-
nents entrapped in the gel during the sol–gel process leads to agglomeration and
electric contact failure. To solve the problem, organically modified sol–gel precursors
were utilized, in which the polyether chains and the silica chains are covalently
bound (the structure is similar to that of ureasil PPG230, which is shown in Fig. 4c,
only n is equal to 68 here) were utilized [102]. The final product, after sol–gel
transformation, was composed of covalently linked SiO2 and poly(alkylene oxide)
networks, well mixed on the nanoscale. Thus the overall efficiency was improved to
about 4%.
As indicated above, RTILs have some qualities that would make them suitable as
a solvent, such as low viscosity, ionic mobility, and negligible volatility. What’s
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 563

more, the big cations in RTILs also facilitate the interaction between the RTILs and
polymer networks [103]. It is natural that many researchers have employed them as a
solvent in polymer electrolytes [36,104,105].
A quasi solid-state DSSC with an ionic polymer electrolyte was reported in 2002
by Kuto et al. [106]. In their previous report [8], poly(hexa(oxyethylene)methacry-
late) (MHH) was utilized to obtain thermoirreversible chemical gel in an organic
liquid electrolyte of a DSSC. However, the problem of stability or sealing still
remained due to the inclusion of organic solvent in the electrolyte. In Kuto’s study,
1-hexyl-3-methylimidazolium iodide was applied as the solvent to form the ionic gel
electrolyte. The DSSCs fabricated with such an electrolyte showed a 3.8%
conversion efficiency under AM 1.5 (100 mW cm2) irradiation, and long-term
stability without any sealing.
In 2003, Shibata et al. [40] also reported a quasi-solid DSSC with a gel electrolyte.
PMII was used to replace the traditional acetonitrile solvent and polyvinylpyridine
(PV) along with tetrabromomethylbenzene (B4Br) as the cross-linkers. In order to
decrease the charge transfer resistance between the gel film and counter electrode, a
conductive polymer, polystyrenesulfonate doped poly (3,4-ethylenedioxy-thiophene)
(PEDOT-PSS), was employed as the counter electrode. Such a polymer electrode has
been investigated by Saito et al. [34] in liquid-state DSSCs, and the performance of
cells utilizing the PEDOT-PSS counter electrode was found to be lower than that of
cells using a Pt counter electrode. However, in Shibata’s report, the photocurrent for
the DSSCs with a PEDOT-PSS electrode was higher than that with a Pt count
electrode when the cells were filled with a gel electrolyte. The charge transfer
resistance between the gel film and Pt was 14.1 O cm2 and that between gel film and
PEDOT-PSS decreased to 1.5 O cm2, which was the reason for the improvement of
the photocurrent. Such an approach is of great importance in pursuing better
performance of solid-state DSSCs.
From the above introduction, nanocomposite electrolytes containing ionic
liquid seems to be a good choice for fabricating highly effective and stable solid-
state DSSCs. Such an attempt has been made by Stathatos et al. [41]. A
nanocomposite organic/inorganic gel electrolyte made by the sol–gel method has
been reported in their previous work [27]. Recently, they employed 1-methyl-3-
propylimidazoliumiodide (MPII) into the nanocomposite electrolyte and the overall
efficiency was 5.4%, which is considered satisfactory for a device containing a quasi-
solid electrolyte. Wang et al. [107] also studied a similar nanocomposite ionic
electrolyte where the cis– bis(isothiocyanato)bis(2,20 -bipyridyl-4,40 -dicarboxylato)r-
uthenium(II) dye was replaced by an amphiphilic Z-907 dye. This new dye has been
used in ionic polymer electrolyte-based DSSCs and a high efficiency of 6.1% was
obtained [44]. When employed with a nanocomposite ionic electrolyte, Z-907 still
showed good performance in solid-state DSSCs. An efficiency as high as 7%
was obtained. It is believed that such an approach will enable the fabrication of
flexible, compact, laminated all solid-state devices free of leakage and available in
varied geometries.
In 2003, a novel nanocomposite ionic gel was reported by Fukushima et al. [108]; a
suspension of single-walled carbon nanotubes (SWCNTs) was ground in an
ARTICLE IN PRESS
564 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

Table 2
Characteristics of DSSCs with different nanocomposite electrolyte [110]

Electrolyte Voc/mV Jsc/mA cm2 Z/% FF

Ionic liquid 648 11.57 4.21 0.56


MWCNTs composite gel 706 12.02 4.79 0.57
CB composit gel 672 11.02 4.83 0.65
TiO2 composite gel 675 11.45 5.00 0.65
SWCVTs composite gel 695 10.78 4.60 0.62
CFs composite gel 688 11.11 4.97 0.65
Graphic composite gel 681 10.60 4.57 0.63

MWCNTs: multi-walled carbon nanotubes, CB: carbon black, SWCVTs: single-walled carbon nanotubes,
CFs: carbon fibers.

imidazolium ion-based RTIL. The gels were thermally stable and did not shrivel,
even under reduced pressure resulting from the nonvolatility of the ionic liquids. This
system has already been utilized in DSSCs. Usui et al. [109] dispersed carbon
nanotubes, other carbon nanoparticles and titanium dioxide nanoparticles into a
1-ethyl-3-methylimidazoliumbis(tribluoromethylsulfonyl)imide (EMIm-TFSI) ionic
liquid electrolyte following the grinding method described by Fukushima et al. [108].
The DSSCs fabricated with the above electrolyte showed superior performance
compared to those with pure ionic liquid electrolytes. Characteristics of DSSCs with
different inorganic nano-materials are summarized in Table 2. The performance of
DSSCs with carbon nanotubes and carbon nanoparticles was improved because
these cells had lower resistance and as such, they included conductive particles that
had lower resistance than ionic liquid electrolytes. However, in the case of TiO2
nanoparticles dispersed into ionic liquid electrolytes, TiO2 is a semiconductor rather
than a conductor. Usui et al.[109] theorized that the presence of TiO2 nanoparticles
enhanced the electron exchange reaction.

4. Mechanisms in solid-state DSSCs

The charge-transfer reactions at the dye-sensitized nanocrystalline TiO2/hole-


transporting material (HTM) interface play a key role in the determination of the
overall efficiency of solar cell devices. However, little was known about the electron-
transfer dynamics in such solid-state sensitized junctions until 1999, when Bach et al.
[110] reported for the first time on the direct observation of photoinduced, interfacial
charge separation across a dye-sensitized solid-state heterojunction by means of
picosecond transient absorption laser spectroscopy. A schematic diagram of the
solid-state device as they presented it is shown in Fig. 5. They concluded that the
dye regeneration by HTM could be time-resolved and it was shown to proceed
primarily on the picosecond time scale with multiphasic kinetics and at least 1 order
of magnitude faster than the dye regeneration process involving iodine/iodide
electrolytes [5].
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 565

Fig. 5. Schematic diagram of charge transporting in solid-state DSSCs with organic HTMs.

Such fast kinetics for hole injection from the dye to the hole conductor and
favorable energetics for this process are crucial for the development of efficient solid-
state DSSCs. In the work of Mahrov et al. [111], the injection of charge carriers from
dye molecules absorbed on hole conductors was investigated by transient and
spectral photovoltage measurements that were performed as previously described
[112]. When specific illumination was applied, a negative photovoltage signal
that arose due to the spatial charge separation between two electrodes was obtained.
This means that the positive charge is injected from the dye molecules into the
hole conductors, while the negative charge remains on the adsorbed dye mole-
cules. What should be emphasized is that the photovoltage technique can be
applied to any system that includes optically induced charge transfer and can
therefore be a useful tool to improve the understanding of DSSCs with solid-state
hole conductors.
In 2003, Haque et al. [113] studied the yield of hole transfer from dye cations
anchored on nanocrystalline TiO2 to a sequence of TPD-based organic p-type
semiconductors using transient absorption spectroscopy. They found that the yield
of hole transfer is controlled not by kinetic competition at the TiO2/dye/HTM
interface but by DG(dyeHTM) , which is defined as DG(dyeHTM) ¼ Em(HTM+/
HTM)Em(Dye+/Dye). They concluded that this free energy difference is
inhomogeneously broadened, which is most probably caused by local variations in
the electrostatics of the interface, and these local variations require a large average
DG(dyeHTM) in order to achieve a high hole-transfer yield.
As the efficiency of solid-state DSSCs is relatively lower than that of liquid-state
DSSCs, much attention has been paid to exploring the internal restriction factors of
ARTICLE IN PRESS
566 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

solid-state cells. In the case of solid-state DSSCs in which HTMs are substituted for
the liquid electrolyte, on the other hand, the carrier transport mechanism in the hole-
transport layer is markedly different from that in liquid-state DSSCs. In addition,
incomplete interfacial contact between the dye attached to the TiO2 electrode and the
hole-transport layer may significantly slow down the hole transfer from the former to
the latter and consequently enhance the back reaction. In order to explore the factors
reducing the energy conversion efficiency in solid-state DSSCs, a performance
simulation was carried out by Tanaka [114] based on the electric model of liquid-
state DSSCs presented by Ferber et al. [115]. By considering that the motion of a
hole (or an electron) in the hole-transport layer is a hopping-like movement between
the lattice sites and may be modeled alternatively in terms of the motion of neutral
and positively charged ions with the mobility of the holes [110], Tanaka [114]
concluded from his simulation that some specific factors such as the detachment of
the hole-transport layer from the TiO2 electrode and the decrease of the shunt
resistance associated with internal leakage are possible causes of the significant
suppression of the efficiency of solid-state DSSCs.
The comparison of electronic properties between liquid-state and solid-state
DSSCs has also been investigated by Kron et al. [116]. They compared the DC and
AC transport
( )
kT J SC E c  E 0 ðE c  E 0  lÞ2
V OC ¼ ln 1=2
þ þ (1)
q qsnth dðkT =plÞ N c cred q 4lq

properties of both types of devices. The temperature dependence of the open-circuit


voltage Voc was given by Eq. (1): where s is the reaction cross section, uth the thermal
velocity of electrons in the semiconductor, d the reaction layer thickness and cred
the concentration of reduced species in the electrolyte. The quantity E0 is related to
the redox potential Eredox via Nernst’s equation for the redox potential of electro-
lytes E redox ¼ E 0 þ kT lnðcox =cred Þ; l the solvation reorganisation energy and Ec
the energy of the conduction band edge of the semiconductor. As presented in Fig. 6,
the open-circuit voltage of solid-state DSSCs is 100–200 mV lower than that of
liquid-state DSSCs. This indicates a difference of the reference energy for
recombination as well as a difference of the recombination probability. The AC
admittance of liquid-state DSSCs under moderate forward bias is dominated by
diffusion of electrons in the TiO2. Solid-state DSSCs exhibit negative reactance
indicating an overall low conductivity of HTMs. The high recombination rate at the
TiO2/HTMs interface is also responsible for the reduction of performance in solid-
state DSSCs.
As mentioned above, the recombination mechanism has been one of the key issues
of solid-state DSSCs. In the work of O’Regan et al. [117], the photocurrent and
photovoltage transient techniques were used to explore the limiting factors in
recombination. It was found that recombination in open-circuit conditions is 10
times faster than in electrolyte cells, and recombination in short-circuit conditions
is 100 times faster. The slow recombination rate in the I 3 /I

electrolyte cells is
determined largely by the chemistry of the recombination reaction. The initial step
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 567

Fig. 6. The extrapolation of the open-circuit voltage Voc toward temperature T ¼ 0 K.

is thought to be the reduction of iodine rather than tri-iodide. Because the


electrolytes contain at least a 5-fold excess of iodide, iodine is present at very low
concentrations, and thus recombination is slow. For solid-state DSSCs, where the
hole moves via electronic rather than ionic conduction, the charge-transport and
recombination rates are different from those of liquid-state DSSCs. In Regan’s solid-
state DSSC, which used CuSCN as the hole conductor, recombination is a one-
electron reduction of a hole at the top of the CuSCN valence band (VB) or possibly
trapped at the TiO2/CuSCN interface. The top of the CuSCN VB is likely to have a
significant contribution from Cu d orbitals. Copper oxidation/reduction is frequently
a fast process. The concentration of free and trapped holes in CuSCN could also
be higher than that of iodine in the electrolyte, where most of the iodine is bound
up in tri-iodide [117]. As a result, the faster recombination in solid-state DSSCs
results in a low fill factor and photocurrent loss, which are both important limiting
factors of cell efficiency. In the further work of O’Regan et al. [118], the surface of
TiO2 was coated with Al2O3, and the fill factor and efficiency were improved.
Transient photovoltage and photocurrent measurements were again utilized to
find the pseudo-first-order recombination rate constant and capacitance as a
function of potential.
Another aspect concerning recombination is the open-circuit voltage (Voc), which
is consistent with the elucidation of Kron et al. [116]. The Voc now is below
the theoretical upper limit, which is the difference between the energy position
of the redox-couple (for liquid-state and quasi solid-state DSSCs) or HOMO of
HTMs (for other solid-state DSSCs) and the conduction band edge of TiO2 [4,38].
ARTICLE IN PRESS
568 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

Recombination reduces the Voc by making it difficult for the quasi-Fermi level to rise
up to the conduction band edge of TiO2 [119]. Considering the fact that traps can
mediate electron leakage from nanocrystallites, a theoretical model was constructed
by Tennakone et al. [119] to quantitatively elucidate the effect of trapping on static
and transient behavior of solid-state DSSCs. It was concluded that the recombina-
tion of electrons injected into the semiconductor with acceptors depends on the
extent of the spread of the wave function of electrons trapped in shallow surface
states.
The charge transfer inside hole conducting media has also been widely
investigated. The movement of electrons on the inside of the electrolyte in liquid-
state DSCCs is performed by diffusion. However, the fundamental properties of the
I/I
3 redox couple in RTILs, including the equilibrium potentials and the charge
transport mechanism, had not been revealed at the time of these studies, although
these properties are crucial for the performance of the RTIL-based DSSCs in terms
of the open-circuit potential and the short-circuit current. In 2003, these properties
were elucidated by Kawano et al. [120] and Wang et al. [107], respectively by employ-
ing a microelectrode technique. The microelectrode technique utilized in electro-
chemical measurements can simplify determination of the equilibrium potential and
the transport property of the redox couple from steady-state voltammetry. It was
found that the charge transport is prone to an exchange reaction of I+I 3 -I3 +I
 
 
especially when the molar ratio of [I ] to [I3 ] is comparable although it fits
the physical diffusion mechanism well when the concentration of [I]+[I 3 ] is
low (0.01 M) [120]. This anomalous transport mechanism can be explained in
terms of a Grotthus-like exchange mechanism and described by the Dahms-Ruff
equation [107].

5. Conclusion

In this review, the basic principle of solid-state dye sensitized solar cells (DSSCs)
was given and the different types of solid or quasi solid-state hole conductors such as
p-type semiconductors, ionic liquid electrolytes and polymer electrolytes were
discussed. Among all of the solid-state cells, the one containing a p-type
semiconductor possesses the advantage of easy preparation and higher stability
while the cells employing polymer electrolytes show higher efficiency and wider
practical future use with the proper encapsulation. In particular, the mechanism
underlying solid-state cells was also introduced, showing a difference between liquid-
state and solid-state DSSCs: the charge transfer mechanisms at the dye/hole
conductor interface and the ionic transportation inside the hole conductor matrix. A
better understanding of these mechanisms provides a better understanding to apply
to research of solid-state DSSCs.
The solid-state dye-sensitized nanocrystalline electrochemical photovoltaic system
is at present considered as the only validated competitor to solid-state junction
devices for the conversion of solar energy into electricity. However, the module
efficiency of such DSSCs is too low to meet the requirements for practical
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 569

application need and the cost is too high. The further development of these new
types of photovoltaic solar cells is based on the optimization of several factors, which
include:
(1) The preparation of nanostructured, high surface area, low-defect, transparent
TiO2 film electrodes. This will be helpful because more sensitizers can be
absorbed, resulting in a higher current and reduced surface trapping to improve
voltage.
(2) The synthesis of light capturing dyes, presenting broad and strong MLCT
absorption bands and bearing functional anchoring groups, and low-cost organic
dye. Efforts should be made to broaden the absorption spectrum, especially in
the near-IR region. Fast charge transfer and carrier dissociation abilities are also
important for such dyes. Recently, an amphiphilic dye [75] was found to satisfy
this requirments.
(3) The use of the I/I
3 redox couple in an appropriate medium, such as RTILs for
quasi solid-state apparatuses. It is of great importance to further improve the
ionic conductivity of this nonvolatile electrolyte.
(4) The utilization of easy pore-filling, high ionic conductivity solid-state hole
conductors for all solid-state DSSCs. A proper soaking or spin-coating process
for the deposition of these hole conductors will facilitate the contact of dye-
sensitized TiO2 and the hole conductors and improve the current.
Several new approaches towards high efficiency and low cost have been
investigated. One of these approaches in optimization of TiO2 film employs meso-
macroporous TiO2 film electrodes [121,122], inside which the mesopores provide a
large surface area that is favorable for efficient light absorption, and the macropores
serve as pathways for diffusion of electroactive agents. A conductive transparent
underlayer of TiO2 porous film consisting of three alternative layers (TiO2/Ag/TiO2)
has also been utilized to suppress reflection losses, thus improving the efficiency
[123]. On the development of light absorbing dye, besides the newly synthesized dyes
[124,125], employing dye multilayers or so-called dye co-sensitization seems to be a
promising method to improve efficiency. This kind of co-sensitization was performed
by introducing a thin layer of CuSCN [39,126] or a second layer of Al2O3 [127] on
the surface of dye-sensitized TiO2 film.
Recently, a 200 m2 DSSC roof has been constructed by the STI Company [128]
and the production of soft DSSCs has been in process in the Konarka Company. It is
probable that in the near future, DSSC panels that have outstanding architectural
appeal will be in wide use in the commercial market. What is more attractive is that
low cost flexible products are expected to become available within a few years.

References

[1] M. Grätzel, Nature 414 (2001) 338.


[2] D.M. Chapin, C.S. Fuller, G.L. Pearson, J. Appl. Phys. 25 (1954) 676.
[3] C.D. Grant, A.M. Schwartzberg, G.P. Smestad, J. Kowalik, L.M. Tolbert, J.Z. Zhang,
J. Electroanal. Chem. 522 (2002) 40.
ARTICLE IN PRESS
570 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

[4] B. Oregan, M. Grätzel, Nature 353 (1991) 737.


[5] M.K. Nazeeruddin, A. Kay, I. Rodicio, R. Humphrybaker, E. Muller, P. Liska, N. Vlachopoulos,
M. Grätzel, J. Am. Chem. Soc. 115 (1993) 6382.
[6] W.C. Sinke, M.M. Wienk, Nature 395 (1998) 544.
[7] K. Tennakone, V.P.S. Perera, I.R.M. Kottegoda, Grra. Kumara, J. Phys. D-Appl. Phys. 32 (1999)
374.
[8] M. Matsumoto, Y. Wada, T. Kitamura, K. Shigaki, T. Inoue, M. Ikeda, S. Yanagida, Bull. Chem.
Soc. Japan 74 (2001) 387.
[9] F. Cao, G. Oskam, P.C. Searson, J. Phys. Chem. 99 (1995) 17071.
[10] K. Tennakone, Grra. Kumara, A.R. Kumarasinghe, K.G.U. Wijayantha, P.M. Sirimanne,
Semicond. Sci. Technol. 10 (1995) 1689.
[11] J. Hagen, W. Schaffrath, P. Otschik, R. Fink, A. Bacher, H.W. Schmidt, D. Haarer, Synth. Met. 89
(1997) 215.
[12] M. Masamitsu, M. Hiromitsu, K. Yoshimasa, Nippon Kagaku Kaishi 7 (1997) 484.
[13] K. Murakoshi, R. Kogure, Y. Wada, S. Yanagida, Chem. Lett. (1997) 471.
[14] U. Bach, D. Lupo, P. Comte, J.E. Moser, F. Weissortel, J. Salbeck, H. Spreitzer, M. Grätzel, Nature
395 (1998) 583.
[15] W. Kubo, K. Murakoshi, T. Kitamura, Y. Wada, K. Hanabusa, H. Shirai, S. Yanagida, Chem. Lett.
(1998) 1241.
[16] K. Tennakone, Grra. Kumara, I.R.M. Kottegoda, V.P.S. Perera, Psrs. Weerasundara,
J. Photochem. .Photobiol: A Chem. 117 (1998) 137.
[17] A.F. Nogueira, N. Alonso-Vante, M.A. De Paoli, Synth. Met. 105 (1999) 23.
[18] A.F. Nogueira, M.A. De Paoli, Sol. Energy Mater. Sol. Cells 61 (2000) 135.
[19] B. O’Regan, D.T. Schwartz, S.M. Zakeeruddin, M. Grätzel, Adv. Mater. 12 (2000) 1263.
[20] D. Gebeyehu, C.J. Brabec, F. Padinger, T. Fromherz, S. Spiekermann, N. Vlachopoulos,
F. Kienberger, H. Schindler, N.S. Sariciftci, Synth. Met. 121 (2001) 1549.
[21] D. Gebeyehu, C.J. Brabec, N.S. Sariciftci, D. Vangeneugden, R. Kiebooms, D. Vanderzande,
F. Kienberger, H. Schindler, Synth. Met. 125 (2002) 279.
[22] J. Kruger, R. Plass, L. Cevey, M. Piccirelli, M. Grätzel, U. Bach, Appl. Phys. Lett. 79 (2001) 2085.
[23] W. Kubo, K. Murakoshi, T. Kitamura, S. Yoshida, M. Haruki, K. Hanabusa, H. Shirai, Y. Wada,
S. Yanagida, J. Phys. Chem. B 105 (2001) 12809.
[24] Grra. Kumara, A. Konno, G.K.R. Senadeera, P.V.V. Jayaweera, Dbra. De Silva, K. Tennakone,
Sol. Energy Mater. Sol. Cells 69 (2001) 195.
[25] A.F. Nogueira, J.R. Durrant, M.A. De Paoli, Adv. Mater. 13 (2001) 826.
[26] Y. Ren, Z. Zhang, E. Gao, S. Fang, S. Cai, J. .Appl. Electrochem. 31 (2001) 445.
[27] E. Stathatos, P. Lianos, C. Krontiras, J. Phys. Chem. B 105 (2001) 3486.
[28] D. Gebeyehu, C.J. Brabec, N.S. Sariciftci, Thin Solid Films 403 (2002) 271.
[29] J. Kruger, R. Plass, M. Grätzel, H.J. Matthieu, Appl. Phys. Lett. 81 (2002) 367.
[30] G.R.A. Kumara, A. Konno, K. Shiratsuchi, J. Tsukahara, K. Tennakone, Chem. Mater. 14 (2002)
954.
[31] C. Longo, A.F. Nogueira, M.A. De Paoli, H. Cachet, J. Phys. Chem. B 106 (2002) 5925.
[32] B. O’Regan, F. Lenzmann, R. Muis, J. Wienke, Chem. Mater. 14 (2002) 5023.
[33] Y.J. Ren, Z.C. Zhang, S.B. Fang, M.Z. Yang, S.M. Cai, Sol. Energy Mater. Sol. Cells 71 (2002) 253.
[34] Y. Saito, T. Kitamura, Y. Wada, S. Yanagida, Chem. Lett. (2002) 1060.
[35] G.K.R. Senadeera, P.V.V. Jayaweera, V.P.S. Perera, K. Tennakone, Sol. Energy Mater. Sol. Cells
73 (2002) 103.
[36] S. Murai, S. Mikoshiba, H. Sumino, T. Kato, S. Hayase, Chem. Commun. 11 (2003) 1534.
[37] H. Paulsson, A. Hagfeldt, L. Kloo, J. Phys. Chem. B 107 (2003) 13665.
[38] V.P.S. Perera, K. Tennakone, Sol. Energy Mater. Sol. Cells 79 (2003) 249.
[39] V.P.S. Perera, Pkddp. Pitigala, P.V.V. Jayaweera, K.M.P. Bandaranayake, K. Tennakone, J. Phys.
Chem. B 107 (2003) 13758.
[40] Y. Shibata, T. Kato, T. Kado, R. Shiratuchi, W. Takashima, K. Kaneto, S. Hayase, Chem.
Commun. (2003) 2730.
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 571

[41] E. Stathatos, R. Lianos, S.M. Zakeeruddin, P. Liska, M. Grätzel, Chem. Mat. 15 (2003) 1825.
[42] T. Taguchi, X.T. Zhang, I. Sutanto, K. Tokuhiro, T.N. Rao, H. Watanabe, T. Nakamori,
M. Uragami, A. Fujishima, Chem. Commun. 19 (2003) 2480.
[43] H. Tokuhisa, P.T. Hammond, Adv. Funct. Mater. 13 (2003) 831.
[44] P. Wang, S.M. Zakeeruddin, J.E. Moser, M.K. Nazeeruddin, T. Sekiguchi, M. Grätzel, Nat. Mater.
2 (2003) 402.
[45] B. Li, L.D. Wang, J.W. Li, Y. Qiu, Chin. Phys. Lett. 21 (2004) 1391.
[46] E. Stathatos, P. Lianos, A.S. Vuk, B. Orel, Adv. Funct. Mater. 14 (2004) 45.
[47] G.P. Smestad, S. Spiekermann, J. Kowalik, C.D. Grant, A.M. Schwartzberg, J. Zhang,
L.M. Tolbert, E. Moons, Sol. Energy Mater. Sol. Cells 76 (2003) 85.
[48] M. Grätzel, J. Photochem. Photobiol. A-Chem. 164 (2004) 3.
[49] B. O’Regan, D.T. Schwartz, Chem. Mater. 10 (1998) 1501.
[50] K. Tennakone, G.K.R. Senadeera, Dbra. De Silva, I.R.M. Kottegoda, Appl. Phys. Lett. 77 (2000)
2367.
[51] K. Tennakone, Grra. Kumara, I.R.M. Kottegoda, K.G.U. Wijayantha, V.P.S. Perera, J. Phys.
D-Appl. Phys. 31 (1998) 1492.
[52] P.M. Sirimanne, T. Jeranko, P. Bogdanoff, S. Fiechter, H. Tributsch, Semicond. Sci. Technol. 18
(2003) 708.
[53] G.R.A. Kumara, S. Kaneko, M. Okuya, K. Tennakone, Langmuir 18 (2002) 10493.
[54] G.R.A. Kumara, M. Okuya, K. Murakami, S. Kaneko, V.V. Jayaweera, K. Tennakone,
J. Photochem. Photobiol. A-Chem. 164 (2004) 183.
[55] Q.B. Meng, K. Takahashi, X.T. Zhang, I. Sutanto, T.N. Rao, O. Sato, A. Fujishima, H. Watanabe,
T. Nakamori, M. Uragami, Langmuir 19 (2003) 3572.
[56] B. Oregan, D.T. Schwartz, Chem. Mater. 7 (1995) 1349.
[57] L. Bandara, H. Weerasinghe, Sol. Energy Mater. Sol. Cells 85 (2005) 385.
[58] L. Sicot, C. Fiorini, A. Lorin, J.M. Nunzi, P. Raimond, C. Sentein, Synth. Met. 102 (1999) 991.
[59] M. Catellani, B. Boselli, S. Luzzati, C. Tripodi, Thin Solid Films 403 (2002) 66.
[60] J.H. Schon, C. Kloc, E. Bucher, B. Batiogg, Nature 403 (2000) 408.
[61] S.E. Shaheen, C.J. Brabec, N.S. Sariciftci, F. Padinger, T. Fromherz, J.C. Hummelen, Appl. Phys.
Lett. 78 (2001) 841.
[62] H. Sirringhaus, P.J. Brown, R.H. Friend, M.M. Nielsen, K. Bechgaard, B.M.W. Langeveld-Voss,
A.J.H. Spiering, R.A.J. Janssen, E.W. Meijer, P. Herwig, D.M. de Leeuw, Nature 401 (1999)
685.
[63] Y.Y. Lin, D.J. Gundlach, S.F. Nelson, T.N. Jackson, IEEE Electron Device Lett. 18 (1997) 606.
[64] H.E. Katz, J.G. Laquindanum, A.J. Lovinger, Chem. Mat. 10 (1998) 633.
[65] C.W. Tang, S.A. VanSlyke, Appl. Phys. Lett. 51 (1987) 913.
[66] S.A. VanSlyke, C.H. Chen, C.W. Tang, Appl. Phys. Lett. 69 (1996) 2160.
[67] L.S. Hung, C.W. Tang, M.G. Mason, Appl. Phys. Lett. 70 (1997) 152.
[68] T. Park, S.A. Haque, R.J. Potter, A.B. Holmes, J.R. Durrant, Chem. Commun. 11 (2003) 2878.
[69] S.A. Haque, T. Park, C.G. Xu, S. Koops, N. Schulte, R.J. Potter, A.B. Holmes, J.R. Durrant, Adv.
Funct. Mater. 14 (2004) 435.
[70] K. Murakoshi, R. Kogure, Y. Wada, S. Yanagida, Sol. Energy Mater. Sol. Cells 55 (1998) 113.
[71] T. Kitamura, M. Maitani, M. Matsuda, Y. Wada, S. Yanagida, Chem. Lett. (2001) 1054.
[72] Y. Saito, T. Kitamura, Y. Wada, S. Yanagida, Synth. Met. 131 (2002) 185.
[73] Y. Saito, N. Fukuri, R. Senadeera, T. Kitamura, Y. Wada, S. Yanagida, Electrochem. Commun. 6
(2004) 71.
[74] H. Matsumoto, T. Matsuda, T. Tsuda, R. Hagiwara, Y. Ito, Y. Miyazaki, Chem. Lett. (2001) 26.
[75] P. Wang, S.M. Zakeeruddin, J.E. Moser, M. Grätzel, J. Phys. Chem. B 107 (2003) 13280.
[76] N. Papageorgiou, Y. Athanassov, M. Armand, P. Bonhote, H. Pettersson, A. Azam, M. Grätzel,
J. Electrochem. Soc. 143 (1996) 3099.
[77] R. Hagiwara, T. Hirashige, T. Tsuda, Y. Ito, J. Fluorine Chem. 99 (1999) 1.
[78] N. Yamanaka, R. Kawano, W. Kubo, T. Kitamura, Y. Wada, M. Watanabe, S. Yanagida, Chem.
Commun. (2005) 740.
ARTICLE IN PRESS
572 B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573

[79] H. Matsui, K. Okadaa, T. Kawashima, T. Ezure, N. Tanabe, R. Kawano, M. Watanabe,


J. Photochem. Photobiol. A-Chem. 164 (2004) 129.
[80] A. Noda, M. Watanabe, Electrochim. Acta 45 (2000) 1265.
[81] R. Kawano, H. Matsui, C. Matsuyama, A. Sato, M.A.B.H. Susan, N. Tanabe, M. Watanabe,
J. Photochem. Photobiol. A-Chem. 164 (2004) 87.
[82] H. Paulsson, M. Berggrund, E. Svantesson, A. Hagfeldt, L. Kloo, Sol. Energy Mater. Sol. Cells 82
(2004) 345.
[83] P. Wang, S.M. Zakeeruddin, J. Moser, R. Humphry-Baker, M. Grätzel, J. Am. Chem. Soc. 126
(2004) 7164.
[84] G. Oskam, B.V. Bergeron, G.J. Meyer, P.C. Searson, J. Phys. Chem. B 105 (2001) 6867.
[85] H. Shirakawa, E.J. Louis, A.G. MacDiarmid, C.K. Chiang, A.J. Heeger, J. Chem. Soc.: Chem.
Commun. 16 (1977) 578.
[86] F. Croce, G.B. Appetecchi, L. Persi, B. Scrosati, Nature 394 (1998) 456.
[87] B. Scrosati, Polym. Int. 47 (1998) 50.
[88] F. Croce, L. Persi, F. Ronci, B. Scrosati, Solid State Ion 135 (2000) 47.
[89] A. Clemente, S. Panero, E. Spila, B. Scrosati, Solid State Ion 85 (1996) 273.
[90] S. Panero, B. Scrosati, M. Baret, B. Cecchini, E. Masetti, Sol. Energy Mater. Sol. Cells 39 (1995)
239.
[91] S. Murai, S. Mikoshiba, H. Sumino, S. Hayase, J. Photochem. Photobiol. A-Chem. 148 (2002) 33.
[92] Z.C. Zhang, S. Fang, J. Appl. Polym. Sci. 77 (2000) 2957.
[93] R. Komiya, L. Han, R. Yamanaka, A. Islam, T. Mitate, J. Photochem. Photobiol. A-Chem. 164
(2004) 123.
[94] M.S. Kang, J.H. Kim, Y.J. Kim, J. Won, N.G. Park, Y.S. Kang, Chem. Commun. (2005) 889.
[95] J. Kang, W. Li, X. Wang, Y. Lin, X. Xiao, S. Fang, Electrochim. Acta 48 (2003) 2487.
[96] S.A. Haque, E. Palomares, H.M. Upadhyaya, L. Otley, R.J. Potter, A.B. Holmes, J.R. Durrant,
Chem. Commun. (2003) 3008.
[97] Masao. Kaneko, Takayuki. Hoshi, Chem. Lett. 32 (2003) 872.
[98] C. Longo, J. Freitas, M.A. De Paoli, J. Photochem. Photobiol. A-Chem. 159 (2003) 33.
[99] B. Scrosati, F. Croce, L. Persi, J. Electrochem. Soc. 147 (2000) 1718.
[100] G. Katsaros, T. Stergiopoulos, I.M. Arabatzis, K.G. Papadokostaki, P. Falaras, J. Photochem.
Photobiol. A-Chem. 149 (2002) 191.
[101] T. Stergiopoulos, I.M. Arabatzis, H. Cachet, P. Falaras, J. Photochem. Photobiol. A-Chem. 155
(2003) 163.
[102] E. Stathatos, P. Lianos, U. Lavrencic-Stangar, B. Orel, Adv. Mater. 14 (2002) 354.
[103] W. Lu, A.G. Fadeev, B.H. Qi, E. Smela, B.R. Mattes, J. Ding, G.M. Spinks, J. Mazurkiewicz,
D.Z. Zhou, G.G. Wallace, D.R. MacFarlane, S.A. Forsyth, M. Forsyth, Science 297 (2002) 983.
[104] S. Sakaguchi, H. Ueki, T. Kato, T. Kado, R. Shiratuchi, W. Takashima, K. Kaneto, S. Hayase,
J. Photochem. Photobiol. A-Chem. 164 (2004) 117.
[105] K. Suzuki, M. Yamaguchi, S. Hotta, N. Tanabe, S. Yanagida, J. Photochem. Photobiol. A-Chem.
164 (2004) 81.
[106] W. Kubo, Y. Makimoto, T. Kitamura, Y. Wada, S. Yanagida, Chem. Lett. (2002) 948.
[107] P. Wang, S.M. Zakeeruddin, P. Comte, I. Exnar, M. Grätzel, J. Am. Chem. Soc. 125 (2003) 1166.
[108] T. Fukushima, A. Kosaka, Y. Ishimura, T. Yamamoto, T. Takigawa, N. Ishii, T. Aida, Science 300
(2003) 2072.
[109] H. Usui, H. Matsui, N. Tanabe, S. Yanagida, J. Photochem. Photobiol. A-Chem. 164 (2004) 97.
[110] U. Bach, Y. Tachibana, J.E. Moser, S.A. Haque, J.R. Durrant, M. Grätzel, D.R. Klug, J. Am.
Chem. Soc. 121 (1999) 7445.
[111] B. Mahrov, G. Boschloo, A. Hagfeldt, L. Dloczik, T. Dittrich, Appl. Phys. Lett. 84 (2004) 5455.
[112] V. Duzhko, V.Y. Timoshenko, F. Koch, T. Dittrich, Phys. Rev. B 64 (2001) 075204.
[113] S.A. Haque, T. Park, A.B. Holmes, J.R. Durrant, Chem. Phys. Chem. 4 (2003) 89.
[114] S. Tanaka, Jpn. J. Appl. Phys. Part 1 40 (2001) 97.
[115] J. Ferber, R. Stangl, J. Luther, Sol. Energy Mater. Sol. Cells 53 (1998) 29.
[116] G. Kron, T. Egerter, G. Nelles, A. Yasuda, J.H. Werner, U. Rau, Thin Solid Films 403 (2002) 242.
ARTICLE IN PRESS
B. Li et al. / Solar Energy Materials & Solar Cells 90 (2006) 549–573 573

[117] B. O’Regan, F. Lenzmann, J. Phys. Chem. B 108 (2004) 4342.


[118] B.C. O’Regan, S. Scully, A.C. Mayer, E. Palomares, J. Durrant, J. Phys. Chem. B 109 (2005) 4616.
[119] K. Tennakone, P.V.V. Jayaweera, P.K.M. Bandaranayake, J. Photochem. Photobiol. A-Chem. 158
(2003) 125.
[120] R. Kawano, M. Watanabe, Chem. Commun. (2003) 330.
[121] S. Ito, K. Ishikawa, C.J. Wen, S. Yoshida, T. Watanabe, Bull. Chem. Soc. Japan 73 (2000) 2609.
[122] S. Ito, S. Yoshida, T. Watanabe, Bull. Chem. Soc. Japan 73 (2000) 1933.
[123] S. Ito, T. Takeuchi, T. Katayama, M. Sugiyama, M. Matsuda, T. Kitamura, Y. Wada, S. Yanagida,
Chem. Mater 15 (2003) 2824.
[124] G. Boschloo, J. Lindstrom, E. Magnusson, A. Holmberg, A. Hagfeldt, J. Photochem. Photobiol.
A-Chem. 148 (2002) 11.
[125] T. Stergiopoulos, S. Karakostas, P. Falaras, J. Photochem. Photobiol. A-Chem. 163 (2004) 331.
[126] V.P.S. Perera, P.K.D.D.P. Pitigala, M.K.I. Senevirathne, K. Tennakone, Sol. Energy Mater. Sol.
Cells 85 (2005) 91.
[127] J.N. Clifford, E. Palomares, M.K. Nazeeruddin, R. Thampi, M. Grätzel, J.R. Durrant, J. Am.
Chem. Soc. 126 (2004) 5670.
[128] R.F. Service, Science 300 (2003) 1219.
[129] C. Jager, R. Bilke, M. Heim, D. Haarer, H. Karickal, M. Thelakkat, Synth. Met. 121 (2001) 1543.
[130] K.R. Haridas, J. Ostrauskaite, M. Thelakkat, M. Heim, R. Bilke, D. Haarer, Synth. Met. 121 (2001)
1573.
[131] S. Spiekermann, G. Smestad, J. Kowalik, L.M. Tolbert, M. Grätzel, Synth. Met. 121 (2001) 1603.
[132] G.K.R. Senadeera, T. Kitamura, Y. Wada, S. Yanagida, J. Photochem. Photobiol. A-Chem. 164
(2004) 61.
[133] S.X. Tan, J. Zhai, M.X. Wan, Q.B. Meng, Y.L. Li, L. Jiang, D.B. Zhu, J. Phys. Chem. B 108 (2004)
18693.

You might also like