Está en la página 1de 35

10/20/2010 TiO2 nanotubes and their application in…

Nanoscale
Home Publishing Journals Nanoscale Contents List DOI: 10.1039/b9nr00131j
Nanoscale, 2010, 2, 45-59 DOI: 10.1039/b9nr00131j Feature Article

TiO2 nanotubes and their application in dye-sensitized solar cells


Poulomi Roy a, Doohun Kim a , Kiyoung Lee a , Erdmann Spiecker b and Patrik Schmuki *a
a Department of Materials Science, WW4-LKO, University of Erlangen–Nürnberg, Martensstrasse 7,
91058 Erlangen, Germany. E-mail: schmuki@ww.uni-erlangen.de; Fax: +49-9131-852-7582; Tel:
+49-9131-852-7575
b Department Werkstoffwissenschaften, Lehrstuhl Mikrocharakterisierung/VII, University of Erlangen–
Nürnberg, Cauerstraße 6, 91058 Erlangen, Germany

Received (in Zürich, Switzerland) 16th June 2009 , Accepted 23rd October 2009

First published on the web 4th December 2009

The present article reviews the current status of using TiO2 nanotubes in Grätzel-type, dye-sensitized solar
cells and extends the overview with the latest results and findings. Critical factors in tube geometry (length,
diameter, top morphology), crystal structure (amorphous, anatase, rutile) as well as factors affecting dye
loading or electron mobility are addressed. The highest solar cell efficiencies today for pure nanotube systems
reach approximately 4% while for some mixed systems, around 7% has been reported. For both systems
significant room for enhancement is anticipated and some key points and strategies for improvement are
outlined.

Poulomi Roy received her MSc degree in Inorganic Chemistry from


Vidyasagar University, India and obtained her PhD degree in Inorganic
Nanomaterials from the Indian Institute of Technology, Kharagpur, in
2007. She joined the group of Prof. Patrik Schmuki at the University of
Erlangen–Nürnberg, Germany in 2008 as a post-doctoral fellow, where
currently she is continuing her research. Her research interests include
the synthesis of semiconductor nanomaterials and their application in
various energy fields, especially in solar cells.

Poulomi Roy

rsc.org/…/DisplayHTMLArticleforfree.cf… 1/35
10/20/2010 TiO2 nanotubes and their application in…
Doohun Kim received his MSc degree from Inha University, Inchon,
South Korea in 1999. He then worked at Solartech Corp., Seoul for four
years. Currently, he is persuing his PhD degree in Prof. Patrik Schmuki
s lab at the University of Erlangen–Nürnberg, Germany. His research
interest includes synthesis of self-organized nanomaterials by
electrochemical means and their applications in different energy
devices.

Doohun Kim

Kiyoung Lee obtained his MSc degree in Chemical Engineering from


Inha University, Korea in 2009. He is currently a PhD student under the
supervision of Prof. Schmuki at the University of Erlangen–Nürnberg
in Germany. His research interests include functional nanomaterial
synthesis by electrochemical methods with the direction of application
in energy conversion and storage devices.

Kiyoung Lee

Erdmann Spiecker received his PhD in physics in 1999 from the


University of Göttingen. In 2000 he joined the Materials Science
Department of the University of Kiel as research associate. He was
awarded a Humboldt Fellowship in 2005 and spent over a year at the
National Center for Electron Microscopy in Berkeley. In 2007 he
received his habilitation from the University of Kiel. Since September
2008 he has been Professor of Materials Science and head of the
Electron Microscopy group at the University of Erlangen–Nürnberg.
His research focuses on electron microscopy techniques for
investigation of nanostructured materials.

Erdmann Spiecker

Patrik Schmuki studied physical chemistry at the University of Basel,


and obtained his PhD degree from ETH-Zurich in 1992. From 1994–97
he worked at Brookhaven National Laboratory, USA and the Institute
rsc.org/…/DisplayHTMLArticleforfree.cf… 2/35
10/20/2010 TiO2 nanotubes and their application in…
for Microstructural Sciences of the NRC, Canada. From 1997–2000 he
was an Associate Professor for Microstructuring Materials at EPFL,
Switzerland, and since 2000 he is Full Professor and head of the
Institute for Surface Science at the Materials Science Department of the
University of Erlangen–Nürnberg. His current research interests cover
electrochemistry and materials at the nanoscale—with particular focus
on functional materials and the control of self-assembly.

Patrik Schmuki

1. Introduction

1.1 General overview

Currently the majority of commercial solar cells are made of different solid-state junction devices, usually from
silicon. These cells reach light-to-electricity conversion efficiencies of 20% for multicrystalline and 25%
for monocrystalline silicon at reasonable cost, and thus represent the milestone of terrestrial applications.1
However, several other concepts are explored that are geared to achieve either higher efficiencies, or
manufacture at more attractive costs, or are expected to be overall more environmentally friendly. A
particularly striking photoelectrochemical concept is based on dye-sensitized TiO2 – it involves neither
expensive semiconductor substrates nor high-level processing steps.
In 1985, Desilvestro et al.2 reported TiO2 nanoparticles onto which suitable organic molecules (Ru-based
complexes) were adsorbed. Such electrodes were active under visible-light irradiation and allowed solar
energy to be converted to electricity with a quantum efficiency of 80%.3 Later on, Grätzel and O'Regan
extended the concept to a full solar cell,4–6 which is nowadays known as a dye-sensitized solar cell (DSSC).
In these DSSCs the hole-transmitting medium is a redox species in an electrolyte, liquid or gel (quasi solid-
state). The principle of such a DSSC is described in Fig. 1. The classic DSSC is composed of a layer of
nanocrystalline TiO2 particles on a conducting substrate, a platinum counter electrode, and an adsorbed Ru-
dye as a sensitizer and an electrolyte. Using such constructions 11% conversion efficiency has been achieved.7
Usually, the TiO2 nanoparticles are deposited on a conducting glass substrate, and sintered to convert the
particles to an anatase crystal phase and to improve the electronic conduction between the conductive glass
and the TiO2 nanoparticles. On the TiO2 surface adsorbed is a dye, which serves as a light absorber
(sensitizer), and is anchored (attached) to the surface of the TiO2 nanoparticles by specific functional groups.
A key requirement for the dye is that the LUMO of the dye molecule is energetically positioned slightly higher
than the conduction band of TiO2 . Under irradiation with sunlight, HOMO–LUMO transitions in the dye
occur. Excited electrons can then be injected from the LUMO into the conduction band of the semiconductor
electrode (Fig. 1B). For typically used Ru-ligand dyes the electron injection from the dye sensitizer to the
conduction band of TiO2 occurs through a metal-to-ligand charge transfer (MLCT) pathway as shown in Fig.
1C.8 The oxidized dye is reduced to its original state by electron donation from the redox couple in the
electrolyte. Usually the electrolyte is an organic solvent or ionic liquid containing an I3 −/I− redox system. After
the I3 − ion is oxidized to I−, it regenerates the dye (refills the HOMO levels). The resulting I− species can
diffuse to the Pt counter electrode where they are regenerated to I3 − by reduction.

rsc.org/…/DisplayHTMLArticleforfree.cf… 3/35
10/20/2010 TiO2 nanotubes and their application in…

Fig. 1 (A) Schematic representation of the principle of a dye-


sensitized solar cell. (B) Schematic diagram of the light-induced
electron excitation (HOMO–LUMO transition) and injection
process from the LUMO of the dye to the conduction band of
TiO2 . (C) Excited electron transfer from Ru-metal to the TiO2
rsc.org/…/DisplayHTMLArticleforfree.cf… 4/35
10/20/2010 TiO2 nanotubes and their application in…
conduction band occurs by the so-called metal-to-ligand
charge transfer pathway (MLCT).

The potential difference under illumination corresponds to the difference between the Fermi level of the
electron in the semiconductor electrode and the redox potential of the electrolyte (i.e., this represents the
achievable Voc).
However, not only energetic (thermodynamic) considerations are crucial but also the reaction kinetics. The
rate constant for electron injection from the conduction band has to be faster than the de-excitation of the dye
and the regeneration time constant for the dye has to be sufficiently fast to avoid depletion effects8,9 (Fig. 2A).
Generally, the rate-limiting factor is the dynamic competition between the electron transport in the TiO2 and
interfacial recombination of electrons. Typically the electron transport time constants in TiO2 nanoparticles are
comparably slow (due to a trapping/de-trapping type of electron transport).10,11 Low electron diffusion
coefficients within the TiO2 nanoparticles can be due to defects, surface states, grain boundaries etc. that act
as electron trapping sites, thus slowing down electron flow, enhancing recombination and thereby diminishing
the electron collection at the back contact.12,13 In order to considerably improve this behaviour, recently
research has been prompted towards the use of one-dimensional nanostructures of TiO2 , such as,
nanotubes,14–19 nanorods,20–23 nanowires24,25 etc. These 1D nanostructures are expected to significantly
improve the electron transport properties due to a directional smooth electron mobility (limiting random walk
in the wide nanocrystalline network) as well as decreased inter-crystalline contacts, thereby accelerating
electron transport and lowering the recombination probability as shown in the schematic diagram of Fig. 2B–
C.14

rsc.org/…/DisplayHTMLArticleforfree.cf… 5/35
10/20/2010 TiO2 nanotubes and their application in…

Fig. 2 (A) The most relevant kinetic constants involved in the


dynamics of the redox processes involving electron injection
rsc.org/…/DisplayHTMLArticleforfree.cf… 6/35
10/20/2010 TiO2 nanotubes and their application in…
from the dye into TiO2 , dye regeneration, electron
transportation and back-transfer of the electrons to the redox
medium. The electron diffusion length (Ln ) in TiO2 , one of the
key parameters in DSSCs, depends on the diffusion constant
(Dn ) and the electron lifetime ( c) (reproduced with permission
from ref. 8. Copyright American Chemical Society, 2005). (B)
Schematic representation of typical solar cell construction using
TiO2 nanotubes grown on a Ti substrate. (C) A comparison of
the electron pathways through nanoparticulate and nanotubular
structured TiO2 (3D random walk vs. 1D transport).

Various techniques for synthesizing TiO2 nanotubes and other 1D structures have been reported in recent
years, such as sol–gel processes, or hydrothermal approaches.26–30 Although comparably high-efficiency solar
cells have been produced with these chemically synthesized 1D morphologies of TiO2 , a multidirectional
orientation of TiO2 nanotubes or nanorods may not ensure the most efficient unidirectional electron flow along
the length of a restricted passage towards the back contact. In principle, as long as the nanotube or nanorod
systems are ideal i.e., they provide a highly conductive, rapid electron transport, their orientation on the
substrate is of minor importance. This, for example, may hold for single crystalline defect-free nanowires.
However, many approaches to grow TiO2 nanotubes lead to a polycrystalline material with substantial
recombination possibilities along the electron path. In this case the optimally increased rate of electron
transport and the higher charge collection efficiencies can only be expected for nanotubes vertically aligned
with the conductive substrate. Therefore approaches to produce self-ordered titania nanotubes by anodic
oxidation of Ti in adequate electrolyte31,32 have become of a significant interest – the present review will focus
on this point in later sections but will first give a brief overview of some general concepts in DSSCs.

1.2 Key factors for DSSCs

More than 2000 papers have been published on dye-sensitized solar cells with the goal of understanding and
optimizing key factors, such as, the TiO2 morphology and structure,33–36 and the dye and the electrolyte.37–41
It is beyond the scope of this article to give a detailed review on these topics. However, some of the most
important points with reference to further reading shall be briefly outlined. While in the beginning, advances in
DSSCs were almost entirely based on experimental findings, for some elementary steps a much higher degree
of understanding is achieved through modelling and detailed optical and electrical characterization.9–16,22 The
development of the dye and the electrolyte are still dominated by general guidelines and empirical approaches.
The dyes used and explored in DSSCs can be broadly characterized as (1) ruthenium complexes (mostly
polypyridyl), or (2) metal-free donor–acceptor systems.37,38,42–45 Some typical examples are shown in Fig. 3.
Both classes of dyes were recently reviewed, for Ru-dyes see ref. 38, for metal-free dyes see refs. 37 and 42.
In addition to Ru, other transition metals, such as, Os, Pt, Cu or Fe have been explored46–51 and although
some metal-free dyes can reach solar cell efficiencies of 9%,42 still the dominating structures are the Ru-
based dyes shown in Fig. 3A–D. For these dyes it is well established that charge transfer from the Ru-dye to
the TiO2 is a two-step process (MLCT).8 Ligands not only influence charge separation in the dye and thus can
control the kinetics of electron transfer from metal to ligand but also affect the electron injection kinetics to the
conduction band of TiO2 .8,38 Attachment (adsorption) of the dye to the TiO2 surface is usually achieved via
carboxylate groups. Although other coordination groups, such as, phosphates or silanes may even lead to a
stronger bond between TiO2 and the dye,38,52,53 they usually do not yield as high an efficiency. The reason
rsc.org/…/DisplayHTMLArticleforfree.cf… 7/35
10/20/2010 TiO2 nanotubes and their application in…
may be that carboxylate ligands attached to the bipyridine molecule lower the energy of the ligand * -orbital –
thus giving charge transfer the desired direction. Some room for improvement is believed to be achieved with
dyes with a significantly higher molar extinction coefficient.37,38 This would allow the use of thinner TiO2
particle layers (as full light absorption could be reached in a shorter path and thus the length for recombination
would be reduced too). In this regard, several metal-free organic dyes were found to be very useful in DSSCs
and for example, coumarin, squaraine, porphyrin and phthalocyanine dyes in DSSCs with solar cell efficiencies
of 6–8% have been reported.54–58 Due to the wide possibilities of synthetic chemistry it is also reasonable to
assume that the absorption spectra of these dyes may be even further tuned and particularly be better
extended into the infrared range. A problem in dye design and application is to avoid multilayer formation or
non-bound dye residues, simply as this material absorbs light but is not able to inject the carriers to TiO2 and
thus leads to recombination losses. Proper monolayer formation may be supported by co-adsorption for
example of deoxycholic acid or by optimized design of the ligands.37

rsc.org/…/DisplayHTMLArticleforfree.cf… 8/35
10/20/2010 TiO2 nanotubes and their application in…

Fig. 3 The most efficient Ru-complex and metal-free dyes with


their maximum achieved solar cell efficiencies. (Reproduced
with permission from ref. 37. Copyright Wiley-VCH, 2009).

Three different types of electrolytes are usually used in DSSCs:39–41,59 liquid electrolytes, quasi-solid-state
electrolytes, and solid electrolytes. Typically, various nitriles, such as, acetonitrile, valeronitrile, 3-
methoxypropionitrile, and esters such as ethylene carbonate, propylene carbonate, -butyrolactone etc. have
been used. The active redox couple is almost exclusively I3 −/I−. The achievable voltage in the solar cell Voc,
depends directly on the redox potential of the redox system. The redox potential should be as positive as
possible but still sufficiently negative to be able to regenerate the dye. With regard to the energy level of the
most common dyes, it can be assumed that with a suitable redox system Voc could be improved by 0.1 V.
Therefore, several researchers tried to change the redox couple. For example, Wang et al.60 investigated the
rsc.org/…/DisplayHTMLArticleforfree.cf… 9/35
10/20/2010 TiO2 nanotubes and their application in…
role of the redox couple Br−/Br2 as well as I3 −/I− on various dyes. And indeed a higher solar cell efficiency
was obtained by replacing the I3 −/I− redox couple with Br−/Br2 if eosin Y was used as a dye. The use of the
Co(II)/Co(III) redox couple as polypyridyl cobalt(II/III) complexes in the electrolyte of the solar cell in
presence of Ru-dye (Z316) showed promising activity and may become a rival to I3 −/I−.61,62 Pseudohalide
redox couples, such as, SCN−/(SCN)2 - and SeCN−/(SeCN)2 -containing electrolytes were also tested in
DSSCs.63,64 Grätzel and his co-workers used a solvent-free, SeCN−/(SeCN)3 -based ionic liquid electrolyte
and achieved 7.5% conversion efficiency under standard AM 1.5 conditions.64 In spite of all of these efforts,
to the date the highest efficiency is observed by using a classic Ru-dye (N719) and I3 −/I− as the redox couple.
A key consideration regarding the redox couple is the back reaction. Probably still the main efficiency loss
mechanism today is interface recombination, i.e., loss by the reaction of electrons already injected into the
TiO2 back to the redox couple in the electrolyte. The kinetics of the back reaction is, among others,
determined by interface states at the TiO2 /electrolyte interface. While certain orientations at TiO2 crystallites
may show a lower defect density, there may also be the possibility of blocking reactive interface states by
adequate chemical or physicochemical treatments. Additives to the electrolyte such as 4-tert-butyl pyridine
are reported to preferentially adsorb at defect sites, thus blocking, to a certain extent, these recombination
sites (reducing the dark current) and as a result a higher Voc is obtained.65
In comparison to conventional organic solvent electrolytes, ionic liquids have several advantages, but
mainly their negligible vapor pressure combined with their high ionic conductivity make them interesting as
alternative electrolytes for DSSCs. Recently, Grätzel achieved 7.6% conversion efficiency using 1-propyl-3-
methylimidazolium iodide and 1-ethyl-3-methylimidazolium tetracyanoborate as an ionic liquid electrolyte in a
TiO2 -based DSSC.66 Solid electrolytes are typically based on hole-conducting polymers. While difficulties in
complete filling and penetration of the nanoparticulate material with the hole-conducting polymer exist,
efficiencies in the range of 5% are reported – recent reviews on this topic are provided in refs. 67 and 68.
Very recently, Grätzel reported 8.2% solar cell efficiency using eutectic melts to produce solvent-free liquid
redox electrolytes.69 As mentioned before, a key factor affecting the overall efficiency of the solar cell is the
competition between charge carrier transport and interface recombination. Therefore, approaches that lead to
a significantly faster (or shorter) overall transport of electrons from the point of injection to the back contact
are highly promising. Therefore nanoscale TiO2 , not present as an extended 3D network, but as a highly
directional 1D nanostructure with potentially very fast carrier kinetics, has attracted considerable attention as
dye-support material.
As the role of the TiO2 in DSSCs is solely to act as a good electron conducting semiconductor, harvesting
the electrons injected from the dye, various other oxide materials may be considered for this role as an
alternative. Semiconducting oxides, having a wide bandgap, such as, ZnO, SnO2, Nb2 O5 70–75 have also been
explored for DSSCs. Particularly in the case of ZnO, 1D nanostructures also seem attractive and are being
explored.75 However, up to now TiO2 is still regarded as by far the best choice.
TiO2 nanorods, nanowires and nanotubes have been prepared by various approaches, such as, template-
assisted growth, sol–gel methods, hydrothermal methods and so on.26–30 The recent findings of self-ordered
growth of vertically aligned TiO2 nanotube arrays by a simple anodic oxidation reaction on titanium metal
attracted wide interest.31,32 It is believed that these nanotubes will be superior to nanorods or nanowires
because of the availability of inner as well as outer wall of tubes for dye adsorption.
In the following we will discuss the current state-of-the-art in dye-sensitized solar cells based on self-
organized TiO2 nanotubes and we will show some results from ongoing research on DSSCs. We will try to
discuss the critical factors of DSSC efficiency when using TiO2 nanotubes, such as, tube morphology,
dimensions and annealing. We will also demonstrate how different secondary treatments can affect DSSC
rsc.org/…/DisplayHTMLArticleforfree.cf… 10/35
10/20/2010 TiO2 nanotubes and their application in…
performance.

2. Anodic self-organized TiO2 nanotube arrays

2.1 Growth of layers

Self-organized TiO2 nanotube layers as shown in Fig. 4 are formed by anodic oxidation of a Ti-metal sheet
(illustrated in Fig. 4A). This simple electrochemical process can be carried out in an electrochemical cell with a
two- or three-electrode system: Ti-foil as the working electrode (the anode) and for example Pt as the counter
electrode, and using a standard voltage source or a high-voltage potentiostat. The key to forming self-ordered
nanotube layers on the metal substrate is that samples are anodized under a set of optimized conditions in
solutions containing fluoride ions. Reviews of this topic have recently been published.31,32,76 The process
originates from work by Zwilling and co-workers published in 1999, on the anodization of Ti and Ti alloys in
chromic acid–HF mixtures.77 These authors could grow the first generation of tubes to layers of a thickness of
few 100 nm. Later on, this process was used by Gong et al.78 and Beranek et al.79 to grow organized TiO2
nanotubular layers in aqueous HF electrolytes or H2 SO4 /HF electrolytes to establish better defined growth
conditions. However, although considerable order in wall thickness, diameter and short-range arrangement
was established, the tube length remained limited to 500 nm and the walls showed considerable disorder
(side-wall ripples and other etch effects). Two key steps in improving the geometry were introduced by
Macak et al.:80–82 (i) the use of neutral electrolytes80 that allowed, under suitable conditions, significantly
longer tubes to be grown,81 and (ii) the use of non-aqueous electrolytes that allowed to significantly improve
the tube wall roughness (smooth, high aspect ratio nanotubes).82 Using such an an anodization procedure in
non-aqueous electrolytes, not only very thick TiO2 nanotube layers (several 100 m) can be grown but they
also show a very smooth wall83 and can be grown to show an almost ideal hexagonal arrangement84 (Fig.
4C).

rsc.org/…/DisplayHTMLArticleforfree.cf… 11/35
10/20/2010 TiO2 nanotubes and their application in…

Fig. 4 (A) Schematic diagram of a electrochemical set-up and

rsc.org/…/DisplayHTMLArticleforfree.cf… 12/35
10/20/2010 TiO2 nanotubes and their application in…
anodic growth of TiO2 nanotubes on a Ti-metal sheet. (B) SEM
images of the self-organized nanotube TiO2 layer formed in
aqueous electrolyte (1 M Na2 SO4/0.14 M NaF electrolyte).
(C) SEM images of TiO2 nanotube layers formed by the anodic
oxidation of Ti-metal in non-aqueous electrolyte (cross section,
top view and bottom view) and also showing a very high-
aspect-ratio nanotube layer (several hundred m).

Nowadays, even more advanced geometries of TiO2 nanotubes can be prepared by altering the voltage
during growth.15,85,86 For example, by regular voltage cycling a bamboo-type morphology of TiO2
nanotubes can be grown in HF-containing ethylene glycol electrolyte15,85 (Fig. 5A). By using different AV
pulse durations the distance between the bamboo rings can be adjusted. Voltage stepping furthermore allows
the growth of branched tubes (tube multilayers) or some 2D TiO2 networks (nanolaces).85,87–89

rsc.org/…/DisplayHTMLArticleforfree.cf… 13/35
10/20/2010 TiO2 nanotubes and their application in…

rsc.org/…/DisplayHTMLArticleforfree.cf… 14/35
10/20/2010 TiO2 nanotubes and their application in…
Fig. 5 Example of advanced geometries of self-organized TiO2
nanotube layers: (A) A bamboo-type nanotube layer (inset:
half broken tube); (Reproduced with permission from ref. 85.
Copyright Wiley-VCH, 2008). (B) Double-walled TiO2
nanotube structure; (Reproduced with permission from ref. 90.
Copyright Wiley-VCH, 2008). (C) Nanograss formation at
TiO2 tube tops under normal anodization conditions (left) and
after pre-conditioning (Reproduced with permission from ref.
16. Copyright Elsevier, 2008). (D) 2 nanotubes grown by rapid
breakdown anodization; (Reproduced with permission from ref.
94. Copyright Elsevier, 2007).

Recently we reported that adequate heat treatment of some TiO2 nanotubes leads to the characteristic
double-wall structure shown in Fig. 5B.90 The smooth tubes are grown in ethylene glycol electrolytes and then,
by rapid thermal annealing, the tubes are separated in two separate entities. As shown in Fig. 5B, the inner
wall has a very high porosity providing a high surface area and the outer wall of the tubes is comparably
smooth. A common feature is observed, if self-organized tubes are grown for extended periods of time: grass
formation at the tube top occurs. Although several times attributed to the formation of nanorods, nanospikes,
or nanowires,91 this feature simply originates from the continuous thinning of the tube walls in the fluoride
electrolyte. These grassy tube tops may interfere with the defined electron transfer through nanotubes but in
any case the disorder at the tube top may hamper optimized solar cell performance. In order to avoid this
problem, various approaches have been developed to achieve ordered and open tube tops (Fig. 5C)15,18,92,93
and effects on the solar cell performance will be discussed in next section.
Another electrochemical approach to form high aspect nanotubes is so called rapid breakdown anodization
(RBA).94–96 Such tubes form very rapidly (within seconds), if Ti is anodized under breakdown conditions in
chloride or perchlorate containing electrolytes. This rapid process leads to formation of flower-like bundles of
nanotubes as shown in Fig. 5D, with very high aspect ratio of tubes having lengths of 3–50 m within some
minutes. These structures have also been reported to show a good performance in DSSC.96,97

2.2. Modification and characterization of TiO2 nanotubes

2.2.1 Annealing. As-prepared TiO2 nanotubes are amorphous in nature and by thermal treatment they can
be converted to anatase or rutile, or mixtures of anatase and rutile phases depending on the annealing
temperature as shown in Fig. 6.31,32,90 The crystallinity of TiO2 has a significant effect on solar cell
performance. If amorphous TiO2 (nanocrystallites or tubes) is annealed at 300–450 °C in air, the anatase
phase of TiO2 can be obtained. At 500 °C, a trace of rutile is typically observed, and on increasing the
annealing temperature the percentage of the rutile phase increases. Typically the material can be converted
completely to its rutile phase at 900 °C. However, at high temperatures (800–900 °C), the structure of the
tubes collapses.90 Typically the anatase phase of TiO2 turns out to be the most efficient for solar cell
performance. A certain amount of rutile phase present in the nanotube layers typically has a detrimental effect
on the solar cell, which may be ascribed to the approximately 10×lower electron mobility in rutile compared
with anatase as reported in ref. 98. Along these lines it is also interesting to note that the group of Frank
reported an approximately 10× higher electron diffusion constant for anatase than for rutile in particulate solar
cells.99 Other than the annealing temperature, the ramping speed also has great influence on the tube
morphology, crystal structure, and finally solar cell performances. There are several findings that are very
rsc.org/…/DisplayHTMLArticleforfree.cf… 15/35
10/20/2010 TiO2 nanotubes and their application in…
relevant considering the annealing procedure. Generally, annealing of the TiO2 nanotubes does not lead to
single crystalline material but very large scale grains ( 20 nm) can be observed (Fig. 6C)100 that do not
depend on the annealing temperature very much. In other words, the grain size as such is considerably larger
than the size of Grätzel-type nanoparticles ( 10 nm) and therefore higher electron mobilities could be
expected. However, at higher annealing temperatures (such as 450 °C) nanoscale cracks in the tube walls
become apparent,100 such as shown in Fig. 6B. Moreover, annealing at temperatures >400 °C will cause
some thermal oxidation of the underlying Ti-metal foil to rutile, i.e., thin rutile layers are present underneath the
nanotube layers as illustrated in Fig. 6C. These layers can only be observed if XRD sufficiently penetrates the
TiO2 nanotube layers (tubes shorter than 10 m length). Some results on the effect of thermal treatment of
TiO2 nanotubes on the solar cell performance will be discussed in section 2.3.4. It may be noteworthy that by
annealing the TiO2 nanotube in NH3 or acetylene successful N- or C-doping can be achieved101–103 or even
conductive TiO2 nanotubes can be formed.104

Fig. 6 (A) XRD analysis of TiO2 nanotube layers at different


annealing temperatures, revealing the amorphous as-prepared
material and anatase A or rutile R formation after the heat
treatment. (B)–(C) SEM and TEM images of TiO2 nanotubes
after annealing at 450 °C showing the crack lines and the
grains present in the tube wall, respectively. (D) SEM image

rsc.org/…/DisplayHTMLArticleforfree.cf… 16/35
10/20/2010 TiO2 nanotubes and their application in…
of TiO2 nanotubes after annealing at 450 °C showing the
bottom rutile layer underneath the nanotubes. Inset: XRD of
short tubes (4 m) and long tubes (20 m), indicating the rutile
peak only for short tubes.

2.2.2 Photoresponse. The typical electrochemical photoresponse of neat TiO2 nanotubes is summarized in
Fig. 7. For illumination with energies larger than the bandgap of TiO2 (rutile, Eg = 3.0 eV; anatase, Eg = 3.2
eV; amorphous, Eg = 3–3.6 eV) clearly non-annealed (amorphous) nanotubes show a low and hardly higher
photoresponse than corresponding compact layers, due to the high recombination rate of carriers in the
amorphous material.105 In other words, in order to switch on the tube walls, the tubes need to be annealed
to a crystallite form. The voltage dependence of the photocurrent seems to follow essentially the Gärtner
model.106 The photocurrent increases from the flatband voltage (ca. −0.2 to −0.5 VSCE in neutral electrolytes)
but reaches saturation under comparably mild anodization conditions (ca. +0.5 VSCE). For anodic TiO2 layers,
frequently comparably high doping densities are found that are attributed to defect structure, to non-
stoichiometry (TixOy) or hydroxide content.107 Even after annealing, doping levels – typically determined by
Mott–Schottky analysis – in the range of 1019 cm−3 are obtained.108,109 This level of doping is likely due to
oxygen vacancies or Ti3+ formation under the present annealing conditions.110 This indicates that the heavily
doped nanotube material [Nd 1019 cm−3 ] can expand a space charge layer with applied voltage but at
voltages of 0.5 V or higher the tube wall is completely depleted in carriers. The situation becomes comparable
to the case of nanoparticulate TiO2 films, where all evidence indicates that band bending effects are negligible,
and it is believed that the charge separation must be under thermal (diffusion) control.111 Under typical
conditions in these TiO2 nanostructures, the reaction of a photogenerated hole with a hole acceptor is assumed
to be fast, whereas the recombination of a photoexcited electron with an electron acceptor is assumed to be
slow. For UV excitation in aqueous electrolytes, this has been recently investigated by Lynch et al.107 Most
striking is that estimates of the electron diffusion length (Lel) for tubes differ strongly from nanoparticulate
layers. For nanoparticulate layers of the Grätzel type, Södergren et al. deduced, Lel 0.9 m.112 For
nanotubular layers for UV illumination Lel depends strongly on their morphology; while the first-generation
tubes (such as in Fig. 4B) yield Lel 2–5 m. For the second-generation smooth tubes (such as in Fig. 4C) Lel
>20 m is obtained.113 These findings on UV-induced electron transport in an aqueous electrolyte support the
high expectations on improved carrier transport properties in a nanotube morphology compared with
nanoparticulate layers.

Fig. 7 Photoelectrochemical behavior of TiO2 nanotubes and


rsc.org/…/DisplayHTMLArticleforfree.cf… 17/35
10/20/2010 TiO2 nanotubes and their application in…
compact oxide layers on annealing (photocurrent spectra taken
in 0.1M Na2 SO4 at 500 mV vs. Ag/AgCl. Inset: voltage
dependence of the photocurrent measured at a wavelength of
350 nm for the annealed samples).

2.3 Tubes in DSSCs

2.3.1 Overview. In 2005, the first report on dye-sensitized TiO2 nanotubes appeared114 and investigated
the first-generation tube layers such as those shown in Fig. 4B. The photoelectrochemical investigations were
carried out with TiO2 nanotubes with a diameter of 100 nm and length of 500 nm to 2.5 m using Ru-dye
(N3) as the sensitizer. The maximum incident photon-to-energy conversion (IPCE) value in the visible range
was of the order of 1.6–3.3%. Such tubes show, in solar cell constructions such as Fig. 2B, efficiencies of
0.036%. However, the use of the geometrically more defined nanotubes of the second generation – either
produced in glycerol or ethylene glycol electrolytes – drastically improved the reported efficiencies. Meanwhile
for plain tubes (not further modified with TiCl4 , TiO2 etc.), several independent articles14–19,115 find overall
solar cell efficiencies of 2–3% under AM 1.5 illumination. Factors that affect the efficiency and means to
improve it will be discussed in the following sections. It should be mentioned, however, that many of the results
reported by different groups are difficult to compare due to two reasons. First, many groups used mixtures of
tubes and nanoparticles,116–118 and in many cases it is not clear that the reported efficiency value indeed can
be ascribed to the presence of TiO2 nanotubes or if similar efficiencies could be obtained simply by using the
nanoparticles alone. Second, in full solar cell constructions (that may look as in Fig. 2A) not only the intrinsic
tube properties determine the efficiency but to a large extent the solar cell engineering parameters such as the
distance between the Pt counter electrode and tubes or the investigated tube area (i.e., active area of the solar
cell). For example it has been shown that with a change in the geometry of the active area, the overall
efficiencies may easily vary by a factor of 2.92,119 This illustrates the need for standardized testing and shows
that absolute efficiency values in literature should be carefully considered when relating them to tube geometry
or material properties. In other words, the discovery of material-specific efficiency enhancing effects can only
be reasonably compared within one work (one solar cell construction approach) and assuming that proper
reference experiments were carried out. Furthermore, one should mention some variations from using self-
organized tubes grown on Ti-metal sheets used in solar cells.
For example, Kang and his group29 fabricated highly ordered TiO2 nanotubes using a nanoporous alumina
templating method. Such nanotubes with 15 m lengths were heat treated at 500 °C for 30 min and soaked in
N3 dye for 24 h, which afterwards showed a solar cell efficiency of 3.5%. Some efforts are directed towards
using completely anodized sputter-deposited layers on conductive glass117,120 but claims such as conversion
efficiency of 2.9% fabricated on an only 360-nm-thick TiO2 nanotube arrays on conductive glass substrates
remain unconfirmed and somewhat questionable.117 In particular, sputtering or evaporating a >10 m thick Ti-
metal layer on conductive glass to convert it to TiO2 nanotubes or lift-off processes seem not economically
competitive. On the other hand, recent reports on hydrothermally fabricated nanorods such as the work by Liu
et al.121 may provide valuable alternatives. Recently, Grätzel and his co-workers19 fabricated a flexible dye-
sensitized solar cell from highly ordered TiO2 nanotube layers further treated with TiCl4 . They used the
conductive polymer Pt/ITO-PEN as the counter electrode in combination with an ionic liquid electrolyte. The
solar cell performance depends on the thickness of the nanotube layers, and the highest energy conversion
efficiency of 3.6% at AM 1.5 simulated full sunlight was obtained for 14 m long tubes.
Also, RBA tubes (shown in Fig. 5D) were used in DSSCs (in a similar arrangement to Fig. 2B). The
rsc.org/…/DisplayHTMLArticleforfree.cf… 18/35
10/20/2010 TiO2 nanotubes and their application in…
growth process leads to the advantage that high-aspect-ratio nanostructures (several dozens of m in length
vs. several dozens of nm in diameter) in the form of nanobundles can be prepared in a very short time
(minutes). Some points regarding these tubes will be discussed in section 2.3.9.

2.3.2 Electron dynamics and lifetime of electrons in nanotubes. In nanoparticulate DSSCs the electron
lifetime in the conduction band of TiO2 after electron injection from the dye and the electron diffusion through
the TiO2 films can be calculated using a multiple trapping model.9,14,107 The electrons can recombine through
surface trap sites or directly from the conduction band with the I−/I3 − redox couple in the electrolyte. In case
of nanoparticulate films, it has been found that the diffusion coefficient (D) of electrons increases and the
electron recombination lifetime ( ) decreases with an increase in the particle size up to 32 nm.12 The value
usually is influenced by the surface area, which determines the effective electron acceptor density for
recombination. The electron dynamics were considered by Frank and his co-workers93 who also compared
tubes to classical nanoparticle layers. It was proposed that the nanotube and nanoparticle films showed similar
transport times, whereas recombination was 10× slower in the nanotube films, leading to a significantly
enhanced charge collection efficiency for the nanotube photoelectrodes. Interestingly, it was reported that
lifetimes do not strongly depend on tube length. Along these lines, Frank et al. reported a significant effect of
tube top geometry on transport times25 as will be discussed more in section 2.3.7.
The most detailed study on charge carrier behaviour in dye-sensitized TiO2 nanotube arrays stems from
Jennings et al.14 It was shown that at about 20 m tube length, IPCE values close to 100% could be obtained
(Fig. 8A) when corrected for reflection losses. The measured and calculated IPCE values for 20 m long
nanotubes show good agreement. Most remarkable is that the electron diffusion length, based on experimental
results to determine the diffusion coefficient and lifetime of the photoinjected electrons is of the order of 100
m in the titania nanotube cells (Fig. 8B). This means a value of L that is significantly larger than 20 m and
100% collection of electrons (without recombination) could be achieved with significantly longer tubes. The
value of Lel = 100 m seems to be in contrast to a value of Lel 30 m obtained from data in the work by
Zhu et al.18,25 However, it seems that tube properties are sensitive to the preparation procedure as mentioned
in the next section, as for the differently prepared nanotubes considerable differences in Lel were obtained.
The study of Jennings et al.14 also revealed a substantial discrepancy between the shapes of the electron trap
distributions measured experimentally using charge extraction techniques and those inferred indirectly from
transient current and voltage measurements.

Fig. 8 (A) Comparison of measured (solid line) and calculated


(dotted line) IPCE values for different thicknesses of TiO2
nanotube layers; dashed line – the calculated IPCE spectrum
for 20 m long nanotubes without including the absorption loss
due to I3 in the electrolyte. (B) Estimation of electron diffusion

rsc.org/…/DisplayHTMLArticleforfree.cf… 19/35
10/20/2010 TiO2 nanotubes and their application in…
length (Ln ) of 20 m long TiO2 nanotubes from the experimental
values of Dn and c by taking the quasi Fermi level (QFL) into
account. (Reproduced with permission from ref. 14. Copyright
American Chemical Society, 2008).

2.3.3 Tube length and diameter. Several authors have investigated the influence of tube length on the
conversion efficiency (IPCE) and on the final solar cell performance.14,16–18 Typical results for the IPCE are
shown in Fig. 8A. As mentioned above, the internal conversion efficiency reaches a maximum for tubes of a
length of approximately 20 m. These findings are consistent if the measurements are performed in a
photoelectrochemical or in a solar cell set-up. Typical more recent results are shown in Fig. 9 and are
obtained with a commercial R50 electrolyte from Solaronix. From the characteristic I–V curve it can be seen
that the short-circuit current (Isc) increases with an increase in the thickness of nanotubes from 8 to 24 m and
for the first-generation tubes (1–2 m long) very small Isc values are observed. This very low efficiency for the
first-generation tubes may be attributed to their highly irregular tube wall structure, which in the non-sensitized
tubes, also has a very detrimental effect on the electron lifetime. For the second-generation tubes, on
increasing the tube thickness, the surface area increases and more dye is adsorbed onto the tube wall surfaces
allowing an increase in light harvesting. The open-circuit voltage (Voc) values decreases from 0.68 V to 0.62
and 0.58 V on increasing the tube thickness from 16 to 24 and 32 m, respectively. This may be due to the
fact that at 20 m an optimum collection is reached, although the electron diffusion length is 100 m.14 This,
in fact, may suggest that for tubes >20 m, the overall properties may not be the same. For example, a lower
collection efficiency may be due to the previously mentioned thinning of the tube walls when tubes are grown
for extended times (grass formation on the tube surface). This factor, that tube top morphology is crucial, has
been addressed several times in literature and will be discussed in section 2.3.7. By considering all the factors,
16–25 m is found to be the optimum tube length for attaining the highest energy conversion efficiency
(typically reaching 2.5%). It should be noted that low short-circuit current values for >30 m long tubes may
also be due to an increasing probability of loss of attachment between tube bottom and the Ti-foil for the long
tubes. Up to now the most comprehensive work on solar cell performances has been provided by Ghicov et
al.17 This is particularly important as measurements on solar cell efficiencies vary strongly from various groups,
i.e., comparable consistent data is only obtained when identical solar cell constructions and measurement
conditions are used. Moreover, different characterization tools and methods are used and the comparison of
tubes is done frequently in different terms, such as, diffusion constants, transport time constants and
recombination lifetimes. The work of Ghicov et al.17 suffers to some extent from comparably poor fill factors
(due to a simple solar cell construction).122,123 Furthermore, in the work of Ghicov et al.,17 a simple
electrolyte such as an 0.1 M lithium iodide +0.05 M iodine was used and a maximum 1.6% efficiency at 20
m tube length was obtained. However, the addition of 0.6M tert-butyl pyridine (BP) increased the efficiency
to 2.4%.17 Tert-butyl pyridine was added as an agent to the electrolyte in order to block surface states by
preferential adsorption, thereby decreasing the chance of recombination.17 However, the overall trend of
maximum efficiency at a tube length 20 m was not changed.

rsc.org/…/DisplayHTMLArticleforfree.cf… 20/35
10/20/2010 TiO2 nanotubes and their application in…

Fig. 9 I–V characteristics measured for the solar cell


configuration of Fig. 2B for different lengths of TiO2
nanotube layers after annealing at 450 °C for 1 h
using commercial R50 as electrolyte and dye soaking
(N719) for 1 day.

The tube diameter also has significant effect on the solar cell performance. It was found that for the same
thickness of nanotube layers, smaller diameter tubes show a higher energy conversion efficiency.17 Fig. 10
compiles a series of results obtained for several different tube geometries. Although some details such as the
electrolyte or solar cell geometry vary slightly in different articles, the clear trend that a smaller diameter (at the
same length) increases the solar cell response is observed as indicated in the inset. These findings overall
correlate with the trend that an increase in the surface area of tubes increases the specific dye loading. This
holds for decreasing the diameter or increasing the length of the nanotubes. In several cases this finding was
confirmed by dye desorption measurements.14,17 As a bottom line for the tubes: though a better electrical
performance is provided than in nanoparticulate layers, the lower specific surface area [BETtubes 30 m2 g−1 ;
BETnanoparticles 80 m2 g−1 ] lowers the dye loading and thus reduces the efficiency. Some measures against
this will be discussed in sections 2.3.5 and 2.3.8.

rsc.org/…/DisplayHTMLArticleforfree.cf… 21/35
10/20/2010 TiO2 nanotubes and their application in…

Fig. 10 Influence of tube geometry: Comparison of


solar cell efficiencies from literature and unpublished
data, illustrating the effect of TiO2 nanotube length and
diameter. All of the new series of experiments were
carried out using R50 as electrolyte and N719 dye
soaking for 1 day. Inset: Schematic trend of solar cell
performances of TiO2 nanotube layers with tube length
and diameter.

2.3.4 Annealing. Other than the tube geometry, the annealing temperature, duration and ramping also has
significant effect on dye loading, electron transport time and in the end the electron collection efficiency. The
as-prepared TiO2 nanotubes are amorphous in nature and show negligible energy conversion efficiency in
solar cells. The nanotubes convert to anatase phase of TiO2 after annealing at 300 °C (compare Fig. 6), and
depending on the annealing temperature anatase, rutile or a mixture of phases is obtained. From the XRD data
can be observed that at 450–500 °C the pure anatase phase of TiO2 is obtained. However, at 550 °C or
above, a mixture of anatase and rutile phases is formed. An overview of results for different annealing
temperatures and times is compiled in Fig. 11. In this case, all the samples were annealed in a rapid thermal
annealing chamber with a ramping speed of 30 °C min−1 . In all the experiments the samples were dye-
sensitized with Ru-based N719 dye for 1 day at room temperature. The I–V characteristic curves for different
annealing temperatures reveal that the short-circuit current and the open-circuit voltage increase with
increasing annealing temperature. This may be due to the increase in crystallinity, thereby promoting the
connectivity between grains and consequent improvement in electron diffusion coefficients and lifetimes.17,90
At higher annealing temperatures the efficiency decreases. This observation can be explained because of the
detrimental effect of rutile phase formation at high annealing temperatures. However, other key points, as
outlined in section 2.2.1, are that annealing of the tubes at higher temperatures leads to cracks visible in the
SEM and TEM images100 and this effect may drastically reduce electron mobility. Additionally, as also
outlined in a previous section, higher temperature annealing of the TiO2 nanotubes on their Ti substrate leads

rsc.org/…/DisplayHTMLArticleforfree.cf… 22/35
10/20/2010 TiO2 nanotubes and their application in…
typically to the formation of a thermal oxide of a rutile type (Fig. 6c) underneath the nanotubes which can also
affect the electron mobility. This would in turn explain why Zhu et al.25 find an electron mobility in TiO2
nanotubes that is similar to particulate systems. A maximum solar cell efficiency of 2.95% is obtained after
annealing at 450 °C for 3 h. Figs. 11B and C show the I–V curves of TiO2 nanotubes for different annealing
durations and different ramping speeds at 450 °C. The higher Isc value for 3 h annealed sample results in the
highest energy conversion efficiency of 3%. This increase in efficiency due to longer annealing times may be
due to better crystallization of the anatase phase. Besides different annealing temperature and durations, also
using a different ramping speed of annealing at 450 °C showed significant effect on solar cell performances. It
is observed that very fast ramping shows a lower solar cell efficiency. This might be due to the fact that very
fast ramping affects the tube morphology drastically as shown in work by Albu et al.90

rsc.org/…/DisplayHTMLArticleforfree.cf… 23/35
10/20/2010 TiO2 nanotubes and their application in…

Fig. 11 Influence of tube annealing: (A) Compilation of


solar cell efficiencies of 16 m TiO2 nanotube layers at
rsc.org/…/DisplayHTMLArticleforfree.cf… 24/35
10/20/2010 TiO2 nanotubes and their application in…
different annealing temperatures (Series I measurements
were carried out with 0.1 M lithium iodide +0.05 M
iodine containing 0.6M tert-butyl pyridine as electrolyte I
and Series II measurements were carried out using R50
as electrolyte II and N719 as dye for both cases). (B)
I–V characteristics of 16 m TiO2 nanotube layers
annealed at 450 °C for different durations; (C) I–V
characteristics of 16 m TiO2 nanotube layers annealed
at 450 °C for different annealing ramping speeds. (D)
Electron transport ( c) and recombination time ( r)
constants calculated from IMPS and IMVS curves for
16 m TiO2 nanotube layer-based DSSCs at different
annealing temperatures (reproduced with permission
from ref. 17. Copyright Wiley-VCH, 2009).

Very remarkable is a recent finding by Ghicov et al.17 that the electron transport properties through
nanotubes can strongly depend on the annealing temperature. The electron recombination ( r) and transport
time constants ( c) can be evaluated by intensity-modulated photocurrent spectroscopy (IMPS) and intensity-
modulated photovoltage spectroscopy (IMVS) measurements as shown in Fig. 11D (from ref. 17). Extracting
the transport time constants shows that c is 10× smaller for tubes annealed at 350 °C compared with tubes
annealed at 450 °C, which means the electron transport is drastically faster in the tubes annealed at 350 °C.
This may be attributed to many fewer cracks induced at this annealing temperature100 or less rutile formation
underneath the tubes. Although the finding of high rate constants at lower annealing temperatures is extremely
promising for further research, the authors point out in their article that for these tubes a significantly lower dye
loading was attained. The reasons for this at the current stage are unclear and further work is required to
possibly exploit these effects.

2.3.5 TiCl4 treatment. The above findings show clearly that for tubes a very crucial factor is simply the
surface area, as it determines the amount of dye adsorbed on the tubes.17 In other words, in order to improve
the overall solar cell efficiency, the amount of dye adsorbed by a unit solar cell volume needs to be higher.
One way to achieve a higher surface area is to use a treatment well established for nanoparticulate DSSCs: the
so-called TiCl4 treatment.124–129 Typical results obtained for nanotube layers before and after this treatment
are shown in Fig. 12(A)–(B).129 The inner as well as the outer wall of TiO2 nanotubes are covered with a
25 nm thick nanoparticle coating of TiO2 nanoparticles having 3 nm, as shown in Fig. 12(C). This leads to a
significant increase in surface area and therefore more dye can be adsorbed to the nanotube walls. Attempts to
carry out TiCl4 treatments directly on as-grown or amorphous nanotube layers resulted in a lift-off of nanotube
layers from the Ti substrate. In order to avoid this problem, double annealing was carried out: The first time to
anneal the tubes, and the second time after TiCl4 tratment of nanotubes to convert the deposited nanoparticles
to a crystalline structure. After decoration with TiO2 nanoparticles, the Isc as well as the Voc values increases
indicating high light harvesting due to the larger amount of dye loading and the increase in the porosity of
material. All these factors help to achieve an energy conversion efficiency of 3.8% after TiCl4 treatment as
shown in Fig. 12(D). The first preliminary results on further optimization of the process, however, show
efficiencies of >5%.130 For less well defined tubes and particle mixtures efficiencies up to 7% have been
reported118 but as mentioned before, the effective contribution of tubes remains unclear.

rsc.org/…/DisplayHTMLArticleforfree.cf… 25/35
10/20/2010 TiO2 nanotubes and their application in…

Fig. 12 Influence of TiCl4 treatment: (A)–(B) SEM images of


TiO2 nanotubes before and after TiCl4 treatment, showing
tube decoration with TiO2 nanoparticles (reproduced with
permission from ref. 129. Copyright Elsevier, 2009). (C)
TEM and HRTEM of TiO2 nanotubes after TiCl4 treatment.
(D) I–V curve for non-treated (NT) and treated tubes (TT).

2.3.6 Dye sensitization duration. Several experiments have been carried out in order to optimize the dye
sensitization duration. In this context it should be noted that also the dye sensitization temperature plays a
significant role. Dye sensitization at 40 °C in a closed container in the dark leads to very fast anchoring of –
COOH group at the TiO2 surface due to possibly a faster dehydration reaction. Some results from various
experimental series are compiled in Fig. 13. First results from Ghicov et al. for TiO2 nanoparticle free
nanotubes showed that classic dye sensitization in the case of tubes can take significantly longer for tubes (3–5
days soaking to reach saturation) than usually needed for nanoparticulate layers (some hours to one day). This
may on the one hand be ascribed to a slightly different surface reactivity of tubes (see e.g., ref. 131) but may
also be due to the fact that the tubes represent a closed porous system, and in some cases a delayed wetting
rsc.org/…/DisplayHTMLArticleforfree.cf… 26/35
10/20/2010 TiO2 nanotubes and their application in…
of the tube insides has been mentioned.132 The fact that for more than 5 days dye sensitization the efficiency
decreases may be associated with dye agglomeration on the tube surface. Clearly, dye sensitization at elevated
temperature such as at 40 °C leads to a faster dye adsorption and thus a significant improvement in solar cell
performance. E.g., improvement from 1.3% to 3% for non-particle treated tubes and up to 4% for TiO2
nanoparticle decorated tubes could be achieved [in short sensitization times (24 h)]. Although these results
give a general idea about important effects in the dye sensitization of nanotubes, and reveal some differences to
nanoparticulate layers, considerable room for optimization seems to exist particularly regarding the interplay
with TiO2 nanoparticle decorated tube walls.

Fig. 13 Influence of dye soaking time: An overview


graph for non-treated (TiCl4 ) as well as treated tubes
for different dye soaking durations.

2.3.7 Top morphology. Most tube tops prepared in conventional ethylene glycol electrolytes are not
entirely ideal. Due to chemical etching during growth of the nanotubes, the overall tube wall thickness is thinner
at the top than at the bottom.84,90 At the top region the tube walls can become very thin so that tube upper
parts may collapse and form bundles, or form needle-like structures apparent as grass on the top of the
tubes. Such a grassy morphology on the top of the tubes may inhibit smooth electron transport in one
direction and thus affect the solar cell performance.
In order to avoid these problems, or at least reduce it, several approaches have been developed by various
researchers. Frank and his co-workers25 improved the solar cell efficiency from 1.6% to 1.9% by modifying
the cleaning and drying step using supercritical CO2 drying of tube layers. This approach leads to bundle-free
and crack-free nanotube films. The positive effect on the solar cell efficiency was ascribed to an enhancement
of the total surface area and making the tube tops more accessible to dye molecules. According to these
authors, the inter-tube connection in the compact nanotube bundle effects the electron transport mechanism
from the expected one-dimensional pathway to a mixture of one- and three-dimensional pathways. By making
bundle-free and crack-free nanotubes, the one-dimensional transport shortens the transport pathway and
reduces multiple electron trapping events.
Another approach to making nanograss-free TiO2 nanotubes is by polishing the Ti-foil prior to
anodization.16 A protective top layer can be formed on polished Ti samples, which delays the chemical attack
of the tube ends and thus the formation of nanograss and tube bundling can be avoided. Additionally, the
layer seems also to act as a stratification against disorder induced by the drying process. Such highly aligned,

rsc.org/…/DisplayHTMLArticleforfree.cf… 27/35
10/20/2010 TiO2 nanotubes and their application in…
ordered, and nanograss -free geometries show higher solar cell efficiencies than the conventional nanotubes
(see Fig. 14). The maximum energy conversion efficiency is obtained again for 20 m thick nanotube layers.
The open-circuit voltage of nanograss-free nanotubes is found to be higher than that for grassy tubes. This
may be due to the fact that in more ordered grass-free tubes the electron lifetime is higher due to the presence
of fewer defects.

Fig. 14 Influence of tube top morphology: A


comparative study of solar cell performances of
nanograss-free and grassy tube tops of TiO2 layers.

Another approach to create top ordered layer was reported by Song et al.93 In this approach a thermal
rutile layer is first formed on the Ti metal surface and tubes are then grown underneath. This treatment was
found to be very effective to produce defined tubetops up to considerable length of >50 m.93 Clearly, such
tubes are very promising in view of achieving defined tube tops. Most recently also direct growth approaches
in optimized electrolytes are being explored to avoid grass formation.92 However, it should be noted that
smooth and regular tube tops may not represent the most ideal case as the disorder has an anti-reflection
effect which helps high light harvesting as evident from the high Isc values.

2.3.8 Advanced geometries. In order to modify the tube geometry, in particular to achieve larger surface
areas, modified morphologies such as bamboo tubes (Fig. 5A)85 or double-walled structures (Fig. 5B)90 are
being explored in DSSCs.15 For example, bamboo-type geometry tubes with rings lead to a more optimized
volume filling of the TiO2 nanostructure with a higher surface area per unit volume of the solar cell. Bamboo
tubes were compared with conventional smooth-walled tubes of the same thickness. The I–V characteristics
revealed that the short-circuit current significantly increases in comparison with the conventional smooth-
walled tubes, due to the chemisorption of a large number of dye molecules that can cover the interior and
exterior of TiO2 nanotube walls. In particular, the rings of bamboo-type tubes may act as spacers at the
exterior of the nanotube walls, which promote the enhanced dye loading.

2.3.9 RBA tubes. As mentioned in section 2.3.1, RBA-type TiO2 nanotubes can be prepared from
fluoride-free electrolytes and long nanotube bundles of 30 m thickness can be prepared in a very short time
(seconds to few minutes). These bundles on the Ti-substrate can be mounted in similar solar cell structures as
shown in Fig. 2B. Due to the high surface area of these bundles of nanotubes and porous structures, already in

rsc.org/…/DisplayHTMLArticleforfree.cf… 28/35
10/20/2010 TiO2 nanotubes and their application in…
very early experiments some comparably good performances have been obtained.95,96 Fig. 15B shows an
I–V curve of RBA tubes with and without ethanol in the electrolyte during growth of the tubes. Ethanol leads
to a higher bundle density. Therefore, the higher surface area available for dye anchoring of highly dense
bundles of nanotubes leads to higher energy conversion efficiencies in RBA-EtOH tubes, yielding overall a
solar cell efficiency of 2.4%.

Fig. 15 Influence of tube morphology modifications:


I–V characteristics for DSSCs of different advanced
geometries of TiO2 nanotubes. (A) Bamboo nanotube
(B-NT) with different ring distance and normal
nanotubes (NT) (reproduced with permission from
ref. 15. Copyright American Chemical Society,
2008). (B) RBA tubes formed in aqueous electrolyte
and in ethanolic electrolyte.

2.4 Conclusions
The present article tries to give an overview of the current state of using TiO2 nanotubes in DSSCs. It is
remarkable that since the first results in 2005 with efficiencies of 0.04%, reproducible results in the order of 4–
5% efficiency have since been obtained. This although a number of parameters, such as, tube annealing, dye
optimization and solar cell geometry, are far from optimized. Exciting results involve the finding of much higher
electron diffusion length in tubes than in particles, and the finding that the tube geometry strongly affects
electron lifetime – in optimized tubes electron diffusion lengths of 100 m have been observed. The fact that
annealing conditions strongly affect electron collection values give room for significant improvement as here
one of the crucial loss mechanisms in DSSCs (recombination with the electrolyte) can be reduced. While some
higher efficiencies up to 7% have been reported in literature for mixing tubes with nanoparticles, one has to

rsc.org/…/DisplayHTMLArticleforfree.cf… 29/35
10/20/2010 TiO2 nanotubes and their application in…
carefully judge if these effects are indeed due to tubes or must be ascribed primarily to the effect of the
nanoparticles.
However, from the current state of investigations it is clear that specific dye loading in the tubes need to be
improved – i.e., TiO2 structures with a larger surface area need to be created. This may be achieved in
various ways by modifying the tube walls with porous materials (double-walled tubes), TiO2 nanoparticle
decoration or creating defined bamboo-type rings. Also, several issues relating to practical solar cell
construction, such as, front- or backside illumination, optimized layout, surface reflectivity, or designed light
absorption are further open research areas and exciting new findings can also be expected in the development
in these technology areas.

References

1 F. Pfisterer, Photovoltaic Cells, Ullmann's Encyclopedia of Industrial Chemistry, 2002,Wiley-VCH,


Germany.
2 J. Desilvestro, M. Grätzel, L. Kavan and J. Moser, J. Am. Chem. Soc., 1985, 107, 2988 [Links].
3 N. Vlachopoulos, P. Liska, J. Augustynski and M. Grätzel, J. Am. Chem. Soc., 1988, 110, 1216
[Links].
4 B. O'Regan and M. Grätzel, Nature, 1991, 353, 737 [Links].
5 M. Grätzel, Nature, 2001, 414, 338 [Links].
6 M. K. Nazeeruddin, F. D. Angelis, S. Fantacci, A. Selloni, G. Viscardi, P. Liska, S. Ito, B. Takeru and
M. Grätzel, J. Am. Chem. Soc., 2005, 127, 16835 [Links].
7 Y. Chiba, A. Islam, Y. Watanabe, R. Komiya, N. Koide and L. Han, Jpn. J. Appl. Phys., 2006, 45,
L638 [Links].
8 M. Grätzel, Inorg. Chem., 2005, 44, 6841 [Links].
9 S. E. Koops, B. C. O'Regan, P. R. F. Barnes and J. R. Durrant, J. Am. Chem. Soc., 2009, 131, 4808
[Links].
10 J. Nelson, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 15374 [Links] LP.
11 J. Bisquert, Phys. Rev. Lett., 2003, 91, 010602 [Links].
12 S. Nakade, Y. Saito, W. Kubo, T. Kitamura, Y. Wada and S. Yanagida, J. Phys. Chem. B, 2003,
107, 8607 [Links].
13 L. M. Peter, J. Phys. Chem. C, 2007, 111, 6601 [Links].
14 J. R. Jennings, A. Ghicov, L. M. Peter, P. Schmuki and A. B. Walker, J. Am. Chem. Soc., 2008, 130,
13364 [Links].
15 D. Kim, A. Ghicov, S. P. Albu and P. Schmuki, J. Am. Chem. Soc., 2008, 130, 16454 [Links].
16 D. Kim, A. Ghicov and P. Schmuki, Electrochem. Commun., 2008, 10, 1835 [Links].
17 A. Ghicov, S. Albu, R. Hahn, D. Kim, T. Stergiopoulos, J. Kunze, C.-A. Schiller, P. Falaras and P.
Schmuki, Chem.–Asian J., 2009, 4, 520 [Links].
18 K. Zhu, N. R. Neale, A. Miedaner and A. J. Frank, Nano Lett., 2007, 7, 69 [Links].
19 D. Kuang, J. Brillet, P. Chen, M. Takata, S. Uchida, H. Miura, K. Sumioka, S. M. Zakeeruddin and
M. Grätzel, ACS Nano, 2008, 2, 1113 [Links].
20 S. H. Kang, S.-H. Choi, M.-S. Kang, J.-Y. Kim, H.-S. Kim, T. Hyeon and Y.-E. Sung, Adv. Mater.,
2008, 20, 54 [Links].
rsc.org/…/DisplayHTMLArticleforfree.cf… 30/35
10/20/2010 TiO2 nanotubes and their application in…

21 M. Y. Song, D. K. Kim, K. J. Ihn, S. M. Jo and D. Y. Kim, Nanotechnology, 2004, 15, 1861


[Links].
22 S. Pavasupree, S. Ngamsinlapasathian, Y. Suzuki and S. Yoshikawa, J. Nanosci. Nanotechnol.,
2006, 6, 3685 [Links].
23 Y. Ohsaki, N. Masaki, T. Kitamura, Y. Wada, T. Okamoto, T. Sekino, K. Niihara and S. Yanagida,
Phys. Chem. Chem. Phys., 2005, 7, 4157 [Links].
24 M. Adachi, Y. Murata, J. Takao, J. Jiu, M. Sakamoto and F. Wang, J. Am. Chem. Soc., 2004, 126,
14943 [Links].
25 K. Zhu, T. B. Vinzant, N. R. Neale and A. J. Frank, Nano Lett., 2007, 7, 3739 [Links].
26 H.-H. Ou and S.-L. Lo, Sep. Purif. Technol., 2007, 58, 179 [Links].
27 X. Chen and S. S. Mao, Chem. Rev., 2007, 107, 2891 [Links].
28 D. V. Bavykin, J. M. Friedrich and F. C. Walsh, Adv. Mater., 2006, 18, 2807 [Links].
29 T.-S. Kang, A. P. Smith, B. E. Taylor and M. F. Durstock, Nano Lett., 2009, 9, 601 [Links].
30 G. R. Patzke, F. Krumeich and R. Nesper, Angew. Chem., Int. Ed., 2002, 41, 2446 [Links].
31 A. Ghicov and P. Schmuki, Chem. Commun., 2009, 2791 [Links].
32 J. M. Macak, H. Tsuchiya, A. Ghicov, K. Yasuda, R. Hahn, S. Bauer and P. Schmuki, Curr. Opin.
Solid State Mater. Sci., 2007, 11, 3 [Links].
33 M. J. Cass, A. B. Walker, D. Martinez and L. M. Peter, J. Phys. Chem. B, 2005, 109, 5100 [Links].
34 N. G. Park, J. van de Lagemaat and A. J. Frank, J. Phys. Chem. B, 2000, 104, 8989 [Links].
35 H. Lindström, A. Holmberg, E. Magnusson, S. E. Lindquist, L. Malmqvist and A. Hagfeldt, Nano
Lett., 2001, 1, 97 [Links].
36 H. Lindström, E. Magnusson, A. Holmberg, S. Sodergren, S. E. Lindquist and A. Hagfeldt, Sol.
Energy Mater. Sol. Cells, 2002, 73, 91 [Links].
37 A. Mishra, M. K. R. Fischer and P. Bauerle, Angew. Chem., Int. Ed., 2009, 48, 2474 [Links].
38 N. Robertson, Angew. Chem., Int. Ed., 2006, 45, 2338 [Links].
39 M. Grätzel, Progr. Photovolt.: Res. Appl., 2006, 14, 429.
40 M. Grätzel, J. Photochem. Photobiol., A, 2004, 164, 3 [Links].
41 P. Bonhôte, A.-P. Dias, N. Papageorgiou, K. Kalyanasundaram and M. Grätzel, Inorg. Chem., 1996,
35, 1168 [Links].
42 S. Hwang, J. H. Lee, C. Park, H. Lee, C. Kim, C. Park, M.-H. Lee, W. Lee, J. Park, K. Kim, N.-G.
Park and C. Kim, Chem. Commun., 2007, 4887 [Links].
43 M. K. Nazeeruddin, A. Kay, I. Rodicio, R. Humpbry-Baker, E. Miiller, P. Liska, N. Vlachopoulos and
M. Grätzel, J. Am. Chem. Soc., 1993, 115, 6382 [Links].
44 M. K. Nazeeruddin, P. Pechy and M. Grätzel, Chem. Commun., 1997, 1705 [Links].
45 M. K. Nazeeruddin, S. M. Zakeeruddin, R. Humphrey-Baker, M. Jirousek, P. Liska, N.
Vlachopoulos, V. Shklover, C.-H. Fischer and M. Grätzel, Inorg. Chem., 1999, 38, 6298 [Links].
46 D. Kuciauskas, M. S. Freund, H. B. Gray, J. R. Winkler and N. S. Lewis, J. Phys. Chem. B, 2001,
105, 392 [Links].
47 R. Argazzi, G. Larramona, C. Contado and C. A. Bignozzi, J. Photochem. Photobiol., A, 2004, 164,
15 [Links].
rsc.org/…/DisplayHTMLArticleforfree.cf… 31/35
10/20/2010 TiO2 nanotubes and their application in…
48 A. Islam, H. Sugihara, K. Hara, L. P. Singh, R. Katoh, M. Yanagida, Y. Takahashi, S. Murata and H.
Arakawa, Inorg. Chem., 2001, 40, 5371 [Links].
49 E. A. M. Geary, L. J. Yellowlees, L. A. Jack, I. D. H. Oswald, S. Parsons, N. Hirata, J. R. Durrant
and N. Robertson, Inorg. Chem., 2005, 44, 242 [Links].
50 N. Alonso-Vante, J.-F. Nierengarten and J.-P. Sauvage, J. Chem. Soc., Dalton Trans., 1994, 1649
[Links].
51 P. M. Jayaweera, S. S. Palayangoda and K. Tennakone, J. Photochem. Photobiol., A, 2001, 140,
173 [Links].
52 P. Wang, C. Klein, J.-E. Moser, R. Humphry-Baker, N.-L. Cevey-Ha, R. Charvet, P. Comte, S. M.
Zakeeruddin and M. Grätzel, J. Phys. Chem. B, 2004, 108, 17553 [Links].
53 H. Park, E. Bae, J.-J. Lee, J. Park and W. Choi, J. Phys. Chem. B, 2006, 110, 8740 [Links].
54 K. Hara, Z.-S. Wang, T. Sato, A. Furube, R. Katoh, H. Sugihara, Y. Dan-oh, C. Kasada, A. Shinpo
and S. Suga, J. Phys. Chem. B, 2005, 109, 15476 [Links].
55 S. Alex, U. Santhosh and S. Das, J. Photochem. Photobiol., A, 2005, 172, 63 [Links].
56 T. Horiuchi, H. Miura, K. Sumioka and S. Uchida, J. Am. Chem. Soc., 2004, 126, 12218 [Links].
57 W. M. Campbell, A. K. Burrell, D. L. Officer and K. W. Jolley, Coord. Chem. Rev., 2004, 248,
1363 [Links].
58 Q. Wang, W. M. Campbell, E. E. Bonfantani, K. W. Jolley, D. L. Officer, P. J. Walsh, K. Gordon, R.
Humphrey-Baker, M. K. Nazeeruddin and M. Grätzel, J. Phys. Chem. B, 2005, 109, 15397 [Links].
59 M. Grätzel, J. Photochem. Photobiol., C, 2003, 4, 145 [Links].
60 Z.-S. Wang, K. Sayama and H. Sugihara, J. Phys. Chem. B, 2005, 109, 22449 [Links].
61 H. Nusbaumer, J.-E. Moser, S. M. Zakeeruddin, M. K. Nazeeruddin and M. Grätzel, J. Phys. Chem.
B, 2001, 105, 10461 [Links].
62 S. A. Sapp, C. M. Elliott, C. Contado, S. Caramori and C. A. Bignozzi, J. Am. Chem. Soc., 2002,
124, 11215 [Links].
63 G. Oskam, B. V. Bergeron, G. J. Meyer and P. C. Searson, J. Phys. Chem. B, 2001, 105, 6867
[Links].
64 P. Wang, S. M. Zakeeruddin, J.-E. Moser, R. Humphry-Baker and M. Grätzel, J. Am. Chem. Soc.,
2004, 126, 7164 [Links].
65 G. Boschloo, L. Haggman and A. Hagfeldt, J. Phys. Chem. B, 2006, 110, 13144 [Links].
66 D. B. Kuang, C. Klein, Z. P. Zhang, S. Ito, J. E. Moser, S. M. Zakeeruddin and M. Grätzel, Small,
2007, 3, 2094 [Links].
67 J.-H. Yum, P. Chen, M. Grätzel and M. K. Nazeeruddin, Chem. Sus. Chem., 2008, 1, 699 [Links].
68 H. J. Snaith and L. Schmidt-Mende, Adv. Mater., 2007, 19, 3187 [Links].
69 Y. Bai, Y. Cao, J. Zhang, M. Wang, R. Li, P. Wang, S. M. Zakeeruddin and M. Grätzel, Nat. Mater.,
2008, 7, 626 [Links].
70 A. B. F. Martinson, J. W. Elam, J. T. Hupp and M. J. Pellin, Nano Lett., 2007, 7, 2183 [Links].
71 N. N. Dinh, M.-C. Bernard, A. H.-L. Goff, T. Stergiopoulos and P. Falaras, C. R. Chim., 2006, 9,
676 [Links].
72 K. Tennakone, G. R. R. Kumara, I. R. M. Kottegoda and V. S. P. Perera, Chem. Commun., 1999,
15 [Links].
rsc.org/…/DisplayHTMLArticleforfree.cf… 32/35
10/20/2010 TiO2 nanotubes and their application in…
73 K. Sayama, H. Suguhara and H. Arakawa, Chem. Mater., 1998, 10, 3825 [Links].
74 F. Lenzmann, J. Krueger, S. Burnside, K. Brooks, M. Grätzel, D. Gal, S. Rühle and D. Cahen, J.
Phys. Chem. B, 2001, 105, 6347 [Links].
75 J. B. Baxter and E. S. Aydil, Appl. Phys. Lett., 2005, 86, 053114 [Links].
76 P. Schmuki, in Electrochemistry at the Nanoscale, ed. P. Schmuki and S. Virtanen, Springer, New
York, 1st edn, 2009, pp. 435–466.
77 V. Zwilling, E. Darque-Ceretti, A. Boutry-Forveille, D. David, M. Y. Perrin and M. Aucouturier, Surf.
Interface Anal., 1999, 27, 629 [Links].
78 D. Gong, C. A. Grimes, O. K. Varghese, Z. Chen and E. C. Dickey, J. Mater. Res., 2001, 16, 3331
[Links].
79 R. Beranek, H. Hildebrand and P. Schmuki, Electrochem. Solid-State Lett., 2003, 6, B12 [Links].
80 J. M. Macak, K. Sirotna and P. Schmuki, Electrochim. Acta, 2005, 50, 3679 [Links].
81 J. M. Macak, H. Tsuchiya, L. Taveira, S. Aldabergerova and P. Schmuki, Angew. Chem., Int. Ed.,
2005, 44, 7463 [Links].
82 J. M. Macák, H. Tsuchiya and P. Schmuki, Angew. Chem., Int. Ed., 2005, 44, 2100 [Links].
83 S. P. Albu, A. Ghicov, J. M. Macak and P. Schmuki, Phys. Status Solidi RRL, 2007, 1, R65.
84 J. M. Macak, S. P. Albu and P. Schmuki, Phys. Status Solidi RRL, 2007, 1, 181.
85 S. P. Albu, D. Kim and P. Schmuki, Angew. Chem., Int. Ed., 2008, 47, 1916 [Links].
86 W. Chanmanee, A. Watcharenwong, C. R. Chenthamarakshan, P. Kajitvichyanukul, N. R. Tacconi
and K. Rajeshwar, J. Am. Chem. Soc., 2008, 130, 965 [Links].
87 Y.-Y. Song, F. Schmidt-Stein, S. Bauer and P. Schmuki, J. Am. Chem. Soc., 2009, 131, 4230
[Links].
88 J. M. Macak, S. Albu, D. H. Kim, I. Paramasivam, S. Aldabergerova and P. Schmuki, Electrochem.
Solid-State Lett., 2007, 10, K28 [Links].
89 K. Yasuda and P. Schmuki, Electrochem. Commun., 2007, 9, 615 [Links].
90 S. P. Albu, A. Ghicov, S. Aldabergenova, P. Drechsel, D. LeClere, G. E. Thompson, J. M. Macak
and P. Schmuki, Adv. Mater., 2008, 20, 4135 [Links].
91 J. H. Lim and J. Choi, Small, 2007, 3, 1504 [Links].
92 K. Lee, D. Kim, P. Roy and P. Schmuki, unpublished work.
93 Y.-Y. Song, R. Lynch, D. Kim, P. Roy and P. Schmuki, Electrochem. Solid-State Lett., 2009, 12,
C17 [Links].
94 R. Hahn, J. M. Macak and P. Schmuki, Electrochem. Commun., 2007, 9, 947 [Links].
95 K. Nakayama in: 208th ECS Meeting, California, 2005, abstract 819 and abstract 843.
96 K. Ishibashi, R. Yamaguchi, Y. Kimura and M. Niwano, J. Electrochem. Soc., 2008, 155, K10
[Links].
97 R. Hahn, T. Stergiopoulus, J. M. Macak, D. Tsoukleris, A. G. Kontos, S. P. Albu, D. Kim, A. Ghicov,
J. Kunze, P. Falaras and P. Schmuki, Phys. Status Solidi RRL, 2007, 1, 135.
98 H. Tang, K. Prasad, R. Sanjinbs, P. E. Schmid and F. Levy, J. Appl. Phys., 1994, 75, 2042 [Links].
99 N.-G. Park, J. van de Lagemaat and A. J. Frank, J. Phys. Chem. B, 2000, 104, 8989 [Links].
100 S. P. Albu, H. Tsuchiya, S. Fujimoto and P. Schmuki, Chem. Phys. Lett., 2009 , to be submitted.
rsc.org/…/DisplayHTMLArticleforfree.cf… 33/35
10/20/2010 TiO2 nanotubes and their application in…
101 R. P. Vitiello, J. M. Macak, A. Ghicov, H. Tsuchiya, L. F. P. Dick and P. Schmuki, Electrochem.
Commun., 2006, 8, 544 [Links].
102 J. M. Macak, A. Ghicov, R. Hahn, H. Tsuchiya and P. Schmuki, J. Mater. Res., 2006, 21, 2824
[Links].
103 R. Hahn, A. Ghicov, J. Salonen, V.-P. Lehto and P. Schmuki, Nanotechnology, 2007, 18, 105604
[Links].
104 R. Hahn, F. Schmidt-Stein, J. Salonen, S. Thiemann, Y. Y. Song, J. Kunze, V.-L. Lehto and Patrik
Schmuki, Angew. Chem., Int. Ed., 2009, 48, 7236 [Links].
105 R. Beranek, H. Tsuchiya, T. Sugishima, J. M. Macak, L. Taveira, S. Fujimoto, H. Kisch and P.
Schmuki, Appl. Phys. Lett., 2005, 87, 243114 [Links].
106 W. W. Gärtner, Phys. Rev., 1959, 116, 84 [Links].
107 J. W. Schultze and M. M. Lohrengel, Electrochim. Acta, 2000, 45, 2499 [Links].
108 L. V. Taveira, A. A. Sagües, J. M. Macak and P. Schmuki, J. Electrochem. Soc., 2008, 155, C293
[Links].
109 H. Tsuchiya, J. M. Macak, A. Ghicov, A. S. Räder, L. Taveira and P. Schmuki, Corros. Sci., 2007,
49, 203 [Links].
110 U. Diebold, Surf. Sci. Rep., 2003, 48, 53 [Links].
111 A. Hagfeldt, U. Björksten and M. Gra1tzel, J. Phys. Chem., 1996, 100, 8045 [Links].
112 S. Södergren, A. Hagfeldt, J. Olsson and S.-E. Lindquist, J. Phys. Chem., 1994, 98, 5552 [Links].
113 R. Lynch, A. Ghicov and P. Schmuki, J. Electrochem. Soc., submitted.
114 J. M. Macak, H. Tsuchiya, A. Ghicov and P. Schmuki, Electrochem. Commun., 2005, 7, 1133
[Links].
115 T. Stergiopoulos, A. Ghicov, V. Likodimos, D. S. Tsoukleris, J. Kunze, P. Schmuki and P. Falaras,
Nanotechnology, 2008, 19, 235602 [Links].
116 K.-M. Lee, V. Suryanarayanan, J.-H. Huang, K. R. J. Thomas, J. T. Lin and K.-C. Ho, Electrochim.
Acta, 2009, 54, 4123 [Links].
117 G. K. Mor, K. Shankar, M. Paulose, O. K. Varghese and C. A. Grimes, Nano Lett., 2006, 6, 215
[Links].
118 J. H. Park, T.-W. Lee and M. G. Kang, Chem. Commun., 2008, 2867 [Links].
119 W. J. Lee, E. Ramasamy and D. Y. Lee, Sol. Energy Mater. Sol. Cells, 2009, 93, 1448 [Links].
120 S. Berger, A. Ghicov, Y.-C. Nah and P. Schmuki, Langmuir, 2009, 25, 4841 [Links].
121 B. Liu, E. S. Aydil and J. Am, J. Am. Chem. Soc., 2009, 131, 3985 [Links].
122 V. Yong, S.-T. Ho and R. P. H. Chang, Appl. Phys. Lett., 2008, 92, 143506 [Links].
123 J. Nelson, The Physics of Solar Cells, Imperial College Press, London, 2003, ch. 1.
124 L. Kavan, B. O'Regan, A. Kay and M. Grätzel, J. Electroanal. Chem., 1993, 346, 291 [Links].
125 M. K. Nazeeruddin, A. Kay, I. Rodicio, R. Humpbry-Baker, E. Miiller, P. Liska, N. Vlachopoulos and
M. Grätzel, J. Am. Chem. Soc., 1993, 115, 6382 [Links].
126 N.-G. Park, G. Schlichthrl, J. van de Lagemaat, H. M. Cheong, A. Mascarenhas and A. J. Frank, J.
Phys. Chem. B, 1999, 103, 3308 [Links].
127 S. Ito, P. Liska, P. Comte, R. Charvet, P. Pechy, U. Bach, L. Schmidt-Mende, S. M. Zakeeruddin, A.

rsc.org/…/DisplayHTMLArticleforfree.cf… 34/35
10/20/2010 TiO2 nanotubes and their application in…
Kay, M. K. Nazeeruddin and M. Grätzel, Chem. Commun., 2005, 4351 [Links].
128 B. C. O'Regan, J. R. Durrant, P. M. Sommeling and N. J. Bakker, J. Phys. Chem. C, 2007, 111,
14001 [Links].
129 P. Roy, D. Kim, I. Paramasivam and P. Schmuki, Electrochem. Commun., 2009, 11, 1001 [Links].
130 P. Roy, D. Kim, K. Lee, E. Spiecker and P. Schmuki, unpublished work.
131 S. Funk, B. Hokkanen, U. Burghaus, A. Ghicov and P. Schmuki, Nano Lett., 2007, 7, 1091 [Links].
132 D. Kim, J. M. Macak, F. Schimidt-Stein and P. Schmuki, Nanotechnology, 2008, 19, 305710
[Links].

This journal is © The Royal Society of Chemistry 2009

rsc.org/…/DisplayHTMLArticleforfree.cf… 35/35

También podría gustarte