Sei sulla pagina 1di 12

Electrochimica Acta 83 (2012) 113124

Contents lists available at SciVerse ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

(Photo)Electrochemical characterization of nanoporous TiO2 and Ce-doped TiO2


solgel lm electrodes
Sutasinee Kityakarn a , Yingyot Pooarporn b , Prayoon Songsiriritthigul b , Attera Worayingyong a ,
Simone Robl c , Andr M. Braun d , Michael Wrner e,
a
Department of Chemistry, Faculty of Science, Kasetsart University, Bangkok 10903, Thailand
b
Synchrotron Light Research Institute (Public Organization), Nakhon Ratchasima 30000, Thailand
c
Institute for Technical Thermodynamics and Refrigeration (ITTK) Karlsruhe Institute of Technology, Karlsruhe D-76131, Germany
d
Engler-Bunte-Institute, Karlsruhe Institute of Technology, Karlsruhe D-76131, Germany
e
Institute of Engineering in Life Sciences, Karlsruhe Institute of Technology, Karlsruhe D-76131, Germany

a r t i c l e i n f o a b s t r a c t

Article history: Focusing on effects of charge separation limitations and doping on the photoactivity of TiO2 , dif-
Received 19 January 2012 ferent nanoporous TiO2 lm electrodes were prepared by the solgel or the suspension methods,
Received in revised form 6 July 2012 respectively. X-ray diffraction (XRD) characterization revealed for all undoped TiO2 electrodes (TiO2 -
Accepted 8 July 2012
SG and TiO2 -P25) mixed phases of anatase and rutile, whereas the 2%CeTiO2 -SG electrode showed only
Available online 16 August 2012
anatase phase pattern, revealing that the incorporation of Ce ions prevented phase transformation. The
(photo)electrochemical characterization of the nanoporous TiO2 lm electrodes was performed by cyclic
Keywords:
voltammetry (CV) and electrochemical impedance spectroscopy (EIS) in the dark and under UV irradi-
Nanoporous TiO2 lm electrode
Ce-doped TiO2
ation, respectively. A photocurrent maximum depending on the lm thickness was observed at about
Photoelectrochemistry 13 m for all TiO2 electrodes investigated. The photocurrent responses of Ce-doped solgel TiO2 elec-
Electrochemical impedance spectroscopy trodes were comparably low, indicating an enhanced electronhole recombination at the Ce ion levels
(EIS) within the band gap of TiO2 . A porous electrode model was adapted for the tting of the experimen-
Porous electrode model tal EIS data. The potential and incident radiant power density dependence of the heterogeneous charge
transfer reaction (RCT ) and the coupled transport of photogenerated electrons to the collector electrode
(ZS ) were ascribed to promoted charge carrier separation and migration, i.e. electron drift dominated
over diffusion. Consequently, the dependence of different discrete impedance elements was discussed,
supposing a macroscopic electric eld across the 3D array of interconnected TiO2 nanoparticles that form
the lm electrodes. Besides photocurrent doubling, addition of methanol as a hole scavenger revealed
the limitation of the overall reaction by the heterogeneous charge transfer reaction, in particular at high
radiant power densities.
2012 Elsevier Ltd. All rights reserved.

1. Introduction contrast to the pure dispersions, the observed increased photo-


catalytic activity was assumed to be due to inter-particle contacts
Titanium dioxide (TiO2 ) is a promising material in photo- formed between anatase and rutile particles in water and, conse-
catalytic applications for environmental technologies, due to its quently, to band bending upon a Fermi level line-up. A potential
semiconductor properties and the generation of electronhole drop within the contact zone would prevent an electron ow from
pairs under UV irradiation. These electronhole pairs migrate or dif- anatase to rutile particles, whereas holes created in anatase may
fuse to the TiO2 surface, where they react with adsorbed molecules. migrate to rutile particles. In addition, for mixed-phase TiO2 -P25
A lot of effort was made to improve the photocatalytic perfor- particles, an intrinsic charge separation was proposed to explain
mance [14]. The photocatalytic properties of titania are sensitive the high photocatalytic activity in particulate dispersive aqueous
to their polymorphism and microstructure. Anatase is widely used systems [8].
as a photocatalyst and considered to be the most active form [5,6], Cerium and transition metals are added to increase the photo-
but high photocatalytic activity was also obtained using mixtures catalytic activity of TiO2 and to extend the absorption range into
of aqueous suspensions of anatase and rutile particles [3,7]. In the visible. The presence of such metal ions provides trap sites
for electrons and holes, in addition to those located at the parti-
cle surface (e.g. adsorbed O2 and OH ), that may lead to higher
Corresponding author. Tel.: +49 721 608 46235; fax: +49 721 608 46240. rates of charge separation and interfacial charge transfer reac-
E-mail address: michael.woerner@kit.edu (M. Wrner). tions. The photocatalytic activity of aqueous TiO2 suspensions for

0013-4686/$ see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.electacta.2012.07.129
114 S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124

mercaptobenzothiazole degradation was signicantly enhanced by diisopropoxide (Aldrich, technical grade) as precursor and cerium
incorporation of Ce(III) [9]. However, contradictory results were nitrate hexahydrate (Fluka, 99.0%) for doping (Table 1).
published, e.g. the doping with different metal ions leading to an The titanium alkoxide precursor solution (0.01 mol) was added
increased rate of recombination of electronhole pairs and to lower drop-wise into a stirred aqueous nitric acid (pH 1.0, 36.0 cm3 )
rates of substrate (4-nitrophenol) degradation [10]. at 296 1 K. The stirred suspension was heated to 358 K and
Photoelectrochemical (PEC) systems were also used to achieve kept at this temperature for 100 min. ITO-coated oat glass (PGO,
an increase of the photocatalytic efciency. Different methods Germany) with SiO2 passivation layers, 2 cm 3 cm (<10 /cm2 )
were used to immobilize TiO2 at conductive substrates in form was used as a support for TiO2 . This support was immersed in boil-
of nanoporous lms to permit large semiconductor surface areas ing acetone (348 K) for 10 min and afterwards in 30% (v/v) HNO3
[2,1113]. The effect of the applied potential was shown for for 30 min. The pre-treated ITO glass was stored in distilled water
degradation of 4-chlorophenol and revealed that depletion layer for 1 h before coating. An area of 2 cm 2 cm of the pretreated ITO
conditions were needed for high rates of photocatalytic degrada- glass was dipped into the sol for 30 s, withdrawn (within 10 s)
tion [2]. and kept at 473 K for 10 min. This procedure was repeated sev-
For the improvement of the performance of PEC systems based eral times to increase the thickness of the resulting TiO2 lm, the
on semiconducting TiO2 , a detailed mechanistic understanding is number of repetitions representing the number of layers. The TiO2 -
needed. Electrochemical impedance spectroscopy (EIS) provides coated substrate was calcinated at 823 K for 4 h in air at ambient
specic information about double layer properties, charge transfer pressure. Compared to TiO2 -P25 electrodes, a higher temperature
phenomena, charge carrier diffusion and charge generation- was required to obtain TiO2 -SG electrodes with the same anatase
recombination processes [14]. The results of EIS analyses are to rutile ratio as for P25 (75:25). For Ce-doping (2% CeTiO2 -SG),
usually discussed in terms of equivalent circuits allowing the eval- the corresponding amount of cerium was added to the nitric acid
uation of different physicochemical parameters of the reaction solution. The procedure was the same as described above for the
system [15]. preparation of TiO2 -SG electrodes.
A number of equivalent circuits for semiconductor/electrolyte
interfaces may be assembled by combinations of resistors and 2.1.2. Nanoparticulate TiO2 -P25 lm electrodes
capacitors (i.e. Debye elements) [16]. However, more recent stud- Vinodgopals method [2] to prepare nanoparticulate thin lm
ies show that it is necessary to employ constant phase elements electrodes was modied as follows: an aqueous suspension of
(CPE) under conditions of the observed frequency dispersion [17]. TiO2 -P25 (Degussa, 1.00 g in 50 cm3 water/methanol, 1/4, v/v) was
CPEs are associated with disorder, either in the electrode structure dispensed on pre-treated ITO glass. The solvent was evaporated at
or in the diffusion dynamics [18,19]. In addition, a transmis- 308 K for 2 h. The procedure was repeated to increase the number
sion line model (TLM) is introduced to describe the small signal of layers, and the electrode was nally calcinated at 723 K for 4 h in
ac-impedance of nanoporous TiO2 lm electrodes in aqueous elec- air at ambient pressure.
trolytes [20,21]. This model accounts for two transport channels, The electrical contact of all prepared electrodes was made by
in the pores (lled with electrolyte) and in the solid (intercon- conductive adhesive bonding of a silver wire over the whole width
nected/sintered TiO2 particles), as well as for the cross reaction of the uncovered upper part of the ITO glass substrate, and all con-
at the interface (interfacial charge transfer at the semiconduc- ductive parts of the electrodes were protected by insulating epoxy
tor/electrolyte boundary). glue, cutting edges included.
Under UV irradiation, several coupled processes such as carrier
generation and diffusion, recombination, charge transfer and mass 2.2. Characterization of TiO2 lm electrodes
transport in the solution have to be considered. Nevertheless, the
analysis of different simultaneous processes by EIS requires ade- 2.2.1. Structure and morphology
quately differing time constants. X-ray diffraction (XRD) analyses were performed with a Philips
The aim of this work was to study the (photo)electrochemical XPert diffractometer using Cu K -radiation with an anode cur-
properties of nanoporous TiO2 lms prepared on ITO (indium tin rent of 30 mA and an accelerating voltage of 40 kV with a scanning
oxide) coated glass substrates for the improvement of the photo- step of 0.02 . The diffraction patterns were indexed by comparison
catalytic performance of such electrodes. A specically developed with the JCPDS (Joint Committee on Powder Diffraction Standards)
solgel (SG) method was used to prepare nanoporous TiO2 -SG lm les number 83-2243, 78-1510 and 81-0792 for anatase, rutile and
electrodes with the same anatase to rutile ratio as the one found cerium oxide, respectively.
in TiO2 -P25 (75/25), providing the possibility to compare different The morphology of the electrodes was analyzed by eld
preparation methods and to evaluate intrinsic charge separation emission scanning electron microscopy (SEM Philips XL30). An
effects. Furthermore, the effect of Ce-doping on the anatase to accelerating voltage in the range of 1030 keV was used. The spec-
rutile transformation and on the photo-electrochemical proper- imens were prepared by Au sputtering to increase the conductivity
ties was studied. A model to describe the impedance behavior of of the samples.
nanoporous TiO2 electrodes (under UV irradiation) is presented in
this work. It provides the possibility to optimize the photocatalytic 2.2.2. (Photo)Electrochemical characterization
activity of TiO2 -electrodes, considering morphology modication The (photo)electrochemical experiments were carried out with
and doping aspects. three electrode technique using the prepared TiO2 lm elec-
trodes as working electrode, a platinum counter electrode (ag
of 1 cm 1 cm) and a Ag/AgCl (in aqueous KCl, 3.0 M) reference
2. Experimental electrode. All electrode potentials are given with respect to the
reference electrode. A rectangular large cuvette (Optical Glass,
2.1. Electrode preparation 40 mm 40 mm 40 mm, Hellma, Germany) equipped with a self-
made Teon top for positioning and xing the electrodes was used
2.1.1. Nanoporous TiO2 -SG lm electrodes (TiO2 -SG and as measuring cell. Cyclic voltammograms (CV, scan rate: 25 mV/s)
2%CeTiO2 -SG) and electrochemical impedance spectra (EIS) were recorded with
Nanoporous TiO2 electrodes were prepared with a solgel an IM6 Electrochemical Workstation (ZAHNER-Elektrik, Germany).
method [22] using titianium(IV)bis(ethyl acetoacetato) An aqueous sodium sulfate solution (Fluka, 99.0%, 0.50 M) was
S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124 115

Table 1
Sample codes and composition of different prepared TiO2 lm electrodes.

Sample code Source Preparation Ti (%w/w) Ce (%w/w)

TiO2 -P25 Degussa Suspension 100


TiO2 -SG Titanium(IV)bis(ethyl acetoacetato)diisopropoxide Solgel 100
2%CeTiO2 -SG Titanium(IV)bis(ethyl acetoacetato)diisopropoxide Solgel 97.7 2.3

used as electrolyte. It was degassed with Argon (5.0) before and 3.2. Electrochemical and photoelectrochemical characterizations
kept under an Argon atmosphere during experiments. Methanol
(Aldrich, 98%) was added to the electrolyte solution (2.5 M of 3.2.1. Cyclic voltammetry (CV)
methanol) to obtain more detailed information about the effect of 3.2.1.1. Experiments in the dark. The voltammetric responses of a
a hole acceptor. bare ITO substrate and of TiO2 -SG electrodes in 0.50 M aqueous
For EIS studies, a 5 mV ac-perturbation was applied within a fre- Na2 SO4 (pH 6.3) are presented in Fig. 3a. In a degassed electrolyte
quency range of 30 mHz to 100 kHz. For comparative experiments solution, the bare ITO electrode could be polarized in the poten-
with undoped and Ce-doped TiO2 -SG electrodes, a 250 W halogen tial range of 1.3 to 1.1 V, the result indicating only double layer
lamp was used as irradiation source (5025AF-S, Braun, Germany). charging.
Further experiments were carried out with a halogen light source For TiO2 -SG coated electrodes, a Faradaic current was observed
equipped with a ber light guide (cold light source, KL 2500 LCD, below 0.3 V (vs. Ag/AgCl) that may be attributed to the reduc-
Schott, Germany) permitting radiant power density variation. Inci- tion/oxidation of the Ti(IV)/Ti(III) sites of the TiO2 layer. The
dent radiant power densities, P0 [23], were determined with a reduction of Ti(IV) at the hydroxylated surface is in accordance with
calibrated ber optic spectroradiometer (USB2000+, Ocean Optics, Eq. (1) [26]:
USA) and a 400 nm short pass lter (Laser 2000 GmbH, Germany)
[Ti(OH)2 ]2+ + H+ + e [Ti(OH)]2+ + H2 O (1)
allowing to quantify the part of the emitted wavelength range that
will be absorbed by the TiO2 lm electrodes. The optical ber was [Ti(OH)2 ]2+
and [Ti(OH)]2+ represent
the Ti(IV) and Ti(III) sur-
placed at the same distance at which the TiO2 electrodes are being face species of TiO2 lms, respectively. The observed current was
irradiated. Therefore, the measured values of P0 are not corrected linearly dependent on the number of layers of nanoporous TiO2 -SG
for adsorption and reection losses in cuvette glass and electrolyte (Fig. 3a). Considering redox reaction occurring only at hydroxylated
solution. surface sites, it may be assumed that the inner surface area of the
porous TiO2 lm is proportional to the amount of deposited TiO2
3. Results and discussion and that the electrolyte solution may penetrate the whole pore
volume of the nanoporous TiO2 lm. To verify, the charge of the
3.1. Structural and morphological characterizations anodic branch was calculated by integration of the corresponding

The XRD patterns of ITO conducting glass, TiO2 -P25, TiO2 -SG
and 2%CeTiO2 -SG lm electrodes are presented in Fig. 1.
The XRD of the TiO2 -P25 lm electrode revealed characteris-
tic peaks of anatase at 2 25.33, 48.10 and 55.14 (JCPDS PDF No.
83-2243) and of rutile at 2 27.51, 36.16 and 54.46 (JCPDS PDF
No. 78-1510). The XRD pattern of the TiO2 -SG lm electrode exhib-
ited the same pattern and almost the same intensity distribution as
TiO2 -P25. The molar ratios of anatase and rutile were determined
by the method described by Spurr and Myers [24]. The rutile phase
of TiO2 -SG lm electrodes was determined to be 23%, in compar-
ison to 24% of that of the TiO2 -P25 lm electrode. It should be
mentioned that calcination of TiO2 -P25 at the same temperature as
for the preparation of TiO2 -SG electrodes (823 K) does not change
the anatase to rutile ratio. The XRD pattern of 2%CeTiO2 -SG lm
electrode was identied to be pure anatase (Fig. 1). The anatase to
rutile transformation requiring temperatures above 873 K for Ce-
doped TiO2 [25], anatase was stabilized by Ce ions at a calcination
temperature of 823 K.
Low resolution SEM images showed signicant differences
between the TiO2 lm electrodes investigated (Fig. 2ac). The SEM
image of the TiO2 -P25 electrode (Fig. 2a) showed a microstructured
lm of TiO2 -P25 at low resolution, but a network of agglomerated
particles, appearing as a nanoparticulate layer, became visible at
high resolution (inset Fig. 2a). The particle diameter was deter-
mined to be about 50 nm. The SEM images of undoped and doped
TiO2 -SG resembled compact lms at low resolution (Fig. 2b and
c), but at higher resolution, tightly packed nanoparticles, forming
nanoporous TiO2 lms were observed (see inset Fig. 2b and c). The
particle sizes at undoped and doped TiO2 -SG electrodes were deter-
mined to be 2030 and 1020 nm, respectively. SEM was also used
to determine the thickness of the lm to be 1215 m for 20 layers Fig. 1. Diffraction patterns of ITO conducting glass and different TiO2 lm electrodes
of TiO2 -SG and for 5 layers of TiO2 -P25, respectively. (TiO2 -P25, TiO2 -SG and 2%CeTiO2 -SG; A: anatase, R: rutile, I: indium oxide).
116 S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124

Fig. 2. SEM images of different TiO2 lm electrodes: (a) TiO2 -P25, (b) TiO2 -SG and (c) 2%CeTiO2 -SG.

i vs. t-plots, and thereby, the charge per mass of deposited TiO2 in IrO0.3 Ti(0.7x) Cex O2 led to a linear increase of the total charge.
was plotted vs. the number of layers (Fig. 3b). Only TiO2 -SG elec- The authors explained their ndings by an increased effective sur-
trodes showed the ideal behavior represented by a constant charge face area due to the decreased probability of homogeneous crystal
per mass (charge density), or in other terms, a linear increase growth. The replacement of TiO2 by CeO2 resulted in fact in the
of the total charge with the number of layers. For TiO2 -P25 and formation of a very ne crystalline structure [27,28]. However, the
2%CeTiO2 -SG, a decrease of the charge density was observed with calculated charge per mass of TiO2 -P25 lm electrodes, consist-
increasing lm thickness. This deviation from the ideal behavior ing of a network of relatively large particles, was (also) found to
may be explained by gradients of active surface sites and/or by be about 10 times higher than that of undoped TiO2 -SG electrodes
different lm morphologies. and is therefore comparable to the charge per mass of Ce-doped
The doping with Ce ions affected the charge per mass of the lm electrodes (Fig. 3b). These ndings might imply that TiO2 -
2%CeTiO2 -SG electrodes considerably. It was found to be about P25 lm electrodes exhibit a signicantly higher number of active
10 times higher than that of the TiO2 -SG electrodes (Fig. 3b). The (hydroxylated) sites per surface area than electrodes prepared by
result might be due to the larger total surface area of 2%CeTiO2 - the solgel method, as might be expected, when a lower calcination
SG as the particles were found to be smaller and forming a temperature is applied [29].
nanoporous lm electrode. The result is also in agreement with
Alves et al. [27] who observed that the replacement of TiO2 by CeO2 3.2.1.2. Experiments under irradiation. Irradiation of TiO2 lm elec-
trodes with energies exceeding the band gap of TiO2 generates
electronhole pairs. The holes diffuse or migrate to the TiO2 surface
and oxidize adsorbed water or surface hydroxide groups, while the
promoted electrons have to cross particle grain boundaries to be
withdrawn at the TiO2 network/ITO interface during anodic polar-
ization. The observed photocurrent depends on the thickness of the
TiO2 lm electrodes, as shown in Fig. 4 for TiO2 -SG electrodes under
chopped irradiation. A photocurrent maximum was observed for
different numbers of layers depending on the type of lm electrode:
20 layers for TiO2 -SG (12 m thick), 15 layers for 2%CeTiO2 -SG and
5 layers for TiO2 -P25 electrodes (15 m thick), respectively. The
CVs of the different TiO2 lm electrodes at maximum photocurrent
are presented in Fig. 5.
During irradiation, a fast hole transfer to the adsorbed accep-
tor and into the solution takes place, whereas the promoted
electrons move through a network of interconnected particles

Fig. 3. (a) Cyclic voltammograms of nanoporous TiO2 -SG lm electrodes in depen-


dence on the number of deposited layers (in aqueous 0.5 M Na2 SO4 supporting
electrode (pH 6.3)). (b) Anodic peak charge of nanoporous TiO2 electrodes in depen- Fig. 4. Anodic photocurrent in dependence on the number of deposited layers of
dence of the number of deposited layers of TiO2 -P25, TiO2 -SG and 2%CeTiO2 -SG. nanoporous TiO2 -SG lm electrodes under chopped irradiation.
S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124 117

Fig. 5. Anodic photocurrent (maximum) of different TiO2 lm electrodes under


chopped irradiation; 20 layers for TiO2 -SG, 15 layers for 2%CeTiO2 -SG and 5 layers
of TiO2 -P25.

to the (ITO) collector electrode, where they are withdrawn as


photocurrent. As long as the depth of penetration of the inci-
dent radiation exceeds the thickness of the TiO2 layer, the
photocurrent increases with the layer thickness due to the
higher number of photons absorbed. The diffusion length of the
promoted electrons being limited mainly by the charge sepa-
ration efciency and the structure of the nanoporous lm, the
probability of recombination processes increases with the lm
thickness related to electron transport across an increased num-
ber of particles and grain boundaries. At highest photocurrents,
the lm thickness (e.g. 20 layers of TiO2 -SG electrode) may
reach the diffusion length of the electrons and/or the distance
of penetration depth of the incident radiation. The decrease of
the photocurrent for thicker lms may therefore be explained
with a higher loss of promoted electrons by recombination and
by charge transfer processes into the solution during diffusion Fig. 6. Impedance spectra (Bode plots) of different nanoporous TiO2 -SG lm elec-
across the interconnected particles to the collector electrode trodes at a potential of 0.3 V (vs. Ag/AgCl ref. electrode) (a) in the dark, (b) under
(back contact). irradiation (symbols: experimental data; solid lines: tted data).
It should be noticed that for 2%CeTiO2 -SG electrodes, the
photocurrent was rather low and a pronounced delay in the pho- uniform cylindrical pores, but assuming that the porous TiO2 lm
tocurrent progression was observed (Fig. 5). Assuming that Ce4+ electrode is a random network of TiO2 nanoparticles, a simplied
interstitial sites would act as effective electron scavenger for model for the simulation procedure may be used [20,32,33]. The
photogenerated electrons of TiO2 [10], the delayed photocurrent total impedance of the porous layer is expressed in terms of differ-
response may be explained by a fast lling of these (deep) trap sites ent macroscopic impedance elements (partial equivalent circuits)
and a subsequent diffusion of electrons via the conduction band to (Fig. 7):
the back contact [30]. Nevertheless, the low photocurrent observed
for 2%CeTiO2 -SG lm electrodes might indicate an enhanced pho- Zq for the porous TiO2 layer/pore electrolyte interface.
tocurrent loss, and it may be assumed that the trapped electrons Zs for the porous TiO2 layer of interconnected TiO2 nanoparticles.
efciently recombine with the photogenerated holes at the Ce ion Zp for the pore electrolyte.
levels within the band gap of TiO2 . Zn the porous TiO2 layer/ITO layer interface.
Zo for the porous TiO2 surface/bulk electrolyte interface.
3.2.2. Electrochemical impedance spectroscopy (EIS)
3.2.2.1. Equivalent circuits for EIS analysis. Impedance spectra of For nanoporous TiO2 electrodes, the impedance of the outer
TiO2 -SG and 2%CeTiO2 -SG electrodes at 0.30 V in the dark and under porous TiO2 /bulk electrolyte interface (Zo ) is thought to be very
irradiation are presented as Bode plots (amplitude, |Z|, and phase, low compared to the impedance of the interior porous TiO2 /pore
||, vs. frequency) in Fig. 6a and b, respectively. electrolyte interface (Zq ) due to the huge size difference of the
For experiments in the dark, the phase angle of both electrodes interfacial areas and can be neglected.
reaches almost 90 in the mid frequency range and remains With this simplication, best tting results of EIS data (recorded
almost constant at low frequencies (Fig. 6a). The ndings indi- in the dark) were obtained by using the equivalent circuit pre-
cate a blocking (fully capacitive) behavior of the nanoporous sented in Fig. 7c. It contains partial schemes for the interfacial
TiO2 electrolyte interface. For the analysis of the EIS results, a impedances Zn and Zq , respectively. Zn comprises two RC circuits
porous electrode model, corresponding to a transmission line in serial connection, the rst representing the conductive ITO bulk
with two transport channels and a crosswise interfacial reac- layer [element 1 (CPE) and 2 (RITO )], the second corresponds to the
tion, was employed. The homogeneous porous electrode model ITO/electrolyte interface at the bottom of the pores [element 3 (C3 )
proposed by Ghr [18,31] is implemented in the simulation and and 4 (R4 )], taking into account the surface oxidation of the ITO
evaluation program of the used software package (Thales, ZAHNER- layer by thermal treatment [22]. The impedance of the (porous)
Elektrik, Germany). This model is based on a porous electrode with TiO2 /pore electrolyte interface (Zq ) is described by a capacitance
118 S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124

Fig. 7. Schematic and equivalent circuits of nanoporous TiO2 lm electrodes. (a) Array of TiO2 nanoparticles and corresponding distributed impedance elements according
to the porous electrode model [18,31]. (b) Impedance of the porous electrode in terms of macroscopic impedance elements (Zs , Zp , Zq , Zo and Zn , as described in the main
text). (c) Equivalent circuit applied for TiO2 electrodes under dark conditions. (d) Equivalent circuit for UV irradiated TiO2 electrodes.

only (element 5, CSC ). Element 6 represents the network of the Under UV irradiation, the generated photocurrent leads to a sig-
differential (distributed) impedances Zpi and Zsi (i = 1, 2, 3, . . ., n, nicant decrease of the total impedance of the porous electrode in
Fig. 7a). The simulation procedure allows determination of the the frequency range below 10 Hz. That was also accompanied by
integral pore electrolyte resistance (Zp ) and the integral solid bulk a drop of the phase angle (Fig. 6b). Both effects were more pro-
resistance, Zs (i.e. resistance of the porous TiO2 lm consisting of nounced for the undoped TiO2 -SG than for the doped 2%CeTiO2 -SG
the interconnected TiO2 particles). Finally, the impedance of the electrode. The higher photoconductivity of the TiO2 -SG electrode
bulk electrolyte is included as a pure ohmic resistance, element 7 hence is indicated by smaller impedance and phase angle at low
(R ). frequencies. These ndings are in accordance with the results of
The results of the tting procedure based on the proposed equiv- CV measurements that revealed a two times higher photocurrent
alent circuit are listed in Table 2. The resulting impedance values for the undoped material. The EIS results were simulated using
for Zn (elements 14) are different for doped and undoped TiO2 - a modied equivalent circuit, where charge carrier transfer and
SG electrodes. This may be explained by the different interfacial recombination processes were taken into account (Fig. 7d). The par-
structures, assuming different particle sizes and particle surface tial equivalent circuit for Zn was adjusted, because the impedance
structures for undoped and doped materials. In addition, ITO sur- element 4 had to be eliminated, i.e. the ITO/pore ground interface
face oxidation during the calcination procedure may be affected by was xed to be insulating (innite resistance). In fact, element
different colloidal structures of undoped and doped gels. 4 could no longer be tted after implementation of additional
Equal amounts of TiO2 being deposited on both electrodes, the elements for Zq . Consequently and in consistence with the the-
higher interfacial capacitance of the pore electrolyte/TiO2 interface ory, a charge transfer reaction at this interface was excluded in
(CSC ) for the 2%CeTiO2 -SG may be related to the Ce-doping and/or the investigated potential range, and the observed current was
to a larger surface area. An increased surface area for the Ce-doped assigned exclusively to the photogenerated electrons crossing the
material would be in accord with the results obtained from the CV TiO2 particle/ITO interface. The interfacial impedance element Zq
experiments (3.2.1). The impedance of the 2%CeTiO2 -SG electrode was modied by addition of elements modeling the photo-induced
is two times higher than that of the undoped material (impedance reactions: (i) an ohmic resistance, element 8 (RCT ) for the heteroge-
element ZS , Table 2). The argument seems to be contradictory, neous charge transfer reaction at the porous TiO2 layer/electrolyte
because the doping of the material should lead to a higher con- interface, (ii) a capacitive element 10 (CIF ) for the accumulation
ductivity and therefore to a smaller overall impedance. However, of photoinduced charge carriers at the semiconductor surface, in
considering that ZS is attributed to a porous TiO2 layer consisting parallel to the heterogeneous charge transfer reaction resistance
of an assembly of interconnected TiO2 nanoparticles, the effective and (iii) an ohmic resistance, element 9 (RREC ) for the charge car-
bulk conductivity of the porous lm depends on the intrinsic con- rier recombination. Under consideration of different models and
ductivity of the TiO2 particles, the geometry of the contact zone and the corresponding theoretical (physicochemical) background, sev-
the packing structure. In addition, the impedance of the sintered eral attempts to lower the number of distinct impedance elements
semiconductor particle network may depend primarily on discon- were made. However, the equivalent circuit presented here was so
tinuities at the grain boundaries. However, the different particle far the only one to yield good tting results, especially for studies
surface structures of undoped and doped TiO2 -SG lm electrodes under electrode potential and radiant power density variation, be
imply that ZS depends mainly on the structure of the particle sinter aware, that a good t does not, in itself, validate the model used.
bridges. However, the use of a simplied partial equivalent circuit for Zq ,
S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124 119

Table 2
Fitting results of different impedance elements for various TiO2 lm electrodes in the dark and under irradiation at an applied potential of 0.3 V (vs. Ag/AgCl ref. electrode).

Dark UV irradiation

Electrode (steady state photocurrent)/impedance element TiO2 -SG (38.7 nA) 2%Ce TiO2 -SG (51.1 nA) TiO2 -SG (51.0 A) 2%Ce TiO2 -SG (39.5 A)

1. CPE (uF) 14.7 11.7 14.5 12.3


0.70 0.74 0.76 0.76
2. RITO ( ) 11.1 16.9 9.5 17.2
3. C3 (F) 30.5 28.1 30.6 31.9
4. R4 (M ) 3.9 2.7
5. CSC (F) 2.4 3.3 45.0 30.4
6. Porous impedance
Zp (Ohm) 8.9 7.6 8.8 8.2
Zs (k ) 240.6 524.5 33.8 146.4
7. R (Ohm) 16.4 14.0 16.5 15.3
8. RCT (k ) 2.1 100.8
9. RREC (k ) 1.2 8.2
10. CIF (mF) 1.58 0.49

often applied for mesoporous semiconductor electrodes in form


of a so called chemical capacitance parallel to a resistance (which
accounts for charge transfer or recombination) [34], yielded simu-
lation results of low signicance and failed. Finally, Zq resulted in an
equivalent circuit representative for coupled electrochemical reac-
tions [35], taking into account, that heterogeneous charge transfer
and recombination may take place at different interfacial electronic
states. The associated capacities are attributed to (i) charge accu-
mulation at the sc surface, CIF (associated to surface trapped holes),
and to (ii) the capacitance of the semiconductor matrix consider-
ing a space charge layer (CSC ). An alternative approach for Zq would
be two parallel RC combinations in serial alignment, regarding to
identical electronic states for charge transfer and recombination
[36].
When comparing the modeling results of electrodes in the dark
and under irradiation, elements for the interface ITO/pore ground
(Zn , including elements 13), pore electrolyte resistance (Zp ) and
electrolyte resistance (element 7, R ) remained constant (Table 2).
Fig. 8. Impedance spectra (Bode plots) of a TiO2 -SG electrode (20 layers) in depen-
As expected, these elements are not affected by UV irradiation, i.e.
dence of the applied electrode potential (vs. Ag/AgCl ref. electrode in 0.5 M Na2 SO4
photoinduced processes. It has to be noticed that the impedances aqueous solution; symbols: experimental data; solid lines: tted data).
RCT , RREC and ZS were found to be signicantly smaller for the
TiO2 -SG electrodes, due to their higher photoactivity. Taking into
account that the same mass of TiO2 was deposited for TiO2 -SG lm electrodes was investigated in the potential range from 0.0
and 2%CeTiO2 -SG electrodes, it may be concluded that the TiO2 to 0.5 V vs. reference electrode, at a xed incident radiant power
nanoparticle network of the undoped lm electrode exhibits a density of 5.2 W/m2 . The results are presented in Fig. 8 and demon-
higher rate constant for the heterogeneous charge transfer reaction, strate the effect of the electrode potential on total impedance and
a smaller recombination rate and a higher conductivity. The most phase angle in the frequency range below 1 Hz.
pronounced difference was found for the charge transfer resistance, Modeling results, presented in Fig. 9, were obtained using the
and it seems that the overall reaction for the doped TiO2 mate- equivalent circuit already described (Fig. 7d). If the anodic polariza-
rial was limited by the rates of the hole reactions at the surface. tion was increased, the heterogeneous charge transfer resistance
However, the charge transfer reaction is linked to the recombina- (RCT ) decreased as well as the resistive element for the recombi-
tion process and, therefore, cannot be seen independently. A more nation (RREC ) and the interconnected TiO2 particle network (ZS ).
detailed discussion of the EIS results would require yet unavailable The observed strong potential dependence of the impedance of
data concerning the inner surface area, the size and the intercon- TiO2 -SG electrode shows rather the characteristics of a compact
nection structure of the particles. solid semiconductor lm electrode under depletion conditions than
of a nanoporous electrode penetrated by the electrolyte solution,
for which only a small effect is expected [37]. The characteristics
3.2.2.2. Photoelectrochemical characterization of nanoporous TiO2 - of a classical electrolytesemiconductor junction under irradi-
SG electrodes by EIS. TiO2 -SG electrodes yielding high photocur- ation can be described by theoretical derivations, as established
rents in CV experiments were prepared to study the potential by Grtner for semiconductor pn junctions or Schottky contacts
dependence, the effects of the incident radiant power density and [3840]. During anodic polarization of solid thin lm semicon-
a hole acceptor/scavenger (addition of methanol), employing EIS. ductors, a space charge layer is formed that leads to an efcient
The irradiation system used for CV experiments was replaced by a separation of the photogenerated charge carrier in the electric eld
halogen lamp coupled to a ber optics to avoid emitting perturbing within the depletion zone. The photo-generated holes migrate to
noise in the low frequency range of EIS measurements. In addition, the TiO2 /electrolyte interface where they can be withdrawn by a
by this measure the emitting power of the installation could be heterogeneous charge transfer reaction (redox reaction), while the
varied and heating of the electrolyte solution could be prevented. remaining electrons are directed to the collector electrode produc-
3.2.2.2.1. Potential dependence of the impedance (under UV irra- ing a photocurrent. When the thickness of the TiO2 lm electrode
diation). The potential dependence of the impedance of TiO2 -SG is larger than the width of the depletion layer (space charge layer),
120 S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124

(ITO layer surface) [37,43]. Alternatively, the observed potential


dependence might be due to an electric eld induced by photogen-
erated charges. Assuming very fast hole removal reactions, negative
charges may shift the band edges to higher energy, due to remaining
electrons trapped in interfacial band-gap states. Taking the spatial
absorption characteristic into account, such a shift of band edges
would result in an electric eld component driving electrons to
the collector electrode, when the nanoparticulate TiO2 lm would
be irradiated from the electrolyte (front) side. Nevertheless, it has
been published that light-induced electric eld effects should not
alter the electron-transport characteristics in a fundamental way
[44]. Both effects, potential drop next to the collector electrode
and light-induced electric eld, should have a strong impact on the
photo-response of the nanoparticulate TiO2 lm, when irradiation
is changed from electrolyte (front) to back side. The experimental
results of this work with nanoparticulate TiO2 -SG lm electrodes
(20 layers) revealed only small changes for all tted impedance ele-
ments (mean error 1.0 0.5%) and cannot support these alternative
approaches [37,43,44].
Compared to dark conditions, the impedance of the particle
network (ZS ) dropped by a factor of about 8 during UV irradia-
tion. This can be explained by a high (excess) concentration of
electrons in the transport state (conduction band) due to exci-
tation of the porous TiO2 electrode in combination with the fast
removal of the holes. In contrast, the observed potential depen-
dence of ZS in the dark was small, and varied only by a factor
of 1.5 in the potential range between 0.0 and 0.5 V vs. ref. elec-
trode. Only biasing TiO2 SG electrodes in the cathodic potential
Fig. 9. Fitting results of different impedance elements in dependence of the applied region towards the conduction band edge potential induced a pro-
potential for a nanoporous TiO2 -SG lm electrode (20 layers) in pure electrolyte nounced drop of ZS as predicted by the theory, which based on
solution (aqueous 0.5 M Na2 SO4 ) under UV irradiation.
an exponential increase of the density of states (DOS) approach-
ing the conduction band [21,45]. Focusing on the photocatalytic
an efcient electron/hole recombination during the diffusion of performance of nanoporous electrodes, we excluded the cathodic
the photogenerated electrons to the back contact must be con- potential region (vs. ref) for the experiments of UV irradiated elec-
sidered. Under these conditions and for the given incident radiant trodes. Hence, under idealized conditions, ZS would only depend on
power density, the photocurrent can be increased with increasing the incident radiant power density, considering the low conductiv-
anodic polarization yielding a thicker depletion layer, where ef- ity (DOS) und the small potential dependency for the dark situation,
cient charge separation may take place. However, in the present respectively. Therefore, the observed potential dependence (see
study, the electrode is composed by very small TiO2 particles Fig. 9a) of the impedance of the particle network (ZS ), i.e. the resis-
forming a nanoporous layer that is penetrated by the electrolyte tive channel of electron transport to the collector electrode, was
solution. Assuming individual nanoparticles, the width of the space unanticipated. It seems that the applied equivalent circuit is over-
charge layer cannot exceed the particle radius, and for maximum simplied with respect to RREC , implemented as an element of the
depletion conditions, the voltage difference between the surface interfacial impedance Zq only. Charge carrier recombination may
and the center of the nanosized TiO2 particles cannot exceed several also take place in the bulk material (nanoparticle array) and at grain
tenths of millivolts [41]. boundaries. In fact, lower charge carrier recombination rates at the
The results obtained with TiO2 -SG electrodes support the grain boundaries would affect the impedance of the nanoparticle
hypothesis of the formation of a macroscopic electric eld between network (ZS ) resulting in the observed decrease of ZS with increas-
the ITO collector electrode and the outer part of the nanoporous ing anodic polarization. This interdependence might be ascribed to
TiO2 lm. It seems that for these nanoporous lm electrodes, an an extended diffusion length of photo-generated electrons.
apparent space charge layer becomes operative. With increas- The partial equivalent circuit for Zq comprises two capacitive
ing anodic polarization (i.e. extended macroscopic electric eld), elements, CIF and CSC (see Fig. 7d). The twofold increase of CIF with
migration of photogenerated electrons to the electrode contact and increasing anodic polarization implies that a higher number of sur-
the redox reactions of holes at the semiconductor/electrolyte inter- face trapped holes cannot be removed by heterogeneous charge
face are enhanced, and a decrease of the charge transfer resistance transfer reactions leading to an accumulation of photogenerated
(RCT ) and of the impedance element for the recombination reaction charge at the semiconductor surface. This seems to contradict the
(RREC ) is expected. In fact, RCT decreased steadily with increasing enhanced charge transfer conditions when increasing the anodic
anodic potential (Fig. 9a). The drop of RREC was less pronounced, bias, but the steady state photocurrent increased by a factor of
most probably due to the characteristics of the applied equivalent 5 simultaneously. CSC also depends on the applied potential and
circuit, as discussed below. decreases steadily with increased anodic biasing. A MottSchottky
A macroscopic electric eld across a semiconductor nanopar- plot yielded a linear dependence of C2 vs. electrode potential, see
ticulate network was also suggested by Hagfeldt et al. to explain Fig. 10. The at band potential under irradiation was determined to
the high photocurrent quantum efciencies of colloidal TiO2 lm be 0.22 V vs. ref. electrode. The results are consistent with respect
electrodes [42]. In contrast to this picture, models of the poten- to the theoretical prediction for a wide band gab bulk semicon-
tial distribution in semiconducting porous electrodes, penetrated ductor. The Schottky behavior of CSC underlines that an apparent
by an electrolyte solution, showed a potential drop within a spa- space charge layer could be responsible for the enhanced charge
tial resolution of few particle layers next to the collector electrode separation efciency by increasing the anodic bias, indicated by
S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124 121

Fig. 10. MottSchottky plot of CSC for a nanoporous TiO2 -SG lm electrode (20
layers) in pure electrolyte solution (aqueous 0.5 M Na2 SO4 ) under UV irradiation.

the decrease of RCT and ZS , and the increase of the steady state
photocurrent by a factor of 5 in the investigated potential range. It
seems that CSC behaves not as a chemical capacitance, often used to
describe the (exponential) potential dependence of charge accumu-
lation in mesoporous semiconductors [34]. Nevertheless, it should
be taken into account that a switch between different mechanism in
dependence of the applied potential may occur, and in addition that
the generation of charge carriers (electron and holes) by excitation
of the semiconductor material itself may require a more complex
model than established for dye sensitized solar cells [21,46].
3.2.2.2.2. Effect of incident radiant power density (P0 ). Photoin-
duced processes at (nanoporous) TiO2 semiconductor electrodes
depend on various parameters such as morphology and absorp-
tion characteristics of the lm, the thickness of space charge layer
with respect to the lm dimension and the spectral emission of
the irradiation source. Investigating the rate determining process,
as e.g. limitations by the heterogeneous charge transfer or by the
recombination process, the incident radiant power density (P0 ) was
varied.
Bode plots of TiO2 -SG electrodes for different incident radiant
power densities applying an electrode potential of 0.3 V are pre-
sented in Fig. 11. Obviously, the variation of the radiant power
density affected only the impedance below 1 Hz. With increas-
ing P0 , the decrease of the overall impedance was accompanied
by a pronounced phase angle bending. Impedance spectra taken
at different incident radiant power densities were simulated with
Fig. 12. Fitting results of different impedance elements and steady state photocur-
the same high accuracy employing the outlined equivalent circuit rent in dependence of the incident radiant power density (P0 ) for a TiO2 -SG electrode
(Fig. 7d). Elements, supposed not to be affected by photo-induced (20 layers) in pure electrolyte solution (aqueous 0.5 M Na2 SO4 ) and after addition
processes, remain almost constant: (i) Zn (elements 13), (ii) Zp of methanol (2.5 M) at an applied potential of 0.3 V (vs. Ag/AgCl ref. electrode).

(pore electrolyte resistance) and (iii) element 7, R (bulk electrolyte


resistance). Impedance elements which showed a corresponding
dependence of the incident radiant power density are presented in
Fig. 12.
The observed photocurrent is due to the coupling of the inter-
facial charge transfer (RCT ) and the resistive electron transport
channel (ZS ). Under idealized conditions, the photoconductivity
of TiO2 -SG lm electrodes is expected to be proportional to the
rate of absorbed photons. Assuming a constant absorption coef-
cient across the TiO2 lm and, a xed correlation between the
incident radiant power density and the rate of absorbed pho-
tons, RCT and ZS should show a linear reciprocal dependence on
the incident radiant power density, if the efciency of charge
separation and the electron diffusion length are not affected by
irradiation. In fact, the charge transfer resistance RCT (element 8)
decreased with increasing incident radiant power density up to
3.0 W/m2 (58% of maximum power). At higher values of P0 , RCT
Fig. 11. Impedance spectra (Bode plots) of a TiO2 -SG electrode (20 layers) in depen-
dence of the incident radiant power density at 0.3 V vs. Ag/AgCl ref. electrode in 0.5 M remained constant within the experimental error of 5%, implying
Na2 SO4 aqueous solution (symbols: experimental data; solid lines: tted data). that the heterogeneous charge transfer reaction starts to limit the
122 S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124

photocurrent (overall reaction). When the electrode is irradiated on radical injecting subsequently an electron into the conduction band
the front side (semiconductor/electrolyte interface), the penetra- of TiO2 (Eq. (4)):
tion depth of the photons, i.e. the space, where the charge carriers +
are produced, has to be related to the width and the position of CH3 OH + h+ CH3 OH (2)
the supposed space charge layer, where an efcient charge sepa-
+
ration can take place. The experimental results may be interpreted CH3 OH CH3 O + H+ (3)
in a way that the penetration depth reaches the thickness of the
apparent space charge layer at an incident radiant power density CH3 O CH2 O + H+ + e (4)
of about 3.0 W/m2 . Based on these idealized assumptions, at low
To determine the impedance elements, the EIS data were pro-
radiant power densities all charge carriers will be produced within
cessed as before. The addition of methanol affected primarily the
the zone of efcient charge separation. Through this, the number
heterogeneous charge transfer resistance (RCT ) that dropped by a
of charge carriers would depend linearly on the incident radiant
factor of about 2.2. This effect is more pronounced at high radiant
power density.
power densities (Fig. 12a). Expectedly, heterogeneous charge trans-
Assuming that holes can be removed quantitatively by surface
fer reactions are, in the case of nanoporous TiO2 -SG lm electrode
redox reactions, RCT should drop reciprocally with the incident
in indifferent aqueous electrolyte solution, a key limiting factor of
radiant power density. For values 3.0 W/m2 the experimental
the efciency of the overall reaction.
results corroborate with the assumed interdependence. Never-
As the hole reaction rate at the semiconductor surface is
theless, this hypothesis does not imply the effect of charge
increased, the coupled charge recombination rate is expected to
recombination reactions (RREC ) that may also depend on P0 . In
drop. In fact, RREC decreased by a factor of about 1.8 due to the
fact, the impedance element for recombination (RREC ) decreases
addition of methanol.
with increasing incident radiant power density, showing a lin-
For the resistive pathway of electrons to the electrode back
ear reciprocal dependence (Fig. 12a). The observed dependence
contact (impedance of the particle array, ZS ), an almost linear radi-
suggests that a certain percentage of photo-generated charge car-
ant power dependence was found. The addition of methanol led
riers are lost by recombination. Nevertheless, this nding has
to experimental results from which a better reciprocal correlation
to be related to the implementation of RREC , see discussion
between ZS and the incident radiant power density (idealized con-
below.
ditions) could be obtained. Compared to the results in the pure
A reciprocal dependence on the incident radiant power density
electrolyte solution, the most pronounced effect of methanol addi-
(for idealized conditions) is also observed for ZS , again limited to
tion was found at high radiant power densities (Fig. 12b).
values 3.0 W/m2 . At higher P0 , the tting procedure resulted in
The evolution of the ZS1 vs. P0 plot for pure electrolyte solutions
higher values of ZS , as displayed by the drop of the slope in the
was explained by a higher loss of electrons due to an enhanced
plot of the reciprocal resistance of the particle network (ZS1 ) vs.
charge carrier recombination under irradiation, where the penetra-
the incident radiant power density (P0 ) (Fig. 12b). Interestingly,
tion depth of light exceeds the width of the (supposed) macroscopic
the deviation of ZS from the idealized behavior was observed at
electric eld. Under such conditions, it may be assumed that an
the same incident radiant power density, where RCT became con-
improved photocurrent is mainly due to a very efcient removal
stant (>3.0 W/m2 ), implying that the same process is decisive for
of holes by addition of methanol, independent of where the charge
the observed dependence on the incident radiant power density. In
carriers are generated, and the effect can be dragged down to the
accord with the discussion concerning RCT , at moderate P0 the pen-
level of individual nanoparticles.
etration depth of the irradiation does not exceed the width of the
Besides the current doubling effect of methanol, the results
supposed space charge layer. Under these considerations, when the
support the hypothesis that the photoelectrochemical response
charge carriers are generated within an electric eld, the expected
of TiO2 -SG electrodes may be increased by hole scavenging. For
reciprocal dependency can be observed. Then again, when at high
environmental applications, e.g. waste water remediation, a strong
P0 the irradiated volume exceeds the zone of supported charge
dependence of the process efciency on chemical nature and
carrier separation, a higher rate of recombination is expected. A
adsorption characteristics of substrates and intermediates is to be
reduced electron diffusion length due to enhanced charge car-
expected.
rier recombination in the bulk and at the grain boundaries would
affect ZS negatively. As a consequence, this suppressed incident
radiant power density dependency would result in a decreased 4. Conclusion
slope in the ZS1 vs. P0 plot at high power densities as displayed
in Fig. 12b. Nanoporous TiO2 -SG and 2%CeTiO2 -SG lm electrodes were
3.2.2.2.3. Effect of methanol addition. The experiments aimed prepared by solgel method, and TiO2 -P25 electrodes were pre-
at possible limitations of the overall reaction (photocurrent) pared by suspension method. XRD characterization revealed for all
by heterogeneous charge transfer processes under conditions undoped TiO2 electrodes (TiO2 -SG and TiO2 -P25) a phase mixture
of high incident radiant power densities. Further, the pos- of anatase and rutile, whereas 2%CeTiO2 -SG electrodes exhibited
sible effect of a hole-scavenger was studied. The addition only anatase phase pattern. The result indicates that the incor-
of e.g. citrate, formic acid or methanol to the supporting poration of Ce ions prevented the transformation from anatase to
electrolyte solution accelerates the hole reaction at the semi- rutile.
conductor surface, leading to a faster removal of the holes The (photo)electrochemical properties of nanoporous TiO2 lm
and, therefore, to a more efcient charge separation process electrodes were studied by CV and EIS. The Faradaic current under
[13]. cathodic polarization in the dark increased with the number of TiO2
In this study, the steady state photocurrent increased by a fac- layers, and the result may be linked to the reduction of hydrated
tor of about two, when 2.5 M methanol in 0.5 M Na2 SO4 aqueous Ti(IV) surface sites within the lm electrodes that are penetrated
solution was used (Fig. 12c). This result may be explained by an by the electrolyte solution. The doped 2%CeTiO2 electrodes showed
effect introduced as current doubling [47,48] that was not a topic higher dark currents per deposited mass compared to TiO2 -SG elec-
of our experimental studies. Current doubling was described to trodes, a result generally explained by a higher interior surface area
a reaction sequence, where methanol is oxidized by a hole (Eq. due a decreased probability of homogeneous crystal growth, caused
(2)) (or by a surface hydroxyl radial), the intermediate C-centered by Ce addition.
S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124 123

For UV irradiated electrodes, a photocurrent maximum was is a feature of individual nanoparticles. The high photocurrent of
observed for lm electrodes consisting of 20 layers of TiO2 -SG, 15 relatively thick lm electrodes at moderate radiant power densi-
layers of 2%CeTiO2 -SG and 5 layers of TiO2 -P25, respectively. The ties would then be the result of nanoparticles surrounded by the
dependence of the photocurrent on the number of deposited layers electrolyte solution which favors a fast hole removal by surface
may be a consequence of a combination of the penetration depth reactions. Alternatively, the observed dependence of the heteroge-
of light and the free diffusion length of electrons in relation to the neous charge transfer reaction (RCT ) on the electrode potential, and
lm thickness. the radiant power density dependence of RCT and the ZS (impedance
In the presence of Ce ions in titania, the observed photocurrent of the TiO2 particle network) could be explained by a 3D array
was comparably small with a delayed signal response. The addition of nanoparticles that behave as a collective like a solid crystalline
of Ce ions may lead to an enhanced charge carrier trapping at the Ce electrode.
levels within the band gap of TiO2 , resulting in an unwanted higher Future work will deal with the validation of the hypotheses by
electronhole recombination rate. EIS characterization of TiO2 -SG electrodes with different lm thick-
Comparing irradiated undoped and doped TiO2 -SG lm elec- ness, investigating potential dependence, direction and power of
trodes, modeling results revealed a higher rate constant for the incident irradiation, as well as hole scavenging reactions. Focusing
heterogeneous charge transfer, a smaller recombination rate and at the possible effect of intrinsic charge separation on the photoac-
a higher conductivity for undoped TiO2 -SG lm electrodes. In tivity of TiO2 , nanoporous lm electrodes with different anatase
addition, the potential dependence, the effect of the incident radi- to rutile ratios will be prepared by solgel method. Furthermore, a
ant power density and methanol addition on the photocurrent new photoelectrochemical set-up featuring monochromatic irradi-
response of TiO2 -SG electrodes (20 layers) were studied. The poten- ation with tunable output will enable investigations on the spectral
tial dependence led to the hypothesis, that a macroscopic electric response of the semiconductor electrodes and the determination of
eld across the nanoporous lm electrode is required to explain photocurrent quantum yields.
the tting results concerning the heterogeneous charge transfer
(RCT ) and electron transportation to the electrode back contact (ZS )
Acknowledgments
which favors drift over diffusion. Such a hypothesis would ask for a
macroscopic space charge layer promoting the drift of electrons to
The authors acknowledge gratefully support from Synchrotron
the back contact that would not be restricted to individual particles
Light Research Institute (Public Organization), Thailand (Grant
or few particle layers next to the contact (ITO collector electrode).
Number: 2548/05) and the former Chair of Umweltmesstechnik,
To receive more information about the rate determining pro-
University of Karlsruhe, now KIT (Karlsruhe Institute of Technol-
cess, the incident radiant power density (P0 ) was varied. Under
ogy), Germany. Further, we acknowledge support from Deutsche
idealized conditions, it may be expected that the photocurrent
Forschungsgemeinschaft (DFG).
density of TiO2 -SG lm electrodes is proportional to the rate of
absorbed photons. In fact, a reciprocal dependence of ZS , the resis-
tive pathway of electron motion to the back contact, on moderate References
values of radiant power densities were determined. At high radiant
power density, the tting procedure yielded higher values for ZS , [1] M.R. Hoffmann, S.T. Martin, W. Choi, D.W. Bahnemann, Chemical Reviews 95
(1995) 69.
explained by enhanced recombination losses of excited electrons [2] K. Vinodgopal, S. Hotchandani, P.V. Kamat, Journal of Physical Chemistry 97
when the penetration depth of light exceeds the zone of promoted (1993) 9040.
charge separation as a consequence of implementation of RREC as [3] B. Sun, P.G. Smirniotis, Catalysis Today 88 (2003) 49.
[4] U.I. Gaya, A.H. Abdullah, Journal of Photochemistry and Photobiology C 9 (2008)
surface related impedance element only. 1.
The same dependence on the incident radiant power density [5] S.-I. Nishimoto, B. Ohtani, H. Kajiwara, T. Kagiya, Journal of the Chemical Society,
was obtained for the heterogeneous charge transfer reaction (RCT ). Faraday Transactions 81 (1985) 61.
[6] Z. Ding, G.Q. Lu, P.F. Greeneld, Journal of Physical Chemistry 104 (2000) 4815.
For P0 of up to 3.0 W/m2 , a drop of RCT with increasing values of
[7] T. Ohno, K. Sarukawa, K. Tokieda, M. Matsumura, Journal of Catalysis 203 (2001)
P0 implies that holes could be removed by surface redox reaction 82.
in an efcient way. At high radiant power densities, the heteroge- [8] B. Sun, A.V. Vorontsov, P.G. Smirniotis, Langmiur 19 (2003) 3151.
[9] F.B. Li, X.Z. Li, M.F. Hou, K.W. Cheah, W.C.H. Choy, Applied Catalysis A 285 (2005)
neous charge transfer reaction rate leveled off. This was ascribed
181.
to conditions, where the electronhole recombination affected [10] K. Nagaveni, M.S. Hegde, G. Madras, Journal of Physical Chemistry B 108 (2004)
the photocurrent response notably. Again, when at high P0 the 20204.
irradiated volume exceeds the zone of supported charge carrier [11] A. Fujishima, K. Honda, Nature 238 (1972) 37.
[12] H. Bartkova, P. Kluson, L. Bartek, M. Drobek, T. Cajthaml, J. Krysa, Thin Solid
separation, a higher rate of recombination is expected, which would Films 515 (2007) 8455.
lead to the observed dependency of RCT . [13] D. Jiang, H. Zhao, Z. Jia, J. Cao, R. John, Journal of Photochemistry and Photobi-
Addition of methanol as hole-scavenger doubled the steady- ology A 144 (2001) 197.
[14] M. Radecka, M. Wierzbicka, M. Rekas, Physica B 351 (2004) 121.
state photocurrent of TiO2 -SG electrodes. The linear reciprocal [15] I.D. Raistrick, D.R. Franceschetti, J.R. Macdonald, in: E. Barsoukov, J.R. Mac-
dependency of RCT on the photon ux was much more consistent, donald (Eds.), Impedance Spectroscopy: Theory, Experiment, and Applications,
in particular at high incident radiant power densities. Under these Second Edition, John Wiley & Sons, Inc, Hoboken, NJ, USA, 2005.
[16] P.J. Boddy, Journal of Electroanalytical Chemistry 10 (1965) 199.
conditions, the heterogeneous charge transfer reaction plays a deci- [17] M. Tomkiewicz, Electrochimica Acta 35 (1990) 1631.
sive role in the limitation of the overall reaction under anodic bias. [18] J.R. MacDonald, Journal of Electroanalytical Chemistry 223 (1987) 25.
The removal of holes by their reaction with methanol suppressed [19] J.-B. Jorcin, M.E. Orazem, N. Pbre, B. Tribollet, Electrochimica Acta 51 (2006)
1473.
the competing electronhole recombination in an efcient way, as
[20] J. Bisquert, G. Garcia-Belmonte, F. Fabregat-Santiago, N.S. Ferriols, P. Bodganoff,
shown by a remarkable drop of RREC , in particular at high radiant E.C. Pereira, Journal of Physical Chemistry B 104 (2000) 2287.
power densities. [21] F. Fabregat-Santiago, G. Garcia-Belmonte, J. Bisquert, A. Zaban, P. Salvador,
Journal of Physical Chemistry B 106 (2002) 334.
In summary, the photochemical characterization gave evidence
[22] Y. Pooarporn, A. Worayingyong, M. Wrner, P. Songsiriritthigul, A.M. Braun,
that nanoporous TiO2 -SG lm electrodes can be considered as a Water Science and Technology 55 (2007) 153.
hybrid of individual nanoparticles and solid crystalline semicon- [23] Radiant power density (P0 ) equates to emitted, transferred, or received radiant
ducting material. Current doubling disregarded, the pronounced energy per time and area.
[24] R.A. Spurr, H. Myers, Analytical Chemistry (Washington, DC, U. S.) 29 (1957)
effect of methanol addition, especially at high incident radiant 760.
power densities, on the heterogeneous charge transfer reaction [25] B. Liu, X. Zhao, N. Zhang, Q. Zhao, X. He, J. Feng, Surface Science 595 (2005) 203.
124 S. Kityakarn et al. / Electrochimica Acta 83 (2012) 113124

[26] J. Zhang, C. Yang, G. Chang, H. Zhu, M. Oyama, Materials Chemistry and Physics [39] M.A. Butler, Journal of Applied Physics 48 (1977) 1914.
88 (2004) 398. [40] P. Lemasson, A. Etcheberry, J. Gautron, Electrochimica Acta 27 (1982) 607.
[27] V.A. Alves, L.A. Da Silva, J.F.C. Boodts, S. Trasatti, Electrochimica Acta 39 (1994) [41] B. ORegan, J. Moser, M. Anderson, M. Grtzel, Journal of Physical Chemistry 94
1585. (1990) 8720.
[28] C. Angelinetta, S. Trasatti, Journal of Electroanalytical Chemistry 214 (1986) [42] A. Hagfeldt, U. Bjrkstn, S.E. Lindquist, Solar Energy Materials and Solar Cells
535. 27 (1992) 293.
[29] A.V. Vorontsov, A.A. Altynnikov, E.N. Savinov, E.N. Kurkin, Journal of Photo- [43] J. Bisquert, G. Garcia-Belmonte, F. Fabregat-Santiago, Journal of Solid State Elec-
chemistry and Photobiology A 144 (2001) 193. trochemistry 3 (1999) 337.
[30] M. Pavlovic, B. Santic, D.I. Desnica-Frankovic, N. Radic, T. Smuc, U.V. Desnica, [44] D. Vanmaekelbergh, P.E. de Jongh, Physical Chemistry B 103 (1999) 747.
Journal of Electronic Materials 32 (2003) 1100. [45] F. Fabregat-Santiago, H. Randriamahazaka, A. Zaban, J. Garcia-Canandas, G.
[31] H. Ghr, Electrochemical Applications 1/97 (1997) 2, <http://www. Garcia-Belmonte, J. Bisquert, Physical Chemistry Chemical Physics 8 (2006)
zahner.de/pdf/ea1997.pdf>. 1827.
[32] J.-P. Candy, P. Fouilloux, M. Keddam, H. Takenouti, Electrochimica Acta 26 [46] S. Gimenez, H.K. Dunn, P. Rodenas, F. Fabregat-Santiago, S.G. Miralles, E.M.
(1981) 1029. Barea, R. Trevisan, A. Guerrero, J. Bisquert, Journal of Electroanalytical Chem-
[33] I.D. Raistrick, Electrochimica Acta 35 (1990) 1579. istry 668 (2012) 119.
[34] J. Bisquert, Physical Chemistry Chemical Physics 5 (2003) 5360. [47] N. Hykaway, W.M. Sears, H. Morisaki, S.R. Morrison, Journal of Physical Chem-
[35] M.E. Orazem, P. Agarwal, L.H. Garcia-Rubio, Journal of Electroanalytical Chem- istry 90 (1986) 6663.
istry 378 (1994) 51. [48] F. Forouzan, T.C. Richards, A.J. Bard, Journal of Physical Chemistry 100 (1996)
[36] J. Schefold, Journal of the Electrochemical Society 142 (1995) 850. 18123.
[37] A. Zaban, A. Meier, B.A. Gregg, Journal of Physical Chemistry B 101 (1997) 7985.
[38] W.W. Grtner, Physical Review 116 (1959) 84.

Potrebbero piacerti anche