You are on page 1of 5

Thin Solid Films 518 (2009) 1294–1298

Contents lists available at ScienceDirect

Thin Solid Films


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / t s f

Nanoporous TiO2 thin film based conductometric H2 sensor


A.Z. Sadek a,⁎, J.G. Partridge b, D.G. McCulloch b, Y.X. Li c, X.F. Yu c, W. Wlodarski a, K. Kalantar-zadeh a
a
School of Electrical and Computer Engineering, RMIT University, Melbourne, VIC 3001, Australia
b
Applied Physics, School of Applied Sciences, RMIT University, Melbourne, VIC 3001, Australia
c
Shanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai, China

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

Available online 9 April 2009 Nanoporous titanium dioxide (TiO2) based conductometric sensors have been fabricated and their sensitivity
to hydrogen (H2) gas has been investigated. A filtered cathodic vacuum arc (FCVA) system was used to
Keywords: deposit ultra-smooth Ti thin films on a transducer having patterned inter-digital gold electrodes (IDTs).
Nanoporous TiO2 Nanoporous TiO2 films were obtained by anodization of the titanium (Ti) thin films using a neutral 0.5% (wt)
Anodization NH4F in ethylene glycol solution at 5 V for 1 h. After anodization, the films were annealed at 600 °C for 8 h to
FCVA
convert the remaining Ti into TiO2. The scanning electron microscopy (SEM) images revealed that the average
H2 sensor
diameters of the nanopores are in the range of 20 to 25 nm. The sensor was exposed to different
concentrations of H2 in synthetic air at operating temperatures between 100 °C and 300 °C. The sensor
responded with a highest sensitivity of 1.24 to 1% of H2 gas at 225 °C.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction foils, preventing their application in the standard mass production of


thin film based sensors.
Gas sensors are increasingly required for ensuring industrial health Many research groups [17–24] have reported formation of
and safety, environmental monitoring, automotive applications and nanoporous and nanotubular surfaces of titanium oxide since Zwilling
manufacturing process control. Recently, much attention has been et al. [25,26] first reported on self-organized nanotubular surfaces by
given to gas sensitive materials in nanostructured forms as they often anodic oxidation of Ti foil. Anodization of Ti foils/films were
exhibit novel and enhanced properties compared to their bulk conducted in diluted fluoride ion containing electrolytes at a fixed
counterparts [1,2]. It is well known that the performance of a gas [23,27] or variable voltage [22,28] for a given time. Recently, it has
sensor is directly related to the granularity, porosity and surface to been reported that nanotubes can be formed in Ti using chloride and
volume ratio of the sensing materials [3,4]. Additionally, gas-sensing bromide containing electrolytes as well [29,30]. A variety of electro-
performances of nanostructured metal oxide films can improve lytes have been used to prepare TiO2 nanostructures such as acidic
remarkably when the size of the nanomaterials becomes comparable [18], neutral [23,31], aqueous or non-aqueous electrolytes
to Debye length [5,6]. Low dimensional nanostructured materials [19,23,32,33]. Typically, the pore/tube diameter is controlled by the
having increased surface to volume ratio facilitate rapid diffusion of applied anodization voltage, and the length can be varied by using
gases into and out of the materials' nanoporosities. This rapid different anodization time and electrolytes [34]. The nanoporous/
diffusion increases the reaction rate, resulting in faster sensor nanotubular layers can be grown with tube or pore dimensions
response and recovery [7,8]. varying from a few hundreds of nanometers to 1 mm in length, and
Titanium dioxide (TiO2) is one of the most widely used functional with diameters varying from 10 to 200 nm [17,23,32–35]. It has been
materials with applications in gas sensors [4,9], dye-sensitized solar reported that the nanopore/tube growth mechanism involves initial
cells [10,11], electro-chromic devices [12] and photocatalysis [5,13]. oxidation of Ti to form an oxide layer and the subsequent localized
TiO2 films have been investigated as sensors for H2 [3,9], CO and NO2 dissolution of this layer by fluoride ions [22]. Details of the suggested
[4,14], O2 [15], and hydrocarbons [16]. The gas-sensing capability of mechanisms for the formation of nanotubes/pores during the
TiO2 is due to changes in the film conductivity in the presence of anodization process can be found elsewhere [22,36,37].
oxidizing or reducing gases. Recent successes in synthesizing TiO2 in The self-organized nanoporous/tubular structures of TiO2 have
nanotubular and nanoporous forms have prompted explorations into already been tailored for various applications with improved
gas-sensing applications. However, most of the nanotubular and performance, such as dye-sensitized solar cells [38–41], sensors
nanoporous TiO2 forms produced to date have been synthesized in [3,42–48], photocatalysts [49,50], electro-chromic displays [51–53]
and wettability applications [54,55]. Recently, Grimes et al. [3,42–44]
⁎ Corresponding author. have reported TiO2 nanotube based conductometric H2 sensors where
E-mail addresses: sadek@ieee.org (A.Z. Sadek), Kourosh.kalantar@rmit.edu.au nanotubes were anodized on Ti foil using acidic electrolyte solution. To
(K. Kalantar-zadeh). obtain experimental gas-sensing responses, they placed platinum

0040-6090/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.tsf.2009.02.151
A.Z. Sadek et al. / Thin Solid Films 518 (2009) 1294–1298 1295

negative bias voltage can be applied to the holder via a high-voltage


feed-through. This bias causes the incident ions to be accelerated
towards the sample and this energetic deposition process assists in
the production of dense flat films. For these experiments the sample
bias was −100 V. The Ti films (with a thickness ranging from 250 nm
to 300 nm) prepared by FCVA were dense, smooth and highly
adherent to the substrates.
Anodization of the Ti films was carried out using a neutral
electrolyte medium of 0.5% (wt) NH4F in ethylene glycol solution at
5 V for 1 h. A platinum foil was used as a counter electrode for the
anodization at room temperature. Anodized samples were washed
with DI water before blow drying with dry nitrogen (N2) gas. After
anodization, the film was annealed at 600 °C for 8 h for conversion to
TiO2.
The sensor was mounted inside an enclosed environmental
chamber. Four mass flow controllers (MFCs) were connected to
form a single output that supplies gas to the cell. The gas delivery
system has the ability to accurately control and vary concentrations
for four different gases simultaneously. In this work, the authors have
employed two channels of MFCs at a time, one for synthetic air and
one for the target analyte (H2). The concentration of the gas was
varied by changing the synthetic air to analyte gas ratio while
Fig. 1. SEM image of the FCVA deposited Ti film supported on a LiTaO3 substrate and maintaining a constant flow rate of 200 sccm.
planar Au contacts (as indicated). Highly pure low concentration H2 (1%) gas balanced in synthetic
air in a cylinder was employed for this work. A high precision Keithley
2001 multimeter was used for measuring the variation of sensor
electrodes on the same side of the nanotube foil and found that the resistance. Custom LabVIEW-based software was utilized to control
device is highly sensitive towards H2. The electrical resistance of the the experimental setup and take measurement data from the sensor.
TiO2 nanotubes changes by almost four-orders of magnitude upon The sensor responses were displayed in real-time and saved for off-
exposure to 100 ppm H2 in N2 against air base line at room line processing and analysis. The gas exposure time was fixed for each
temperature [44]. Grime et al. [48] also developed thin film gas pulse of analyte gas and the cell was purged with synthetic air for
sensors based on TiO2 nanotube arrays and reported a four-order fixed periods of time between each pulse to allow full recovery of the
magnitude drop in resistance on exposure to 1000 ppm H2 gas in N2 sensor.
and negligible sensitivity to other reducing gases like methane, carbon A planar micro-heater with dimensions of 25 × 25 mm2 was placed
monoxide and ammonia. Zhang et al. [56] developed TiO2 nanotube beneath the sensor. The micro-heater was fabricated on a sapphire
based humidity sensors using anodization of Ti foil. They observed substrate with a patterned platinum resistive element. A regulated DC
nearly two orders of magnitude change in resistance during the power supply was connected to the heater to control the operating
relative humidity variation from 11 to 95%. temperature of the sensor between the range of 100 and 300 °C in
In this paper, we report a conductometric H2 sensor based on increments of approximately 30 °C. A thermocouple was employed to
nanoporous TiO2 thin films. Ultra-smooth Ti thin films were deposited
by filtered cathodic vacuum arc (FCVA) onto inter-digital transducers
(IDTs) supported on LiTaO3 substrates. The Ti films were then
anodized using a neutral NH4F/ethylene glycol electrolyte solution.
The completed nanoporous TiO2 films exhibited high surface to
volume ratio, homogenous pore growth and good adherence to the
substrate/contacts. Measurements were conducted for various gas
concentrations of H2 mixed in synthetic air.

2. Experimental

The sensor was designed and fabricated to operate as a resistive


element. The device comprises a sensitive TiO2 nanoporous layer
anodized over gold (Au) sputtered inter-digital electrodes on a
12 × 12 mm2 LiTaO3 substrate. The IDTs were formed by patterning
an 80 nm layer of gold (Au) and a 20 nm titanium layer. The titanium
layer is added as an under-layer to improve adhesion of the gold layer.
After formation of the contacts, the thin film Ti was deposited onto
the substrate using a FCVA system operated with a 70 mm diameter
cathode of 99.99% purity Ti. The cathode is struck by a grounded
mechanical striker to initiate the plasma. An arc current of 120 A (and
average arc power 3 kW) was found to be sufficient to produce a stable
plasma for the deposition experiments. A double-bend magnetic filter
was incorporated into the system and this prevented deposition of
macro-particles onto the sample, thereby minimizing surface rough-
ness. The samples themselves were mounted on a metallic holder Fig. 2. SEM image of the anodized TiO2 film deposited around the boundary of an Au
which slides onto an electrically isolated arm inside the chamber. A electrode and the LiTaO3 substrate.
1296 A.Z. Sadek et al. / Thin Solid Films 518 (2009) 1294–1298

Fig. 5. XRD patterns of an FCVA deposited Ti film on a LiTaO3 substrate (bottom); and of
an anodized film annealed at 600 °C (top).

tures, nanotubes are obtained [48,57]. These films exhibit XRD


reflections corresponding to (101) and (002). The anodization of
FCVA deposited films only resulted in nanopores. It is therefore
possible that the initial (002) orientation of Ti films promotes the
formation of nanopores. A (101) reflection peak was observed in the
XRD spectra after annealing the anodized films at 600 °C for 8 h,
Fig. 3. SEM image of the anodized TiO2 film supported on LiTaO3. indicating the formation of rutile TiO2 (Fig. 5).
The dynamic response of a TiO2 sensor to a sequence of different H2
obtain a real-time reading of the sensor surface temperature with an concentrations in synthetic air is shown in Fig. 6. The sensor response
accuracy of 1 °C. for a reducing gas, defined as the ratio of the baseline resistance to the
sensor resistance after exposure to gas (Rair/Rgas), is 1.24 for 1% of H2 in
3. Results synthetic air at 225 °C. The highest sensitivity with the fastest
response and recovery were measured at 225 °C for 1% H2. The sensor
An SEM image of the FCVA deposited Ti film supported on the response as a function of the operating temperature in the presence of
conductometric transducer is shown in Fig. 1. Since the FCVA 1% H2 is shown in Fig. 7.
deposited films are ultra smooth, the SEM image reveals little detail. The repeatability of the sensor's response towards H2 was
SEM images of the nanoporous TiO2 film are shown in Fig. 2 (over observed with two 0.12% H2 pulses on the sequence shown in Fig. 6.
film-contact interface), Fig. 3 (over LiTaO3 surface), and Fig. 4 (over It was observed that magnitude of the second pulse decreased slightly
gold electrode surface). The SEM images revealed that the morphol- (~2%) with a downward baseline drift due to slower desorption rate.
ogy of the anodized Ti film is very similar, irrespective of whether the The chemisorption of hydrogen ions (H2f2H+) onto the TiO2
film is over the Au or the LiTaO3. The average diameter of the nanopores is the basic mechanism for the sensor response. The
nanopores is 23 nm ± 3 nm. An XRD pattern of a FCVA deposited Ti efficient interaction area between the gas molecules and the sensitive
film supported on a LiTaO3 substrate is given in Fig. 5. The XRD pattern nanoporous TiO2 surface determines the performance of the gas
shows that for the FCVA deposited Ti thin film, only the (002) sensor. Thus, increasing the surface to volume ratio of the TiO2 films
reflection was produced. It has been previously reported that after the introducing nanoporosity improves the sensitivity as well as the
anodization of Ti films which were sputtered at elevated tempera- dynamic performance of the sensor. Although the interaction of a gas
with a sensing material is primarily a surface phenomenon, the gas-
sensing responses of the TiO2 nanopores/tubes cannot be explained
on the basis of surface interaction alone [43,44,47]. The nanoporous/
tubular forms of TiO2 have greatly reduced diameters and wall
thickness whose dimensions are comparable to the depletion layer
depth. The space charge layers are thus strongly modulated by the
dimensional features. Accordingly, oxygen adsorption to the nano-
pore/tubes' wall can lead to a complete depletion of conduction-band
electrons with a large variation in resistance.

Fig. 4. SEM image of the anodized TiO2 film supported on Au. Fig. 6. Dynamic response of the sensor to different H2 gas concentrations at 225 °C.
A.Z. Sadek et al. / Thin Solid Films 518 (2009) 1294–1298 1297

References

[1] A.P. Alivisatos, P.F. Barbara, A.W. Castleman, J. Chang, D.A. Dixon, M.L. Klein, G.L.
McLendon, J.S. Miller, M.A. Ratner, P.J. Rossky, S.I. Stupp, M.E. Thompson, Adv.
Mater. 10/16 (1998) 1297.
[2] A. Thiaville, J. Miltat, Science 284/5422 (1999) 1939.
[3] O.K. Varghese, D.W. Gong, M. Paulose, K.G. Ong, C.A. Grimes, Sens. Actuators B-
Chem. 93/1-3 (2003) 338.
[4] V. Guidi, M.C. Carotta, M. Ferroni, G. Martinelli, L. Paglialonga, E. Comini, G. Sberveglieri,
Sens. Actuators B-Chem. 57/1-3 (1999) 197.
[5] Z.B. Zhang, C.C. Wang, R. Zakaria, J.Y. Ying, J. Phys. Chem. B 102/52 (1998) 10871.
[6] S. Shukla, S. Patil, S.C. Kuiry, Z. Rahman, T. Du, L. Ludwig, C. Parish, S. Seal, Sens.
Actuators B-Chem. 96/1-2 (2003) 343.
[7] A.Z. Sadek, W. Wlodarski, Y. Li, W. Yu, X. Li, X. Yu, K. Kalantar-Zadeh, Thin Solid
Films 515/24 (2007) 8705.
[8] A.Z. Sadek, S. Choopun, W. Wlodarski, S.J. Ippolito, K. Kalantar-zadeh, Ieee Sens. J.
7/5-6 (2007) 919.
[9] G.C. Mather, F.M.B. Marques, J.R. Frade, J. Eur. Ceram. Soc. 19/6-7 (1999) 887.
Fig. 7. Sensor sensitivity to 1% H2 gas versus the operating temperature. [10] B. Oregan, M. Gratzel, Nature 353/6346 (1991) 737.
[11] U. Bach, D. Lupo, P. Comte, J.E. Moser, F. Weissortel, J. Salbeck, H. Spreitzer, M. Gratzel,
Nature 395/6702 (1998) 583.
In order to clarify the performances of the developed nanoporous [12] C. Bechinger, S. Ferrer, A. Zaban, J. Sprague, B.A. Gregg, Nature 383/6601 (1996)
TiO2 based devices for H2 sensing, a brief comparative study of 608.
previous works on TiO2 and other metal oxide based H2 sensors is [13] M. Gratzel, Nature 414/6861 (2001) 338.
[14] J. Tan, W. Wlodarski, K. Kalantar-Zadeh, Thin Solid Films 515 (2007) 8738.
given here. The authors previously reported [8] a conductometric H2, [15] U. Kirner, K.D. Schierbaum, W. Gopel, B. Leibold, N. Nicoloso, W. Weppner, D. Fischer,
NO2 and hydrocarbon gas sensor based on single-crystalline zinc W.F. Chu, Sens. Actuators B-Chem. 1/1-6 (1990) 103.
oxide (ZnO) nanobelts deposited using a radio frequency (RF) [16] J. Trimboli, P.K. Dutta, Sens. Actuators B-Chem. 102/1 (2004) 132.
[17] S.P. Albu, A. Ghicov, J.M. Macak, P. Schmuki, Phys. Status Solidi-Rapid Res. Lett. 1/2
magnetron sputterer. The sensitivity of the developed sensor to 1% (2007) R65.
H2 was calculated to be 14.3 at 385 °C. Although the sensitivity of the [18] D. Gong, C.A. Grimes, O.K. Varghese, W.C. Hu, R.S. Singh, Z. Chen, E.C. Dickey,
ZnO nanobelt based sensor towards H2 is higher than the reported J. Mater. Res. 16/12 (2001) 3331.
[19] J.M. Macak, P. Schmuki, Electrochim. Acta 52/3 (2006) 1258.
nanoporous based TiO2 sensor, the optimum operating temperature [20] J.M. Macak, H. Tsuchiya, P. Schmuki, Angew. Chem., Int. ed. 44/14 (2005) 2100.
for the latter is much lower (225 °C). The author also recently reported [21] J.M. Macak, H. Tsuchiya, L. Taveira, S. Aldabergerova, P. Schmuki, Angew. Chem.,
[7] hydro-thermally grown ZnO nanorod based surface acoustic wave Int. ed. 44/45 (2005) 7463.
[22] G.K. Mor, O.K. Varghese, M. Paulose, N. Mukherjee, C.A. Grimes, J. Mater. Res. 18/11
(SAW) sensors and found that the optimal operating temperature of
(2003) 2588.
the devices for H2 sensing was 265 °C with a response of 274 kHz [23] M. Paulose, K. Shankar, S. Yoriya, H.E. Prakasam, O.K. Varghese, G.K. Mor, T.A.
towards 0.15% of H2. Wang et al. [58] developed self-assembled SnO2 Latempa, A. Fitzgerald, C.A. Grimes, J. Phys. Chem. B 110/33 (2006) 16179.
[24] W.Z. Wang, O.K. Varghese, M. Paulose, C.A. Grimes, Q.L. Wang, E.C. Dickey, J. Mater.
nanowire sensors for H2 sensing and the device response was
Res. 19/2 (2004) 417.
measured to be 3.25 towards 1000 ppm H2 at 300 °C. [25] V. Zwilling, M. Aucouturier, E. Darque-Ceretti, Electrochim. Acta 45/6 (1999) 921.
Grimes' group reported room temperature hydrogen sensors based [26] V. Zwilling, E. Darque-Ceretti, A. Boutry-Forveille, D. David, M.Y. Perrin, M. Aucouturier,
on anodic TiO2 nanotube arrays in foil [44] and in thin films [48]. Surf. Interf. Anal. 27/7 (1999) 629.
[27] R. Beranek, H. Hildebrand, P. Schmuki, Electrochem. Solid State Lett. 6/3 (2003)
These TiO2 nanotube array sensors showed a change in electrical B12.
resistance upon exposure to 100 ppm hydrogen in N2 of four orders of [28] J.M. Macak, S. Albu, D.H. Kim, I. Paramasivam, S. Aldabergerova, P. Schmuki,
magnitude against the air base line at 25 °C. In this study, the Electrochem. Solid State Lett. 10/7 (2007) K28.
[29] Q.A. Nguyen, Y.V. Bhargava, T.M. Devine, Electrochem. Commun. 10/3 (2008) 471.
sensitivity seems to be lower than the sensitivity of the devices [30] N.K. Allam, C.A. Grimes, J. Phys. Chem. C 111/35 (2007) 13028.
previously fabricated by Grimes group. However, we tested the [31] C.M. Ruan, M. Paulose, O.K. Varghese, G.K. Mor, C.A. Grimes, J. Phys. Chem. B 109/33
nanoporous TiO2 against target H2 gas in synthetic air ambient and not (2005) 15754.
[32] M. Paulose, H.E. Prakasam, O.K. Varghese, L. Peng, K.C. Popat, G.K. Mor, T.A. Desai,
N2 which is the simulation of a real field application. The presence of C.A. Grimes, J. Phys. Chem. C 111/41 (2007) 14992.
ambient oxygen (21%) affects the performance of the sensor and thus, [33] H.E. Prakasam, K. Shankar, M. Paulose, O.K. Varghese, C.A. Grimes, J. Phys. Chem. C
results are not comparable with the results obtained by Grimes group. 111/20 (2007) 7235.
[34] S. Bauer, S. Kleber, P. Schmuki, Electrochem. Commun. 8/8 (2006) 1321.
In the future work, we intend to enhance the performance of these
[35] K. Shankar, G.K. Mor, H.E. Prakasam, S. Yoriya, M. Paulose, O.K. Varghese, C.A.
nanoporous sensors by modifying the pore diameters, layer thickness Grimes, Nanotechnology 18/6 (2007).
and doping with noble catalytic metals (Pt, Pd or Au). [36] J.M. Macak, H. Tsuchiya, A. Ghicov, K. Yasuda, R. Hahn, S. Bauer, P. Schmuki, Curr.
Opin. Solid State Mater. Sci. 11/1-2 (2007) 3.
[37] K. Yasuda, J.M. Macak, S. Berger, A. Ghicov, P. Schmuki, J. Electrochem. Soc. 154/9
4. Conclusion (2007) C472.
[38] J.M. Macak, H. Tsuchiya, A. Ghicov, P. Schmuki, Electrochem. Commun. 7/11 (2005)
Conductometric hydrogen sensors based on nanoporous titanium 1133.
[39] K. Shankar, J. Bandara, M. Paulose, H. Wietasch, O.K. Varghese, G.K. Mor, T.J.
dioxide (TiO2) thin films have been developed and studied. Nano- LaTempa, M. Thelakkat, C.A. Grimes, Nano Lett. 8/6 (2008) 1654.
porosity on the films was obtained by the anodization of Ti thin films [40] G.K. Mor, O.K. Varghese, M. Paulose, K. Shankar, C.A. Grimes, Solar Energ. Mater.
deposited onto conductometric transducer substrates using a FCVA Solar Cells 90/14 (2006) 2011.
[41] G.K. Mor, K. Shankar, M. Paulose, O.K. Varghese, C.A. Grimes, Nano Lett. 6/2 (2006)
system. Anodization of the Ti films was performed using a neutral 0.5% 215.
(wt) NH4F in ethylene glycol solution at 5 V for 1 h. The films were [42] M. Paulose, O.K. Varghese, G.K. Mor, C.A. Grimes, K.G. Ong, Nanotechnology 17/2
annealed at 600 °C for 8 h for conversion to the rutile form of TiO2. The (2006) 398.
[43] O.K. Varghese, D.W. Gong, M. Paulose, K.G. Ong, E.C. Dickey, C.A. Grimes, Adv.
SEM images showed that the average diameter of the nanopores was Mater. 15/7-8 (2003) 624.
23 nm ± 3 nm. The XRD patterns suggest that the presence of (002) [44] O.K. Varghese, G.K. Mor, C.A. Grimes, M. Paulose, N. Mukherjee, J. Nanosci.
preferred orientation on the FCVA deposited films may be the reason Nanotechnol. 4/7 (2004) 733.
[45] S. Yoriya, H.E. Prakasam, O.K. Varghese, K. Shankar, M. Paulose, G.K. Mor, T.J.
behind the nanopores (instead of nanotubes) formation during
Latempa, C.A. Grimes, Sensor Lett. 4/3 (2006) 334.
anodization. The sensor was exposed to different concentrations of [46] G.K. Mor, M.A. Carvalho, O.K. Varghese, M.V. Pishko, C.A. Grimes, J. Mater. Res. 19/2
H2 in synthetic air at operating temperatures in a range of 100 °C to (2004) 628.
[47] G.K. Mor, O.K. Varghese, M. Paulose, C.A. Grimes, Sensor Lett. 1/1 (2003) 42.
300 °C and the sensor response of 1.24 towards 1% of H2 at 225 °C was
[48] G.K. Mor, O.K. Varghese, M. Paulose, K.G. Ong, C.A. Grimes, Thin Solid Films 496/1
measured. The developed sensors are worthy of further investigation (2006) 42.
for commercial applications. [49] I. Paramasivalm, J.M. Macak, P. Schmuki, Electrochem. Commun. 10/1 (2008) 71.
1298 A.Z. Sadek et al. / Thin Solid Films 518 (2009) 1294–1298

[50] O.K. Varghese, M. Paulose, K. Shankar, G.K. Mor, C.A. Grimes, J. Nanosci. [56] Y.Y. Zhang, W.Y. Fu, H.B. Yang, Q. Qi, Y. Zeng, T. Zhang, R.X. Ge, G. Zou, Appl. Surf. Sci.
Nanotechnol. 5/7 (2005) 1158. 254/17 (2008) 5545.
[51] N.K. Allam, K. Shankar, C.A. Grimes, J. Mater. Chem. 18/20 (2008) 2341. [57] G.K. Mor, O.K. Varghese, M. Paulose, C.A. Grimes, Adv. Funct. Mater. 15/8 (2005)
[52] R. Beranek, H. Tsuchiya, T. Sugishima, J.M. Macak, L. Taveira, S. Fujimoto, H. Kisch, 1291.
P. Schmuki, Appl. Phys. Lett. 87/24 (2005). [58] B. Wang, L.F. Zhu, Y.H. Yang, N.S. Xu, G.W. Yang, J. Phys. Chem. C 112/17 (2008)
[53] G.K. Mor, K. Shankar, O.K. Varghese, C.A. Grimes, J. Mater. Res. 19/10 (2004) 2989. 6643.
[54] E. Balaur, J.M. Macak, H. Tsuchiya, P. Schmuki, J. Mater. Chem. 15/42 (2005) 4488.
[55] E. Balaur, J.M. Macak, L. Taveira, P. Schmuki, Electrochem. Commun. 7/10 (2005)
1066.

You might also like