You are on page 1of 5

Sensors and Actuators B 156 (2011) 505–509

Contents lists available at ScienceDirect

Sensors and Actuators B: Chemical


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

Letter

A selective room temperature formaldehyde gas sensor using TiO2 nanotube arrays

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

Keywords: A new gas sensor using TiO2 nanotube arrays was fabricated and explored for formaldehyde detection at
TiO2 room temperature. Highly ordered vertically grown TiO2 nanotube arrays were synthesized by using the
Room temperature gas sensor conventional electrochemical anodization process. The sensor using the fabricated nanotube arrays as the
Formaldehyde
sensing elements demonstrated a good response to different concentrations of formaldehyde from 10 to
50 ppm and a very good selectivity over other reducing gas species such as ethanol and ammonia at room
temperature. While the exact sensing mechanism is unclear, some possibilities are briefly discussed.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction ysis and chemical sensing [24–32]. The flexibility in adjustment


of the microstructure in terms of tube length, pore diameter and
Formaldehyde (HCHO) is a colorless, pungent-smelling gas and chemical composition has offered an excellent platform that can
can cause several symptoms such as watery eyes, burning sensa- be tailored to meet different requirements [26]. Particularly, in
tion in the eyes and throat, nausea, and difficulty in breathing in gas sensing, it has been studied for the detection of ppm-level
some humans exposed at elevated levels (above 0.1 parts per mil- oxygen, H2 , acetone, and humidity at relatively low temperatures
lion). High concentrations may trigger heart attack in people with [24–26,29–32]. The basic working principle involves adsorption
asthma [1–6]. HCHO is also a possible carcinogen and may cause and desorption of the dissociated oxygen due to the large specific
central nervous system damage, immune system disorders, as well surface area of the nanotubes for oxygen, humidity and acetone
as blindness and respiratory diseases [1–3]. Such highly health- detection. In the case for H2 detection in pure nitrogen background,
threatening gas species can exist in substantial concentrations both the splitting of H2 and direct interaction between hydrogen and the
indoors and outdoors. Sources of formaldehyde in the home include nanotube surface is involved [26,33].
building materials and household products [1–6]. Therefore, there In this work, TiO2 nanotube arrays have been studied for the
is a high demand on deployment of portable, cheaper gas detec- detection of formaldehyde at room temperature for the first time
tors at home and work place for real-time monitoring the levels of to the best of our knowledge.
formaldehyde so that precautions can be taken in the presence of
this toxic gas. 2. Experimental
Metal oxide semiconductor gas sensors to detect toxic gas
species are attractive due to their simple working principle, high Titanium sheets (purity: 99.4%) were successively sonicated
sensitivity, portability and low cost [7–15]. For formaldehyde in acetone, ethanol and deionized water for 10 min at each step
detection, various metal oxides have been investigated including to remove any grease on the surface before the electrochemical
SnO2 or doped SnO2 [4–6,8,9], WO3 [9], LaFe1−x Znx O3 [10], NiO anodization. The anodization was performed using a two-electrode
thin film [11,12], CdO-mixed In2 O3 [13] and doped ZnO [14,15]. cell. The as-treated Ti plates were used as the working electrodes
These sensors indicate good sensitivity to ppm or sub-ppm levels of and a stainless steel served as the counter electrode. The samples
formaldehyde and have potential to be employed in real situation. were anodized in a solution containing 0.27 M NH4 F consisting of
However, one of the obvious disadvantages is that they operate mixtures of DI water and glycerol (1,2,3-propanetriol) prepared in
at elevated temperatures usually around 200–400 ◦ C [7–15] and a volumetric ratio of 50:50. The anodization consisted of a poten-
require a heater integrated with the sensor. This would raise the tial ramp from the open-circuit potential to 30 V, and held at that
operating cost and also add requirements on packaging materials, potential for another 3 h. Subsequently, the as-fabricated oxidized
further complicating the sensor design. Therefore, recent research samples were rinsed in sequence with ethanol and deionized water
has focused on modifying the conventional materials including and dried in an air stream. The samples were then calcined at 400 ◦ C
utilization of nanosized particles or nanostructural architectures, and 500 ◦ C for 3 h, respectively.
sometimes aided with the UV–visible illumination during the sen- The phase formation of the fabricated samples was analyzed
sor operation [16–22]. using a powder X-ray diffractometer (XRD: D8 Advance, Bruker-
Since the highly ordered vertically grown TiO2 nanotubes have AXS, Germany). The XRD patterns were collected using Ni-filtered
been successfully synthesized by Grimes et al. [23], they have been Cu K␣ radiation at 40 kV and 25 mA between 2 of 15–65◦ at a scan-
explored for various applications such as in solar cells, photocatal- ning speed of 12◦ /min. The surface morphology of the fabricated

0925-4005/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.snb.2011.02.046
506 Letter / Sensors and Actuators B 156 (2011) 505–509

Fig. 2. X-ray diffraction patterns of TiO2 nanotube arrays annealed at different tem-
peratures: (a) as-fabricated, (b) annealed at 400 ◦ C and (c) annealed at 500 ◦ C.
Fig. 1. Photograph of the electrode geometry of the sensor using TiO2 nanotube
array.

crystalline phase, thermal treatment was carried out above 300 ◦ C.


nanotubes was examined using a scanning electron microscope The clear sharp peaks shown in the XRD patterns of the two samples
(SEM: S-3000N, Hitachi, Japan) on gold-coated specimens. calcined at 400 ◦ C and 500 ◦ C indicate the presence of a crystalline
The sensing measurements were conducted under a static phase. As calcination temperature increased, the samples crystal-
condition in a sealed chamber with a total volume of 50 L. The lized into two phases with anatase dominant at lower temperature
temperature inside the chamber remained constant at 22 ± 1 ◦ C and then the more stable rutile appeared when the temperature
monitored by a mercury thermometer. The relative humidity (R.H.) was raised up to 500 ◦ C. The titanium peaks labeled in figure are
was uniform throughout at around 40% which was measured by from the top edge of the titanium sheet which was not immersed
an electronic hygrometer fixed inside the chamber. The corre- in the electrolyte during the anodization process and thus was not
sponding liquid solution containing formaldehyde was placed in converted to oxide.
a small crucible and then evaporated into the chamber during Fig. 3 shows the surface morphology of the fabricated nan-
each sensing measurement. The concentration of the formaldehyde otubes. The measured pore diameter from Fig. 3(a) was around
was calculated according to the volume of the evaporated vapor 150 nm and the tube length was 1.7 micrometers determined
and total volume of the chamber assuming the total pressure was by the cross-sectional observation (result not shown here). The
kept constant at one atmosphere and at 25 ◦ C. Therefore, differ- as-fabricated amorphous nanotube array was packed closely and
ent concentrations of formaldehyde were obtained by controlling uniformly, while some fuzzy deposits were also observed on the
the amount of liquid in the small crucible. Ethanol and ammonia surface. Increasing the calcination temperature to 400 ◦ C (Fig. 3(b))
gas vapors were obtained in a similar way. The background gas was indicates signs of stress in the thin-film nanotubes; cracks are
the ambient air. After each test, the gas was released to the ambient clearly observed among the nanotubes. This became more obvi-
by opening the front window of the chamber and simultaneously, ous when the annealing temperature further increased up to 500 ◦ C
the ambient air was re-introduced. Since the process for the ana- (Fig. 3(c)). Appearance of the rutile phase usually accompanied the
lyte clearance out of the chamber would take a longer time, the crystal growth as the calcination temperature increased [35]. It is
recovery process of the sensor was not measured in this work. possible that such expansion in the crystal size of the rutile induced
The electrode connections of the TiO2 nanotube array based sen- a certain stress in the nanotube array producing the small cracks
sors are shown in Fig. 1. The sensing element is the TiO2 tube grown observed.
on a Ti sheet. Two stainless steel clips were clamped on the TiO2 Fig. 4(a) shows the response of the TiO2 nanotube-based sensors
nanotubes and served as the electrodes during the sensor tests. to different concentrations of formaldehyde in the background of
The distance between the two electrodes was around 2–3 mm. The ambient air with R.H. ∼40% at room temperature. All three sen-
electrical resistance of the nanotubes was measured by an Agi- sors indicated responses to formaldehyde within a concentration
lent digital electrometer (34401A) with data acquisition capability range from 10 to 50 ppm. The resistance of the TiO2 nanotubes
using the IntuiLink software. The response of the sensor is defined increased with exposure to formaldehyde. The sensor using the
as the relative change of the resistance of the nanotubes in ambi- TiO2 nanotubes annealed at 400 ◦ C showed the best response
ent air and in the analytes: S = (Rg − Ro )/Ro × 100%, where Rg is the to formaldehyde among the three sensors tested. The detection
resistance of the sensor in formaldehyde with air background and limit is estimated to be around 0.04 ppm using the calculation
Ro is the resistance in air. method in reference [36]. The response curve of the sensor using
TiO2 nanotubes annealed at 400 ◦ C to different concentrations of
3. Results and discussion formaldehyde is shown in Fig. 4(b). The response time, defined
as the time taken for the sensor’s resistance to reach the 90% of
Fig. 2 shows XRD patterns of the as-fabricated and calcined the steady-state resistance, is around 3 min. Due to the limit of the
TiO2 nanotubes at 400 ◦ C and 500 ◦ C. The as-fabricated sample was testing apparatus, the recovery curves were not measured.
amorphous. It has been reported that this amorphous structure Fig. 5(a) shows a comparison of the responses of the sensor
would persist until it is calcined above 300 ◦ C [30,34]. To obtain the using the TiO2 nanotube array annealed at 400 ◦ C to formalde-
Letter / Sensors and Actuators B 156 (2011) 505–509 507

40
As-fabricated (a)
o
Calcined at 400 C
o
30 Calcined at 500 C

Response (S%)
20

10

10 20 30 40 50
Cformaldehyde/ppm

50
4.5
40

30

R (MOhm)
4.0
20

[ppm] HCHO
10
3.5
Ambient

3.0
0 150 300 450 600 750
Time/Min

Fig. 4. (a) Response of the fabricated TiO2 nanotube array based sensors and (b) a
response curve of the sensor using the TiO2 nanotubes annealed at 400 ◦ C to different
concentrations of formaldehyde at room temperature (22 ± 1 ◦ C) with R.H.∼40%.

the metal oxides [37–39]. The ionized species in the water molecule
would ensure a change in the pH value on the oxide surface and thus
change the electrons/holes densities at the conduction or valence
bands of the oxide semiconductor. This consequently would result
in the sensor signals observed [37,39]. However, as pointed out
by Helwig et al. [37,39], this “dissociative gas sensing mechanism”
requires that the analyte gases should have a good water solubility
and be easily ionized in water. Moreover, the sensor signals induced
by such type of sensing mechanism are usually small. Therefore, it
seems the response of the sensor to ammonia and ethanol should
fall into this type of sensing theory.
Fig. 3. Scanning electron micrographs of TiO2 nanotube arrays (a) as-prepared, (b) Usually, pure TiO2 shows an n-type behavior when exposed to
annealed at 400 ◦ C and (c) annealed at 500 ◦ C.
reducing gases [24,26]. However, iron dopant has been reported
to be present in the fabricated nanotubes when stainless steel is
used as the cathode material [27]. The small amount of the triva-
hyde and other possible interference gases such as ammonia (NH3 ) lent Fe introduced could substitute on the Ti site in TiO2 according
TiO2 
and ethanol. The sensor showed much higher response to 50 ppm to: Fe2 O3 −→2FeTi + 3Oo + 2h· , consequently generating electronic
formaldehyde than that at 50 ppm NH3 and 1000 ppm ethanol. holes and making the TiO2 nanotube a p-type semiconductor.
Moreover, it can be observed from the corresponding response Therefore, according to the “dissociative gas sensing mechanism”
curves as shown in Fig. 5(b) that the sensor also indicates a shorter [37–39], ammonia and ethanol could be ionized in the water film
response time to formaldehyde. adsorbed on the surface of TiO2 nanotubes and thereby increase
The sensing mechanism for the metal oxides working in the the pH value, i.e. a decrease in the proton concentration (CH+ ), at
humidified air at room temperatures (usually 0–30 ◦ C) is often the surface. Consequently, more positively charged holes at the
proposed to be related to the electrolytic dissociations of the gas surface of the nanotube oxide have to be extracted to neutralize
species in the adsorbed water molecules covered on the surface of the adsorbed water molecule film. However, because of the strong
508 Letter / Sensors and Actuators B 156 (2011) 505–509

a when exposed to formaldehyde would lead to a decrease in the pH


50 ppm value consequently resulting in a decrease in the resistance of the
TiO2 nanotubes [37–39]. Obviously, this is contrary to the increase
in the sensor resistance when exposed to formaldehyde observed
30
in this work (Fig. 4). With the limited data we have, it is not possible
to explain this behavior and further studies are needed.
Response (S%)

20
4. Conclusions
1000 ppm
In summary, a new room-temperature formaldehyde gas
10 sensor using the TiO2 nanotube array is reported in this
50 ppm work. The sensor using TiO2 nanotube array annealed at
400 ◦ C shows the best response among the samples investi-
0 gated to different concentrations of formaldehyde from 10 to
Ethanol Formaldehyde Ammonia 50 ppm and good selectivity toward formaldehyde over 50 ppm
ethanol and 1000 ppm ammonia in humidified air at room
b temperature.
2
50 ppm NH3
1 Acknowledgements
0
Dr. Shiwei Lin appreciates the financial support from the
12 Program for New Century Excellent Talents in University
8 (NCET-09-0110) and National International Cooperation Program
S%

1000 ppm Ethanol


4 (2009DFA92551). Dr. Xiaogan Li would like to thank the foun-
0 dations from the Fundamental Research Funds for the Central
Universities. The authors thank Prof. Jing Wang for providing the
30 50 ppm HCHO sensor testing facility. Help rendered by Mr. Pengjun Yao and Mr.
Yangong Zheng on the experiments is also appreciated.
15
0
References
0 50 100 150 200 250 300
Time/Min [1] K.C. Gupta, A.G. Ulsamer, P.W. Preuss, Formaldehyde in indoor air: sources and
toxicity, Environ. Int. 8 (1982) 349–358.
[2] J.A. Pickreil, B.V. Mokier, L.C. Griffis, Formaldehyde release rate coefficients from
Fig. 5. (a) Selectivity of the sensor using the TiO2 nanotubes annealed at 400 ◦ C over
selected consumer products, Environ. Sci. Technol. 17 (1983) 753–757.
NH3 and ethanol, and (b) the corresponding response curves to 50 ppm formalde- [3] Y. Herschkovitz, I. Eshkenazi, C.E. Campbell, J. Rishpon, An electrochemical
hyde, 50 ppm NH3 and 1000 ppm ethanol at room temperature (22 ± 1 ◦ C) with biosensor for formaldehyde, J. Electroanal. Chem. 491 (2000) 182–187.
R.H.∼40%. [4] J. Wang, L. Liu, S.-Y. Cong, J.-Q. Qi, B.-K. Xu, An enrichment method to detect
low concentration formaldehyde, Sens. Actuators B 134 (2008) 1010–1015.
[5] J. Wang, P. Zhang, J.-Q. Qi, P.-J. Yao, Silicon-based micro-gas sensors for detect-
ing formaldehyde, Sens. Actuators B B136 (2009) 399–404.
coordination of the water molecule with the charged protons, the [6] P. Lv, Z.A. Tang, J. Yu, F.T. Zhang, G.F. Wei, Z.X. Huang, Y. Hu, Study on a micro-
neutralisation force would be expected to be very weak. This com- gas sensor with SnO2 –NiO sensitive film for indoor formaldehyde detection,
Sens. Actuators B 132 (2008) 74–80.
bined effects possibly led to the small increase in the resistance of [7] N. Yamazoe, Toward innovations of gas sensor technology, Sens. Actuators B
the oxides observed in Fig. 5. B108 (2005) 2–14.
However, for formaldehyde, it could not be easily ionized in [8] J. Wang, P. Zhang, J.-Q. Qi, P.-J. Yao, Silicon-based micro-gas sensors for detect-
ing formaldehyde, Sens. Actuators B 136 (2009) 399–404.
water although it has a good water solubility. Evidence has also
[9] L. Daza, S. Dassy, B. Delmon, Chemical sensors based on SnO2 and WO3 for the
indicated that formaldehyde would not be decomposed on the detection of formaldehyde: cooperative effects, Sens. Actuators B 10 (1993)
nanocrystalline TiO2 in the humidified air at room temperature 99–105.
[40]. Therefore, both the room temperature “dissociative sensing [10] S. Huang, H. Qin, P. Song, X. Liu, L. Li, R. Zhang, J. Hu, H. Yan, M. Jiang, The.
formaldehyde sensitivity of LaFe1−x Znx O3 -based gas sensor, J. Mater. Sci. 42
mechanism” discussed above and the conventional combustion (2007) 9973–9977.
theory for metal oxides based gas sensors working at high tem- [11] J.A. Dirksen, K. Duval, T.A. Ring, NiO thin film formaldehyde gas sensor, Sens.
peratures do not seem to be valid for the response of the sensor to Actuators B 80 (2001) 106–115.
[12] C.Y. Lee, C.M. Chiang, Y.H. Wang, R.H. Ma, A self-heating gas sensor with inte-
formaldehyde observed in Fig. 4. A new theory has to be considered. grated NiO thin film for formaldehyde detection, Sens. Actuators B 122 (2007)
The adsorbed water molecule on the surface of the nanocrys- 503–510.
talline TiO2 could be dissociated into two types of hydroxyl (OH− ) [13] T. Chen, Z. Zhou, Y. Wang, Effects of calcining temperature on the phase struc-
ture and the formaldehyde gas sensing properties of CdO-mixed In2 O3 , Sens.
groups: terminal hydroxyl (TH) and bridging hydroxyl (BH) groups Actuators B 135 (2008) 219–223.
even at room temperature in the humidified air [40]. These OH− [14] N. Han, Y. Tian, X. Wu, Y. Chen, Improving humidity selectivity in formaldehyde
groups are believed to contribute to the pH value and density of gas sensing by a two-sensor array made of Ga-doped ZnO, Sens. Actuators B 138
(2009) 228–235.
the charge carriers in the thin adsorbed water film on the surface [15] N. Han, L. Chai, Q. Wang, Y. Tian, P. Deng, Y. Chen, Evaluating the doping effect
of the oxide [37–39]. Moreover, according to the work reported in of Fe, Ti and Sn on gas sensing property of ZnO, Sens. Actuators B 147 (2010)
[40], with the introduction of formaldehyde on the surface of the 525–530.
[16] G. Shen, P.-C. Chen, K. Ryu, C. Zhou, Devices and chemical sensing applications
nanocrystalline TiO2 in the humidified atmosphere at room tem-
of metal oxide nanowires, J. Mater. Chem. 19 (2009) 828–839.
perature, the amount of these two types of OH− groups adsorbed on [17] N. Zhang, K. Yu, Q. Li, Z.Q. Zhu, Q. Wan, Room-temperature high-sensitivity
the TiO2 surface was significantly decreased due to the hydrogen H2 S gas sensor based on dendritic ZnO nanostructures with macroscale in
bonding between the molecularly adsorbed formaldehyde and the appearance, J. Appl. Phys. 103 (2008) 104305–104314.
[18] Z. Liu, T. Yamazaki, Y. Shen, T. Kikuta, N. Nakatani, T. Kawabata, Room tem-
hydroxyl groups. Therefore, one would expect that the significant perature gas sensing of p-type TeO2 nanowires, Appl. Phys. Lett. 90 (2007)
decrease in OH− groups on the surface of the nanocrystalline TiO2 173119–173125.
Letter / Sensors and Actuators B 156 (2011) 505–509 509

[19] M. Law, H. Kind, B. Messer, F. Kim, P. Yang, Photochemical sensing of NO2 with Biographies
SnO2 nanoribbon nanosensors at room temperature, Angew. Chem. Int. Ed. 41
(2002) 2405–2408.
[20] L. Peng, Q. Zhao, D. Wang, J. Zhai, P. Wang, S. Pang, T. Xie, Ultraviolet-assisted Dr. Shiwei Lin currently is a professor at Hainan University, PR China. He received
gas sensing: a potential formaldehyde detection approach at room temperature his Ph.D. degree from the University of Manchester in Electrical and Electronic Engi-
based on zinc oxide nanorods, Sens. Actuators B 136 (2009) 80–85. neering in the U.K. in 2006. His research interests include self-organized nanotubes,
[21] L. Peng, J. Zhai, D. Wang, Y. Zhang, P. Wang, Q. Zhao, T. Xie, Size- and photo- metal-oxide semiconductors and their photoelectric properties.
electric characteristics-dependent formaldehyde sensitivity of ZnO irradiated
with UV light, Sens. Actuators B 148 (2010) 66–73. Ms. Dongrong Li is a final-year undergraduate student in Hainan University, Haikou,
[22] K. Gopal, A.M. Maria, O.K. Carvalho, M. Varghese, V. Pishko, C.A. Grimes, A room- PR China.
temperature TiO2 -nanotube hydrogen sensor able to self-clean photoactively
Mr. Jian Wu is a final-year undergraduate student in the School of Electronic Science
from environmental contamination, J. Mater. Res. 19 (2004) 628–634.
and Technology in Dalian University of Technology, Dalian, Liaoning, PR China.
[23] D. Gong, C.A. Grimes, O.K. Varghese, W. Hu, R.S. Singh, Z. Chen, E.C. Dickey,
Titanium oxide nanotube arrays prepared by anodic oxidation, J. Mater. Res. 16 Dr. Xiaogan Li is an associate professor in the School of Electronic Science and Tech-
(2001) 3331–3335. nology in Dalian University of Technology, Dalian, Liaoning, PR China. He received
[24] O.K. Varghese, D. Gong, M. Paulose, K.G. Ong, C.A. Grimes, Hydrogen sensing his Ph.D. in Materials Science and Engineering from University of Leeds, U.K. in 2007.
using titania nanotubes, Sens. Actuators B 93 (2003) 338–344. Then he conducted a two-year postdoctoral research in chemical gas sensors at The
[25] H. Nakagawa, N. Yamamoto, S. Okazaki, A room-temperature operated hydro- Ohio State University in USA from 2007 to 2009. His current research interests are
gen leak sensor, Sens. Actuators B 93 (2003) 468–474. in chemical sensors, inorganic materials chemistry and physics, and biosensors.
[26] C.A. Grimes, Synthesis and application of highly ordered arrays of TiO2 nan-
otubes, J. Mater. Chem. 17 (2007) 1451–1457. Prof. Sheikh Akbar received his Ph.D. in materials and engineering from Purdue
[27] N.K. Allam, C.A. Grimes, Effect of cathode material on the morphology and pho- University. He is currently a professor of materials science and engineering at The
toelectrochemical properties of vertically oriented TiO2 nanotube arrays, Sol. Ohio State University. Prof. Akbar also founded the NSF Center for Industrial Sensors
Energy Mater. Sol. Cells 92 (2008) 1468–1475. and Measurements (CISM) at Ohio State, with primary focus on the R&D of sensors
[28] L.X. Yang, W.Y. Yang, Q.Y. Cai, Well-dispersed Pt/Au nanoparticles loaded for applications in hostile industrial environments.
into anodic titania nanotubes: a high antipoison and stable catalyst sys-
tem for methanol oxidation in alkaline media, J. Phys. Chem. C 111 (2007)
16613–16617. Shiwei Lin
[29] C.A. Grimes, K.G. Ong, O.K. Varghese, X. Yang, G. Mor, M. Paulose, E.C. Dickey, Dongrong Li
C. Ruan, M.V. Pishko, J.W. Kendig, A.J. Mason, A sentinel sensor network for Key Laboratory of Ministry of Education for
hydrogen sensing, Sensors 3 (2003) 69–82.
[30] Y. Zhang, W. Fu, H. Yang, Q. Qi, Y. Zeng, T. Zhang, R. Ge, G. Zou, Synthesis and Application Technology of Chemical Materials in
characterization of TiO2 nanotubes for humidity sensing, Appl. Surf. Sci. 254 Hainan Superior Resources, Hainan University,
(2008) 5545–5547. Haikou 570228, PR China
[31] H.F. Lu, F. Li, G. Liu, Z.G. Chen, D.-W. Wang, H.-T. Fang, G.Q. Lu, Z.H. Jiang, H.-M.
Cheng, Amorphous TiO2 nanotube arrays for low-temperature oxygen sensors, Jian Wu
Nanotechnology 19 (2008) 405504–405600.
[32] C.H. Han, D.W. Hong, I.J. Kim, Synthesis of Pd or Pt/titanate nanotube and its
Xiaogan Li ∗
application to catalytic type hydrogen gas sensor, Sens. Actuators B 128 (2007) Institute of Sensor Technologies, School of Electronic
320–325. Science and Technology, Dalian University of
[33] O.K. Varghese, D. Gong, M. Paulose, K.G. Ong, C.A. Grimes, Extreme changes
Technology, Dalian, Liaoning 116024, PR China
in the electrical resistance of titania nanotubes with hydrogen exposure, Adv.
Mater. 15 (2003) 624–627.
S.A. Akbar
[34] G.K. Mor, O.K. Varghese, R.H.T. Wilke, S. Sharma, K. Shankar, T.J. Latempa, K.-S.
Choi, C.A. Grimes, p-Type Cu–Ti–O nanotube arrays and their use in self-biased Center for Industrial Sensors and Measurements,
heterojunction photoelectrochemical diodes for hydrogen generation, Nano Department of Materials Science and Engineering,
Lett. 8 (2008) 1906–1911. The Ohio State University, Columbus, OH 43210, USA
[35] P.I. Gouma, P.K. Dutta, M.J. Mills, Structural stability of titania thin films, Nanos-
truct. Mater. 11 (1999) 1231–1237.
[36] J. Li, Y. Lu, Q. Ye, M. Cinke, J. Han, M. Meyyappan, Carbon nanotube sensors for ∗ Corresponding author at: School of Electronic
gas and organic vapor detection, Nano Lett. 3 (2003) 929–933. Science and Technology, Institute of Sensor
[37] A. Helwig, G. Müller, G. Sberveglieri, M. Eickhoff, On the low-temperature
response of semiconductor gas sensors, J. Sens. 2009 (2009) 1–17. Technologies, Chuan Xin Yuan Building A1226,
[38] B. Ostrick, J. Muhlsteff, M. Fleischer, H. Meixner, T. Doll, C.-D. Kohl, Adsorbed Linggong RD, No. 2 Gan Jing Zi, Dalian, Liaoning
water as key to room temperature gas-sensitive reactions in work function type 116024, PR China. Tel.: +86 0411 8470 6660.
sensors: the carbonate–carbon dioxide system, Sens. Actuators B 57 (1999)
115–119.
E-mail address: lixg@dlut.edu.cn (X. Li)
[39] A. Helwig, G. Müller, M. Eickhoff, G. Sberveglieri, Dissociative gas sensing at
metal oxide surfaces, IEEE Sens. J. 7 (2007) 1675–1679. 1 July 2010
[40] S. Sun, J.J. Ding, J. Bao, C. Gao, Z. Qi, C. Li, Photocatalytic oxidation of gaseous
Available online 2 March 2011
formaldehyde on TiO2: an in situ DRIFTS study, Catal. Lett. 137 (2010) 239–246.

You might also like