Vous êtes sur la page 1sur 4

Electrochemical and Solid-State Letters, 6 8 G108-G111 2003

G108

0013-4651/2003/68/G108/4/$7.00 The Electrochemical Society, Inc.

High Sensitivity Semiconductor NO2 Gas Sensor Based on


Mesoporous WO3 Thin Film
L. G. Teoh,a I. M. Hung,a J. Shieh,b W. H. Lai,a and M. H. Hona,z
a

Department of Materials Science and Engineering, National Cheng Kung University, Tainan 70101, Taiwan
National Nano Device Laboratories, Hsinchu 30050, Taiwan

A NO2 gas sensor based on mesoporous WO3 thin film with low operating temperatures and its sensing characteristics are
reported. The mesoporous WO3 thin film exhibits regular pores with an average pore size of 5 nm and specific surface area of 151
m2/g. Excellent sensing properties are found upon exposure to 3 ppm of NO2 at 35-100C for mesoporous WO3 thin film. The
sensor response is 180 for 3 ppm NO2 at 100C. The ability to sense NO2 at such low temperatures is attributed to the large surface
area 151 m2/g that offers many active sites for reaction with NO2 molecules.
2003 The Electrochemical Society. DOI: 10.1149/1.1585252 All rights reserved.
Manuscript submitted December 1, 2002; revised manuscript received March 21, 2003. Available electronically May 27, 2003.

A semiconductor-type metal oxide can respond to the presence of


a specific gas with a change of its electronic conductivity.1 Semiconductor sensors based on this property have been used extensively
for chemical and gas detection, which include toxic gases (NO2 ,
CO, H2 S) and combustible gases (H2 , CH4 , alcohol.2 Metal oxide
materials, such as WO3 , 3,4 SnO2 , 5,6 and TiO2 , 7 have been examined for such gas sensing.
NO2 is a toxic gas, often produced in automotive emissions,
combustion, or natural gas manufacturing. The threshold limited
value TLV for NO2 is 3 ppm. When the concentration of NO2 is
higher than 20 ppm, it is dangerous to the human body and is
thought to cause asthma at significantly lower levels. Reliable, lowcost NO2 sensors with high sensitivity, selectivity, and low energy
consumption are in high demand for environmental safety and industrial control.
WO3 -based gas sensors have been studied extensively since the
1960s.8 In 1967, Shaver9 discovered that the conductivity of WO3
thin film is increased under a low H2 concentration at a high temperature. WO3 thin film was used as the H2 gas detector thereafter.
Recently, it has been reported that WO3 materials have a good selectivity for low concentration of NO2 gas10 and sensor response
increases steeply with a decrease of grain size.11 The advantages of
the sensor fabricated by WO3 are high sensitivity, simple design,
and low cost. However, these sensors have good sensitivity at high
temperature between 200 and 500C12,13 due to the enhancement of
the chemical reaction between gas molecules and WO3 . One current
trend of the WO3 sensor is directed toward fabrication of a high
sensitivity sensor at lower temperatures.
Since the discovery of mesoporous MCM-41 materials by
Kresge et al.,14 the self-assembly mechanisms utilized to synthesize
the mesoporous materials with the advantage of increasing specific
surface areas have been widely investigated. Stucky et al.15,16 has
established that surfactant templating methods can be extended to
other metal oxides such as WO3 . Recently, we have prepared a
WO3 thin-film gas sensor with nanocrystallitc characteristic by solgel dip-coating.17 In this study, the method developed by Stucky
et al. was adopted to add block copolymer surfactant in our previous
sol-gel process to increase the specific surface area further and to
improve the sensing properties at operating temperatures below
100C.
Experimental
Syntheses.In a typical synthesis, 0.5 g of polyalkylene oxide
block copolymer BASF Pluronic EO133PO50EO133 F108 was dissolved in 10 g of ethanol. To this solution, 0.01 mol of the anhydrous inorganic chloride precursor, WCl6 Aldrich was added with

E-mail: mhhon@mail.ncku.edu.tw

vigorously stirring. The resulting sol solution was gelled in an open


Petri dish at 60C in air. The sol solution can be used to prepare thin
films on Al2 O3 substrate 6 4 mm which coated with Pt electrode by dip coating. The thickness of the thin films was 0.8 m.
Then, the thin films were dried at 60C for several hours. The asmade bulk samples or thin films were calcined at 250C for 5 h and
then washed by ethanol to remove the residual block copolymer.
Characterization.X-ray powder diffraction XRD patterns
were obtained on a Rigaku D/max-IV diffractometer using Cu K
radiation with Ni filter. Transmission electron microscopy TEM
studies were carried out on a Hitachi model HF- 2000 electron microscope operating at 200 keV. The samples for TEM were prepared
by directly dispersing the fine powders of the product onto 200 mesh
Cu grids. The morphology of mesoporous WO3 films was observed
by scanning electron microscope SEM, Philips XL-40 FEG. The
nitrogen adsorption and desorption isotherms at 77 K were measured using a Micromeritics ASAP 2010 system after the samples
were vacuum-dried at 150C for 10 h in N2 atmosphere. BrunauerEmmett-Teller BET surface areas were estimated over a relative
pressure ( P/ P 0 ) range from 0 to 1.0. Pore size distribution was
obtained from the analysis of the adsorption branch of the isotherms
using the Barrett-Joyner-Halenda BJH model. The pore volume
was taken at the P/ P 0 0.983 signal point.
Gas-sensing properties.The resistance of the films was obtained by measuring the current through the film at a constant voltage of 1 V and recorded by a multimeter HP 3458A. The samples
under test were placed in a quartz chamber 85 cm3 and exposed to
3 ppm NO2 gas. Gas sensing properties of the films were studied at
various operating temperatures T g in the 35C T g 100C
range. The sensor response is defined as R gas /R air , where R gas and
R air are the electric resistance in the NO2 gas and in clean air, respectively.
Results and Discussion
Figure 1 shows XRD patterns of mesoporous WO3 thin film
calcined at 250C for 5 h on Al2 O3 substrate. The diffraction peaks
of the WO3 thin films are assigned based on monoclinic structure
JCPDS card no. 83-0951. The perfect crystal structure of WO3 is
ReO3 type with the corner-sharing packing of WO6 6 octahedra,
which is distorted as the fabrication temperature is low. Mesoporous
WO3 film is strongly 020 oriented and this preferred orientation
may be derived from the disordered cubic phase. The relative intensity of 002, 020, and 200 peaks different from the monoclinic
phase with a more symmetrical profile indicates that the monoclinic
lattice is distorted. Depero et al.18 demonstrated that this cubic-like
XRD pattern may result from a static cubic disorder introduction as
edge-sharing octahedra randomly distributed in the lattice by oxygen
deficiency. For mesoporous materials, the lattice distortion may also

Downloaded 12 Nov 2009 to 140.116.208.44. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

Electrochemical and Solid-State Letters, 6 8 G108-G111 2003

G109

Figure 3. TEM image of the mesoporous WO3 film calcined at 250C for
5 h.
Figure 1. XRD pattern for the mesoporous WO3 thin film calcined at 250C
for 5 h.

arise from the short-range order in the pore wall of worm-like morphology. After employing Scherrers formula, the calculated grain
size of the thin film calcined at 250C for 5 h is approximately 4
nm. Yamazoe and Miura19 had shown that small grain size twice
the Debye length, L D , the smaller grain size increases gas sensor
response because the diameter is comparable with or less than the
space charge region of the grain is required for sensor applications.
Figure 2 shows the SEM micrograph of the WO3 thin film calcined at 250C for 5 h. The thin film exhibits a very porous structure
with a spherical powder of approximately 1 m. The larger scale
compared to TEM and XRD results may be attributed to the agglomeration of grains during the gelation and drying processes. Such a

Figure 2. SEM micrograph of the mesoporous WO3 thin film calcined at


250C for 5 h.

film morphology clearly can facilitate film electron flow. Because


NO2 gas can diffuse easily on WO3 thin film with voids, it is believed that this type of film offers a good sensor response to NO2
gas. Figure 3 shows a TEM image of the WO3 mesophase, which
indicates that the pores are not ordered worm-like morphology,
although they do exhibit a fairly regular pore diameter and have a
mean size of 5 nm. The size of the mesopores estimated by TEM
is in agreement with the values determined from adsorption data
see BET. Selected-area electron diffraction patterns recorded on
mesoporous WO3 confirm that the walls of our material are made up

Figure 4. N2 adsorption -desorption isotherms and BJH pore size


distribution plots for the calcined mesoporous WO3 .

Downloaded 12 Nov 2009 to 140.116.208.44. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

G110

Electrochemical and Solid-State Letters, 6 8 G108-G111 2003


recovery time time for 90% of resistance change for NO2 /air mixture for 3 ppm concentration at 50C is about 4-6 min with good
reproducibility. The longer recovery time needed indicates that the
residual gas is captured in the mesopores. For comparison, a thick
film type WO3 20 by screen printing operated at 100C for detecting
100 ppm of NO2 with a sensor response of 200 and a thermally
evaporated WO3 21 thin films operated at 350C for detecting 3 ppm
of NO2 with a sensor response 10. Thus, the mesoporous WO3
sensors have the advantage of 100C temperature operation for detecting 3 ppm with sensor response up to 180 over these materials.
Good sensor response is observed with a mesoporous WO3 gas
sensor having high surface area and small grain size. This result can
be attributed to two probable reasons.
1. Due to the large surface area of the mesoporous WO3 thin
film, the sensors fabricated in this work possess many active sites.
Li and Kawi22 had proposed a viewpoint, where a linear relationship
was found between the surface areas of SnO2 sensors and their
sensitivities to 500 ppm of H2 .
2. The carrier depletion is suggested to correspond to the sensing
mechanism of the semiconductor.23 As the grain size is reduced to
the Debye length, the carrier is depleted in the whole grain and the
sensor response can be improved according to the grain control
model. Therefore, the grain size of 4 nm for the mesoporous WO3
thin film used in the studies, a size closer to the Debye length of
WO3 , may contribute to the improvement of the sensor response.
It is also considered that NO2 gas adsorbed on the WO3 surface
and penetrated through the pore network of worm-like arrangement
may be a continuous network. According to Pauly et al.,24 the
wormhole channel motif is a potentially important structural feature
for favorable catalytic reactivity because channel branching within
the framework can facilitate access to reactive sites on the framework walls.
Conclusions

Figure 5. a Sensor response to 3 ppm NO2 from 35 to 100C of the


mesoporous WO3 thin film calcined at 250C for 5 h, b electrical response
of the sensor to 3 ppm NO2 at 50C.

of nanocrystalline oxides that show characteristic diffuse electron


diffraction rings. The results lead to a conclusion that the crystallized WO3 is essential for obtaining good sensor response or expected to afford a better sensor response toward gas sensing reactions.
Nitrogen adsorption-desorption isotherms exhibiting a type IV
curve are shown in Fig. 4 which is characteristic of mesoporous
WO3 . 15 The calculated pore diameters by the BJH method are in the
mean pore size of 5 nm Fig. 4 inset, which fits well to the observed
pore size from TEM. The surface area calculated for the worm-like
thin film is 151 m2/g, demonstrating the accessibility of the mesophase porosity.
Sensor response of a mesoporous WO3 thin film calcined at
250C for 5 h to 3 ppm NO2 from 35 to 100C is shown in Fig. 5a
in which the sensor response increases with increasing operating
temperature. The characteristic behavior of the resistance increase of
the sensor upon detecting NO2 is typical for n-type semiconductor
oxide gas sensors. Figure 5b presents the response and recovery
transients of the sensor to 3 ppm NO2 at 50C. The response and

Sol-gel dip-coated mesoporous WO3 gas sensors with high specific surface area, nanocrystallite characteristic, and disordered cubic structure were obtained as identified by XRD, SEM, TEM, and
BET analyses. The mesoporous WO3 film calcined at 250C for 5 h
and operated at 100C exhibits a sensor response of 180 for 3 ppm
NO2 . The large surface area and small grain size are the main reasons corresponded to this sensor response. The experimental results
also indicate the potential of using mesoporous WO3 for gas sensing.
Acknowledgments
This work was financially supported by the National Science
Council of Taiwan, ROC, grant no. NSC 90-2216-E-006-064, which
is gratefully acknowledged.
National Cheng Kung University assisted in meeting the publication
costs of this article.

References
1. P. B. Weisz, J. Chem. Phys., 21, 1531 1953.
2. T. G. Nevol and S. P. Yordanov, Ceramic Sensors-Technology and Applications,
Technomic Publishing, Lancaster, PA 1996.
3. J. I. Yang, H. Lim, and S. D. Han, Sens. Actuators B, 60, 71 1999.
4. M. D. Antonik, J. E. Schneider, E. L. Wittman, K. Snow, J. F. Vetelino, and R. J.
Lad, Thin Solid Films, 256, 247 1995.
5. T. Mochida, K. Kikuchi, T. Kondo, H. Ueno, and Y. Matsuura, Sens. Actuators B,
24, 433 1995.
6. G. Sberveglieri, S. Groppelli, P. Nelli, and C. Perego, Sens. Actuators B, 15, 86
1993.
7. H. M. Lin, T. Y. Hsu, C. Y. Tung, and C. M. Hsu, Nanostruct. Mater., 6, 1001
1995.
8. T. Seiyama, A. Kato, K. Fujiishi, and M. Nagatani, Anal. Chem., 34, 1502 1962.
9. P. J. Shaver, Appl. Phys. Lett., 11, 255 1967.
10. M. Akiyama, J. Tamaki, N. Miura, and N. Yamazoe, Chem. Lett., 1991, 1611.
11. Y. Shimizu and M. Egashira, MRS Bull., 24, 18 1999.
12. J. L. Solis, A. Hoel, L. B. Kish, and C. G. Granqvist, J. Am. Ceram. Soc., 84, 588
2001.

Downloaded 12 Nov 2009 to 140.116.208.44. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

Electrochemical and Solid-State Letters, 6 8 G108-G111 2003


13. H. T. Sun, C. Cantalini, L. Lozzi, M. Passacantando, and S. Santucci, Thin Solid
Films, 287, 258 1996.
14. C. T. Kresge, M. E. Leonowicz, W. J. Roth, J. C. Vartuli, and J. S. Beck, Nature
(London), 359, 710 1992.
15. P. Yang, D. Zhao, D. I. Margolese, and G. D. Stucky, Nature (London), 396, 152
1998.
16. D. Zhao, J. Feng, Q. Huo, N. Melosh, G. H. Fredrickson, B. F. Chmelka, and G. D.
Stucky, Science, 279, 548 1998.
17. J. Shieh, H. M. Feng, M. H. Hon, and H. Y. Juang, Sens. Actuators B, 86, 75
2002.

G111

18. L. E. Depero, S. Groppelli, G. Sberveglieri, and E. Tondello, J. Solid State Chem.,


121, 379 1996.
19. N. Yamazoe and N. Miura, in Chemical Sensor Technology, Vol. 4, S. Yamauchi,
Editor, Kodansha, Tokyo 1992.
20. Y. K. Chung, M. H. Kim, and K. W. Chung, Sens. Actuators B, 60, 49 1999.
21. H. T. Sun, C. Cantalini, and M. Pelino, Sens. Actuators B, 287, 258 1996.
22. G. J. Li and S. Kawi, Mater. Lett., 34, 99 1998.
23. C. Xu, J. Tamaki, N. Miura, and N. Yamazoe, Sens. Actuators B, 3, 147 1991.
24. T. R. Pauly, T. J. Pinnavaia, and S. J. L. Billinge, J. Am. Chem. Soc., 121, 8835
1999.

Downloaded 12 Nov 2009 to 140.116.208.44. Redistribution subject to ECS license or copyright; see http://www.ecsdl.org/terms_use.jsp

Vous aimerez peut-être aussi