Sei sulla pagina 1di 9

Sensors and Actuators B 158 (2011) 199207

Contents lists available at ScienceDirect

Sensors and Actuators B: Chemical


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

Tincopper mixed metal oxide nanowires: Synthesis and sensor response to chemical vapors
Xiaopeng Li a , Zhiyong Gu a, , JungHwan Cho b , Hongwei Sun c , Pradeep Kurup b
a b c

Department of Chemical Engineering, University of Massachusetts Lowell, One University Ave., Lowell, MA 01854, USA Department of Civil and Environmental Engineering, University of Massachusetts Lowell, One University Ave., Lowell, MA 01854, USA Department of Mechanical Engineering, University of Massachusetts Lowell, One University Ave., Lowell, MA 01854, USA

a r t i c l e

i n f o

a b s t r a c t
Tincopper mixed metal oxide nanowires were successfully prepared by thermally oxidizing electrodeposited metallic nanowires (Sn8 at.% Cu, Sn43 at.% Cu and Sn86 at.% Cu). The structure and composition of these nanowires before and after thermal oxidation were characterized by scanning electron microscopy (SEM), energy dispersive X-ray spectrometry (EDS), and X-ray diffraction (XRD). Dielectrophoresis was utilized to align the nanowires in contact with pre-fabricated interdigitated electrodes to form a chemiresistive gas sensor circuit. The sensitivity variation of the nanowires with different compositions was tested with acetone, ethanol and ethyl acetate vapors at different concentration levels, and the temperature effect was studied at ve operating temperatures, ranging from 200 C to 440 C. All the three mixed metal oxide nanowire sensors exhibited higher sensitivity than that of pure tin oxide nanowire sensor. The sensor performance was also investigated in terms of response/recovery time and repeatability. An interesting positive/negative response was observed by varying the element composition of the mixed oxide nanowires. 2011 Elsevier B.V. All rights reserved.

Article history: Received 30 December 2010 Received in revised form 27 May 2011 Accepted 1 June 2011 Available online 12 June 2011 Keywords: Nanowire Mixed metal oxide Synthesis Chemical gas sensor

1. Introduction The concept of using semiconducting metal oxide material as a detector for gaseous component was proposed as early as in 1966 by Seiyama and Kagawa [1]. Ever since then, various metal oxides such as SnO2 [2], CuO/CuO2 [3], TiO2 [4], In2 O3 [5], NiO [6], Co3 O4 [7] and ZnO [8], have been synthesized and the semiconductor typed gas sensors were used for automotive [9], indoor air quality control [10] and homeland security [11]. However, these sensing materials are normally used in the form of bulk materials or thin lms in conventional sensors, which have draw backs in terms of large power consumption and long response time [12]. With the rising tide of nanotechnology, researchers nd that the sensing properties of nanostructured metal oxide materials differ substantially from bulk materials of the same element and composition, and may lead to sensors with high sensitivity, fast response/recovery time, low energy consumption and small in size making it ideal for portable devices. To date, semiconducting metal oxide nanostructures have been successfully synthesized and intensively studied. In addition, in order to further improve the sensor performance, many efforts have been devoted to the development of modied metal oxide materials featuring multiple elements instead of traditional single element metal oxide material.

Corresponding author. Tel.: +1 978 934 3540; fax: +1 978 934 3047. E-mail address: Zhiyong Gu@uml.edu (Z. Gu). 0925-4005/$ see front matter 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.snb.2011.06.004

Among them, p-type copper oxide and n-type tin oxide have drawn much attention due to their desirable electrical, ferromagnetic [13], and chemical properties exhibiting high sensitivity [14,15] toward gas sensing, with promising applications in catalyst [16], bio/chemical sensors [17,18] and eld effect transistors [1921]. Currently the major methods of creating tincopper (SnCu) mixed oxide materials include: (1) doping a small amount of copper into tin oxide [22,23] and (2) fabricating heterostructured copper oxide/tin oxide composites [2426]. In most cases above, tin oxide is chosen to be the primary component while the molar percentage of copper is controlled at a relatively low level. However, the high mutual solubility of copper in tin makes it possible to expand the copper fraction, even more than 50%. This may signicantly expand the sensing window and provide more freedom for sensing ability. In this article, we report a unique method to synthesize mixed metal oxide nanowires with a wide range of composition which started with synthesis of electrodeposited tincopper metallic nanowires, followed by a thermal oxidation process. Tincopper mixed metal oxide nanowires with different compositions were investigated, and dramatic surface morphology difference was observed from samples with varying compositions. Further exploration on their sensor performance was carried out by both static and dynamic method, to reveal the sensitivity and repeatability of these materials toward chemical exposure. The result also showed an interesting positive/negative response by varying the element composition of the mixed oxide nanowires.

200

X. Li et al. / Sensors and Actuators B 158 (2011) 199207

Fig. 1. Schematic diagram of sensor chip fabricating procedure and top-view of the sensor chip surface (SEM image, bottom left).

2. Experimental procedures 2.1. Nanowire synthesis and characterization Polycarbonate (PC) membrane with porous structure of 50 nm each in diameter (nominal size) was utilized to fabricate metallic nanowires. One side of this membrane was coated with a thin layer of silver (200 nm) and put into contact with a copper plate as cathode, while the other side was immersed in the plating solution with a platinum anode. The growth of nanowires inside the membrane was achieved by applying current between the electrodes a typical electrodeposition process. Later the PC membrane was dissolved in dichloromethane to release nanowires. The detailed fabrication process has been described in a previous paper by our group [27]. The plating solution was prepared by mixing commercial plating solution of copper (Cu) (Copper U Bath RTU, Technic, Inc.) and tin (Sn) (Techni Tin) together with a corresponding make-up solution and antioxidant (obtained from Techni, Inc. too). Because copper and tin can be co-deposited at any ratio [28], we made appropriate adjustment to the electroplating bath solution to control the composition. As a result, mixed metallic nanowires of Sn8 at.% Cu, Sn43 at.% Cu and Sn86 at.% Cu were fabricated under the conditions of applied current: 18 mA and time duration: 70 s. The next step was to convert tincopper metallic nanowires into their corresponding oxides. For that purpose, a thermal treatment was carried out in a tube furnace (Thermo Scientic Lindberg Blue M) at 600700 C for 5 h with two ends of the quartz tube open to the air to ensure sufcient oxygen circulation. After nanowire fabrication and treatment, the nanowires were characterized using a JEOL JSM-7401F eld emission scanning electron microscope (FE-SEM) and an EDAX Genesis V4.61 energy disperse X-ray detector before and after thermal oxidation, for morphology information and elementary analysis. X-ray diffraction (XRD) patterns were obtained using Oxford diffraction Xcalibur PX Ultra with ONYX detector for crystal phase study.

2.2. Nanowire sensor integration A schematic diagram is given in Fig. 1 to depict the nanowire sensor fabrication process. In detailed description, after the nanowires were synthesized and released from the PC membrane they were subjected to several rinsewash cycles and then stored in ethanol. A drop of this nanowire suspension was transferred to a commercially available micro-electrode substrate (Platinum based materials) with an embedded heater (Heraus, MSP332) under the inuence of dielectrophoresis (DEP). The interdigitated electrodes were connected to a function generator (Tektronix FG502) which gave a sine wave output as the source of the non-uniform AC electric eld and dielectrophoretic force. The parameters for this DEP alignment were tuned up rst and nally xed at 6 MHz frequency and 6 V peak to peak amplitude to make sure that most of the nanowires were aligned, bridging between the interdigitated electrodes after ethanol evaporated. Then thermal oxidation was conducted by placing the sensor chip in the tube furnace at 600700 C for 5 h. The inset picture in Fig. 1 shows an SEM image of the top-view of the sensor chip surface with nanowires assembled on the interdigitated electrodes. 2.3. Gas sensing experiments The gas sensing measurements were conducted in the static and dynamic modes. The experimental set-up for the static method consists of a 4.4 L glass chamber, a sensor board mounting four sensor chips with different oxide compositions, and a data acquisition device (USB-1608). The sensor board was sealed inside the chamber with all the connection wires going out through an airtight port and connected to the power supply and data acquisition system. The chamber was purged with dry air before testing and during testing. A certain volume of chemical vapor was extracted from a sample container using a syringe. The sample container was placed in a water bath at constant temperature to maintain a constant equi-

X. Li et al. / Sensors and Actuators B 158 (2011) 199207

201

Fig. 2. (A) SEM images of as-fabricated tincopper metallic nanowires and a zoomin view of a single nanowire (inset); (B) EDS spectra layout of tincopper metallic nanowires with three different compositions.

librium vapor pressure. The extracted vapor sample was injected into the testing chamber. Repeating this injecting operation caused a series of step increases in concentration, and the sensor response was recorded by a computer. Dynamic test set-up had two gas ow paths controlled by two mass ow controllers (MFCs). One was directly connected to a carrier gas (dry air) tank and the other went through a bubbler with the liquid chemical inside. The dynamic sensor responses were obtained from a measurement sequence consisting of vapor delivery and dry-air purging by controlling two three-way valves connected to the bubbler and MFCs. The total measurement sequences and their operation times were controlled by a virtual instrument (VI) programmed using LabVIEW. According to the measurement sequence, the dynamic sensor responses were acquired and sent to computer using a data acquisition and control system developed using a microcontroller. 3. Results and discussion Morphological characteristics of as-fabricated tincopper metallic nanowires are shown in Fig. 2A. Nanowires appear to be very uniform, and the typical dimensions of nanowires are 56 m in length and 5070 nm in diameter. The co-existence of tin and copper elements in the original metallic nanowires was conrmed

by energy-dispersive X-ray spectroscopy (EDS) analysis as a spectra layout shown in Fig. 2B. The EDS analysis also provides quantitative information of copper and tin in the nanowires. According to their atomic fractions of each element, the three types of nanowires were labeled as Sn14Cu86, Sn57Cu43 and Sn92Cu8 according to their atomic fractions. Compositional information after thermal oxidation is shown in Fig. 3. The EDS mapping results (Fig. 3A) reveal three major elements colored differently (O, Cu, Sn) and the spectrum (Fig. 3B) shows a strong oxygen peak in addition to the peaks of the metal elements. In general, by performing EDS analysis in spot mode along the entire length of the nanowire and then averaging the measurements, all three samples have an oxygen fraction slightly lower than the highest amount (stoichiometric ratio) that we can expect for tin dioxide (SnO2 ) and copper oxide (CuO), indicating a close-to-complete oxidation. In addition to EDS, the X-ray diffraction patterns of nanowire samples are shown in Fig. 4, where crystallographic information was studied to verify the compounds existing in this mixed metal/metal oxide system. The peak intensity and 2 angle of possible compounds obtained from Powder Diffraction File (International Centre for Diffraction Data (ICDD)) are displayed at the bottom as reference. Fig. 4A lists the XRD spectra of nanowire samples before oxidation, from top to bottom: sample Sn92Cu8, sample Sn57Cu43, and sample Sn14Cu86. In sample Sn92Cu8, except for the characteristic peaks of tin, only slight amount of copper and Cu3 Sn appear. As the atomic fraction of copper goes up, in sample Sn57Cu43, peak intensity of copper and intermetallic compound Cu3 Sn increases, while that of tin lowers in comparison to sample Sn92Cu8. It was also found that small peaks of intermetallic compound Cu6 Sn5 showed up, indicating the second intermetallic compound formation in this sample. For sample Sn14Cu86, tin peaks further decrease, while copper peak intensity shows a noticeable increase as well as intermetallic compound Cu6 Sn5 and Cu3 Sn. To sum up, for metallic nanowires, major peaks are identied to be the characteristic peaks of tin and copper, and trend of their intensity differing is in good accordance with tincopper composition ratio in each sample. Intermetallic compound Cu3 Sn was observed at low copper concentration; while both Cu3 Sn and Cu6 Sn5 were observed at medium to high copper concentration. The intermetallic compound peak intensity increases along with the increase of copper ratio in the samples, indicating a higher concentration of both intermetallic compounds. In the layout of XRD spectra after thermal oxidation (Fig. 4B), strong peaks of SnO2 were observed in oxidized sample Sn92Cu8. Small amount of intermediate oxides SnO and Sn2 O3 were found co-existing. The evidence of CuO in the mixed metal oxide system shows up as peaks with relatively small intensity. No observable Cu2 O peak was found. In oxidized sample Sn57Cu43, the characteristic peaks of Cu2 O appear in small intensity, and the height of CuO peaks increases due to higher copper ratio. In oxidized sample Sn14Cu86, the peak intensity of two copper oxides (CuO and Cu2 O) further increases. The existence of Cu2 O, Sn2 O3 and SnO help explain the measured oxygen deciency from the EDS (none stoichiometric, complete oxidation). It is very common to nd intermediate oxides co-existing since they are the main contributors to the electron donor/acceptor of bulk metal oxide; for example, in some studies tin oxide may also be referred as SnOx /SnO2x [29,30], or Snx Oy [31]. Morphological details of each type of nanowire surface after oxidation are shown in Fig. 5. These high-magnication images show that after oxidation, the nanowires have bumpy surfaces decorated with island structures, which are very different from the smooth surfaces before oxidation. Similar bumpy surfaces have been reported for several metal oxide nanowires [32,33]. We believe that the growth and relief of stresses generated during oxidation

202

X. Li et al. / Sensors and Actuators B 158 (2011) 199207

Fig. 3. (A) SEM image of tincopper mixed oxide nanowires (sample Sn92Cu8) and EDS mapping of three major elements (Sn, Cu, and O) on nanowires, and (B) EDS spectrum of the same sample.

may have contributed to the formation of the irregular surfaces. The stresses may come from recrystallization, non-stoichiometric growth [34] and the volume difference between the oxide and the metal from which it forms. From the phase diagram, tin and copper alloy system has various phases such as Cu6 Sn5 , Cu3 Sn, Cu10 Sn3 , Cu41 Sn11 , etc.; however, there are only four phases Sn, Cu, Cu6 Sn5 and Cu3 Sn at the room temperature, which compare well with our experimental observation. Due to the limited solubility, the co-existence of multiple solid phases is inevitable during the electroplating, and there has to be an excessive amount of tin or copper not incorporated with the other element so that grains of only tin or copper atoms may form all over the nanowire. The existence of some specic sites on the surface where copper containing compound or pure copper grains may lead to the formation of oxide islands after thermal oxidation. This is a very common phenomenon for copper [35,36] and copper containing composites [37]. Also, for those incorporated atoms, the element which is minority in atomic fraction would be considered a substitutional impurity in the fcc structure [38], and such impurities, especially the one on or close to the surfaces may greatly inuence the oxide nucleation, growth and hence the surface roughness [39]. As a comparison, bare tin nanowire synthesized through similar procedures is presented as a control sample and its surface after oxidation is relatively smooth (Fig. 5D). Obviously, the introduction of copper element indeed has a dramatic effect during the reconstruction and conversion of metal oxides. The surface change becomes vigorous comparing to the mildness of tin oxide surface. It should be noted that oxidized sample Sn92Cu8 has island structures of largest size

and oxidized sample Sn14Cu86 has a higher number of uniform island structures. Although the oxidation is an important process and the famous CabreraMott model [40,41] has shown considerable predictive power for projecting chemical order of oxidation of metals and alloys on a macro-scale, yet no atomic level details of the growth mechanism are available [42]. So the fully understanding of what exactly causes the surface change in such a manner still needs to be further investigated. However, it should be noted that such surface roughness features will increase the total surface area of the nanowires, and this may be in favor of their sensing performance. The response or sensitivity of sensor chips toward chemical vapors is measured in terms of nanowires resistance change and is dened as (Rg R0 )/R0 , also expressed as R/R0 , where R0 is the overall resistance measured from the sensor chip in dry air condition and Rg is the resistance when a chemical vapor is in presence. Our research began with investigating how the sensors behave upon exposure to ethanol (Fig. 6A) at a xed operating temperature of 380 C and the summarized data of these nanowires with three compositions were plotted as a function of concentration (Fig. 6B). Fig. 6A is the static testing result from sample Sn14Cu86 when responding to a series of concentration step changes from 1 ppm to 500 ppm. Apparently, the change of resistance and the concentration is not in a linear relationship, which resembles the general trend of metal oxide sensors with increasing concentration. At lower concentration region (<100 ppm), the response caused by injection of ethanol/ppm is much higher than that in higher concentration region (>100 ppm). Noticeably, this nanowire had

X. Li et al. / Sensors and Actuators B 158 (2011) 199207

203

Fig. 4. X-ray diffraction patterns of nanowire samples (A) before and (B) after oxidation, with reference patterns of possible compounds from Powder Diffraction File.

about 20% augments in resistance at a few ppm level which indicated that those sensor can measure ethanol concentration well below ppm level. Fig. 6B shows the sensor responses of the three mixed metal oxide nanowire sensors and that of tin oxide nanowire sensor to ethanol ranging from 1 ppm to 500 ppm. The tin oxide nanowire sensor, as a control sample, was synthesized using the same method, and the response data were gained from our previous work [43]. All of three copper-containing metal oxides have exhibited larger response than pure SnO2 nanowires at the same concentration level. The sensor response enhancement is signicant for sample Sn92Cu8 and Sn14Cu86 and moderate for sample Sn57Cu43. Interestingly, both positive and negtive responses were observed here. For tin dominated compositions, the resistance of nanowires decreased as the concentration went up, and the response was larger with more copper element. However, for sample Sn14Cu86, copper dominated nanowire, its resistance increased with increasing concentration. Another critical parameter working temperature was also investigated in the testing toward three chemicals including ethanol, acetone, and ethyl acetate. Five temperature points in the range of 200440 C were chosen to help understand how the sensitivity changes with the working temperature. The response data were compared in absolute values and put into radar charts to provide a straightforward view of the results. In Fig. 7, radar charts have been adopted in sensor related studies for displaying multiple variables in a single gure [44,45]. Here in our study, each chart consists of ve scaled spokes representing the operating temperature, and the length of the data point is proportional to the magnitude of the sensor response. In Fig. 7A, the response of three nanowire sensors to ethanol at 500 ppm are presented. Sample Sn57Cu43 displays the lowest sensitivity among the three nanowire sensors, while in most temperature points, sample Sn14Cu86 gives the highest, except for 440 C, where the sensor response of sample Sn92Cu slightly

Fig. 5. SEM images of surface morphology of mixed metal oxide nanowire samples (A) Sn92Cu and (B) Sn57Cu43 and (C) Sn14Cu86 and (D) SnO2 nanowires.

204

X. Li et al. / Sensors and Actuators B 158 (2011) 199207

Fig. 6. (A) Static testing results of sample Sn14Cu86 nanowire sensor responding to a series of step changes with increasing ethanol concentration and (B) sensor response of three types of mixed metal oxide nanowire sensors and that of tin oxide nanowire sensor for ethanol from 1 ppm to 500 ppm.

exceeds sample Sn14Cu86. It is interesting to notice that Sn92Cu8 metal oxide nanowires behaves most like tin oxide nanowire with respect to the temperature dependence [43]. The sensitivity of sample Sn92Cu8 changes positively with the increased working temperatures up to 440 C. The other two tincopper oxide nanowires have a preferable working temperature region from 320 C to 380 C. By conducting the same experiments to acetone and ethyl acetate exposure, these three nanowire sensors were found to have similar trends of sensor response. For all three chemicals tested, the highest sensor response was evoked by exposing sample Sn14Cu86 (working temperature 380 C) to acetone at 500 ppm, due to its p-type sensor response, and there is a resistance increase over 100%. For n-type response, sample Sn92Cu8 (working temperature 440 C) gives an over 80% change when exposed to acetone. Generally speaking, the sensitivity order of each nanowire sensor toward these three chemicals is: for sample Sn92Cu8 and Sn57Cu43, acetone > ethyl acetate > ethanol; for sample Sn14Cu86, acetone > ethanol > ethyl acetate. Despite the difference in sensitivity, the sensing mechanisms of these three chemicals are very similar [4648]; it basically involved the reactions of chemical molecules with pre-adsorbed oxygen species, regardless the materials [4951]. For these three reducing analytes, when oxidized by the oxygen ions, the electrical disturbance is mainly from the OH bond-breaking process [52,53]. From our testing results, acetone appeared to be the most active over all, and ethanol caused higher response than ethyl acetate in 200380 C, and lower in 380440 C. It is also well known that in low temperature region, physisorption is dominant, while chemisorption of oxygen gradually takes charge

Fig. 7. The response variation of three types of mixed metal oxide nanowire sensors for ethanol, acetone and ethyl acetate in correlation with ve working temperatures within the range of 200440 C.

when temperature increases [54]. The combination of them determines the sensitivity and resistivity of the material. Another way to look at the temperature effect and composition effect on sensitivity is that, for three mixed metal oxide materials, the surface adsorption of oxygen species (O2 , O2 and O ) [55] might be different. However, the detailed mechanism behind the preference of each chemical at different temperatures and the exact chemi- vs. physisorption competition of oxygen species on metal oxide material surface need to be further investigated. In the case of metal oxide chemiresistive sensors, the most critical property measured in testing is their resistance/conductance change when exposed to various chemicals. It is well known that the resistance of semiconducting materials is determined by the number of charge carriers, either electrons (n-type) or holes (p-type). The reason that tin oxide is considered to be an ntype semiconductor is because of its oxygen-deciency a type of intrinsic defects which does not change the overall composition but may cause the semiconducting behavior [56]. On the other hand, oxygen-excess copper oxide shows p-type conductivity by holes [57]. In dry air condition, the absorbed surface oxygen plays a very important role in trapping electrons from the semiconductor [54,58]. With this extra amount of oxygen getting

X. Li et al. / Sensors and Actuators B 158 (2011) 199207

205

The response kinetics and the repeatability of the nanowire sensors were investigated through the dynamic testing. A full test cycle includes exposure of all three sensors to the gas ow of the same condition, then back to initial state. This cycle was repeated for 5 times. Fig. 8A shows the testing results of sample Sn92Cu8 upon exposure to three analytes at a concentration of roughly 500 ppm and a working temperature of 380 C. Fast responses were recorded and the typical response time (90% of the peak value) was calculated to be around 8 s, with a small variation of 3 s toward different chemicals and concentrations. Fig. 8B is a comparison of three types of nanowires responding to 500 ppm of ethyl acetate at a working temperature of 380 C. The intrinsic positive and negative resistance change were observed for different metal oxides, and good repeatability was demonstrated. The recovery process appeared to be relatively slow comparing to the fast response time, and it usually took about 6 min. The recovery time of samples with higher tin ratio (Sn92Cu8 and Sn57Cu43) was faster than that of the sample Sn14Cu86. Besides sensing property of nanowires, the box-shaped geometry of our test chamber may also affect the recovery time by providing a fraction of dead volume when it gets ushed with blank carrier gas. Optimization and better design of the testing chamber may decrease the response/recovery time of the nanowire sensors synthesized. 4. Conclusion Tincopper mixed metal oxide nanowires were synthesized by thermally oxidizing the tincopper metallic nanowires that have been obtained by electrodeposition method using nanoporous templates. The composition of prepared tin and copper elements can be tailored in a wide range through changing the electroplating solutions. Oxidation process causes a tremendous change on the surface condition of the nanowires with different compositions. XRD measurement indicated that two intermetallic compounds Cu6 Sn5 and Cu3 Sn exist in the mixed metal nanowires, together with metallic tin and copper; while for mixed metal oxide, the existence of SnO2 and CuO was found to be majority and intermediate oxides Cu2 O, SnO and Sn2 O3 were minority co-existing in the oxidized samples. Testing of these nanowires response to chemical vapors of acetone, ethanol and ethyl acetate revealed that the supplement of copper into the tin nanowires improved the sensitivity as compared to the pure tin oxide nanowires. An interesting n-type/p-type sensing properties was observed with the variation of tincopper composition. Signicant enhancement on sensitivity is observed on Sn92Cu8 and Sn14Cu86 metal oxide nanowires and both positive and negative resistance change were observed. Acknowledgements Financial support from the National Science Foundation (award number ECCS-0731125) is greatly acknowledged. We thank Ying Wang and Dr. Yu Lei at University of Connecticut for the help on XRD measurements. References
[1] T. Seiyama, S. Kagawa, Study on a detector for gaseous components using semiconductive thin lms, Anal. Chem. 38 (1966) 10691073. [2] N. Yamazoe, J. Fuchigami, M. Kishikawa, T. Seiyama, Interactions of tin oxide surface with O2 , H2 O and H2 , Surf. Sci. 86 (1979) 335344. [3] P. Samarasekara, N. Kumara, N. Yapa, Sputtered copper oxide (CuO) thin lms for gas sensor devices, J. Phys. Condens. Matter 18 (2006) 2417. [4] C. Garzella, E. Comini, E. Tempesti, C. Frigeri, G. Sberveglieri, TiO2 thin lms by a novel solgel processing for gas sensor applications, Sens. Actuators B 68 (2000) 189196. [5] A. Gurlo, M. Ivanovskaya, N. Barsan, M. Schweizer-Berberich, U. Weimar, W. Gopel, A. Dieguez, Grain size control in nanocrystalline In2 O3 semiconductor gas sensors, Sens. Actuators B 44 (1997) 327333.

Fig. 8. (A) The response of sample Sn92Cu8 for acetone, ethanol and ethyl acetate of 500 ppm obtained from the dynamic test and (B) the response of three types of mixed metal oxide nanowires toward ethyl acetate of 500 ppm from the dynamic test.

electrons from metal oxide materials, the concentration of holes (n-type)/electrons (p-type) essential charge carriers in the semiconductor will increase/decrease, hence the material become more/less conductive. It is generally accepted that, with the introduction of analytes, oxygen ions tend to react with the hydrogen dissociated from adsorbed organic molecules, releasing electrons back to the surface of metal oxide [59]. Thus, for n-type semiconducting metal oxide, the resistance drops when exposed to chemicals. The opposite holds for p-type semiconducting metal oxide. From our observation, among mixed metal oxide nanowire samples, two of them (Sn92Cu8, Sn57Cu43) with higher tin fraction behave like n-type while the sample Sn14Cu86 has a typical p-type response to chemical exposure. Small amount of copper incorporated in the crystal lattice is considered to be electroactive dopant to tin oxide [60], and its possessing acceptor action also contributes to the process of energy barrier formation by changing the electron concentration in the bulk of tin dioxide due to compensation mechanism [61]. It has been reported that the incorporation of Cu2+ ions onto Sn4+ sites leads to the creation of oxygen vacancies, and hence a decrease in the free electron concentration and an increase in sensitivity of the material [62]. This mechanism may also apply to the situation of small amount of tin incorporated in copper, as the substitution of Cu2+ by Sn4+ decreases the number of holes. Excessive amount of tin oxide is responsible for sensing, showing a similar but damped sensor response like the sample Sn92Cu8.

206

X. Li et al. / Sensors and Actuators B 158 (2011) 199207 [40] D. Cocke, T. Hess, D. Mencer, D. Naugle, Prediction of alloy component oxidation order from a kinetic model using free energies of oxide formation, J. Chem. Soc., Chem. Commun. 1994 (1994) 27652766. [41] N. Cabrera, N. Mott, Theory of the oxidation of metals, Rep. Prog. Phys. 12 (1949) 163184. [42] D. Cocke, G. Chuah, N. Kruse, J. Block, Copper oxidation and surface copper oxide stability investigated by pulsed eld desorption mass spectrometry, Appl. Surf. Sci. 84 (1995) 153161. [43] X. Li, E. Chin, H. Sun, P. Kurup, Z. Gu, Fabrication and integration of metal oxide nanowire sensors using dielectrophoretic assembly and improved postassembly processing, Sens. Actuators B 148 (2010) 404412. [44] S. Capone, P. Siciliano, N. Brsan, U. Weimar, L. Vasanelli, Analysis of CO and CH4 gas mixtures by using a micromachined sensor array, Sens. Actuators B 78 (2001) 4048. [45] T. Maekawa, K. Suzuki, T. Takada, T. Kobayashi, M. Egashira, Odor identication using a SnO2 -based sensor array, Sens. Actuators B 80 (2001) 5158. [46] D. Barreca, E. Comini, A. Gasparotto, C. Maccato, C. Sada, G. Sberveglieri, E. Tondello, Chemical vapor deposition of copper oxide lms and entangled quasi1D nanoarchitectures as innovative gas sensors, Sens. Actuators B 141 (2009) 270275. [47] D. Barreca, E. Comini, A. Ferrucci, A. Gasparotto, C. Maccato, C. Maragno, G. Sberveglieri, E. Tondello, First example of ZnOTiO2 nanocomposites by chemical vapor deposition: structure, morphology, composition, and gas sensing performances, Chem. Mater. 19 (2007) 56425649. [48] Q. Qi, T. Zhang, L. Liu, X. Zheng, Q. Yu, Y. Zeng, H. Yang, Selective acetone sensor based on dumbbell-like ZnO with rapid response and recovery, Sens. Actuators B 134 (2008) 166170. [49] X. Gou, G. Wang, J. Yang, J. Park, D. Wexler, Chemical synthesis, characterisation and gas sensing performance of copper oxide nanoribbons, J. Mater. Chem. 18 (2008) 965969. [50] C. Wang, X. Fu, X. Xue, Y. Wang, T. Wang, Surface accumulation conduction controlled sensing characteristic of p-type CuO nanorods induced by oxygen adsorption, Nanotechnology 18 (2007) 145506. [51] N. Funazaki, A. Hemmi, S. Ito, Y. Asano, Y. Yano, N. Miura, N. Yamazoe, Application of semiconductor gas sensor to quality control of meat freshness in food industry, Sens. Actuators B 25 (1995) 797800. [52] T. Jinkawa, G. Sakai, J. Tamaki, N. Miura, N. Yamazoe, Relationship between ethanol gas sensitivity and surface catalytic property of tin oxide sensors modied with acidic or basic oxides, J. Mol. Catal. A: Chem. 155 (2000) 193200. [53] M. Calatayud, J. Andres, A. Beltran, A theoretical analysis of adsorption and dissociation of CH3 OH on the stoichiometric SnO2 (1 1 0) surface, Surf. Sci. 430 (1999) 213222. [54] M. Batzill, U. Diebold, The surface and materials science of tin oxide, Prog. Surf. Sci. 79 (2005) 47154. [55] S. Chang, Oxygen chemisorption on tin oxide: correlation between electrical conductivity and EPR measurements, J. Vac. Sci. Technol. 17 (2009) 366369. [56] L. Smart, E. Moore, Solid State Chemistry: An Introduction, third ed., CRC Press, Boca Raton, 2005. [57] M. Jayalakshmi, K. Balasubramanian, Hydrothermal synthesis of CuOSnO2 and CuOSnO2 Fe2 O3 mixed oxides and their electrochemical characterization in neutral electrolyte, Int. J. Electrochem. Sci. 4 (2009) 571581. [58] N. Barsan, C. Simion, T. Heine, S. Pokhrel, U. Weimar, Modeling of sensing and transduction for p-type semiconducting metal oxide based gas sensors, J. Electroceram. (2009) 19. [59] H. Gong, Y. Wang, S. Teo, L. Huang, Interaction between thin-lm tin oxide gas sensor and ve organic vapors, Sens. Actuators B 54 (1999) 232235. [60] M. Rumyantseva, O. Safonova, M. Boulova, L. Ryabova, A. Gaskov, Dopants in nanocrystalline tin dioxide, Russ. Chem. Bull. 52 (2003) 12171238. [61] O. Safonova, M. Rumyantseva, L. Ryabova, M. Labeau, G. Delabouglise, A. Gaskov, Effect of combined Pd and Cu doping on microstructure, electrical and gas sensor properties of nanocrystalline tin dioxide, Mater. Sci. Eng. B 85 (2001) 4349. [62] M. Wagh, L. Patil, T. Seth, D. Amalnerkar, Surface cupricated SnO2 ZnO thick lms as a H2 S gas sensor, Mater. Chem. Phys. 84 (2004) 228233.

[6] I. Hotovy, V. Rehacek, P. Siciliano, S. Capone, L. Spiess, Sensing characteristics of NiO thin lms as NO2 gas sensor, Thin Solid Films 418 (2002) 915. [7] W. Li, L. Xu, J. Chen, Co3 O4 nanomaterials in lithium-ion batteries and gas sensors, Adv. Funct. Mater. 15 (2005) 851857. [8] J. Xu, Q. Pan, Y. Shun, Z. Tian, Grain size control and gas sensing properties of ZnO gas sensor, Sens. Actuators B 66 (2000) 277279. [9] L. Francioso, D. Presicce, A. Taurino, R. Rella, P. Siciliano, A. Ficarella, Automotive application of solgel TiO2 thin lm-based sensor for lambda measurement, Sens. Actuators B 95 (2003) 6672. [10] H. Yamaura, K. Moriya, N. Miura, N. Yamazoe, Mechanism of sensitivity promotion in CO sensor using indium oxide and cobalt oxide, Sens. Actuators B 65 (2000) 3941. [11] D. Lee, H. Jung, J. Lim, M. Lee, S. Ban, J. Huh, D. Lee, Explosive gas recognition system using thick lm sensor array and neural network, Sens. Actuators B 71 (2000) 9098. [12] P. Bhattacharyya, S. Sen, A. Chatterjee, S. Das, K. Basu, A. Pal, H. Saha, MEMS based nanocrystalline metal oxide gas sensors for coalmine environment, in: International Conference on MEMS and Semiconductor Nanotechnology, Kharagpur, India, 2005. [13] L. Li, K. Yu, Z. Tang, Z. Zhu, Q. Wan, Room-temperature ferromagnetism properties of Cu-doped SnO2 nanowires, J. Appl. Phys. 107 (2010) 014303. [14] V. Kumar, S. Sen, K. Muthe, N. Gaur, S. Gupta, J. Yakhmi, Copper doped SnO2 nanowires as highly sensitive H2 S gas sensor, Sens. Actuators B 138 (2009) 587590. [15] R. Niranjan, K. Patil, S. Sainkar, I. Mulla, High H2 S-sensitive copper-doped tin oxide thin lm, Mater. Chem. Phys. 80 (2003) 250256. [16] X. Zheng, S. Wang, S. Zhang, W. Huang, S. Wu, Copper oxide catalysts supported on ceria for low-temperature CO oxidation, Catal. Commun. 5 (2004) 729 732. [17] J. Tamaki, T. Maekawa, N. Miura, N. Yamazoe, CuOSnO2 element for highly sensitive and selective detection of H2 S, Sens. Actuators B 9 (1992) 197203. [18] E. Reitz, W. Jia, M. Gentile, Y. Wang, Y. Lei, CuO nanospheres based nonenzymatic glucose sensor, Electroanalysis 20 (2008) 24822486. [19] M. Arnold, P. Avouris, Z. Pan, Z. Wang, Field-effect transistors based on single semiconducting oxide nanobelts, J. Phys. Chem. B 107 (2003) 659663. [20] L. Liao, Z. Zhang, B. Yan, Z. Zheng, Q. Bao, T. Wu, C. Li, Z. Shen, J. Zhang, H. Gong, Multifunctional CuO nanowire devices: p-type eld effect transistors and CO gas sensors, Nanotechnology 20 (2009) 085203. [21] S. Kalinin, J. Shin, S. Jesse, D. Geohegan, A. Baddorf, Y. Lilach, M. Moskovits, A. Kolmakov, Electronic transport imaging in a multiwire SnO chemical eldeffect transistor device, J. Appl. Phys. 98 (2005) 044503. [22] Q. Wan, E. Dattoli, W. Lu, Doping-dependent electrical characteristics of SnO2 nanowires, Small 4 (2008) 451. [23] A. Galdikas, A. Mironas, Copper-doping level effect on sensitivity and selectivity of tin oxide thin-lm gas sensor, Sens. Actuators B 26 (1995) 2932. [24] R. Vasiliev, M. Rumyantseva, N. Yakovlev, A. Gaskov, CuO/SnO2 thin lm heterostructures as chemical sensors to H2 S, Sens. Actuators B 50 (1998) 186 193. [25] J. Tamaki, K. Shimanoe, Y. Yamada, Y. Yamamoto, N. Miura, N. Yamazoe, Dilute hydrogen sulde sensing properties of CuOSnO2 thin lm prepared by lowpressure evaporation method, Sens. Actuators B 49 (1998) 121125. [26] W. Yuanda, T. Maosong, H. Xiuli, Z. Yushu, D. Guorui, Thin lm sensors of SnO2 CuOSnO2 sandwich structure to H2 S, Sens. Actuators B 79 (2001) 187191. [27] F. Gao, S. Mukherjee, Q. Cui, Z. Gu, Synthesis, characterization, and thermal properties of nanoscale lead-free solders on multisegmented metal nanowires, J. Phys. Chem. C 113 (2009) 95469552. [28] V. Averkin, Electrodeposition of Alloys, Israel Program for Scientic Translations, Jerusalem, 1964. [29] J.N. Zemel, Theoretical description of gas-lm interaction on SnOx , Thin Solid Films 163 (1988) 189202. [30] K. Yoo, N. Cho, H. Song, H. Jung, Surface morphology and gas-sensing characteristics of SnO2x thin lms oxidized from Sn lms, Sens. Actuators B 25 (1995) 474477. [31] P. Jackson, K. Fisher, G. Willett, Monitoring reagent modication of charged Snx Oy nanoclusters using Fourier transform ion cyclotron mass spectrometry, Phys. Chem. Chem. Phys. 7 (2005) 16871693. [32] C. Kim, Y. Myung, Y. Cho, H. Kim, S. Park, J. Park, J. Kim, B. Kim, Electronic structure of vertically aligned Mn-doped CoFe2 O4 nanowires and their application as humidity sensors and photodetectors, J. Phys. Chem. C 113 (2009) 70857090. [33] H. Kim, S. Shim, Fabrication and structural characterization of Mg2 SiO4 nanowires, Appl. Phys. A 86 (2007) 361364. [34] N. Birks, G. Meier, F. Pettit, Introduction to the High-Temperature Oxidation of Metals, Cambridge University Press, Cambridge, 2006. [35] G. Zhou, J. Yang, Initial oxidation kinetics of copper (1 1 0) lm investigated by in situ UHV-TEM, Surf. Sci. 531 (2003) 359367. [36] G. Zhou, J. Yang, Temperature effects on the growth of oxide islands on Cu (1 1 0), Appl. Surf. Sci. 222 (2004) 357364. [37] K. Ernst, A. Ludviksson, R. Zhang, J. Yoshihara, C. Campbell, Growth model for metal lms on oxide surfaces: Cu on ZnO (0 0 0 1)O, Phys. Rev. B 47 (1993) 1378213796. [38] M. Tiwari, B. Agrawal, Enhanced Lattice Specic Heat Due to Heavy Tin Impurities in Copper, J. Phys. F: Met. Phys. 3 (1973) 1313. [39] J. Rakowski, G. Meier, F. Pettit, F. Dettenwanger, E. Schumann, M. Ruehle, The effect of surface preparation on the oxidation behavior of gamma TiAl-base intermetallic alloys, Scripta Mater. 35 (1996) 14171422.

Biographies
Xiaopeng Li is currently a Ph.D. student in the Department of Chemical Engineering at the University of Massachusetts Lowell, USA. He received his B.S. in Chemistry and B.Eng. in Chemical Engineering in a dual bachelors program from Shandong University, China, in 2007. His current project is focused on the synthesis and assembly of metal oxide nanowires and conducting polymer nanowires for sensor applications. Zhiyong Gu is an Assistant Professor in the Department of Chemical Engineering at the University of Massachusetts Lowell. He is also afliated with the CHN/NCOE Nanomanufacturing Center. He received his B.E. from Qingdao Institute of Chemical Technology, China, in 1996, his M.S. from the University of Notre Dame in 2001, and his Ph.D. from the State University of New York at Buffalo in 2004, respectively. He was a Postdoctoral Fellow at the Johns Hopkins University from 2004 to 2006. He has published 4 book chapters and over 40 refereed papers. His current research interests include synthesis of nanoparticles and nanowires, self-assembly, block copolymers, nanocomposites, and nanoscale-integration for electronics and sensors.

X. Li et al. / Sensors and Actuators B 158 (2011) 199207 JungHwan Cho is currently a PostDoc. at the University of Massachusetts Lowell, U.S.A. He received a B.S. in Instrumentation and Control Engineering from Gyeongsang National University, Jinju, South Korea in 2001, then received an M.S. and a Ph.D. in Electronic Engineering from Kyungpook National University, Daegu, South Korea in 2003 and 2008, respectively. His research interests include pattern recognition techniques, fuzzy systems, and articial neural networks applied to electronic noses and gas detection devices. Hongwei Sun is an Assistant Professor in the Department of Mechanical Engineering at the University of Massachusetts Lowell (UML). He graduated with a Ph.D. from Institute of Engineering Thermophysics at Chinese Academy of Science in 1998. Prior to joining UML in 2005, he was a postdoctoral researcher at the University of Rhode Island and later a research scientist at the Massachusetts Institute of Technology. His research interest is on power microelectromechanical systems (Power

207

MEMS), MEMS acoustic sensors, microscale cooling systems. His other interest is in micro/nano fabrication technology, fundamental understanding of micro/nanoscale uidics and their applications in biological analysis and energy areas. Pradeep U. Kurup is a Professor in the Department of Civil and Environmental Engineering at the University of Massachusetts Lowell. He received his B.Tech. in Civil Engineering in 1985 from the University of Kerala and obtained his M.Tech. from the Indian Institute of Technology Madras (1987). He holds a Ph.D. in Civil Engineering (1993) from Louisiana State University (LSU). Dr. Kurup has conducted extensive research in sensor integration; pattern recognition using intelligent models; multi-sensor data fusion; articial olfaction; and geotechnical & environmental site characterization. Dr. Kurup is a member of several professional societies, and is a registered Professional Engineer in the State of Louisiana.

Potrebbero piacerti anche