Selectively enhanced sensing performance for oxidizing gases based on ZnO nanoparticle-loaded electrospun SnO2 nanotube heterostructures

Kaidi Diao ab, Yunpeng Huangc, Minjie Zhoub, Jicheng Zhangb, Yongjian Tangb, Shuxia Wanga, Tianxi Liuc and Xudong Cui*b
aDepartment of Applied Physics, Chongqing University, Chongqing, 400044, China
bScience and Technology on Plasma Physics Laboratory, Research Center of Laser Fusion, CAEP, Mianyang, 621900, China. E-mail: xudcui@gmail.com
cState Key Laboratory of Molecular Engineering of Polymers, Department of Macromolecular Science, Fudan University, Shanghai, 200433, China

Received 2nd February 2016 , Accepted 10th March 2016

First published on 11th March 2016


Abstract

In this work, we present gas sensors based on ZnO nanoparticle-loaded electrospun SnO2 nanotube (ZnO/SnO2) n–n heterostructures (HSs) synthesized by electrospinning combined with facile thermal decomposition. The sensing properties of the pristine SnO2 nanotubes (NTs) and ZnO/SnO2 HSs were investigated toward the representative oxidizing (NO2) and reducing (H2, CO) gases. Results show that the as-prepared ZnO/SnO2 HSs exhibit selectively enhanced and diminished sensing performances for oxidizing and reducing gases, respectively. These phenomena are closely associated with the modulation of the local depletion layer on the surface of SnO2 nanoparticles (NPs) caused by charge transfer at the heterojunctions due to work function difference. A modified grain boundary-controlled sensing mechanism is proposed to describe charge transport in sensing layers based on the contact potential barriers between nanoparticles. Our study indicates that the selection of material system and their synergism are keys to the effective design of gas sensors with semiconducting metal oxide HSs.


1. Introduction

Metal oxide semiconductor (MOS) resistive gas sensors have been widely used in the detection of toxic, flammable and explosive gases,1,2 energy efficiency and emission in combustion processes,3 healthcare,4 environmental monitoring,5 and food processing.6,7 With the urgent requirements of sustainable development all over the world, gas sensing techniques are now facing challenges towards high sensitivity and selectivity, and short response and recovery times at low gas concentrations. For instance, the U.S. Environmental Protection Agency (EPA) has set a primary standard of 53 parts per billion (ppb) for NO2,8 above which may cause possible health problems especially for those sensitive populations including children, elderly and people with asthma, just naming a few critical challenges. It is therefore of crucial importance to develop high-performance gas sensors to fulfill those future missions.

Currently, various approaches have been used to approach above-mentioned goals with MOS gas sensors. These include doping or surface decorating with noble metal or metal oxide nanoparticles, compositing with other oxides, or plasma exposure etc.9–16 Among them, the HSs formed by surface coating with NPs, particularly metal oxide NPs with special functionalities is a widely accepted route to improve the sensing performance.

The metal oxide NPs are considered to be electronically sensitive and can modulate the conduction channel through charge transfer at the interface between the host materials and the attached NPs.17 Due to the modulation, the electric conduction in the sensing layer of a gas sensor becomes more sensitive to the process of gas adsorption and desorption, remarkably influencing the sensing performance. Generally, two approaches have been employed to realize charge transfer between the host materials and the attached NPs. One is the use of p–n heterojunction.16,18–20 Due to carrier concentration gradient, charge transfer will occur at the interface until the built-in electric field is formed to prevent carrier flow and finally reaches equilibrium in the p–n heterojunction.18 The other is the n–n or p–p isotype heterojunction, in which charge transfer occurs due to the work function difference between the host materials and the attached NPs.19,20 From the engineering's point of view, isotype heterojunctions are preferred since constructing the work function difference is relatively easy to be implemented with a variety of materials. In this work, we then aim to n–n HSs configuration and carefully investigate the contribution of heterojunctions to the sensing performance.

One-dimensional (1D) metal oxides (i.e., nanowires, nanobelts, nanorods, NTs and nanofibers, etc.) have been attracting intensive attentions as potential sensing materials due to their high surface-to-volume ratio and efficiencies for carrier transport.21,22 Particularly, 1D electrospun nanostructures consisting of many nanograins would result in potential barriers between the adjacent NPs, quite different from other types of 1D nanostructures.23,24 More importantly, the potential barrier can be easily modulated by the process of gas adsorption and desorption, which ensures that the structures exhibit excellent sensing performance.25 It has been demonstrated that the 1D electrospun nanostructures with NPs coatings can effectively improve sensor performance toward some target gases (NH3, H2S, H2 and NO2) due to the formation of local heterojunctions.26–30 Up to date, there still have been some critical challenges left to further exploit the potentials for these structures. (1) Are the NPs-coated 1D electrospun nanostructures efficient for the target gases, regardless of their oxidizing or reducing nature? (2) How does the local heterojunction impact the potential barriers between the adjacent nanograins? (3) What is the possible sensing mechanism for the NPs-coated 1D electrospun nanostructures? To address these issues, here we select two common and typical sensing materials, SnO2 and ZnO as examples. Although the single 1D electrospun nanostructures of those two materials have exhibited high sensing performance, their synergism is expected to further improve the performance and extend the applications of devices at low concentration and complex atmospheres.

In this work, we present the gas sensors based on ZnO nanoparticle-loaded electrospun SnO2 NT n–n HSs synthesized by the electrospinning and thermal decomposition. The sensing properties of the pristine SnO2 NTs and ZnO/SnO2 HSs are investigated toward the representative oxidizing (NO2) and reducing (H2, CO) gases. Results show that compared with the pristine SnO2 NTs, ZnO/SnO2 HSs exhibit selectively enhanced and diminished sensing performances for oxidizing and reducing gases, respectively. A modified grain boundary-controlled sensing mechanism is proposed to interpret these phenomena based on the modulation of local depletion layers on surface of SnO2 NPs. Our study indicates that the selection of material system and their synergism are keys to the effective design of gas sensors with semiconducting metal oxide HSs.

2. Experimental

2.1. Materials

Zinc acetate dihydrate (Zn(CH3COO)2·2H2O), and anhydrous ethanol were purchased from Chengdu Kelong Chemical Co. Ltd., China. Polyvinylpyrrolidone (PVP, Mw = 1[thin space (1/6-em)]300[thin space (1/6-em)]000 g mol−1) was obtained from Sigma-Aldrich. Tin(II) chloride dihydrate (SnCl2·2H2O), and N,N-dimethylformamide (DMF) were supplied by Shanghai Chemical Reagent Company, China. All the chemicals were of analytical reagent grade and used without further purification. Deionized water with a resistivity of 18.25 MΩ cm was obtained using an up-water purification system (Chengdu YouPu Technology Co. Ltd., China) at room temperature.

2.2. Preparation of ZnO/SnO2 HSs

The SnO2 NTs were synthesized by electrospinning as reported in our previous work.31 In experiments, 1.2 g SnCl2·2H2O was dissolved in a mixture of ethanol (5 mL) and of DMF (5 mL) with vigorous stirring at room temperature. Afterwards, 1.2 g PVP was added to the above mixture and stirred overnight to form a uniformly transparent solution. Then, the solution was transferred into a plastic syringe and injected in a rate of 0.25 mL min−1 using a no. 21 stainless steel needle connected to a high voltage DC power supply. Here, a rotating aluminum drum was selected as the collector with a distance of ∼18 cm to the needle tip. When a fixed voltage of 21 kV was applied to the system, the electrospun SnCl2/PVP nanofibers were produced and deposited on the aluminum drum. Finally, the SnCl2/PVP nanofibers were peeled off from the collector and calcinated for 3 h at 600 °C in air to remove the polymer and form SnO2 NTs.

The ZnO/SnO2 HSs were synthesized by a one-step thermal decomposition according to Zhao's method.32 Briefly, a certain amount of saturated ethanol solution of Zn(CH3COO)2·2H2O were dropped onto the dispersed SnO2 NTs and dried in air. Then the zinc acetate covered SnO2 NTs were baked at 450 °C for 20 min in air to obtain the ZnO/SnO2 HSs.

2.3. Characterization

The morphology of resulting products was investigated using a FEI field emission scanning electron microscopy (FESEM) equipped with an Oxford energy dispersive X-ray spectroscopy (EDS) operated at 15 kV. The detailed microstructure of the ZnO/SnO2 HSs was studied by a transmission electron microscopy (TEM) and a high-resolution transmission electron microscopy (HRTEM) performed with a FEI TEM under an accelerating voltage of 200 kV. The powder X-ray diffraction (XRD) patterns were recorded on a Bruker X-ray diffractometer with Cu Kα radiation (λ = 0.154 nm) operating at 40 kV in the 2θ range from 20° to 80°.

2.4. Fabrication and measurement of gas sensors

The gas sensors were fabricated by a similar method as reported in our previous work.33 First, the as-synthesized sensing materials, including SnO2 NTs and ZnO/SnO2 HSs, were mixed with deionized water in a weight ratio of 4[thin space (1/6-em)]:[thin space (1/6-em)]1 and slightly grinded in an agate mortar to form homogeneous pastes. Then, the pastes were coated on a ceramic substrate of Au interdigitated electrode. After drying at room temperature for 24 h in air, the sensing devices were sintered at 400 °C for 2 h to improve the mechanical strength (Fig. 1 inset).
image file: c6ra03061k-f1.tif
Fig. 1 Photograph of the intelligent gas sensing analysis system of CGS-4TP. The inset is a top view of Au interdigitated electrode and a sample sensor.

The gas-sensing properties of the sensors were investigated with a CGS-4TP intelligent gas sensing analysis system (Beijing Elite Tech Co. Ltd., China) as shown in Fig. 1. This multifunctional system consists of a heating system, a probe adjustment system, a measurement and data acquisition system, as well as a set of built-in measurement control software. An external temperature control provided by the system can conductively adjust the sensor temperature from 20 °C to about 500 °C with a precision of 1 °C. Two probes were pressed on sensor substrates to export electrical signals. All sensors were pre-heated at different operating temperatures for about 30 min. When the sensor resistances were stable, the target gas with a certain concentration was injected into the test chamber (1.8 L in volume) in 1000 standard cubic centimeters per minute (sccm) by a mass flow controller (MFC). After the sensor resistances reached new constant values, the dry air was injected into the test chamber in 1000 sccm by MFC. The operating temperature, the sensor resistance and response values of these sensors were automatically reported by the analysis system. The response value (R) was defined as R = Rg/Ra or R = Ra/Rg for the oxidizing and inducing gases, respectively, where Ra was the sensor resistance in the ambient air and Rg was the sensor resistance in the detected gas. The response time and recovery time were defined as the time taken for the sensor to reach 90% of its maximum response and decrease to 10% of its minimum response, respectively.

3. Results and discussion

3.1. Morphology and structure

The morphologies of the as-prepared electrospun SnO2 NTs and ZnO/SnO2 HSs were observed using FESEM. Fig. 2a shows a hollow tubular structure of the SnO2 NTs with an outer diameter ranging from ∼470 nm to ∼580 nm. The surface is extremely rough, covered with nanosized SnO2 grains with an average diameter of ∼90 nm. The thickness of the tube wall is about 130 nm. Fig. 2b shows a low-magnification SEM image of the ZnO/SnO2 HSs after thermal decomposition process. It can be seen that, the tubular structure is maintained, and the surface becomes much rougher. At higher magnification (Fig. 2c), it can be observed that many tiny ZnO NPs with a mean size of ∼40 nm are loaded on the surface of SnO2 NTs. The ZnO/SnO2 HSs comprising of two different components are further confirmed by the EDS mapping analysis (Fig. 2d–g), confirming the existence of Zn (3.37 wt%) element besides Sn (75.45 wt%) and O (21.18 wt%) elements, where the SnO2 NTs are uniformly covered with ZnO NPs. According to the EDS data, the percentage of ZnO in ZnO/SnO2 HSs is 4.19 wt%. The tubular structure and rough surface of ZnO/SnO2 HSs are expected to significantly increase the active sites for gas adsorption, and provide an effective way to improve gas–solid interaction.
image file: c6ra03061k-f2.tif
Fig. 2 FESEM images of (a) the electrospun SnO2 NTs and (b and c) ZnO/SnO2 HSs. (d–g) shows the EDS elemental mapping images of the ZnO/SnO2 HSs.

To further analyze the detailed microstructure and morphology of ZnO/SnO2 HSs, TEM and HRTEM observations were carried out. As shown in Fig. 3a and b, the TEM images confirm that the ZnO/SnO2 HSs present rough tubular structure, and the tube walls are uniformly coated with tiny ZnO NPs, being consistent with the observations in FESEM images. In the HRTEM images of ZnO/SnO2 HSs (Fig. 3c and d), uniform lattice fringe are observed, and the lattice fringes of d = 0.34 nm and 0.26 nm match well with the crystallographic planes of rutile SnO2 (110) and hexagonal wurtzite ZnO (002), respectively.


image file: c6ra03061k-f3.tif
Fig. 3 Bright-field (a and b) TEM and (c and d) HRTEM images of the ZnO/SnO2 HSs.

In order to analyze the composition and crystalline structure, XRD measurement was then performed. As shown in Fig. 4a, high intensity diffraction peaks are indexed to pure SnO2 crystals with a rutile structure, well agreeing with the reported values from the Joint Committee on Powder Diffraction Standards card (JCPDS no. 41-1445). No characteristic peaks from other impurities are observed in the XRD pattern, and the sharp diffraction peaks confirm the good crystallinity of SnO2 NTs. Fig. 4b shows the XRD pattern taken from ZnO/SnO2 HSs. Compared to the pure SnO2 NTs, two new but relatively weak diffraction peaks at 2θ = 31.8° and 36.3° are detected, which can be respectively assigned as the (100) and (101) planes of the hexagonal wurtzite ZnO (JCPDS no. 41-1451), further indicating the formation of ZnO NPs coated on the surface of SnO2 NTs. According to Scherrer formula, the crystallite size is about 51.3 nm, 22.5 nm for SnO2 and ZnO nanoparticles, respectively. Note that in the XRD patterns, the peak at 2θ = 33.9° indexing to the (101) plane of SnO2 is rather intense and slightly broaden, the diffraction peak at 2θ = 34.4° assigning to the (002) plane of ZnO is overlapped and cannot be observed. Nevertheless, the (002) plane of ZnO is clearly observed in HRTEM images (Fig. 3c and d).


image file: c6ra03061k-f4.tif
Fig. 4 XRD patterns of (a) the electrospun SnO2 NTs and (b) the ZnO/SnO2 HSs.

3.2. Gas-sensing performance

The operating temperature of metal oxide gas sensors plays a vital role on their sensing performance.34,35 Generally, by elevating operating temperature, one could achieve an optimal response with improved gas–solid interaction and relatively large surface adsorption. In our work, to determine the optimal operating temperature, the temperature dependent sensing behaviors of the SnO2 NTs and ZnO/SnO2 HSs sensors to the target gases NO2 (10 ppm), H2 (50 ppm) and CO (50 ppm) are investigated at temperatures ranging from 150 °C to 400 °C. As shown in Fig. 5, the response of all sensors exhibits a trend of “increase-maximum-decay” with the operating temperature, and the largest response is found at 300 °C for all sensors. We can find that, in the initial stage, the gas chemisorption is higher than the gas desorption on the oxide surfaces at low temperature range, thus the adsorption quantity increases with the elevated temperature. As the temperature rises, the gas chemisorption will gradually decrease and the gas desorption will increase, until a balance of these two processes is reached. At this balance point, the adsorption quantity reaches a maximum, or in other words, the sensor exhibits the best response. After this critical point, the exothermic gas chemisorption becomes difficult and the gas desorption dominates, leading to a reduced response.36 Thus 300 °C is selected as the optimal operating temperature and all the following measurements were performed at this temperature.
image file: c6ra03061k-f5.tif
Fig. 5 Sensor responses of the electrospun SnO2 NTs and ZnO/SnO2 HSs at different temperatures (150–400 °C) toward (a) NO2 (10 ppm), (b) H2 (50 ppm) and (c) CO (50 ppm).

The dynamic resistance behaviors of the SnO2 NTs and ZnO/SnO2 HSs were investigated towards a typical oxidizing gas NO2. Fig. 6a shows the real-time response–recovery curves of both sensors to NO2 with the concentration of 0.5–100 ppm. Once the sensors are exposed to a certain concentration of NO2 in each dynamic cycle, the resistance of both sensors will increase; when the gas supply is stopped, the sensor resistance decreases and finally returns to its baseline, confirming that both sensors are with desired reversibility and repeatability. This is the characteristics of n-type semiconducting oxides to oxidizing gases, and has been well interpreted.37 Note that from the inset of Fig. 6a, the ZnO/SnO2 HSs exhibit a smaller baseline (∼2.4 kΩ) than that of the SnO2 NTs (∼4.9 kΩ), being closely related to the modulation of ZnO NPs on the depletion layer of SnO2 NPs at the local heterojunctions.9,38 We will discuss this in next section. When focus on a certain dynamic cycle, the ZnO/SnO2 HSs present a larger resistance change compared with the case of SnO2 NTs, implying that a larger sensor response is achieved, especially at high gas concentrations. Fig. 6b shows the response of both sensors as a function of gas concentration from 0.5 ppm to 100 ppm. The responses of both sensors increase with gas concentration, roughly obeying linear relationship. Taking the 10 ppm NO2 as an example, the ZnO/SnO2 HSs has a response of 65.9, being six times higher than that of the SnO2 NTs (10.6). Clearly, surface functionalization would assist to improve the response of SnO2 NTs based gas sensors. Nevertheless, the response and the recovery times of the ZnO/SnO2 HSs are prolonged (Fig. 6c and d). Based on the above results, we can see that the ZnO/SnO2 HSs can remarkably enhance the sensor response toward the representative oxidizing gas (NO2). We then look back to the first issue: is this structure also efficient for the reducing gases?


image file: c6ra03061k-f6.tif
Fig. 6 Gas sensing performance of the electrospun SnO2 NTs and ZnO/SnO2 HSs toward NO2 in concentration range of 0.5–100 ppm: (a) typical dynamic curves; (b) sensor response as a function of gas concentration; (c) response time; (d) recovery time. Inset of panel (a) shows dynamic curves in concentration range of 0.5–3 ppm. Inset of panel (b) shows sensor responses as a function of gas concentration (0.5–10 ppm).

Accordingly, the sensing performance of the SnO2 NTs and ZnO/SnO2 HSs were investigated to the typical reducing gases H2 and CO in the range of 10–50 ppm. From Fig. 7a and b, it is observed that the sensor resistances decrease when H2 or CO are injected, and then reach steady values; once the gas supply is stopped, the resistances will increase and finally return to their baselines. Although the dynamic behavior of the sensors to reducing gases is completely different from the case of NO2, the phenomenon can also be well explained by the sensing mechanism of n-type semiconducting oxides to reducing gases.37 The baseline resistance of the ZnO/SnO2 HSs is lower than that of SnO2 NTs, being consistent with the results in the case of NO2. More interestingly, the ZnO/SnO2 HSs exhibit much smaller resistance change than that of SnO2 NTs, being opposite to the case of NO2. The resistance change of ZnO/SnO2 HSs could hardly be detected at the concentration of 10 ppm for H2, even for CO at 25 ppm. The change of sensor resistance is expressed as sensor response and summarized as a function of gas concentration shown in Fig. 7c and d. There exist two features: first, the sensor response roughly increases with the gas concentration; the ZnO/SnO2 HSs exhibit smaller sensor response compared with the SnO2 NTs. Even at high concentration of 100 ppm, the responses of ZnO/SnO2 HSs are only 1.9 and 1.1 for H2 and CO, respectively, indicating that the ZnO/SnO2 HSs can greatly decline the sensor response to reducing gases.


image file: c6ra03061k-f7.tif
Fig. 7 Gas sensing performances of the electrospun SnO2 NTs and ZnO/SnO2 HSs toward H2 and CO in the range of 10–100 ppm: typical dynamic curves toward (a) H2 and (b) CO; sensor responses as a function of gas concentration toward (c) H2 and (d) CO.

In order to intuitively compare the sensing performances of the SnO2 NTs and the ZnO/SnO2 HSs to oxidizing and reducing gases, the sensor responses of both sensors to CO (50 ppm), H2 (50 ppm) and NO2 (5 ppm) are displayed in Fig. 8. For the reducing gases, the ZnO/SnO2 HSs exhibit declined device response. Particularly, the response of ZnO/SnO2 HSs (1.44) for H2 becomes nearly one-fifth of the SnO2 NTs (6.89). While for the oxidizing gas (NO2), the ZnO/SnO2 HSs present remarkably enhanced effect, showing six times higher response (30.84) than that of the SnO2 NTs (5.09). Thanks to the opposite trend, the ZnO/SnO2 HSs exhibit a excellent selectivity to NO2, although the concentration of NO2 is only one-tenth of the reducing gases. On the basis of this effect, we believe that our devices can be used to selectively detect oxidizing gases under complex atmospheres, in which reducing gases and oxidizing gases coexist, for instance, in chemical industries, laboratories, mines etc.


image file: c6ra03061k-f8.tif
Fig. 8 Comparison of the sensor responses based on the electrospun SnO2 NTs and ZnO/SnO2 HSs toward CO (50 ppm), H2 (50 ppm) and NO2 (5 ppm).

3.3. Gas-sensing mechanism

It is well known that the gas-sensing property is heavily dependent on surface area and porosity.39 To examine these parameters, nitrogen adsorption and desorption measurements are performed on the sensor materials (Fig. S1). The isotherms roughly belong to the type III isotherm in the IUPAC classification, and exhibit type H3 loop, indicating that all samples have similar pore size distribution, i.e., average pore size of ∼2–50 nm in diameter.40 The BET surface areas of ZnO/SnO2 HSs is ∼5.65 m2 g−1, only a litter higher than that of the electrospun SnO2 NTs (∼5.54 m2 g−1). Since the difference of the BET surface area of both the samples may be negligible, we reasonably believe that the main factors leading to sensing difference should be other reasons. In addition, after surface loading of ZnO NPs on SnO2 NTs, the original tubular structure is still maintained and the mean size of ZnO nanoparticles is ∼40 nm, so the morphology of the ZnO/SnO2 HSs is almost the same as SnO2 NTs. Accordingly, the true factor resulting in the sensing difference may be the modulation of depletion layer on surface of SnO2 NPs caused by charge transfer at the local heterojunctions.

Compared with 1D single-crystal nanowires, the tube walls of both SnO2 NTs and ZnO/SnO2 HSs are constructed with nanosized SnO2 grains, resulting in a large amount of homojunctions at the boundaries between the adjacent nanograins. Due to the structure characteristic of tube walls, there would be some interspaces between nanograins in sensing layers.41–43 As shown in Fig. 9a1, when the SnO2 NTs sensor is exposed to the air, the oxygen molecules will diffuse into the tube walls through interspaces and surround the nanograins. Then the oxygen molecules will be chemisorbed on the surface of nanograins and become oxygen species (O2, O or O2−) by extracting electrons from the conduction band of SnO2 grains, resulting in an electron depletion layer on the surface of oxides.37 At the same time, the energy band upward bends at the grain boundaries. The generated potential barriers between homojunctions will obstruct the transport of electrons under the applied voltage. Under this circumstance, the sensor will maintain the produced initial resistance, i.e. the baseline in air. If the sensor is exposed to the atmosphere including reducing gases (Fig. 9a2), the chemisorbed oxygen species will interact with the reducing gases and being removed from the surface of the oxides. The captured electrons will return to the conduction band, and the potential barriers will decrease, resulting in a depressed sensor resistance. When the sensor is exposed to an oxidizing gas (Fig. 9a3), the oxidizing gas (NO2) will diffuse into the tube walls through interspaces and be chemisorbed on the surface of the nanograins, further trapping electrons from the conduction band. As a consequence, the potential barrier will increase, resulting in a larger sensor resistance.25,44,45 In the sensing process, the modulation of potential barriers between the adjacent nanograins is key to determine the change of sensor resistance, i.e. the sensor response. The potential barrier is induced by the electron depletion whose width can be denoted by the Debye length (λD). The Debye length of an oxide material can be expressed by the following equation:46

image file: c6ra03061k-t1.tif
where ε is the static dielectric constant, k is the Boltzmann's constant, T is the absolute temperature, q is the electrical charge and nc is the carrier concentration. For the SnO2 NTs, the Debye length is calculated to be ∼3.2 nm with ε = 13.5 × 8.85 × 10−12 F m−1, T = 573 K, nc = 3.6 × 1018 cm−3.47 In principle, for the pure SnO2 NTs, the width of the depletion layer could be expanded or compressed due to the change of electron concentration caused by oxidizing and reducing gases, which eventually determine the potential barriers between the adjacent nanograins.


image file: c6ra03061k-f9.tif
Fig. 9 Schematically showing the modulation of potential barriers between SnO2 NPs to reducing and oxidizing gases based on grain boundary-controlled sensing mechanism: (a) the electrospun SnO2 NTs; (b) the ZnO/SnO2 HSs. (c) Energy-band diagram of ZnO/SnO2 heterojunction.

When ZnO NPs are attached on the surface of SnO2 NTs (Fig. 9b), there exist two physical phenomena. One occurs on the surface of ZnO NPs due to gas adsorption, and the other is at the interface between the ZnO NPs and SnO2 NPs. When the ZnO/SnO2 HSs is exposed in air (Fig. 9b1), oxygen molecules are chemisorbed on the surface of ZnO NPs, then the gas molecules will capture electrons from the conduction band of ZnO NPs, resulting in electron depletion.37 The width of electron depletion layer can also be indicated by the Debye length. For ZnO NPs, ε = 8.75 × 8.85 × 10−12 F m−1,48 T = 573 K, the carrier concentration varies in the range of 1016 to 1019 cm−3,49 thus the Debye length changes from ∼22 nm to ∼0.7 nm, smaller than the mean size (40 nm) of ZnO NPs. This indicates that the region of ZnO NPs close to SnO2 NPs is not affected by the gas chemisorption. On the other hand, since the work function of ZnO NPs (5.2 eV) is larger than that of SnO2 NPs (4.9 eV), electrons will transfer from ZnO NPs to SnO2 NPs until a same Fermi level is reached (Fig. 9c). Thus, the width of depletion layer at the local heterojunctions reduces on the side of SnO2. Particularly, the ZnO NPs on or close to the boundaries between the adjacent SnO2 NPs can heavily decrease the potential barriers (Fig. 9b1). The relatively weak potential barriers cause a small sensor baseline for the ZnO/SnO2 HSs. When the sensor is exposed to the oxidizing gas (Fig. 9b3), the electron depletion layer of the SnO2 NPs expands to a width as wide as the pure SnO2 NTs. The sensor experiences from a less resistive state to a more resistive one, resulting in an increased resistance change, which is responsible for the enhanced sensing performance for oxidizing gases with the ZnO/SnO2 HSs. On the contrary, when reducing gas is supplied (Fig. 9b2), the captured electrons are released, and the depletion layer becomes narrow. However, since the depletion layer has already been shortened by the electron transfer due to the n–n heterojunction with different work functions, the change of sensor resistance is decreased, leading to a smaller sensor response.20 This is the reason why the ZnO/SnO2 HSs exhibit increased and decreased sensor responses respectively for oxidizing and reducing gases.

Interestingly, as previously reported,20,38 gas sensors developed by SnO2–ZnO core–shell (C–S) nanofibers or nanowires exhibit opposite sensing performance, showing increased response for reducing gases and decreased response for oxidizing gases. In C–S structures, the model suggests that as the shell thickness increases, the dominant electrical transport pathway is localized from the core (SnO2) to the shell layer (ZnO). When the shell thickness increases to be comparable to the Debye length of the shell material, the shell layer is completely depleted in air due to the chemisorption of oxygen species, resulting in a relatively high initial resistance. If the reducing gas is supplied, a large amount of electrons will return to the shell layer, leading to a low sensor resistance. The process from a more resistive state to a less one contributes to a large response. While for oxidizing gases, the depleted shell layer can hardly release free electrons to further adsorb gas molecules, only causing a small resistance change. In our case, the ZnO NPs are dispersively attached on the surface of SnO2 NTs, not forming a continuous conducting shell, and the transport pathway is still confined to the SnO2 NTs. Duo to electron transfer from ZnO NPs to SnO2 NTs at the interface, the SnO2 NTs become electron accumulated, beneficial to adsorb oxidizing gases, leading to an enhanced sensor response. The opposite sensing performances caused by the SnO2–ZnO C–S structures and the ZnO/SnO2 HSs demonstrate that the structure of heterojunctions can tremendously impact the sensing properties of sensor devices.

4. Conclusions

In summary, the ZnO nanoparticle-loaded electrospun SnO2 nanotube n–n HSs were synthesized by electrospinning combined with facile thermal decomposition. The sensing properties of the pristine SnO2 NTs and ZnO/SnO2 HSs were investigated toward the representative oxidizing (NO2) and reducing (H2, CO) gases. The response of ZnO/SnO2 HSs is 30.84 for 5 ppm NO2, six times higher than that of SnO2 NTs (5.09). While for 50 ppm H2, the response of ZnO/SnO2 HSs (1.44) becomes nearly one-fifth of the SnO2 NTs (6.89). Results indicate that ZnO/SnO2 HSs exhibit selectively enhanced and diminished sensing performances respectively for oxidizing and reducing gases. These phenomena are closely associated with the modulation of the local depletion layer on surface of SnO2 NPs caused by charge transfer at the heterojunctions due to work function difference. A modified grain boundary-controlled sensing mechanism is proposed to describe charge transport in sensing layers based on the contact potential barriers between nanoparticles. Our study indicates that the selection of material system and their synergism are keys to the effective design of gas sensors with semiconducting metal oxide HSs.

Acknowledgements

This work was supported by the fund from the Science and Technology on Plasma Physics Laboratory at Research Center of Laser Fusion, CAEP and the Science and Technology Development Fund of CAEP (2014A0302014).

References

  1. N. Barsan, D. Koziej and U. Weimar, Sens. Actuators, B, 2007, 121, 18–35 CrossRef CAS.
  2. A. Tricoli, M. Righettoni and A. Teleki, Angew. Chem., Int. Ed., 2010, 49, 7632–7659 CrossRef CAS PubMed.
  3. D. J. Wales, J. Grand, V. P. Ting, R. D. Burke, K. J. Edler, C. R. Bowen, S. Mintova and A. D. Burrows, Chem. Soc. Rev., 2015, 44, 4290–4321 RSC.
  4. N. Shehada, G. Brönstrup, K. Funka, S. Christiansen, M. Leja and H. Haick, Nano Lett., 2014, 15, 1288–1295 CrossRef PubMed.
  5. G. F. Fine, L. M. Cavanagh, A. Afonja and R. Binions, Sensors, 2010, 10, 5469–5502 CrossRef CAS PubMed.
  6. A. Berna, Sensors, 2010, 10, 3882–3910 CrossRef CAS PubMed.
  7. A. Ponzoni, E. Comini, I. Concina, M. Ferroni, M. Falasconi, E. Gobbi, V. Sberveglieri and G. Sberveglieri, Sensors, 2012, 12, 17023–17045 CrossRef CAS PubMed.
  8. http://www3.epa.gov/ttn/naaqs/criteria.html.
  9. S.-W. Choi, A. Katoch, J.-H. Kim and S. S. Kim, ACS Appl. Mater. Interfaces, 2014, 6, 17723–17729 CAS.
  10. S.-W. Choi and S. S. Kim, Sens. Actuators, B, 2012, 168, 8–13 CrossRef CAS.
  11. P. Rai, Y.-S. Kim, H.-M. Song, M.-K. Song and Y.-T. Yu, Sens. Actuators, B, 2012, 165, 133–142 CrossRef CAS.
  12. X.-j. Wang, W. Wang and Y.-L. Liu, Sens. Actuators, B, 2012, 168, 39–45 CrossRef CAS.
  13. D. J. Yang, I. Kamienchick, D. Y. Youn, A. Rothschild and I. D. Kim, Adv. Funct. Mater., 2010, 20, 4258–4264 CrossRef CAS.
  14. H. W. Kim, S.-W. Choi, A. Katoch and S. S. Kim, Sens. Actuators, B, 2013, 177, 654–658 CrossRef CAS.
  15. G. Lu, L. E. Ocola and J. Chen, Adv. Mater., 2009, 21, 2487–2491 CrossRef CAS.
  16. M. Mashock, K. Yu, S. Cui, S. Mao, G. Lu and J. Chen, ACS Appl. Mater. Interfaces, 2012, 4, 4192–4199 CAS.
  17. D. R. Miller, S. A. Akbar and P. A. Morris, Sens. Actuators, B, 2014, 204, 250–272 CrossRef CAS.
  18. D. Neamen, Semiconductor physics and devices, McGraw-Hill, Inc., 2002 Search PubMed.
  19. S. Sen, P. Kanitkar, A. Sharma, K. Muthe, A. Rath, S. Deshpande, M. Kaur, R. Aiyer, S. Gupta and J. Yakhmi, Sens. Actuators, B, 2010, 147, 453–460 CrossRef CAS.
  20. S.-W. Choi, A. Katoch, G.-J. Sun, J.-H. Kim, S.-H. Kim and S. S. Kim, ACS Appl. Mater. Interfaces, 2014, 6, 8281–8287 CAS.
  21. J. Huang and Q. Wan, Sensors, 2009, 9, 9903–9924 CrossRef PubMed.
  22. M. J. Spencer, Prog. Mater. Sci., 2012, 57, 437–486 CrossRef CAS.
  23. R. Ab Kadir, Z. Li, A. Z. Sadek, R. Abdul Rani, A. S. Zoolfakar, M. R. Field, J. Z. Ou, A. F. Chrimes and K. Kalantar-zadeh, J. Phys. Chem. C, 2014, 118, 3129–3139 CAS.
  24. S. Bai, S. Chen, Y. Zhao, T. Guo, R. Luo, D. Li and A. Chen, J. Mater. Chem. A, 2014, 2, 16697–16706 CAS.
  25. A. Katoch, S.-W. Choi and S. S. Kim, Nanotechnology, 2014, 25, 455504 CrossRef PubMed.
  26. Q. Qi, P.-P. Wang, J. Zhao, L.-L. Feng, L.-J. Zhou, R.-F. Xuan, Y.-P. Liu and G.-D. Li, Sens. Actuators, B, 2014, 194, 440–446 CrossRef CAS.
  27. S.-W. Choi, J. Zhang, K. Akash and S. S. Kim, Sens. Actuators, B, 2012, 169, 54–60 CrossRef CAS.
  28. Z. Wang, Z. Li, T. Jiang, X. Xu and C. Wang, ACS Appl. Mater. Interfaces, 2013, 5, 2013–2021 CAS.
  29. L. Xu, R. Zheng, S. Liu, J. Song, J. Chen, B. Dong and H. Song, Inorg. Chem., 2012, 51, 7733–7740 CrossRef CAS PubMed.
  30. B.-H. Jang, O. Landau, S.-J. Choi, J. Shin, A. Rothschild and I.-D. Kim, Sens. Actuators, B, 2013, 188, 156–168 CrossRef CAS.
  31. Y. Huang, Y.-E. Miao, L. Zhang, W. W. Tjiu, J. Pan and T. Liu, Nanoscale, 2014, 6, 10673–10679 RSC.
  32. X. Zhao, P. Wang, Y. Gao, X. Xu, Z. Yan and N. Ren, Mater. Lett., 2014, 132, 409–412 CrossRef CAS.
  33. K. Diao, M. Zhou, J. Zhang, Y. Tang, S. Wang and X. Cui, Sens. Actuators, B, 2015, 219, 30–37 CrossRef CAS.
  34. L. Wang, Z. Lou, J. Deng, R. Zhang and T. Zhang, ACS Appl. Mater. Interfaces, 2015, 7, 13098–13104 CAS.
  35. T. Kida, S. Fujiyama, K. Suematsu, M. Yuasa and K. Shimanoe, J. Phys. Chem. C, 2013, 117, 17574–17582 CAS.
  36. Y.-F. Sun, S.-B. Liu, F.-L. Meng, J.-Y. Liu, Z. Jin, L.-T. Kong and J.-H. Liu, Sensors, 2012, 12, 2610–2631 CrossRef CAS PubMed.
  37. C. Wang, L. Yin, L. Zhang, D. Xiang and R. Gao, Sensors, 2010, 10, 2088–2106 CrossRef CAS PubMed.
  38. A. Katoch, S.-W. Choi, G.-J. Sun and S. S. Kim, J. Mater. Chem. A, 2013, 1, 13588–13596 CAS.
  39. K. Sahner and H. Tuller, J. Electroceram., 2010, 24, 177–199 CrossRef CAS.
  40. S. J. Gregg and K. S. W. Sing, Adsorption, Surface Area, and Porosity, Academic Press, London, 1991 Search PubMed.
  41. S.-W. Choi, J. Y. Park and S. S. Kim, Chem.–Eng. J., 2011, 172, 550–556 CrossRef CAS.
  42. S.-W. Choi, J. Y. Park and S. S. Kim, J. Mater. Res., 2011, 26, 1662–1665 CrossRef CAS.
  43. C. Jin, S. Park, H. Kim and C. Lee, Sens. Actuators, B, 2012, 161, 223–228 CrossRef CAS.
  44. C. Xu, J. Tamaki, N. Miura and N. Yamazoe, Sens. Actuators, B, 1991, 3, 147–155 CrossRef CAS.
  45. N. Barsan and U. Weimar, J. Electroceram., 2001, 7, 143–167 CrossRef CAS.
  46. S. M. Sze and K. K. Ng, Physics of semiconductor devices, John Wiley & Sons, 2006 Search PubMed.
  47. H. Ogawa, M. Nishikawa and A. Abe, J. Appl. Phys., 1982, 53, 4448–4455 CrossRef CAS.
  48. D. C. Look, J. W. Hemsky and J. Sizelove, Phys. Rev. Lett., 1999, 82, 2552 CrossRef CAS.
  49. Ü. Özgür, Y. I. Alivov, C. Liu, A. Teke, M. Reshchikov, S. Doğan, V. Avrutin, S.-J. Cho and H. Morkoc, J. Appl. Phys., 2005, 98, 041301 CrossRef.

Footnotes

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra03061k
K. D. and Y. H. contribute equally to this work.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.