DOI:
10.1039/C6RA21672B
(Communication)
RSC Adv., 2016,
6, 105317-105321
Enhanced gas-sensing performance of SnO2/Nb2O5 hybrid nanowires†
Received
30th August 2016
, Accepted 22nd October 2016
First published on 24th October 2016
Abstract
Semiconductor tin oxide (SnO2) is acknowledged to be one of the semiconductor oxides with most potential for utilization as nitrogen dioxide (NO2)-sensor materials. However, the challenge to improve the gas sensitivity still remains, and this limits the full realization of the SnO2 sensor potential. In this paper, enhanced gas sensitivity obtained by the formation of a SnO2/niobium pentoxide (Nb2O5) hybrid structure is reported. As the NO2 concentration is lower to 3 ppm, the measured resistance in the SnO2/Nb2O5-based sensor is still 2.5 times larger than that in a pristine SnO2-based sensor. With an increase in the NO2 concentration, the response time of the SnO2/Nb2O5-based sensor decreases sharply whereas the pristine SnO2-based sensor response time increases. The enhanced performance is attributed to the presence of more oxygen vacancies and a higher specific surface area in the hybrid structure induced by the introduction of Nb2O5. These results provide some general guidelines for the selection of compositions to enhance the sensor performance.
Introduction
More and more environmental concerns such as acid rain, ozone depletion and the greenhouse effect result from gaseous pollutants, which severely influence health and society. Therefore, the gas sensor utilized for monitoring these pollutants becomes a part of environmental protection. Recently, the semiconducting metal oxides [such as tin oxide (SnO2), indium oxide (In2O3) and tungsten trioxide] become more preferable for use on account of their easily measurable resistance changes during monitoring of pollutants.1–5 When exposed to the reducing gases such as hydrogen sulfide, ammonia and so on, the surface of the metal oxide forms a charge depletion layer because of the adsorbed oxygen trapping electrons from the conduction band of the metal oxide. Subsequently, the oxygen species formed react with the adsorbed reducing gas molecules.6–9 As a result, the reducing gas can be detected by monitoring the change of the resistance in the metal oxide. In the detection of the oxidizing gas, the oxygen vacancies play a crucial role as chemisorption sites on the surface of the metal oxide.10,11 Generally, the oxygen vacancies carry a positive charge, which can trap the electrons. During the reactions with adsorbed gas molecules, the electron transfer occurring in the conduction band of the metal oxide induces a resistance change. Therefore, the oxidizing gas can be detected by exploring the resistance change in the metal oxide. Compared to other semiconductors, SnO2 has less barrier potential to adsorb nitrogen dioxide (NO2) molecules onto its surface.12–14 In addition, the natural formation of oxygen vacancies in SnO2 facilitates the electron transfer during the process of adsorption and desorption.15–17 Therefore, SnO2 has become one of the most popular materials for commercial NO2 gas sensor manufacture. Considering the sensor performance, the material surface structure plays a key role in the sensitivity, selectivity, and stability.18,19 Generally increase of the active sites on the surface is a prevalent strategy to enhance its sensitivity performance as a gas sensor.20,21 It has been reported that the microstructure of SnO2 as well as its specific surface area can be tuned by mixing it with other different elements.22–24 Keying Shi et al. have demonstrated an enhanced sensitivity performance of SnO2/In2O3 nanorods by increasing the specific surface area.24 With the presence of the strong oxidizing element cerium, an increased amount of oxygen vacancies were observed in SnO2/cerium oxide spheres, leading to an enhanced gas sensitivity performance.25 However, a challenge to enhance both the specific surface area and oxygen vacancies in SnO2 still remains. In this paper, an enhanced gas sensitivity performance in a SnO2/niobium pentoxide (Nb2O5) (SN) hybrid nanowire structure with both enhanced oxygen vacancies and specific surface area induced by the introduction of Nb2O5 is demonstrated. Scanning electron microscopy (SEM), transmission electron microscopy (TEM), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS) and nitrogen (N2) adsorption–desorption isotherms (Brunauer–Emmett–Teller; BET) were utilized for characterizing the structure and chemical composition of the hybrid nanowires. The sensitivity performance of the hybrid structure based gas sensor was examined in both NO2 and an ambient atmosphere environments.
Materials and methods
Chemicals and materials
Poly(vinyl pyrrolidone) (PVP, molecular weight (Mw) = 1
300
000) and niobium chloride (NbCl5) were purchased from Alfa Aesar (China) Chemical Co. Ltd. Stannous chloride (SnCl2), ethanol and dimethyl formamide (DMF) were purchased from Sinopharm Chemical Reagent Co. Ltd.
Method
The SN nanowires were prepared using an electrospinning method. The precursor sources of SnO2 and Nb2O5 were SnCl2 and NbCl5, respectively. In a typical experiment, 0.5 g of SnCl2·H2O was added into a mixture of 2.5 ml ethanol and 2.5 ml DMF which was then stirred for ∼2 h to ensure that the precursors were dissolved. Then 0.15 g of NbCl5 was added into the solution to dissolve. When the solution became transparent, 0.5 g of PVP was added and the mixture was stirred for about 12 h. The solution was then transferred into a syringe. A high voltage power supply was connected with a 22G sized needle. The flow rate was 0.75 ml l−1 and a voltage of 16 kV was applied. The distance between the needle and a collector (a piece of aluminum foil) was fixed at a value of 16 cm. Then the prepared, electrospun SnCl2/NbCl5/PVP nanowires were dried in an oven at a temperature of 80 °C for 12 h in ambient air. To remove the polymer matrix and oxidize two kinds of chloride, the nanowires were annealed in the air at 500 °C for 1 h and the SN hybrid nanowire was obtained. For comparison, the pristine SnO2 nanowires were prepared using the same method without the addition of NbCl5.
Characterization
Field-emission scanning electron microscopy (JSM-6700F, Jeol) was used to explore the morphology of the as-prepared SN nanowires. TEM (Talos F200X, FEI) and XRD (MXPAHF, Mac Science Co. Ltd., Japan) were utilized for the microstructure investigations. In addition, these nanowires were also examined using XPS (ESCALAB 250Xi, Thermo Fisher Scientific) with Al Kα radiation (1486.6 eV).
Gas sensing experiments
The samples were ground with deionized water at a volume ratio of 4
:
1 and were painted uniformly on the interdigital electrode. After drying, the device was kept at 400 °C for one day and then it was used for gas sensor investigation.
Results and discussion
The morphologies and microstructures of the as-synthesized nanowires were examined using SEM and TEM as shown in Fig. 1. The rough surface indicates that the nanowires are composed of polycrystalline nanoparticles. During the annealing process, the polymer was oxidized and overflowed from the surface of the nanowires. In the meanwhile, SnCl2 and NbCl5 were also oxidized and the SN nanowires were formed internally. The energy dispersive spectroscopy (EDS) characterization confirms the chemical composition of the SN nanowires (Fig. 1c–f). This reveals that niobium (Nb) atoms are mainly located in the nanowire core whereas more tin (Sn) atoms are distributed on the surface of the SN nanowires. Because the ionic radius of Sn4+ is smaller than Nb5+, the Sn4+ ionic becomes easier to research the surface of the nanowires during an annealing process.19,24 Therefore, a core/shell-like structure is observed in the SN nanowires. Two types of lattice fringes (Fig. 1g) were also observed at the surface and core of the hybrid structure. They are indexed to be (110) plane of SnO2 shell and (100) plane of the Nb2O5 core as shown in the Fig. 1h and i. The XRD pattern of the hybrid structure (Fig. 1j) shows the SnO2 representative planes of (110), (101) and (211). At 22.3°, the Nb2O5 representative peak of (100) was also observed. Because the molar ratio of the Sn and Nb in the synthesis precursors is 5
:
1, the peak intensity of SnO2 is much stronger than that of Nb2O5.
 |
| | Fig. 1 (a and b) The SEM images of SnO2/Nb2O5 and pure SnO2 nanowires. (c) The transmission electron microscope image. (d–f) The corresponding energy-filtered transmission electron microscopy (EFTEM) elemental mapping of Sn, Nb and O. (g) The high-resolution TEM (HRTEM image of SnO2/Nb2O5). (h and i) Partially enlarged area of the HRTEM image shown in (g). (j) The XRD patterns of SnO2/Nb2O5 (red line) and SnO2 (black line) nanowires. The peak of the Nb2O5 is labeled with red stars and the others are the peaks of the SnO2 phase. | |
To further confirm the structure and the chemical composition of the hybrid structure, XPS was used. The high-resolution spectra (Fig. 2a–c) shows the binding energy of SnO2 and Nb2O5 in the hybrid structure. The O 1s spectra (Fig. 2a) can be divided into two contributions from 529.5 eV and 530.5 eV, which are from the O 1s levels of tin(II) oxide (SnO) and SnO2, respectively. Compared to the pure SnO2, the peak intensity of SnO in the hybrid structure is much higher. In general, the existence of the SnO phase in the SnO2 implies that the formation of oxygen vacancies occurs.3,26,27 Therefore, the amount of oxygen vacancies in the hybrid structure is much higher than that of pure SnO2. This is attributed to the strong oxidizing property of Nb, which causes more of the SnO phase to be formed in the hybrid structure.28 As shown in Fig. 2b, there are four different states in the Sn 3d spectra. The peaks labeled by the red line and green lines represent Sn4+ 3d3/2, and the peaks labeled by the blue and pink lines represent Sn2+ 3d3/2. This further confirms the existence of the SnO phase in both the SN and SnO2 nanowires. The XPS spectra of the Nb binding energy (Fig. 2c) also prove that there is Nb2O5 phase existing in the hybrid structure, which can be utilized to tune the surface state of the SnO2.
 |
| | Fig. 2 The high-resolution XPS spectra showing the binding energy of (a) O 1s and (b) Sn 3d and (c) Nb 3d in SnO2/Nb2O5 (lower spectra) and SnO2 (upper spectra) nanowires, respectively. | |
To explore the specific surface area of the samples, N2 adsorption–desorption isotherms were measured and characterized using the BET pore size distribution. These measurements revealed that the BET surface area of the SN hybrid nanowire is 31.7476 m2 g−1, which is twice that of the pristine SnO2 (14.6543 m2 g−1) as shown in Fig. 3. The pore size of the SN hybrid nanowires are mainly distributed around 110 nm whereas the pristine SnO2 has a four times larger value of the pore size as shown in the insets of Fig. 3.
 |
| | Fig. 3 BET results showing the specific surface area and pore size distributions of (a) SnO2/Nb2O5 and (b) SnO2 nanowires. | |
In order to explore the gas sensing properties of the SN hybrid nanowires, a gas sensor device was fabricated by painting the nanowires onto an interdigital electrode. The resistance change of the device was monitored as it was exposed to NO2 of different concentration ranges from 0.5 ppm to 50 ppm. These experiments were carried out under a dry air diluted NO2 environment. The parabolic shaped curve shown in Fig. 4a shows the response of the SN nanowires as a function of temperature, indicating that the best working temperature for the device is 200 °C. The dynamic sensing response versus time curves is shown in Fig. 4b. The sensor sensitivity was defined as R = Rg/Ra. Ra and Rg represent the resistance in air and NO2, respectively. At low concentrations, the sensitivity of the SN nanowires is much better than that of pure SnO2, and there is 3.5 times larger resistance change at a NO2 concentration of 3 ppm. As demonstrated, in the range of 0.5 ppm to 50 ppm, the SN nanowire-based devices maintain a strong sensitivity performance compared to that of the pristine SnO2. The response time and recovery time are another two important parameters used to evaluate the performance of a sensor. The response time of the hybrid structure based device decreases (Fig. 4c) whereas the response time of the SnO2-based device exhibits certain fluctuations as the NO2 concentration increases. Compared to the recovery time of SnO2, the hybrid structure based device under different NO2 concentrations presents a more stable behavior as shown in Fig. 4d.
 |
| | Fig. 4 (a) Resistance change response of SnO2/Nb2O5 at different temperatures; (b) the resistance change response of SnO2/Nb2O5 (red line) and SnO2 (black line) under an NO2 environment; (c and d) the response time and recovery time of SnO2/Nb2O5 (red line) and SnO2 (black line) nanowires. | |
In the hybrid structure, the SnO2 phase is mixed with the Nb2O5 phase as discussed previously. Therefore, the strong oxidizing property of Nb traps more oxygen surrounding the Nb atoms and causes less oxygen to be surrounding the Sn atoms in the hybrid nanowires during the synthesis process.28 As a result, there are more oxygen vacancies formed in the hybrid structure compared to the pristine SnO2. This is consistent with XPS results obtained in this research, which indicate more of the SnO phase is observed in the hybrid structure. As shown in Fig. 1, the core of the hybrid nanowire is mostly occupied by the aggregated Nb2O5 nanoparticles and therefore the surface of the hybrid structure was mainly coated with SnO2 nanoparticles. Compared to the pristine SnO2, there are more active reaction sites exposed to the NO2 gas in the hybrid structure. As a result, the SN nanowires exhibit a better sensitivity performance as a NO2 gas sensor with the introduction of Nb. Selectivity and stability are two of the important parameters used for evaluating the sensing performance. These results reveal that the SnO2/Nb2O5 exhibits an apparently higher sensing performance in NO2 compared to other different gas environments. In addition, the SN nanowires exhibit an excellent stable sensing performance (see the ESI†). To explore the effect of gas on the morphology, structure, vacancy concentration, SEM, XPS characterizations were used on the sample of SN nanowires after the sensing measurement (see the ESI†). The results show that there is rarely any effect on the SN nanowires during the sensor performance. Although the response and recovery time of the sample was slightly longer, the best working temperature was lower than that of the reported SnO2 and In2O3 sensor devices (Table S1, ESI†).
Conclusion
In this paper, a hybrid structure of SN nanowires is reported which can be applied in a NO2 gas sensor. The strong oxidizing property of Nb introduced into the hybrid structure caused a large amount of oxygen vacancies. This indicated that the aggregation behavior of Nb2O5 nanoparticles led to more SnO2 being exposed to the surface of the hybrid structure compared to with pristine SnO2. Therefore, the hybrid structure based sensor device exhibited an enhanced sensitivity performance. These results provide an effective method to tune the material surface state with a selection of compositions.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (21373196, 11434009, 11474249), the National Program for Thousand Young Talents of China and the Fundamental Research Funds for the Central Universities (WK2340000050, WK2060140014).
References
- B. Y. Kim, J. S. Cho, J. W. Yoon, C. W. Na, C. S. Lee, J. H. Ahn, Y. C. Kang and J. H. Lee, Sens. Actuators, B, 2016, 234, 353–360 CrossRef CAS.
- A. Stanoiua, S. Somacescub, J. M. C. Morenob, V. S. Teodorescua, O. G. Floreaa, A. Sackmannc and C. E. Simion, Sens. Actuators, B, 2016, 231, 166–174 CrossRef.
- J. S. Jang, S. J. Kim, S. J. Choi, N. H. Kim, M. Hakim, A. Rothschild and I. D. Kim, Nanoscale, 2015, 7, 16417–16426 RSC.
- M. W. G. Hoffmann, J. D. Prades, L. Mayrhofer, F. H. Ramirez, T. T. Järvi, M. Modeler, A. Waag and H. Shen, Adv. Funct. Mater., 2014, 24, 595–602 CrossRef CAS.
- S. Maeng, S. W. Kim, D. H. Lee, S. E. Moon, K. C. Kim and A. Maiti, ACS Appl. Mater. Interfaces, 2014, 6, 357–363 CAS.
- C. S. Rout, M. Hegde, A. Govindaraj and C. N. R. Rao, Nanotechnology, 2007, 18, 205504–205513 CrossRef.
- C. S. Rout, M. Hegde and C. N. R. Rao, Sens. Actuators, B, 2008, 128, 488–493 CrossRef CAS.
- C. S. Rout, A. Govindaraj and C. N. R. Rao, J. Mater. Chem., 2006, 16, 3936–3941 RSC.
- C. S. Rout, G. U. Kulkarni and C. N. R. Rao, J. Phys. D: Appl. Phys., 2007, 40, 2777–2782 CrossRef CAS.
- C. S. Rout, K. Ganesh, A. Govindaraj and C. N. R. Rao, Appl. Phys. A, 2006, 85, 241–246 CrossRef CAS.
- A. Sharma, M. Tomar and V. Gupta, Sens. Actuators, B, 2011, 156, 743–752 CrossRef CAS.
- D. J. Yang, I. Kamienchick, D. Y. Youn, A. Rothschild and I. D. Kim, Adv. Funct. Mater., 2010, 20, 4258–4264 CrossRef CAS.
- J. Y. Liu, M. J. Dai, T. S. Wang, P. Sun, X. S. Liang, G. Y. Lu, K. Shimanoe and N. Yamazoe, ACS Appl. Mater. Interfaces, 2016, 8, 6669–6677 CAS.
- S. Cui, Z. H. Wen, X. K. Huang, J. B. Chang and J. H. Chen, Small, 2015, 11, 2305–2313 CrossRef CAS PubMed.
- Y. L. Wei, C. L. Chen, G. Z. Yuan and S. Gao, J. Alloys Compd., 2016, 681, 43–49 CrossRef CAS.
- M. Epifani, J. D. Prades, E. Comini, E. Pellicer, M. Avella, P. Siciliano, G. Faglia, A. Cirera, R. Scotti, F. Morazzoni and J. R. Morante, J. Phys. Chem. C, 2008, 112, 19540–19546 CAS.
- M. Chen, Z. H. Wang, D. M. Han, F. H. Gu and G. S. Guo, J. Phys. Chem. C, 2011, 115, 12763–12773 CAS.
- R. Fiz, F. H. Ramirez, T. Fischer, L. L. Conesa, S. Estrade, F. Peiro and S. Mathur, J. Phys. Chem. C, 2013, 117, 10086–10094 CAS.
- M. Bagheri, M. Y. Masoomi, A. Morsali and A. Schoedel, ACS Appl. Mater. Interfaces, 2016, 9, 21472–21479 Search PubMed.
- J. Zhao, W. N. Wang, Y. P. Liu, J. M. Ma, X. W. Li, Y. Du and G. Y. Lu, Sens. Actuators, B, 2011, 160, 604–608 CrossRef CAS.
- R. B. Wang, S. Yang, R. Deng, W. Chen, Y. L. Liu, H. Zhang and G. S. Zakharov, RSC Adv., 2015, 5, 41050–41058 RSC.
- L. Xiao, S. M. Shu and S. T. Liu, Sens. Actuators, B, 2015, 221, 120–126 CrossRef CAS.
- Y. Qu, H. Wang, H. Chen, J. Xiao, Z. D. Lin and K. Dai, RSC Adv., 2015, 5, 16446–16449 RSC.
- S. Xu, J. Gao, L. L. Wang, K. Kan, Y. Xie, P. K. Shen, L. Li and K. Y. Shi, Nanoscale, 2015, 7, 14643–14651 RSC.
- J. S. Cho and Y. C. Kang, Small, 2015, 11, 4673–4681 CrossRef CAS PubMed.
- M. Batzill, K. Katsiev, J. M. Burst and U. Diebold, Phys. Rev. B: Condens. Matter Mater. Phys., 2015, 72, 165414 CrossRef.
- Y. Nagasawa, T. Choso, T. Karasuda, S. Shimomura, F. Ouyang, K. Tabata and Y. Yamaguchi, Surf. Sci., 1999, 226, 433–435 Search PubMed.
- M. Romero, L. Huerta, T. Akachi, J. L. S. Llamazares and R. Escamilla, J. Alloys Compd., 2013, 579, 516–520 CrossRef CAS.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra21672b |
|
| This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.