DOI:
10.1039/C6RA09033H
(Communication)
RSC Adv., 2016,
6, 65146-65151
Surface defect engineering: gigantic enhancement in the optical and gas detection ability of metal oxide sensor†
Received
8th April 2016
, Accepted 22nd June 2016
First published on 23rd June 2016
Abstract
In this work, the detection ability of nanosensors can be improved extraordinarily via surface defect engineering. A kinked SnO2−X/SnO2 nanostructure was fabricated by tuning the oxygen flow, and this kinked SnO2−X/SnO2 nanostructure was used to study the mechanism of surface defect (oxygen vacancy, VO) effects via electric measurements. For UV light sensing, the response of the SnO2−X NW device is always better than the SnO2 NW device, and is two orders higher under pure O2 surrounding conditions. The detection mechanism can be clarified by changing the detection environment (oxygen concentration) and the UV light detection sensitivity can be improved by increasing the surface VO density. Furthermore, the SnO2−X NW device is very sensitive to its surrounding environment due to the high surface VO density. Hence, CO/O2 alternate-detection was used to verify our hypothesis; the results show that the SnO2−X NW device presents great detection abilities, compared with the SnO2 NW device. The sensitivity of the SnO2−X NW device is two orders higher and the reset/response time is faster, compared with the SnO2 NW device. To verify this hypothesis, the polycrystalline structure was fabricated to prove that the detection ability of metal oxide nanosensors can be improved gigantically by increasing surface defect amounts.
Introduction
Recently, nanomaterials, such as nanowires,1–4 nanobelts5–7 and nanotubes,8–10 and polymer materials,11 have been extensively studied and utilized to form nano-devices for different applications.12–18 Metal oxide nanomaterials, because of the high surface-to-volume ratio and surface defects of their nanostructures, are quite sensitive to variation in the environment.19–26 Many articles reported that Schottky contact mechanisms can enhance the performance of metal oxide nanomaterials, such as in piezotronic devices,27,28 nanogenerators29–32 and nanosensors.20,33–39
For sensor applications, using Schottky contacts as control gates can improve the sensitivity, and the response and reset time.39–42 Besides, the defect amount of metal oxide nanomaterials is the main parameter for sensor applications.43–49 The surface defects50 and the oxygen vacancies51 of metal oxide nanomaterials have been reported and discussed widely. Many research works tried to control the defect amount and level of metal oxides in chemical or physical ways to create other possibilities or applications.52–60 Besides, the signal output of Schottky contact device has strong relation to the surface defect.
Tin dioxide, SnO2, has potential for ultraviolet (UV) light detection applications, due to its large band gap.61 Otherwise, SnO2 is also a candidate for gas detection because of its oxygen vacancies.18,39,48,57,58 Besides, SnO2 can also be used as a photocatalyst and in solar-cell applications.55,62–64 In this research work, tin oxide nanomaterials were fabricated with different levels of defects, and the devices were formed as Schottky gate devices for mechanism investigation. The relationship between the detection ability and defect amount of the tin oxide nanosensor can be figured out by using the Schottky contact devices made with this kinked structure.
Results and discussion
The detailed analyses of the tin oxide nanostructure (NS) can be obtained from the TEM images, as shown in Fig. 1. The growth direction of the tin oxide NS can be identified from the diffraction pattern and high resolution TEM images, as shown in Fig. 1(a)–(c). Due to our control of the synthesis, the oxygen composition of both sides would be different; a lower oxygen composition was detected for the NS growth without oxygen flow, as shown in Fig. 1(d)–(f). The schematic and SEM images of the nanodevice can be seen in Fig. 1(g). The Pt was deposited via focused ion beam (FIB) to form ohmic contacts; the free contacts on both sides can form Schottky contacts as detection units.65 The composition of each side of the nanodevice can be identified via EDS analysis of the system, as illustrated in Fig. 1(h).
 |
| Fig. 1 (a) Low magnification TEM image of a kinked SnO2−X/SnO2 nanostructure with the growth orientation changed from (200) to (002); the growth orientation is indicated by the selected area electron diffraction (SAED) pattern in the inset. (b) and (c) High resolution images and lattice constants of SnO2 and SnO2−X, respectively. (d) The Sn and O concentrations of a kinked SnO2−X/SnO2 nanowire analyzed via EDS element line scan. (e) and (f) The elemental mapping of Sn and O, respectively. (g) Schematic diagram and SEM image of the SnO2−X/SnO2 nanowire Schottky contacted device. (h) EDS analysis and the atomic percentage data of SnO2−X/SnO2 NWs. | |
For photodetection, the sensitivity of the SnO2−X NW device for UV (254 nm) light detection is higher than the SnO2 NW device, as shown in Fig. 2(a), because the dark current (ID) of the SnO2−X NW device is smaller than the ID of the SnO2 NW device and the photo current (IP) of the SnO2−X NW device is larger than the IP of the SnO2 NW device. The measurement was taken at room temperature, 25 °C. The UV detection ability of both of the devices is related to the oxygen concentration of the sensing environment, as illustrated in Fig. 2(b). The sensitivity of the SnO2−X NW device is always higher than the SnO2 NW device, no matter what the oxygen concentration percentage is. The UV detection sensitivity of both of the NW devices will be increased when the oxygen concentration is increasing. The sensing mechanism is illustrated in Fig. 2(c) and (d) for each device.
 |
| Fig. 2 (a) 254 nm UV detection performance in a 40% oxygen gas environment (with 40% oxygen and 60% nitrogen). (b) The sensitivities at different oxygen concentrations of the SnO2−X and SnO2 NW devices. (c) and (d) The mechanism diagrams of the SnO2−X and SnO2 NW devices and oxygen molecule interaction at the interface. Step (i) and (ii) represent the adsorption of oxygen molecules. Step (iii) shows the desorption of oxygen molecules via UV illumination. The SnO2−X NW device has higher sensitivity because its VO is higher. When the UV light was off, more oxygen molecules were trapped at the Schottky contact interface to form O2− and raise the Schottky barrier height (SBH) to reduce the current, so the ID of SnO2−X NW device was lower. But turn on the UV light, and electron–hole pairs would be generated and O2− would be desorbed by the hole. Because the hole would combine with the O2− to form O2(g) and desorb, that would reduce the SBH to increase the IP. (e) The variation of SBH from a vacuum environment to pure oxygen environment (with 100% oxygen) of the SnO2−X and SnO2 NW devices. | |
The SnO2−X NW device has an impressive improvement in sensitivity because the density of surface oxygen vacancies (VO) is higher. When the UV light is off, the SnO2−X NW device would trap more oxygen molecules at the Schottky contact interface to form O2− and raise the Schottky barrier height (SBH) to reduce the current, so the ID of the SnO2−X NW device would be lower than the SnO2 NW device. But when the UV light is on, electron–hole pairs would be generated and O2− would be desorbed by the hole. Because the hole would combine with the O2− to form O2(g) and desorb, that would reduce the SBH to increase the IP. Besides, the lone pair electrons will increase the IP of the SnO2−X NW device, as compared with the SnO2 NW device, the amount of lone pair electrons in the SnO2−X NW device is larger, as shown in Fig. 2(c) and (d). Due to the above-mentioned results, the response can be improved no matter what the oxygen concentration by using the SnO2−X NW device, compared with the SnO2 NW device. The SBH variation also can be analyzed from the electrical measurements, the ΔΦ (SBH variation) would be stable for the SnO2 NW device after 40% oxygen concentration. But for the SnO2−X NW device, the ΔΦ would increase with the increasing concentration of oxygen, as shown in Fig. 2(e). The SBH can be approximately presented by the following equation:66
where Ir is the reverse bias we set, A is the contact area between the SnO2 NW and Pt electrode, A* is the effective Richardson constant, q is the electric charge, Φeff is the SBH, and k and T are the Boltzmann constant and system temperature, respectively.
The ΔΦ of both of the devices is almost the same for an oxygen concentration of around 20%; but when the oxygen concentration increases to above 40%, the ΔΦ would be different. This result symbolizes that the surface VO density of the SnO2−X NW device is higher than that of the SnO2 NW device based on the mechanism illustrated in Fig. 2(c) and (d), which is also consistent with the TEM and SEM analyses. The SBH differences can be measured and analyzed between the SnO2−X and SnO2 NW devices via oxygen concentration variation, and the difference will be enlarged as the oxygen concentration increases, as shown in Table 1 (detailed measurements can be seen in Fig. S1†). The above result shows that the tin oxide based NW device would be very sensitive to its surrounding environment, especially for oxygen concentration. So we can presume that the tin oxide based NW device will have great detection ability for gas, especially for the SnO2−X NW device.
Table 1 The variation and difference of the two Schottky barrier heights in different O2 concentration environments
Oxygen concentration (y%) |
ΔΦy–(y-20%) |
ΦSnO2−X− ΦSnO2 |
SnO2−X |
SnO2 |
0% |
— |
— |
— |
20% |
85 meV |
83 meV |
2 meV |
40% |
27 meV |
21 meV |
8 meV |
60% |
6 meV |
3 meV |
11 meV |
80% |
18 meV |
5 meV |
25 meV |
100% |
26 meV |
4 meV |
47 meV |
From our data, the gas detection ability of the SnO2−X NW device is better than the SnO2 NW device, as shown in Fig. 3. For gas detection, the O2-detection current density (JO2) was used for the base current (J0 = JO2) and the CO-detection current density (JCO) was used for the reaction current (J = JCO). The sensitivity (S) can be defined as S = ΔJ/J0, and ΔJ = JCO − JO2. For sensing, the CO detection signal output of the SnO2−X NW device is larger than the SnO2 NW device, and the J0 of the SnO2−X NW device is smaller than that of the SnO2 NW device, as can be seen in Fig. 3(a). This means that the sensitivity of the SnO2−X NW device is higher than the SnO2 NW device for the detection of each individual CO concentration, as shown in Fig. 3(b). The repetition of both devices for 2 ppm CO detection can be seen in Fig. S3.† The detection mechanism is described in Fig. 3(c). Because of the high surface VO density of the SnO2−X NW device, the JO2 can be reduced due to the SBH raising on oxygen adsorption. Considering CO/O2 alternate-detection, the SBH variation of the SnO2−X NW device is larger than the SnO2 NW device. Based on this surface VO density difference, the improvements in the gas detection abilities (the sensitivity, and the response and reset time) of the SnO2−X NW device are gigantic, compared with the SnO2 NW device. From Table 2, we can see higher sensitivity and a faster reset time on increasing the CO concentration by using the SnO2−X NW device (detailed data can be seen in Fig. S2†). That is because a higher CO concentration can reduce the SBH more, leading to a higher current (JCO); so when the O2 flows in to raise the SBH, the current (JO2) will be decreased immediately. But the response time is related to the influence of gas-surface interactions, and the interactions most depend on the temperature effect. So the response time will be unaffected by the different CO concentrations. The results can support our hypothesis that increasing surface defects can improve the response. So we synthesized another polycrystalline nanostructure to reconfirm our hypothesis. The polycrystalline structure we use is SnO2 (p-SnO2) and ZnO (p-ZnO) nanowire to compare with single-crystalline SnO2 (s-SnO2). The sensitivity enhancements of p-SnO2 and p-ZnO are two orders larger than that of s-SnO2 for UV light detection; the sensitivities of s-SnO2, p-SnO2 and p-ZnO are 9.4%, 132.3% and 130.7%, as seen in Fig. S4.†
 |
| Fig. 3 (a) Detection performance of O2 for sensing different CO concentrations at 200 °C operation temperature. (b) The sensitivities of the SnO2−X NW device are higher than the SnO2 NW device for CO detection at different concentrations. (c) The detection mechanism of the SnO2−X and SnO2 NW devices. Considering CO/O2 alternate-detection, the SBH variation of the SnO2−X NW device is larger than the SnO2 NW device (ΔΦSnO2−X > ΔΦSnO2), due to the high surface VO density of the SnO2−X NW device. | |
Table 2 The sensitivity, response time and reset time of CO gas sensing at different concentrations
CO concentration |
|
Sensitivity |
τresponse |
τreset |
2 ppm |
SnO2−X |
490 ± 43% |
169 s |
221 s |
SnO2 |
75 ± 7% |
180 s |
238 s |
50 ppm |
SnO2−X |
27 549 ± 1521% |
234 s |
7 s |
SnO2 |
254 ± 53% |
451 s |
9 s |
100 ppm |
SnO2−X |
43 657 ± 2232% |
420 s |
2 s |
SnO2 |
541 ± 124% |
630 s |
3 s |
200 ppm |
SnO2−X |
46 726 ± 780% |
225 s |
<1 s |
SnO2 |
551 ± 76% |
341 s |
2 s |
Experiment
Asymmetric kinked SnO2/SnO2−X nanowires
The kinked SnO2−X/SnO2 NWs were synthesized on silicon substrates at 850 °C via the general catalyst free thermal evaporation method using SnO2 and carbon powders as sources in a horizontal quartz tube connected to a vacuum pump and a programmable mass flow controller (MFC). The source was placed in an alumina boat located at the high temperature zone of 1000 °C. The substrates were then positioned in the source downstream. After the tube had been sealed and evacuated to the base pressure, a carrier gas, Ar/O2, was kept flowing through the tube to direct the deposition process. Initially, SnO2 NWs were grown at 100 standard cubic centimeters per minute (sccm) Ar/O2 mixture gas with the volume ratio of 5
:
1 for 15 min to create the stoichiometric SnO2 segment, as shown in Fig. S5.† After reaction, the oxygen input was turned off, and the background pressure was kept at 4 torr to create the SnO2−X segment. The growth plane of the kinked NW shifts from (200) to (002) owing to the external pressure perturbation.
Conclusions
In this research work, we have demonstrated that a kinked SnO2−X/SnO2 nanostructure can be formed by controlling the oxygen flow. We used this kinked SnO2−X/SnO2 nanostructure to form Schottky contact devices and to study the effects of the surface VO density via optical and gas condition electric measurements. For UV light detection, the gigantic enhancement in the sensitivity of the SnO2−X NW device is because the VO can generate more electron–hole pairs to improve the photo response. We also studied the mechanism by controlling the detection environment (oxygen concentration) and we figured out that the ΔΦ is related to the oxygen concentration. Based on this result, we found that the SnO2−X NW device is quite sensitive to its surrounding environment. So we used CO/O2 alternate-detection to verify our hypothesis, and the results show that the SnO2−X NW device presents great detection abilities, compared with the SnO2 NW device. The sensitivity of the SnO2−X NW device is two orders larger than that of the SnO2 NW device when the CO concentration is over 50 ppm. The response and reset time both improved by using the SnO2−X NW device. We proved that increasing the surface defects and using Schottky contacts can improve the detection ability of metal oxide nanosensors extraordinary. Based on the above idea we also designed another, polycrystalline, nanostructure to reconfirm the hypothesis. The sensitivities of p-SnO2 and p-ZnO devices are 132.3% and 130.7%, and all perform better than s-SnO2, at 9.4%. We can use the technologies of material science engineering and physics to design high-resolution and fast monitor speed nanosensors.
Acknowledgements
This research was supported by the Ministry of Science and Technology, Taiwan, under grants NSC-101-2112-M-032-004-MY3 and MOST-104-2112-M-032-003-006.
References
- S. Xu and Z. Wang, Nano Res., 2011, 4, 1013–1098 CrossRef CAS.
- P. Gao and Z. Wang, in Scanning Microscopy for Nanotechnology, Springer New York, 2007, ch. 13, pp. 384–426 Search PubMed.
- Ü. Özgür, Y. I. Alivov, C. Liu, A. Teke, M. A. Reshchikov, S. Doğan, V. Avrutin, S. J. Cho and H. Morkoç, J. Appl. Phys., 2005, 98, 041301 CrossRef.
- Z. Wang, Chin. Sci. Bull., 2009, 54, 4021–4034 CrossRef CAS.
- J. Wang, D. N. Tafen, J. P. Lewis, Z. Hong, A. Manivannan, M. Zhi, M. Li and N. Wu, J. Am. Chem. Soc., 2009, 131, 12290–12297 CrossRef CAS PubMed.
- K. Shankar, J. I. Basham, N. K. Allam, O. K. Varghese, G. K. Mor, X. Feng, M. Paulose, J. A. Seabold, K.-S. Choi and C. A. Grimes, J. Phys. Chem. C, 2009, 113, 6327–6359 CAS.
- O. Lupan, T. Braniste, M. Deng, L. Ghimpu, I. Paulowicz, Y. K. Mishra, L. Kienle, R. Adelung and I. Tiginyanu, Sens. Actuators, B, 2015, 221, 544–555 CrossRef CAS.
- G. She, X. Huang, L. Jin, X. Qi, L. Mu and W. Shi, Small, 2014, 10, 4685–4692 CrossRef CAS PubMed.
- S. Y. Yang, W. Choi and H. Park, ACS Appl. Mater. Interfaces, 2015, 7, 1907–1914 CAS.
- D. H. Kim, Y.-S. Shim, J.-M. Jeon, H. Y. Jeong, S. S. Park, Y.-W. Kim, J.-S. Kim, J.-H. Lee and H. W. Jang, ACS Appl. Mater. Interfaces, 2014, 6, 14779–14784 CAS.
- Y. Gao, Y. Zhao, L. Qiu, Z. Guo, D. O’Hare and Q. Wang, Polym. Compos., 2015 DOI:10.1002/pc.23764.
- X. Wang, N. Aroonyadet, Y. Zhang, M. Mecklenburg, X. Fang, H. Chen, E. Goo and C. Zhou, Nano Lett., 2014, 14, 3014–3022 CrossRef CAS PubMed.
- L. Peng, L. Hu and X. Fang, Adv. Funct. Mater., 2014, 24, 2591–2610 CrossRef CAS.
- S. Jeong, M. Choe, J.-W. Kang, M. W. Kim, W. G. Jung, Y.-C. Leem, J. Chun, B.-J. Kim and S.-J. Park, ACS Appl. Mater. Interfaces, 2014, 6, 6170–6176 CAS.
- A. Pescaglini, A. Martín, D. Cammi, G. Juska, C. Ronning, E. Pelucchi and D. Iacopino, Nano Lett., 2014, 14, 6202–6209 CrossRef CAS PubMed.
- X. Yang, R. Liu, C. Du, P. Dai, Z. Zheng and D. Wang, ACS Appl. Mater. Interfaces, 2014, 6, 12005–12011 CAS.
- N. A. Kyeremateng, ChemElectroChem, 2014, 1, 1442–1466 CrossRef CAS.
- M. Law, H. Kind, B. Messer, F. Kim and P. Yang, Angew. Chem., 2002, 114, 2511–2514 CrossRef.
- G. F. Fine, L. M. Cavanagh, A. Afonja and R. Binions, Sensors, 2010, 10(6), 5469–5502 CrossRef CAS PubMed.
- Z. Yang, X. Dou, S. Zhang, L. Guo, B. Zu, Z. Wu and H. Zeng, Adv. Funct. Mater., 2015, 25, 4039–4048 CrossRef CAS.
- H. S. Woo, C. H. Kwak, J. H. Chung and J. H. Lee, ACS Appl. Mater. Interfaces, 2014, 6, 22553–22560 CAS.
- 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.
- A. Katoch, J. H. Kim and S. S. Kim, ACS Appl. Mater. Interfaces, 2014, 6, 21494–21499 CAS.
- H. Zhou, R. Deng, Y.-F. Li, B. Yao, Z.-H. Ding, Q.-X. Wang, Y. Han, T. Wu and L. Liu, J. Phys. Chem. C, 2014, 118, 6365–6371 CAS.
- F. Hernández-Ramírez, J. Rodríguez, O. Casals, E. Russinyol, A. Vilà, A. Romano-Rodríguez, J. R. Morante and M. Abid, Sens. Actuators, B, 2006, 118, 198–203 CrossRef.
- F. Hernández-Ramírez, A. Tarancón, O. Casals, J. Arbiol, A. Romano-Rodríguez and J. R. Morante, Sens. Actuators, B, 2007, 121, 3–17 CrossRef.
- Z. L. Wang, in Piezotronics and Piezo-Phototronics, Springer, Berlin, Heidelberg, 2012, ch. 1, pp. 1–17 Search PubMed.
- R. Yu, C. Pan, J. Chen, G. Zhu and Z. L. Wang, Adv. Funct. Mater., 2013, 23, 5868–5874 CrossRef CAS.
- W. Seung, M. K. Gupta, K. Y. Lee, K.-S. Shin, J.-H. Lee, T. Y. Kim, S. Kim, J. Lin, J. H. Kim and S.-W. Kim, ACS Nano, 2015, 9, 3501–3509 CrossRef CAS PubMed.
- L. W. Yamin, in Piezoelectric ZnO Nanostructure for Energy Harvesting, John Wiley & Sons, Inc., 2015, ch. 4, pp. 65–103 Search PubMed.
- S. Garain, T. K. Sinha, P. Adhikary, K. Henkel, S. Sen, S. Ram, C. Sinha, D. Schmeißer and D. Mandal, ACS Appl. Mater. Interfaces, 2015, 7, 1298–1307 CAS.
- Y. Wu, Q. Jing, J. Chen, P. Bai, J. Bai, G. Zhu, Y. Su and Z. L. Wang, Adv. Funct. Mater., 2015, 25, 2166–2174 CrossRef CAS.
- T.-Y. Wei, P.-H. Yeh, S.-Y. Lu and Z. L. Wang, J. Am. Chem. Soc., 2009, 131, 17690–17695 CrossRef CAS PubMed.
- C. Pan, R. Yu, S. Niu, G. Zhu and Z. L. Wang, ACS Nano, 2013, 7, 1803–1810 CrossRef CAS PubMed.
- Y. Hu, J. Zhou, P. H. Yeh, Z. Li, T. Y. Wei and Z. L. Wang, Adv. Mater., 2010, 22, 3327–3332 CrossRef CAS PubMed.
- P. Chinnamuthu, J. C. Dhar, A. Mondal, A. Bhattacharyya and N. K. Singh, J. Phys. D: Appl. Phys., 2012, 45, 135102 CrossRef.
- S. Lenaerts, J. Roggen and G. Maes, Spectrochim. Acta, Part A, 1995, 51, 883–894 CrossRef.
- E. R. Viana, J. C. González, G. M. Ribeiro and A. G. de Oliveira, J. Phys. Chem. C, 2013, 117, 7844–7849 CAS.
- C.-M. Chang, C.-H. Hsu, Y.-W. Liu, T.-C. Chien, C.-H. Sung and P.-H. Yeh, Nanoscale, 2015, 7, 20126–20131 RSC.
- Y.-H. Zhang, L.-F. Han, Y.-H. Xiao, D.-Z. Jia, Z.-H. Guo and F. Li, Comput. Mater. Sci., 2013, 69, 222–228 CrossRef CAS.
- C.-H. Sung, T.-C. Chien, C.-M. Chang, C.-M. Chang and P.-H. Yeh, RSC Adv., 2015, 5, 16769–16773 RSC.
- J. K. Hsu, T. Y. Lin, C. Y. Lai, T. C. Chien, J. H. Song and P. H. Yeh, Appl. Phys. Lett., 2013, 103, 123507 CrossRef.
- N. O. Savage, S. A. Akbar and P. K. Dutta, Sens. Actuators, B, 2001, 72, 239–248 CrossRef CAS.
- C. S. Moon, H.-R. Kim, G. Auchterlonie, J. Drennan and J.-H. Lee, Sens. Actuators, B, 2008, 131, 556–564 CrossRef CAS.
- Q. H. Li, T. Gao, Y. G. Wang and T. H. Wang, Appl. Phys. Lett., 2005, 86, 123117 CrossRef.
- S. Mathur, S. Barth, H. Shen, J.-C. Pyun and U. Werner, Small, 2005, 1, 713–717 CrossRef CAS PubMed.
- M. K. Nowotny, L. R. Sheppard, T. Bak and J. Nowotny, J. Phys. Chem. C, 2008, 112, 5275–5300 CAS.
- S. Das and V. Jayaraman, Prog. Mater. Sci., 2014, 66, 112–255 CrossRef CAS.
- J. Nisar, Z. Topalian, A. De Sarkar, L. Österlund and R. Ahuja, ACS Appl. Mater. Interfaces, 2013, 5, 8516–8522 CAS.
- G. Pacchioni, Phys. Chem. Chem. Phys., 2013, 15, 1737–1757 RSC.
- X. Pan, M.-Q. Yang, X. Fu, N. Zhang and Y.-J. Xu, Nanoscale, 2013, 5, 3601–3614 RSC.
- C. Shoou-Jinn, H. Ting-Jen, I. C. Chen and H. Bohr-Ran, Nanotechnology, 2008, 19, 175502 CrossRef PubMed.
- J. M. Wu, Y.-R. Chen and W. T. Kao, ACS Appl. Mater. Interfaces, 2014, 6, 487–494 CAS.
- A. Kar, S. Kundu and A. Patra, J. Phys. Chem. C, 2010, 115, 118–124 Search PubMed.
- Y. Zhang, A. Kolmakov, S. Chretien, H. Metiu and M. Moskovits, Nano Lett., 2004, 4, 403–407 CrossRef CAS.
- S. Bhaumik, A. K. Sinha, S. K. Ray and A. K. Das, IEEE Trans. Magn., 2014, 50, 1–6 CrossRef.
- J. Pan, R. Ganesan, H. Shen and S. Mathur, J. Phys. Chem. C, 2010, 114, 8245–8250 CAS.
- F. Hernandez-Ramirez, J. D. Prades, A. Tarancon, S. Barth, O. Casals, R. Jimenez-Diaz, E. Pellicer, J. Rodriguez, J. R. Morante, M. A. Juli, S. Mathur and A. Romano-Rodriguez, Adv. Funct. Mater., 2008, 18, 2990–2994 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.
- H. Zeng, G. Duan, Y. Li, S. Yang, X. Xu and W. Cai, Adv. Funct. Mater., 2010, 20, 561–572 CrossRef CAS.
- W. Zhou, Y. Liu, Y. Yang and P. Wu, J. Phys. Chem. C, 2014, 118, 6448–6453 CAS.
- Y. Li, J. Zhu, Y. Huang, F. Liu, M. Lv, S. Chen, L. Hu, J. Tang, J. Yao and S. Dai, RSC Adv., 2015, 5, 28424–28429 RSC.
- X. Li, Q. Yu, C. Yu, Y. Huang, R. Li, J. Wang, F. Guo, Y. Zhang, S. Gao and L. Zhao, J. Mater. Chem. A, 2015, 3, 8076–8082 CAS.
- J. Song, E. Zheng, J. Bian, X.-F. Wang, W. Tian, Y. Sanehira and T. Miyasaka, J. Mater. Chem. A, 2015, 3, 10837–10844 CAS.
- P. H. Yeh, Z. Li and Z. L. Wang, Adv. Mater, 2009, 21, 4975–4978 CrossRef CAS PubMed.
- J. M. Shannon, Solid-State Electron., 1976, 19, 537–543 CrossRef.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra09033h |
‡ The first two authors contributed to this work equally. |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.