Next Article in Journal
An ECCD—Electronic Charge Compensation Device—As a Quantum Dissipative System
Previous Article in Journal
Advanced Adaptive Cruise Control Based on Operation Characteristic Estimation and Trajectory Prediction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel X-Ray Radiation Sensor Based on Networked SnO2 Nanowires

1
Department of Materials Science and Engineering, Inha University, Incheon 22212, Korea
2
The Research Institute of Industrial Science, Hanyang University, Seoul 04763, Korea
3
Department of Materials Science and Engineering, Shiraz University of Technology, Shiraz 71557-13876, Iran
4
Division of Materials Science and Engineering, Hanyang University, Seoul 04763, Korea
5
Department of Physics, Kyungpook National University, 80 Daehakro, Bukgu, Daegu 41566, Korea
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2019, 9(22), 4878; https://doi.org/10.3390/app9224878
Submission received: 15 October 2019 / Revised: 5 November 2019 / Accepted: 11 November 2019 / Published: 14 November 2019
(This article belongs to the Section Nanotechnology and Applied Nanosciences)

Abstract

:
X-Ray radiation sensors that work at room temperature are in demand. In this study, a novel, low-cost real-time X-ray radiation sensor based on SnO2 nanowires (NWs) was designed and tested. Networked SnO2 NWs were produced via the vapor–liquid–solid technique. X-ray diffraction (XRD), transmission electron microscopy (TEM) and field emission scanning electron microscopy (SEM) analyses were used to explore the crystallinity and morphology of synthesized SnO2 NWs. The fabricated sensor was exposed to X-rays (80 kV, 0.0–2.00 mA) and the leakage current variations were recorded at room temperature. The SnO2 NWs sensor showed a high and relatively linear response with respect to the X-ray intensity. The X-ray sensing results show the potential of networked SnO2 NWs as novel X-ray sensors.

1. Introduction

Radiation can be broadly divided into two categories: uncharged radiation, including X-rays, gamma rays, and neutrons; charged radiation, including electrons, positrons (beta particles), and alpha particles [1]. Radiation sensors are able to transform the energy lost by an incident particle into an electrical signal, which is then processed by electronic techniques. Metal oxides (MOs) are among the most commonly used materials for advanced applications due to their availability, low cost, simple synthesis procedures and flexible morphology and composition [2]. Accordingly, many researchers have used these materials for different applications, ranging from catalysts [3] to sensors [4,5,6,7].
In particular, they can be used for high-energy radiation sensors. In fact, the radiations with a high energy like gamma rays and X-rays can alter the concentration of oxygen vacancies and also create point defects and other structural defects in MOs [8]. Therefore, metal oxides can be used for the detection of high-energy radiation [9].
Today, low-cost radiation sensors based on MOs operating at room temperature are sought after for applications in science, industry, medicine and security [10]. Additionally, unpredictable accidents may occur during the storage or transportation of radioactive materials, which highlights the needs for metal oxide radiation sensors.
A radiation sensor should have a high sensitivity, a linear performance and an on-line response, low levels of noise and high reliability [11]. Based on the above criteria, a number of MOs were suggested for gamma ray sensing [12,13]. For example, it has been reported that the electrical properties of TeO2 and many other MOs greatly change after gamma ray bombardment [14]. Although X-rays can also affect the structural properties of MOs, there are few reports about X-ray radiation sensors based on MOs in the literature [15,16,17].
SnO2 is a well-known n-type semiconducting metal oxide (Eg = 3.8 eV). It has a high mobility of electrons (160 cm2/Vs), high physical and chemical stability, a high availability and a low price [18]. Therefore, it has been used in many applications, including gas sensors [19], lithium ion batteries [20], photodetectors [21] and electrodes [22]. Therefore, different methods, such as the vapor–liquid–solid (VLS) technique [23,24], colloidal synthesis [25], chemical vapor deposition [26,27] and hydrothermal [28] have been used for the preparation of SnO2 nanowires (NWs). Among them, the VLS method has advantages, such as simplicity, low price and the possibility of control over the SnO2 NWs. Generally, the length of NWs depends on the growth time, and the NW’s diameter is determined by the size of the metal catalyst droplets [29].
In this study, we prepared networked SnO2 NWs and measured their capability to detect X-ray radiation at room temperature. Although networked NWs are very common for gas sensing studies [30,31], there is no report on their X-ray sensing capabilities. To the best of our knowledge, this is the first study reporting X-ray radiation sensing using networked SnO2 NWs. From a sensing response and performance standpoint, networked NWs sensors are potential candidates for radiation sensors because their carrier lifetime is enhanced significantly due to the reduction in defect density on the surface passivated NWs [32]. The results obtained demonstrate the effective sensing capability of SnO2 NWs as X-ray sensors. The sensing mechanism is explained in detail.

2. Materials and Methods

2.1. Synthesis of Networked SnO2 NWs

We used the vapor–liquid–solid (VLS) technique for the selective growth of networked SnO2 NWs, described in detail in earlier reports of the present authors [33,34,35]. To this end, tri-layered interdigitated electrodes comprising of Au (3 nm), Pt (100 nm) and Ti (100 nm) were deposited by a sputtering process on SiO2-grown Si substrates. The Au top layer served as a catalytic layer for the selective growth of SnO2 NWs. Then, the substrates were put into a quartz-tube furnace, where an Al2O3 crucible containing highly pure metallic Sn powders (Sigma-Aldrich, St. Louis, MO, USA, 99.9%) was placed. Afterwards, the furnace was then evacuated using a rotary pump down to a pressure of 1 × 10−3 Torr and was heated (10 °C/min) to 900 °C in the presence of N2 and O2 gases with flow rates of 300 and 10 sccm, respectively. After keeping at this temperature for 15 min, networked SnO2 NWs were successfully grown on the substrate.

2.2. Characterization

The morphology of the SnO2 NWs was studied by field-emission scanning electron microscopy (FE-SEM, Hitachi-S-4200, Hitach, Ltd., Tokyo, Japan). Transmission electron microscopy (TEM) was also used to further investigate the morphology of the synthesized product. X-ray diffraction (XRD, Xpert MPD PRO, Philips, The Netherlands) with CuKα radiation (λ = 1.540 Å) was employed to study the phase and crystalline structure of the synthesized SnO2 NWs.

2.3. X-Ray Sensing studies

An X-ray tube from DRGEM Co., Ltd, Gwangmyeong, Korea. with different currents was used to irradiate the fabricated sensor. The X-ray-induced current was measured by a high-resistance electrometer, a KEITHLEY 6517, versus bias applied voltage which was applied between the interdigitated electrodes, before and after exposure to X-rays. The fabricated sensor was held near the outlet of the X-ray tube where the X-ray intensity was highest. During the X-ray irradiation experiments, temperature and humidity were held constant. Figure 1a shows a digital photograph of the set up for sensing and Figure 1b schematically shows the X-ray sensing procedure.

3. Results and Discussion

3.1. Growth of Networked SnO2 NWs

SnO2 NWs were fabricated using the Au-catalyzed vapor–liquid–solid (VLS) growth method. In this process, ultrafine Au nanoparticles (NPs) were deposited on the Si substrate on which SnO2 NWs were grown. Sn vapors coming from the Sn source were first carried and condensed on the substrate and with the Au formed an alloy and liquefied. With a decrease in temperature and in the presence of flowing oxygen, the SnO2 crystals nucleated at the liquid–solid interface and further condensation/dissolution of Sn vapor increased the amount of SnO2 crystal precipitation from the alloy. The incoming Sn species preferred to diffuse to and condense at the existing solid–liquid interface because of the lower levels of energy involved with the crystal step growth as compared with secondary nucleation events in a finite volume. As a result, no new solid–liquid interfaces were formed, and the interface was pushed to form a SnO2 NW. After the system completely cooled, the alloy droplets solidified on the SnO2 NWs tips [36,37,38].

3.2. Morphological and Structural Analyses

Figure 2 shows a representative XRD pattern taken from networked SnO2 NWs. All the peaks located at 2θ = 26.55, 33.85, 51.85, 54.75, 61.75, 64.85 and 65.95° can be indexed as the (110), (101), (211), (220), (301), (112), and (311) planes of SnO2, respectively, with a tetragonal rutile structure (JCPDS Card No. 88-0287) [39]. The high intensity of the SnO2 Bragg peaks demonstrates the highly crystalline nature of the synthesized SnO2 NWs.
In Figure 3a–c, FE-SEM micrographs with different magnifications are provided. It is evident that NWs are grown selectively on interdigitated electrodes due to the catalytic activity of Au. In particular, Figure 3a indicates a plan-view low-magnified FE-SEM micrograph of networked SnO2 NWs, showing a selective growth of SnO2 NWs. Figure 3b shows a higher magnification image, where long SnO2 NWs with a relatively dense structure and a straight-line morphology can be seen. Figure 3c clearly shows the formation of ultrathin SnO2 NWs, where the diameter of the SnO2 NWs is estimated to be 50–60 nm.
To further examine the crystalline quality of synthesized NWs, TEM analysis was performed. A low-magnification bright-field image of a single NW is presented in Figure 3d. Obviously, the NW is very smooth at the surface and has a single-crystalline nature without any defects, including stacking faults or dislocations. The high-resolution lattice micrograph, presented in Figure 3e, again demonstrates the single-crystalline quality. The spacing between the parallel fringes is ~0.340 nm, which can correspond to the (110) plane of SnO2 [40].

3.3. Leakage Current and X-Ray Sensing Studies

Similarly to visible-light photodetectors, which need a low leakage current to have a high signal-to-noise ratio, semiconductor-based X-ray sensors must have low leakage currents to ensure their high sensitivity. Accordingly, before testing the X-ray radiation sensing capability of the SnO2 NWs, in order to have insight into the leakage current of a fabricated sensor, the typical I-V characteristics of SnO2 NWs were studied. Figure 4a–d shows the FE-SEM images of SnO2 NWs with different densities of SnO2 NWs from the highest to the lowest densities, respectively, and corresponding leakage currents versus applied voltages are shown in Figure 4e–h, respectively. As can be seen, as the density of the SnO2 NWs decreases, the leakage current increases. Therefore, for practical applications, the use of highly dense SnO2 NWs will result in a better sensing performance and will enhance the signal to noise ratio. In particular, the sample with the highest density of SnO2 NWs showed the lowest leakage current. This sample had a very low dark leakage current; for example, for bias voltages of 1 and 5 V, the leakage currents were ~0.25 and 0.9 pA respectively. These very low leakage currents enhance the signal to noise ratio for semiconductor radiation sensors. The leakage current of a semiconductor sensing material is strongly governed by the band gap energy and the structure of the sensing material. Large band gap materials, >1.5 eV, are needed to maintain a low intrinsic carrier concentration and a low leakage current during sensor operation at room temperature. Additionally, the amount of carrier trapping centers must be extremely low. For example, regions near and along the cracks show much higher leakage current than regions away from the cracks [41]. Therefore, the low current leakages obtained for the fabricated sensor demonstrate the high quality of the synthesized SnO2 NWs in this study.
To evaluate the X-ray sensing capability, the X-ray current response of the SnO2 NWs sensor with respect to X-ray intensity was measured at room temperature, as shown in Figure 5. During the measurements, the current of the X-ray source was varied from 0 to 2 mA while the applied voltages for the X-ray generator and sensor were kept constant at 80 kV and 5 V, respectively. The signal intensity was strong as compared to dark current and it was quite linear, as shown in Figure 5. Based on these results, we have a meaningful response of the SnO2 NWs to X-ray radiation. For example, for X-ray currents of 1 and 2 mA, the sensor currents were 65 and 120 pA respectively, demonstrating the near-linearity of the response of the SnO2 NWs sensor.
We also tested the conductivity of SnO2 NWs. Figure 6 shows current passing through the SnO2 NWs under a fixed voltage of 1 V over time. As can be seen, the SnO2 NW sensor shows a stable signal over time, demonstrating its high stability for sensing applications.

3.4. Sensing Mechanism

In semiconductor radiation sensor operation, incident radiation creates a charge pulse consisting of electrons and holes within the detector, which are then separated using an applied electric field and the current is sensed by an external circuit. Electromagnetic radiation can interact with a material via different mechanisms [1]: (i) elastic scattering, (ii) photo-electric absorption, (iii) Compton scattering, and (iv) pair production. However, the dominant mechanism and the result of the interaction strongly depend on the incident energy of the interacting X-ray beam and the composition of the sensor [42]. In the case of elastic scattering, the energy of the incident photon is not changed and this process does not contribute to the deposition of energy on the sensor. Photoelectric absorption is the ideal process for sensor operation, where the incident photon loses all of its energy upon incident with the sensor material to one of the orbital electrons of the atoms of the sensor material. Subsequently, this photoelectron loses its kinetic energy through Columbic interactions and creates plenty of electron–hole pairs, which are subsequently separated by the electric field, and a current is produced [41,43]. During Compton scattering, which is a collision between an incident photon and an orbital electron, both the energy and the direction of the photons change. In fact, some of the energy will be lost to the electrons of the sensing material. These electrons will then lose their energy through the generation of electron–hole pairs. In Compton scattering, a photon does not transfer all of its energy to an electron. Accordingly, the number of electron–hole pairs generated varies remarkably between different Compton events [1]. In the pair production process, a photon with high energy interacts within the Coulomb field of the nucleus and directly produces an electron–positron pair. Generally, the photoelectric effect dominates in the energy range to ~200 keV, the Compton effect to a few MeV and pair production above ~6 MeV [44].
The SnO2 NW is a semiconductor material which can directly convert ionization radiation into an electrical signal. In this study, X-ray irradiation was used to demonstrate the radiation sensing capability of SnO2 NW. When an X-ray traverses the sensor, it may partly or fully transfer its energy to electrons via the photoelectric effect or Compton scattering [45]. These primary energetic electrons may cause an ionization cascade until their energy is in equilibrium with the energy band gap of the media. This leads to the creation of several electron–hole pairs in the sensor. Signal intensity strongly depends on the collection efficiency of these electron–hole pairs. The collection probability is in competition with electron–hole recombination, which is related to carrier lifetime or defect density.

4. Conclusions

In this study, networked SnO2 NWs were fabricated by a VLS technique and were used for the detection of X-ray radiation at room temperature. XRD, TEM and FE-SEM characterizations approved the successful formation of NWs with the desired composition and morphology. The non-proportional response of the X-ray source was tested. The X-ray radiation sensing results show that the SnO2 NWs sensors have promising sensing capability, where low leakage current was observed for the SnO2 NWs sensor. Another strategy to reduce the defect density of the oxide NWs is to perform high-temperature annealing [45]. Stabilization studies of these sensors will also be conducted in future work.

Author Contributions

S.S.K., H.W.K. and H.J.K. conceived and designed the experiments and completed the paper; J.-H.K. and P.Q.V. performed the experiments; J.-H.K., Q.V.P. and A.M. analyzed the data and prepared the draft of the paper. All authors approved the final version of the paper.

Funding

This research was funded by the National Research Foundation of Korea (NRF) via a grant number 2018R1A6A1A06024970 and 2016R1A6A1A03013422 and APC was 100% discounted.

Acknowledgments

This study was supported byInha University.

Conflicts of Interest

The authors declare no conflict of interest and the founding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the decision to publish the results.

References

  1. Schlesinger, T.E.; James, R.B. Semiconductors for Room Temperature Nuclear Applications, 1st ed.; Academic Pres: New York, NY, USA, 1995. [Google Scholar]
  2. Arshak, K.; Korostynska, O.; Fahim, F. Various Structures Based on Nickel Oxide Thick Films as Gamma Radiation Sensors. Sensors 2003, 3, 176–186. [Google Scholar] [CrossRef]
  3. Wei, Z.; Wei, X.; Wang, S.; He, D. Preparation and Visible-Light Photocatalytic Activity of α-Fe2O3/γ-Fe2O3 Magnetic Heterophase Photocatalyst. Mater. Lett. 2014, 118, 107–110. [Google Scholar] [CrossRef]
  4. Kwon, Y.J.; Mirzaei, A.; Kang, S.Y.; Choi, M.S.; Bang, J.H.; Kim, S.S.; Kim, H.W. Synthesis, Characterization and Gas Sensing Properties of ZnO-Decorated MWCNTs. Appl. Surf. Sci. 2017, 413, 242–252. [Google Scholar] [CrossRef]
  5. Gancarz, M.; Nawrocka, A.; Rusinek, R. Identification of Volatile Organic Compounds and Their Concentrations Using a Novel Method Analysis of MOS Sensors Signal. J. Food Sci. 2019, 84, 2077–2085. [Google Scholar] [CrossRef]
  6. Rusinek, R.; Siger, A.; Gawrysiak-Witulska, M.; Rokosik, E.; Malaga-Toboła, U.; Gancarz, M. Application of an Electronic Nose for Determination of Pre-Pressing Treatment of Rapeseed Based on the Analysis of Volatile Compounds Contained in Pressed Oil. Int. J. Food Sci. Technol. 2019. [Google Scholar] [CrossRef]
  7. Liu, H.; Zhu, W.; Han, Y.; Yang, Z.; Huang, Y. Single-Nanowire Fuse for Ionization Gas Detection. Sensors 2019, 19, 4358. [Google Scholar] [CrossRef]
  8. Lavanya, N.; Sekar, C.; Anithaa, A.; Sudhan, N.; Asokan, K.; Bonavita, A.; Leonardi, S.; Neri, G. Investigations on the Effect of Gamma-Ray Irradiation on the Gas Sensing Properties of SnO2 Nanoparticles. Nanotechnology 2016, 27, 385502. [Google Scholar] [CrossRef]
  9. Arshak, K.; Korostynska, O. Effect of Gamma Radiation onto the Properties of TeO2 Thin Films. Microelectron. Int. 2002, 19, 30–34. [Google Scholar] [CrossRef]
  10. Arshak, K.; Korostynska, O.; Clifford, S. Screen Printed Thick Films of NiO and LaFeO3 as Gamma Radiation Sensors. Sens. Actuators A 2004, 110, 354–360. [Google Scholar] [CrossRef]
  11. Arshak, K.; Korostynska, O. Preliminary Studies of Properties of Oxide Thin/Thick Films for Gamma Radiation Dosimetry. Mater. Sci. Eng. B 2004, 107, 224–232. [Google Scholar] [CrossRef]
  12. Arshak, K.; Korostynska, O.; Henry, J. Thick Film pn-Junctions Based on Mixed Oxides of Indium and Silicon as Gamma Radiation Sensors. Microelectron. Int. 2004, 21, 19–27. [Google Scholar] [CrossRef]
  13. Arshak, K.; Korostynska, O. Thin Films of (TeO2)1−x(In2O3)x as Gamma Radiation Sensors. Sens. Rev. 2003, 23, 48–54. [Google Scholar] [CrossRef]
  14. Arshak, K.; Korostynska, O. Gamma Radiation Dosimetry Using Tellurium Dioxide Thin Film Structures. In Proceedings of the IEEE Sensors 2002, Orlando, FL, USA, 12–14 June 2002; Volume 1, pp. 547–551. [Google Scholar]
  15. Lu, X.; Zhou, L.; Chen, L.; Ouyang, X.; Liu, B.; Xu, J.; Tang, H. Schottky X-ray Detectors Based on a Bulk β-Ga2O3 Substrate. Appl. Phys. Lett. 2018, 112, 103502. [Google Scholar] [CrossRef]
  16. Zhou, L.; Huang, Z.; Zhao, X.; He, Y.; Chen, L.; Xu, M.; Zhao, K.; Zhang, S.; Ouyang, X. A High-Resistivity ZnO Film-Based Photoconductive X-ray Detector. IEEE Photonics Technol. Lett. 2019, 31, 365–368. [Google Scholar] [CrossRef]
  17. Katsumata, T.; Takeuchi, H.; Komuro, S.; Aizawa, H. X-ray Detector Based on Mn Doped MgAl2O4 and Si Photodiode. Rev. Sci. Instrum. 2018, 89, 095104. [Google Scholar] [CrossRef]
  18. Kim, J.-H.; Lee, J.-H.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Optimization and Gas Sensing Mechanism of n-SnO2-p-Co3O4 Composite Nanofibers. Sens. Actuators B 2017, 248, 500–511. [Google Scholar] [CrossRef]
  19. Kwon, Y.J.; Kang, S.Y.; Mirzaei, A.; Choi, M.S.; Bang, J.H.; Kim, S.S.; Kim, H.W. Enhancement of Gas Sensing Properties by the Functionalization of ZnO-Branched SnO2 Nanowires with Cr2O3 Nanoparticles. Sens. Actuators B 2017, 249, 656–666. [Google Scholar] [CrossRef]
  20. Liang, G.; Sun, X.; Lai, J.; Wei, C.; Huang, Y.; Hu, H. Adding Lithium Fluoride to Improve the Electrochemical Properties SnO2@C/MWCNTs Composite Anode for Lithium-Ion Batteries. J. Electroanal. Chem. 2019, 853, 113401. [Google Scholar] [CrossRef]
  21. He, J.; Wu, J.; Hu, S.; Shen, H.; Hu, X. A Low-Cost Flexible Broadband Photodetector Based on SnO2/CH3NH3PbI3 Hybrid Structure. Opt. Mater. 2019, 88, 689–694. [Google Scholar] [CrossRef]
  22. Cho, C.J.; Pyeon, J.J.; Hwang, C.S.; Kim, J.S.; Kim, S.K. Atomic Layer Deposition of Ta-Doped SnO2 Films with Enhanced Dopant Distribution for Thermally Stable Capacitor Electrode Applications. Appl. Surf. Sci. 2019, 497, 143804. [Google Scholar] [CrossRef]
  23. Kim, J.-H.; Mirzaei, A.; Bang, J.H.; Kim, H.W.; Kim, S.S. Selective H2S Sensing without External Heat by a Synergy Effect in Self-Heated CuO-Functionalized SnO2-ZnO Core-Shell Nanowires. Sens. Actuators B 2019, 300, 126981. [Google Scholar] [CrossRef]
  24. Zhong, X.; Shen, Y.; Zhao, S.; Li, T.; Lu, R.; Yin, Y.; Han, C.; Wei, D.; Zhang, Y.; Wei, K. Effect of Pore Structure of the Metakaolin-Based Porous Substrate on the Growth of SnO2 Nanowires and Their H2S Sensing Properties. Vacuum 2019, 167, 118–128. [Google Scholar] [CrossRef]
  25. Gao, N.; Li, H.Y.; Zhang, W.; Zhang, Y.; Zeng, Y.; Zhixiang, H.; Liu, J.; Jiang, J.; Miao, L.; Yi, F.; et al. QCM-Based Humidity Sensor and Sensing Properties Employing Colloidal SnO2 Nanowires. Sens. Actuators B 2019, 293, 129–135. [Google Scholar] [CrossRef]
  26. Ngoc, T.M.; Van Duy, N.; Hoa, N.D.; Hung, C.M.; Nguyen, H.; Van Hieu, N. Effective Design and Fabrication of Low-Power-Consumption Self-Heated SnO2 Nanowire Sensors for Reducing Gases. Sens. Actuators B 2019, 295, 144–152. [Google Scholar] [CrossRef]
  27. Chen, Y.; Qiu, W.; Wang, X.; Liu, W.; Wang, J.; Dai, G.; Yuan, Y.; Gao, Y.; Sun, J. Solar-Blind SnO2 Nanowire Photo-Synapses for Associative Learning and Coincidence Detection. Nano Energy 2019, 62, 393–400. [Google Scholar] [CrossRef]
  28. Singh, M.K.; Pandey, R.K.; Prakash, R. High-Performance Photo Detector Based on Hydrothermally Grown SnO2 Nanowire/Reduced Graphene Oxide (rGO) Hybrid Material. Org. Electron. 2017, 50, 359–366. [Google Scholar] [CrossRef]
  29. Suh, D.-I.; Byeon, C.C.; Lee, C.-L. Synthesis and Optical Characterization of Vertically Grown ZnO Nanowires in High Crystallinity Through Vapor–Liquid–Solid Growth Mechanism. Appl. Surf. Sci. 2010, 257, 1454–1456. [Google Scholar] [CrossRef]
  30. Park, S.; Hong, T.; Jung, J.; Lee, C. Room Temperature Hydrogen Sensing of Multiple Networked ZnO/WO3 Core–Shell Nanowire Sensors under UV Illumination. Curr. Appl. Phys. 2014, 14, 1171–1175. [Google Scholar] [CrossRef]
  31. Chinh, N.D.; Van Toan, N.; Van Quang, V.; Van Duy, N.; Hoa, N.D.; Van Hieu, N. Comparative NO2 Gas-Sensing Performance of the Self-Heated Individual, Multiple and Networked SnO2 Nanowire Sensors Fabricated by a Simple Process. Sens. Actuators B 2014, 201, 7–12. [Google Scholar] [CrossRef]
  32. Kar, A.; Stroscio, M.A.; Meyyappan, M.; Gosztola, D.J.; Wiederrecht, G.P.; Dutta, M. Tailoring the Surface Properties and Carrier Dynamics in SnO2 Nanowires. Nanotechnology 2011, 22, 285709. [Google Scholar] [CrossRef]
  33. Abideen, Z.U.; Kim, J.-H.; Kim, S.S. Optimization of Metal Nanoparticle Amount on SnO2 Nanowires to Achieve Superior Gas Sensing Properties. Sens. Actuators B 2017, 238, 374–380. [Google Scholar] [CrossRef]
  34. Choi, S.-W.; Katoch, A.; Sun, G.-J.; Wu, P.; Kim, S.S. NO2-Sensing Performance of SnO2 Microrods by Functionalization of Ag Nanoparticles. J. Mater. Chem. C 2013, 1, 2834–2841. [Google Scholar] [CrossRef]
  35. Park, J.Y.; Choi, S.-W.; Kim, S.S. Junction-Tuned SnO2 Nanowires and Their Sensing Properties. J. Phys. Chem. C 2011, 115, 12774–12781. [Google Scholar] [CrossRef]
  36. Calestani, D.; Zha, M.; Salviati, G.; Lazzarini, L.; Zanotti, L.; Comini, E.; Sberveglieri, G. Nucleation and Growth of SnO2 Nanowires. J. Cryst. Growth 2005, 275, e2083–e2087. [Google Scholar] [CrossRef]
  37. Wu, Y.; Yang, P. Direct Observation of Vapor−Liquid−Solid Nanowire Growth. J. Am. Chem. Soc. 2001, 123, 3165–3166. [Google Scholar] [CrossRef]
  38. Kim, S.; Bang, J.H.; Choi, M.S.; Oum, W.; Mirzaei, A.; Lee, N.; Kwon, H.C.; Lee, D.; Jeon, H.; Kim, S.S.; et al. Synthesis, Characterization and Gas-Sensing Properties of Pristine and SnS2 Functionalized TeO2 Nanowires. Met. Mater. Int. 2019, 25, 805–813. [Google Scholar] [CrossRef]
  39. Kim, J.-H.; Kim, S.S. Realization of ppb-Scale Toluene-Sensing Abilities with Pt-Functionalized SnO2-ZnO Core-Shell Nanowires. ACS Appl. Mater. Interfaces 2015, 7, 17199–17208. [Google Scholar] [CrossRef]
  40. Katoch, A.; Kim, J.-H.; Kwon, Y.J.; Kim, H.W.; Kim, S.S. Bifunctional Sensing Mechanism of SnO2-ZnO Composite Nanofibers for Drastically Enhancing the Sensing Behavior in H2 Gas. ACS Appl. Mater. Interfaces 2015, 7, 11351–11358. [Google Scholar] [CrossRef]
  41. Schlesinger, T.; Toney, J.; Yoon, H.; Lee, E.; Brunett, B.; Franks, L.; James, R. Cadmium Zinc Telluride and Its Use as a Nuclear Radiation Detector Material. Mater. Sci. Eng. R Rep. 2001, 32, 103–189. [Google Scholar] [CrossRef]
  42. Ahmadi, M.; Yeow, J.T. Fabrication and Characterization of a Radiation Sensor Based on Bacteriorhodopsin. Biosens. Bioelectron. 2011, 26, 2171–2176. [Google Scholar] [CrossRef]
  43. Nava, F.; Bertuccio, G.; Cavallini, A.; Vittone, E. Silicon Carbide and Its Use as a Radiation Detector Material. Meas. Sci. Technol. 2008, 19, 102001. [Google Scholar] [CrossRef]
  44. Owens, A.; Peacock, A. Compound Semiconductor Radiation Detectors. Nucl. Instrum. Methods Phys. Res. Sect. A 2004, 531, 18–37. [Google Scholar] [CrossRef]
  45. Knoll, G.F. Radiation Detection and Measurement, 3rd ed.; John Wiley & Sons. Inc.: New York, NY, USA, 2000. [Google Scholar]
Figure 1. (a) Digital photograph of the set up for sensing. (Inset: X-ray sensor). (b) Schematic showing the set up for X-ray sensing.
Figure 1. (a) Digital photograph of the set up for sensing. (Inset: X-ray sensor). (b) Schematic showing the set up for X-ray sensing.
Applsci 09 04878 g001
Figure 2. Representative X-ray diffraction (XRD) pattern of SnO2 nanowires (NWs).
Figure 2. Representative X-ray diffraction (XRD) pattern of SnO2 nanowires (NWs).
Applsci 09 04878 g002
Figure 3. (ac) Field-emission scanning electron microscopy (FE-SEM) micrographs of SnO2 NWs with different magnifications. Transmission electron microscopy (TEM) analyses results; (d) low-magnification bright-field TEM image; (e) high-resolution lattice image.
Figure 3. (ac) Field-emission scanning electron microscopy (FE-SEM) micrographs of SnO2 NWs with different magnifications. Transmission electron microscopy (TEM) analyses results; (d) low-magnification bright-field TEM image; (e) high-resolution lattice image.
Applsci 09 04878 g003
Figure 4. Scanning electron microscopy (SEM) images of SnO2 NWs with different densities of NWs from (a), (b), (c) to (d) the highest density to the lowest density. (e)–(h) Corresponding leakage currents of SnO2 NWs versus applied voltage.
Figure 4. Scanning electron microscopy (SEM) images of SnO2 NWs with different densities of NWs from (a), (b), (c) to (d) the highest density to the lowest density. (e)–(h) Corresponding leakage currents of SnO2 NWs versus applied voltage.
Applsci 09 04878 g004
Figure 5. Dependence of the current response with respect to the intensity of the X-ray source under a constant applied voltage of 5 V.
Figure 5. Dependence of the current response with respect to the intensity of the X-ray source under a constant applied voltage of 5 V.
Applsci 09 04878 g005
Figure 6. Current passing the SnO2 NWs under a fixed voltage of 1 V versus the time.
Figure 6. Current passing the SnO2 NWs under a fixed voltage of 1 V versus the time.
Applsci 09 04878 g006

Share and Cite

MDPI and ACS Style

Kim, J.-H.; Mirzaei, A.; Kim, H.W.; Kim, H.J.; Quoc Vuong, P.; Kim, S.S. A Novel X-Ray Radiation Sensor Based on Networked SnO2 Nanowires. Appl. Sci. 2019, 9, 4878. https://doi.org/10.3390/app9224878

AMA Style

Kim J-H, Mirzaei A, Kim HW, Kim HJ, Quoc Vuong P, Kim SS. A Novel X-Ray Radiation Sensor Based on Networked SnO2 Nanowires. Applied Sciences. 2019; 9(22):4878. https://doi.org/10.3390/app9224878

Chicago/Turabian Style

Kim, Jae-Hun, Ali Mirzaei, Hyoun Woo Kim, Hong Joo Kim, Phan Quoc Vuong, and Sang Sub Kim. 2019. "A Novel X-Ray Radiation Sensor Based on Networked SnO2 Nanowires" Applied Sciences 9, no. 22: 4878. https://doi.org/10.3390/app9224878

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop