Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

ZnO subwavelength wires for fast-response mid-infrared detection

Open Access Open Access

Abstract

Room temperature operating thermal detection for mid-infrared light based on ZnO subwavelength wires has been demonstrated. Electric resistance in ZnO wires increases linearly with the intensity of incident light. Noise equivalent power (NEP) of 5.8 μW/Hz1/2 (at 1 kHz) with typical response times as fast as 1.3 ms is obtained at 10.6-μm wavelength. The sensitivity and response time of the detector are also found to be insensitive to the ambient.

©2009 Optical Society of America

1. Introduction

Mid infrared (IR) detection, typically relying on thermal or photo response [1], has wide applications in the fields of medicine [2], remote sensing [3], environmental monitoring [4], and telecommunications [5]. Generally, IR photon detectors, including photoconductors, quantum well photodetectors, quantum dots photodetectors and superconductor detectors [1,6], offer advantages of fast response and high sensitivity, but usually require low-temperature operation with complex cooling equipments [3]. Thermal detectors, such as thermocouple detectors and resistance thermal detectors, provide the possibility for room-temperate operation with broadband response, but usually suffer from slow response times due to relatively large thermal inertia of the sensitive elements [1,3]. One way to speed up the response of a thermal detector is reducing the thermal inertia or equivalently the size of the sensitive element, through adopting air-bridge microstructure [7]. Recently, low-dimensional micro- or nanostructures have attracted extensive attentions in optical detection [810], among which zinc oxide (ZnO) micro- and nanowires show great commercial potential owing to their low cost and easy fabrication. So far most of the researches on ZnO nanowire detection are focused on UV photodetection [1121]. Considering the relatively strong absorption in mid-IR regime (about 1 mm−1 from 8 to 30 μm) [22] and the excellent chemical and thermal stabilities of ZnO microwire, here we propose mid-IR thermal detection based on ZnO subwavelength wires. In the work, we explore the photothermal response of ZnO subwavelength wires, and demonstrate a room-temperature-operation thermal detector with response time down to 1.3 ms at 10.6-μm wavelength.

2. Detector configuration

ZnO subwavelength wires were synthesized via a chemical vapor transport process [23]. Figure 1 shows a scanning electron microscope (SEM) image of a typical ZnO microwire with 1.0-μm-diameter, smooth surface and hexagonal section. To assemble the detection structure, a ZnO wire, with diameter less than 3 μm, was first transferred to a grooved glass plate by micromanipulation. The wire was placed across two Ti/Au electrodes sputtered on the plate, as schematically illustrated in Fig. 2(a) . To improve the electrical contact properties, a microdrop of In/Ga liquid alloy was used to cover the contact area to ensure an ohmic contact. The middle part of the wire was suspended above the groove (several hundred micrometers in width) to get isolation from the substrate. For reference, a typical as-fabricated detection structure, consisting of a 2.0-μm-diameter, 760-μm-length ZnO wire, is shown in Fig. 2(b).

 figure: Fig. 1

Fig. 1 The SEM image of a 1.0-μm-diameter ZnO wire.

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 The schematic structure (a) and SEM image of a ZnO wire on a grooved glass plate (b).

Download Full Size | PDF

Light from a Coherent K-250 CO2 laser, centered at the wavelength of 10.6 μm, was used to irradiate the ZnO wire. The laser beam was focused by a ZnSe lens (focus length = 5.0 cm) to a 220-μm-diameter spot on the ZnO wire. Since the dark resistance of the ZnO wire is very large, a constantly illumination from a halogen lamp (about 6800 lx) is applied on the ZnO wire for stable and reliable measurement of the response of the ZnO wire.

3. ZnO wire for mid-IR detection

When a ZnO wire absorbs mid-IR light, its temperature rises, leading to the change in the resistance that can be used to retrieve the intensity of the incident light. Figure 3 shows some typical I-V characteristics of a 2.9-μm-diameter, 520-μm-length ZnO wire under irradiation of 10.6-μm-wavelength light at various intensities, for reference, the I-V curve of the ZnO wire without IR irradiation is provided. The linear shape of the curves in the range of −10 to 10 V reveals good ohmic contacts. The current decreases with the increase of irradiation power: as the light intensity increases to 22.8 mW, the current reduces by 38%. Supposing only the 226-μm-length irradiated part of ZnO wire elevating temperature, the resistance of this part increases by 144%.

 figure: Fig. 3

Fig. 3 I-V characteristics of a 2.9-μm-diameter, 520-μm-length ZnO wire as a function of 10.6-μm-wavelength light intensity.

Download Full Size | PDF

Figure 4 shows the irradiation-intensity-dependent resistance of a 2.0-μm-diameter, 760-μm-length ZnO wire. As can be seen, the resistance increases linearly with intensity of the irradiation. The inset image in Fig. 4 shows the repetition-frequency dependence of the resistance change of the ZnO wire to the incident pulses (500 pJ /pulse, 25 μs pulse width). The amplitude of the resistance changes decreases with the increasing repetition frequency. By means of frequency domain analysis [24], we obtain the noise intensity of 700 μV/Hz1/2 (at 1kHz), which corresponds to a noise equivalent power (NEP) of 5.8 μW/Hz1/2.

 figure: Fig. 4

Fig. 4 Resistance of a single 2.0-μm-diameter, 760-μm-length ZnO wire device measured as a function of irradiation intensity. Inset is the frequency-dependent resistance changes of the ZnO wire.

Download Full Size | PDF

Recent study shows that the response of ZnO nanowire for UV photodetection is significantly influenced by ambient [17]. To investigate the influence of ambient on ZnO wire for mid-IR detection, we measured resistance of a 2.0-μm-diameter, 760-μm-length ZnO wire (the same ZnO wire shown Fig. 4) in typical ambient gases including air, argon, nitrogen, and oxygen, with results shown in Fig. 5 . The 10.6-μm-wavelength light was irradiated on the wire with a pulse period of 9 ms and a pulse width of 0.9 ms. Though the background resistance varies in different atmospheres, the amplitude and response time are insensitive to the ambient gases. The estimated response time of the ZnO wire mid-IR detection is about 1.3 ms when the resistance of ZnO wire falls from 37.2 to 34.5 MΩ in the air, which is much faster than other types of room-temperature-operated microbolometers or thermocouples [9,2531], and three orders of magnitude faster than that in ZnO nanowire UV photodetectors [11,14,21].

 figure: Fig. 5

Fig. 5 Response of a 2.0-μm-diameter, 760-μm-length ZnO wire to 10.6-μm-wavelength light irradiation measured in different atmospheres.

Download Full Size | PDF

In UV and mid-IR spectral ranges, ZnO wires have different response mechanisms. When mid-IR photon is absorbed by ZnO wires, the photon energy is converted to the thermal energy, heating up the ZnO wire. When temperature rises, thermal lattice vibration becomes stronger, leading to stronger scattering of carriers in ZnO wire, which is in turn, reduces the mean free paths of carriers, resulting in the increasing of the resistance. On the other hand, the temperature rising may excite electrons to the conduction band and increase the density of carriers, resulting in the decreasing of the resistance. However, at room temperature and above, with the halogen lamp illumination, electrons on the shallow doping levels have already been excited, and the free-carrier density is almost saturated. Therefore, heating of the ZnO wires would not increase the density of carriers obviously, and the resistance of the ZnO wire would increase when it is irradiated by mid-IR light.

As a kind of thermal detection, the response time of the ZnO wire detection can be theoretically estimated using the time constant [7]

τ=H/G,
in which the heat capacity (of the irradiated part of ZnO wire) H is about 5.5 × 10−9 J/K, and the thermal conductivity G is about 5.0 × 10−6 W/K [32,33]. Calculated τ is about 1.1 ms, which coincides well with the measured value of 1.3 ms.

4. Conclusion

In conclusion, we demonstrate fast thermal detection of mid-IR light based on ZnO wires. A NEP of 5.8 μW/Hz1/2 (at 1kHz) and a typical response time of 1.3 ms are obtained at 10.6-μm wavelength. The sensitivity and response time of the detector are found to be insensitive to the ambient. The low thermal inertia of ZnO wire allows the response time down to the order of millisecond. Although the light used in this work is a monochromatic 10.6-μm-wavelength laser, the fast and sensitive response of the ZnO wire can be extended to a wider mid-IR spectrum owning to the broadband absorption of ZnO in the mid-IR spectral range, and will be promising for fast-response mid-infrared detection.

Acknowledgements

This work is supported by the National Natural Science Foundation of China (Grant No. 60706020) and the National Basic Research Program of China (Grant No. 2007CB307003).

References and links

1. A. Rogalski, “Infrared detectors: an overview,” Infrared Phys. Technol. 43(3-5), 187–210 ( 2002). [CrossRef]  

2. M. Jackson, M. G. Sowa, and H. H. Mantsch, “Infrared spectroscopy: a new frontier in medicine,” Biophys. Chem. 68(1-3), 109–125 ( 1997). [CrossRef]   [PubMed]  

3. P. G. Datskos, P. I. Oden, T. Thundat, E. A. Wachter, R. J. Warmack, and S. R. Hunter, “Remote infrared radiation detection using piezoresistive microcantilevers,” Appl. Phys. Lett. 69(20), 2986–2988 ( 1996). [CrossRef]  

4. P. Werle, F. Slemr, K. Maurer, R. Kormann, R. Mucke, and B. Janker, “Near- and mid-infrared laser-optical sensors for gas analysis,” Opt. Lasers Eng. 37(2-3), 101–114 ( 2002). [CrossRef]  

5. K. Karstad, A. Stefanov, M. Wegmuller, H. Zbinden, N. Gisin, T. Aellen, M. Beck, and J. Faist, “Detection of mid-IR radiation by sum frequency generation for free space optical communication,” Opt. Lasers Eng. 43(3-5), 537–544 ( 2005). [CrossRef]  

6. B. Cabrera, R. M. Clarke, P. Colling, A. J. Miller, S. Nam, and R. W. Romani, “Detection of single infrared, optical, and ultraviolet photons using superconducting transition edge sensors,” Appl. Phys. Lett. 73(6), 735–737 ( 1998). [CrossRef]  

7. D. P. Neikirk, W. W. Lam, and D. B. Rutledge, “Far-infrared microbolometer detectors,” Int. J. Infrared Millim. Waves 5(3), 245–278 ( 1984). [CrossRef]  

8. Y. W. Heo, D. P. Norton, L. C. Tien, Y. Kwon, B. S. Kang, F. Ren, S. J. Pearton, and J. R. LaRoche, “ZnO nanowire growth and devices,” Mater. Sci. Eng. Rep. 47(1-2), 1–47 ( 2004). [CrossRef]  

9. V. R. Mehta, S. Shet, N. M. Ravindra, A. T. Fiory, and M. P. Lepselter, “Silicon-integrated uncooled infrared detectors: perspectives on thin films and microstructures,” J. Electron. Mater. 34(5), 484–490 ( 2005). [CrossRef]  

10. S. Herminjard, L. Sirigu, H. P. Herzig, E. Studemann, A. Crottini, J. P. Pellaux, T. Gresch, M. Fischer, and J. Faist, “Surface Plasmon Resonance sensor showing enhanced sensitivity for CO2 detection in the mid-infrared range,” Opt. Express 17(1), 293–303 ( 2009), http://www.opticsinfobase.org/abstract.cfm?URI=oe-17-1-293. [CrossRef]   [PubMed]  

11. H. Kind, H. Yan, B. Messer, M. Law, and P. Yang, “Nanowire ultraviolet photodetectors and optical switches,” Adv. Mater. 14(2), 158–160 ( 2002). [CrossRef]  

12. J. Suehiro, N. Nakagawa, S. Hidaka, M. Ueda, K. Imasaka, M. Higashihata, T. Okada, and M. Hara, “Dielectrophoretic fabrication and characterization of a ZnO anowire-based UV photosensor,” Nanotechnology 17(10), 2567–2573 ( 2006). [CrossRef]   [PubMed]  

13. S. Kumar, V. Gupta, and K. Sreenivas, “Synthesis of photoconducting ZnO nano-needles using an unbalanced magnetron sputtered ZnO/Zn/ZnO multilayer structure,” Nanotechnology 16(8), 1167–1171 ( 2005). [CrossRef]  

14. Q. H. Li, T. Gao, Y. G. Wang, and T. H. Wang, “Adsorption and desorption of oxygen probed from ZnO nanowire films by photocurrent measurements,” Appl. Phys. Lett. 86(12), 123117 ( 2005). [CrossRef]  

15. Y. W. Heo, L. C. Tien, D. P. Norton, B. S. Kang, F. Ren, B. P. Gila, and S. J. Pearton, “Electrical transport properties of single ZnO nanorods,” Appl. Phys. Lett. 85(11), 2002–2004 ( 2004). [CrossRef]  

16. C. Soci, A. Zhang, B. Xiang, S. A. Dayeh, D. P. R. Aplin, J. Park, X. Y. Bao, Y. H. Lo, and D. Wang, “ZnO nanowire UV photodetectors with high internal gain,” Nano Lett. 7(4), 1003–1009 ( 2007). [CrossRef]   [PubMed]  

17. E. Schlenker, A. Bakin, T. Weimann, P. Hinze, D. H. Weber, A. Gölzhäuser, H.-H. Wehmann, and A. Waag, “On the difficulties in characterizing ZnO nanowires,” Nanotechnology 19(36), 365707 ( 2008). [CrossRef]   [PubMed]  

18. J. Goldberger, D. J. Sirbuly, M. Law, and P. D. Yang, “ZnO nanowire transistors,” J. Phys. Chem. B 109(1), 9–14 ( 2005). [CrossRef]   [PubMed]  

19. P. J. Li, Z. M. Liao, X. Z. Zhang, X. J. Zhang, H. C. Zhu, J. Y. Gao, K. Laurent, Y. Leprince-Wang, N. Wang, and D. P. Yu, “Electrical and photoresponse properties of an intramolecular p-n homojunction in single phosphorus-doped ZnO nanowires,” Nano Lett. 9(7), 2513–2518 ( 2009). [CrossRef]   [PubMed]  

20. J. B. K. Law and J. T. L. Thong, “Simple fabrication of a ZnO nanowire photodetector with a fast photoresponse time,” Appl. Phys. Lett. 88(13), 133114 ( 2006). [CrossRef]  

21. S. E. Ahn, J. S. Lee, H. Kim, S. Kim, B. H. Kang, K. H. Kim, and G. T. Kim, “Photoresponse of sol-gel-synthesized ZnO nanorods,” Appl. Phys. Lett. 84(24), 5022–5024 ( 2004). [CrossRef]  

22. V. Srikant and D. R. Clarke, “On the optical band gap of zinc oxide,” J. Appl. Phys. 83(10), 5447–5451 ( 1998). [CrossRef]  

23. J. Wang, J. Sha, Q. Yang, X. Y. Ma, H. Zhang, J. Yu, and D. R. Yang, “Carbon-assisted synthesis of aligned ZnO nanowires,” Mater. Lett. 59(21), 2710–2714 ( 2005). [CrossRef]  

24. S. K. Mitra, Digital Signal Processing: A Computer Based Approach (McGraw-Hill, New York, 2001).

25. H. Wang, X. Yi, G. Huang, J. Xiao, X. Li, and S. Chen, “IR microbolometer with self-supporting structure operating at room temperature,” Infrared Phys. Technol. 45(1), 53–57 ( 2004). [CrossRef]  

26. M. Garcia, R. Ambrosio, A. Torres, and A. Kosarev, “IR bolometers based on amorphous silicon germanium alloys,” J. Non-Cryst. Solids 338-340, 744–748 ( 2004). [CrossRef]  

27. E. Iborra, M. Clement, L. V. Herrero, and J. Sangrador, “IR uncooled bolometers based on amorphous GexSi1-xOy on silicon micromachined structures,” J. Microelectromech. Syst. 11(4), 322–329 ( 2002). [CrossRef]  

28. P. G. Datskos, N. V. Lavrik, and S. Rajic, “Performance of uncooled microcantilever thermal detectors,” Rev. Sci. Instrum. 75(4), 1134–1148 ( 2004). [CrossRef]  

29. K. Kim, J. Y. Park, Y. H. Han, H. K. Kang, H. J. Shin, S. Moon, and J. H. Park, “3D-feed horn antenna-coupled microbolometer,” Sens. Actuator. A110,196–205 ( 2004). [CrossRef]  

30. J. P. Ploteau, P. Glouannec, and H. Noel, “Conception of thermoelectric flux meters for infrared radiation measurements in industrial furnaces,” Appl. Therm. Eng. 27(2-3), 674–681 ( 2007). [CrossRef]  

31. J. Fonollosaa, M Carmona, J Santander, L Fonseca, M Moreno, and S. Marco, “Limits to the integration of filters and lenses on thermoelectric IR detectors by flip-chip techniques,” Sens. Actuator A149,65–73 ( 2009). [CrossRef]  

32. R. A. Robie, H. T. Haselton Jr, and B. S. Hemingway, “Heat capacities and energies at 298.15 K of MgTiO3 (geikielite), ZnO (zincite), and ZnCO3 (smithsonite),” J. Chem. Thermodyn. 21(7), 743–749 ( 1989). [CrossRef]  

33. T. Olorunyulemi, A. Birnboim, Y. Carmel, O. C. Wilson, and I. K. Lloyd, “Thermal conductivity of zinc oxide: from green to sintered state,” J. Am. Ceram. Soc. 85, 1249–1253 ( 2002). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 The SEM image of a 1.0-μm-diameter ZnO wire.
Fig. 2
Fig. 2 The schematic structure (a) and SEM image of a ZnO wire on a grooved glass plate (b).
Fig. 3
Fig. 3 I-V characteristics of a 2.9-μm-diameter, 520-μm-length ZnO wire as a function of 10.6-μm-wavelength light intensity.
Fig. 4
Fig. 4 Resistance of a single 2.0-μm-diameter, 760-μm-length ZnO wire device measured as a function of irradiation intensity. Inset is the frequency-dependent resistance changes of the ZnO wire.
Fig. 5
Fig. 5 Response of a 2.0-μm-diameter, 760-μm-length ZnO wire to 10.6-μm-wavelength light irradiation measured in different atmospheres.

Equations (1)

Equations on this page are rendered with MathJax. Learn more.

τ = H / G ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.