Sie sind auf Seite 1von 3

APPLIED PHYSICS LETTERS 86, 011923 (2005)

Adsorption effects of NO2 at ppm level on visible photoluminescence response of SnO2 nanobelts
G. Faglia,a) C. Baratto, and G. Sberveglieri
INFM and University of Brescia, Sensor Laboratory, Department of Chemistry and Physics for Engineering and for Materials, Via Valotti 9, I-25123 Brescia, Italy

M. Zha and A. Zappettini


IMEM-CNR Parco Area delle Scienze 37/A, I-43010 Parma, Italy

(Received 24 May 2004; accepted 9 November 2004; published online 28 December 2004) The visible photoluminescence (PL) of tin oxide nanobelts is quenched by nitrogen dioxide at ppm level in a fast (time scale order of seconds) and reversible way. Besides, the response seems highly selective toward humidity and other polluting species, such as CO and NH3. We believe that adsorbed gaseous species that create surface states can quench PL by creating competitive nonradiative paths. A comparison between conductometric and PL response suggests that the two responses are ascribable to different adsorption processes. 2005 American Institute of Physics. [DOI: 10.1063/1.1849832] In the past few years, progress has been achieved in the synthesis, structural characterization, and physical properties investigation of nanostructures. Due to their peculiar characteristics and size effects, these materials often show some physical properties that are different from those of the bulk, and are of great interest both for fundamental studies and for potential nanodevice applications. Among them, metal oxide nanobelts and nanorings are potential candidates for the fabrication of nanoscale devices. SnO2 is a direct wide-gap semiconductor (3.6 eV). Bulk oxygen vacancies are responsible for two donor states localized at 30 meV and 150 meV from the bottom of the conduction band.1 The transport properties of SnO2 depend strongly on surface states produced by gaseous adsorption that produces space-charge regions and band modulation. The extraordinary gas sensing properties to NO2 of tin oxide nanobeltsthe electrical conductance decreases by orders of magnitude upon NO2 adsorptionhave recently been reported.2,3 Theoretical calculations4 showed that NO2 can adsorb in many forms over nanobelts grown along the 101 direction with different binding energies and net charge transfer. Regarding metal oxide nanoparticles photoluminescence (PL), in literature much attention has been devoted to study the optical properties of ZnO nanostructuresa high gap metal oxide semiconductor in which oxygen vacancies (as in tin oxide) is deemed responsible for doping. The interest was mainly driven by ZnO PL in the UV range and in its lasing activity.5 Only recently has much attention been paid to the PL response in the visible-green range. Still, there is no consensus on the positions of the peaks in the PL spectrum of ZnO nanostructures and thin lm, and their origin. The green emission has been assigned to the transition between the photoexcited holes and a single ionized oxygen vacancy,6 attributed to an antisite oxygen7 and donor-acceptor complexes.8,9 Surface states have also been identied as a possible cause of the visible emission in ZnO nanowires.10 It was determined that the ratio of UV-to-visible emission is dependent on fabrication conditions.11 A very interesting rea)

Electronic mail: gfaglia@tab.ing.unibs.it

sult is the progressive increase of the green-light emission intensity determined as the diameter of the wires decreases.12 This suggest the possibility that reduced dimensionality of the system could play a role in the onset of visible PL spectra of high gap metal oxides, both due to quantum size effects or by the increase of the surface, rich in oxygen vacancies, to volume ratio. Comparatively few studies have been carried out on tin oxide nanostructures: The onset of visible PL in tin oxide shbonelike nanoribbons has been recently shown13 and luminescence properties of SnO2 nanoparticles studied,14 reporting broad PL optical bands from tin oxide nanobelts in the visible range from 400 to 600 nm. Semiconducting tin oxide nanobelts were prepared by thermal evaporation on alumina and silica substrates.15 Inside the tubular reactor, it was possible to control the ow of different gas sources. 99.9% pure SnO powders were used as the source material. The depositions were performed in the 8001000 C temperature range. In these conditions, SnO2 nanobelts were reproducibly obtained. The typical nanobelts were a few tens of nanometers thick, a few hundreds of nanometers wide, and several hundreds of micrometers long. Transmission electron microscopy measurement revealed that nanobelts grew in the tetragonal rutile structure parallel to the 101 direction. Devices were placed inside a test chamber kept at T = 20 C, featuring a large quartz window granting easy access for optical measurements and a microfurnace to heat up the sample in the range room temperature (RT)400 C. A constant ux of synthetic air equal to 0.3 l/min, mixed with the desired amount of gaseous species, owed through the chamber at 20 C, atmospheric pressure. The examined species were NO2 (0.310 ppm), NH3 (50 ppm), and CO (10 1000 ppm) in dry and humid synthetic air and normal ambient pressure conditions. PL measurements were performed using a HeCd laser as the light source at 326 nm. PL spectra were acquired perpendicular to the sample surface using a single spectrograph and a Peltier-cooled charge coupled device camera. Figure 1 reports the PL spectra in steady-state conditions keeping the device at 120 C as a function of the chamber

0003-6951/2005/86(1)/011923/3/$22.50 86, 011923-1 2005 American Institute of Physics Downloaded 05 Nov 2007 to 61.190.88.139. Redistribution subject to AIP license or copyright, see http://apl.aip.org/apl/copyright.jsp

011923-2

Faglia et al.

Appl. Phys. Lett. 86, 011923 (2005)

FIG. 1. PL spectra at 120 C as a function of the gas environment: PL reversible quenching after NO2 introduction (10 ppm) is observable.

environment composition. PL spectra were independent from the substrate. The broad peak was the result of the convolution of at least two emission lines: One dominating yellow emission with maximum at about 587 nm and a shoulder at about 517 nm. Data were taken in air, 1 h after introducing 10 ppm of NO2 and 1 h after restoring dry air. NO2 at the ppm level strongly quenches the spectral intensity of the PL emission band by adsorbing over the surface and creating competitive nonradiative recombination paths. Taking into account the optical path between the sample and the optical window in our experimental setup (about 4 cm), we can denitely exclude that the effect was due to the optical absorption of the NO2 gas both at the laser wavelength and at the luminescence wavelength. In fact, although it is known that NO2 at RT absorbs light in the 300600 nm spectral range,16 even at the maximum concentration of the gas employed in the presented experiments (10 ppm) the absorption coefcient due to NO2 that can be calculated from the absorption cross sections is of the order of 105 cm1. Further evidence that the quenching is produced by NO2 adsorption over the surfacea well-known activated process17is the behavior of quenching with temperature. As a matter of fact, while PL is reduced overall with temperatureas expected due to increased thermal

FIG. 3. Steady-state response (dened as relative change of peak area) toward NO2 concentration. The sensor is kept at 120 C in dry air at T = 20 C.

disorderPL quenching by NO2 is not effective at a temperature lower than 80 C. To examine the kinetics of the involved processes, PL data were taken every 5 s and the broad PL band peak area was extracted as a feature for every sampling. Figure 2 reports the dynamic response of the peak area toward 1 ppm of NO2 in dry air, relative humidity RH = 70% and RH = 30% at T = 20 C. The sensor is kept at T = 120 C. Compared to electrical results3 performed on the sample produced through the same technique, the dynamic is much faster: PL response time to NO2 needed to reach 90% of the steadystate step response is equal to about 30 s (RH= 70% at T = 20 C), while the conductivity one is of the order of minutes. The adsorption process is reversible: 90% of PL baseline is recovered after restoring air in about 600 s. Response as a function of NO2 concentration is reported in Fig. 3; it increases as a logarithm of concentration. The same relation in a conductance-type sensor follows a powertype law.18 Humidity is considered the most important interfering species in gas sensing, since water vapor concentration can

FIG. 2. Kinetic response (peak area intensity) toward 1 ppm of NO2 at FIG. 4. Kinetic response (peak area intensity) toward 1000 ppm of CO and 120 C in dry air in RH= 70% and at T = 20 C in RH= 30%. Dynamic is 50 ppm of NH3 at 120 C in dry air. The sensor is selective toward these fast, reversible, and unaffected by humidity changes. interfering species. Downloaded 05 Nov 2007 to 61.190.88.139. Redistribution subject to AIP license or copyright, see http://apl.aip.org/apl/copyright.jsp

011923-3

Faglia et al.

Appl. Phys. Lett. 86, 011923 (2005)

span in normal conditions many orders of magnitude in a short time. Therefore, a very important feature is that water vapor almost does not affect the PL response (see Fig. 2). On the contrary, it is well known that its adsorption over metal oxides produces an hydroxylated surface and modies electrical properties by acting as an electron donor.18 Besides, we investigated the effect on PL of NH3 (50 ppm) and CO (1000 ppm) without observing any appreciable quenching effect, as shown in Fig. 4. The same gaseous species are active in affecting electrical transport.3 All of the above mentioned differences among PL and conductivity responses prompt us to conclude that they are produced by heterogeneous processes: Surface states active in introducing nonradiative recombination centers are not active in modifying electrical transport. NO2 involved surface states responsible for PL quenching are probably the ones with lower binding energies and negligible net charge transfer that can easily and quickly be adsorbed and desorbed from the oxide surface, such as the species NO2 2,Sn,O identied by Maiti et al.4 We showed that PL spectra of tin oxide nanobelts are reversible quenched by NO2. The response is highly selective toward humidity and other polluting species, such as CO and NH3. Optical results are completely different from conductivity ones, in which response to all examined species is present at a much slower dynamic. The results foresee the development of a class of selective metal oxide gas sensors. This work has been partly supported by the European Commission in the project NMP4-CT-2003-001528 Nano-

structured sold-state gas sensors with superior performance (NANOS4).


S. Samson and C. Fonstad, J. Appl. Phys. 44, 4618 (1973). M. Law, H. Kind, B. Messer, F. Kim, and P. Yang, Angew. Chem., Int. Ed. 41, 2405 (2002). 3 E. Comini, G. Faglia, G. Sberveglieri, Z. Pan, and Z. Wang, Appl. Phys. Lett. 81, 1869 (2002). 4 A. Maiti, J. Rodriguez, M. Law, P. Kung, J. McKinney, and Y. Peidong, Nano Lett. 3, 1025 (2003). 5 M. H. Huang, S. Mao, H. Feick, Y. H, Y. Wu, H. Kind, E. Weber, R. Russo, and P. Yang, Science 292, 1897 (2001). 6 K. Vanhausden, W. Warren, C. H. Seager, D. R. Tallant, J. A. Voigt, and B. E. Gnade, J. Appl. Phys. 79, 7983 (1996). 7 J. W. Hu and Y. Bando, Appl. Phys. Lett. 82, 1401 (2003). 8 B. Lin, Z. Fu, and Y. Jia, Appl. Phys. Lett. 79, 943 (2001). 9 S. A. Studenikin and M. Cocivera, J. Appl. Phys. 91, 5060 (2002). 10 B. D. Yao, Y. F. Chan, and N. Wang, Appl. Phys. Lett. 81, 757 (2002). 11 V. A. L. Roy, A. B. Djurisic, W. K. Chan, J. Gao, H. F. Lui, and C. Surya, Appl. Phys. Lett. 83, 141 (2003). 12 M. H. Huang, Y. Wu, H. Feick, N. Tran, E. Weber, and P. Yang, Adv. Mater. (Weinheim, Ger.) 13, 113 (2001). 13 J. Hu, Y. Bando, and D. Golberg, Chem. Phys. Lett. 372, 758 (2003). 14 F. Gu, S F. Wang, C. Song, M. Lu, Y. Qi, G. Zhou, D. Xu, and D. Yuan, Chem. Phys. Lett. 372, 451 (2003). 15 D. Calestani, M. Zha, G. Salviati, L. Lazzarini, E. Comini, and G. Sberveglieri, J. Cryst. Growth (to be published). 16 M. H. Harwood and R. L. Jones, J. Geophys. Res., [Space Phys.] 99, 22955 (1994). 17 J. Tamaki, N. Nagaishi, Y. Teraoka, N. Miura, and Y. Yamazoe, Surf. Sci. 221, 183 (1989). 18 M. Madou and R. Morrison, Chemical Sensing with Solid State Devices (Academic, San Diego, 1989).
2 1

Downloaded 05 Nov 2007 to 61.190.88.139. Redistribution subject to AIP license or copyright, see http://apl.aip.org/apl/copyright.jsp

Das könnte Ihnen auch gefallen