Photoluminescence and High-Temperature Persistent

The persistent photoconductivity (PPC) effect was studied in individual tin oxide (SnO2) nanobelts as a function of ... Applied Physics A 2018 124 (6)...
0 downloads 0 Views 4MB Size
Article pubs.acs.org/JPCC

Photoluminescence and High-Temperature Persistent Photoconductivity Experiments in SnO2 Nanobelts E. R. Viana,* J. C. González, G. M. Ribeiro, and A. G. de Oliveira Universidade Federal de Minas Gerais, Instituto de Ciências Exatas, Departamento de Física, Belo Horizonte/MG, 31720-901, Brazil ABSTRACT: The persistent photoconductivity (PPC) effect was studied in individual tin oxide (SnO2) nanobelts as a function of temperature, in air, helium, and vacuum atmospheres, and low temperature. Photoluminescence measurements were carried out to study the optical transitions and to determine the acceptor/donors levels and their best representation inside the band gap. Under ultraviolet (UV) illumination and at temperatures in the range of 200−400 K, we observed a fast and strong enhancement of the photoconductivity, and the maximum value of the photocurrent induced increases as the temperature or the oxygen concentration decreases. By turning off the UV illumination, the induced photocurrent decays with lifetimes up to several hours. The photoconductivity and the PPC results were explained by adsorption and desorption of molecular oxygen at the surface of the SnO2 nanobelts. On the basis of the temperature dependence of the PPC decay, an activation energy of 230 meV was found, which corresponds to the energy necessary for thermal ionization of free holes from acceptor levels to the valence band, in agreement with the photoluminescence results presented. The molecular-oxygen recombination with holes is the origin of the PPC effect in metal oxide semiconductors, so that the PPC effect is not related to the oxygen vacancies, as commonly presented in the literature.

1. INTRODUCTION One-dimensional nanostructures have stimulated significant contributions in scientific research in the recent years. Because of its small diameter, d ≤ 100 nm, the surface effects and the quantum confinement can influence their optical1−5 and electrical6,7 properties which are important for the functionality and performance of nanodevices.8 In particular, tin oxide (SnO2) nanowires and nanobelts9 have large potential for applications as gas sensors10,11 and as ultraviolet light detectors12 because of their inherent high surface-to-volume ratio properties. However, a deeper understanding of the optoelectronic properties of the SnO2 nanostructures is still needed. In this work, the photoconductivity (PC) and the photoluminescence (PL) of SnO2 nanobelts was studied. PC measurements were carried out as a function of temperature and at different atmospheres, namely, in air, helium, and vacuum. Low-temperature PL experiments were also carried out to study the optical transitions that will be used for the determination of the acceptor/donor energy levels involved and to support the photoconductivity studies. When exposed to UV illumination, the SnO2 devices show a fast and strong enhancement of the photoinduced current, and the maximum value obtained increases with decreasing temperature. On the other hand, when the UV illumination is turned off, the current decays for hours, even at temperatures as high as 400 K which characterizes the phenomenon named persistent photoconductivity (PPC). © 2013 American Chemical Society

The photoconductivity and the PPC results were explained in terms of oxygen adsorption and desorption at the surface of the SnO2 nanobelts that promote the creation of a band bending, a spatial charge separation, and a decrease in the free carriers’ recombination time. The temperature dependence of PPC was found to exhibit an activation energy of 230 meV, which corresponds to the energy necessary for thermal ionization of free holes from acceptor levels to the valence band, determined from the PL experiment. So, holes from acceptor levels are responsible for the PPC effect in SnO2, when a light source with energy below the band gap is used. Some authors have reported and proposed models to explain the PPC effect in metal oxide semiconductor nanowires, such as in ZnO3,4,13,14 and in WO3.15 However, there is still no conclusive proposal about the origin of PPC on oxide semiconductors,3,4,13 and that is the subject of the present work.

2. SNO2 NANOBELTS: SYNTHESIS AND CHARACTERIZATION The SnO2 nanobelts used in this work were synthesized by a gold-catalyst-assisted vapor−liquid−solid (VLS) method.16 A suitable amount of pure Sn powder (1 g, 99.99% purity) was placed on top of a Si:SiO2 substrate, previously coated with a Received: December 11, 2012 Revised: February 14, 2013 Published: March 6, 2013 7844

dx.doi.org/10.1021/jp312191c | J. Phys. Chem. C 2013, 117, 7844−7849

The Journal of Physical Chemistry C

Article

thin Au film. The substrate was placed in a horizontal quartz tube furnace that was then heated to 800 °C for 2 h and after that let to cool naturally. SnO2 nanobelts were found on the substrate surface in a cotton-wool-like form. Transmission electron microscopy (TEM) examinations (see Figure 1) show that nanobelts present a prismatic cross-section

Figure 1. (a) TEM image of a SnO2 nanobelt covered by a thick amorphous layer. A SEM image of several SnO2 nanobelts is shown in the inset.

and are covered by a 10−20 nm thick amorphous tin oxide layer. Scanning electron microscopy (SEM) examinations (see inset Figure 1) reveal a large quantity of nanobelts with typical widths in the range of 50−500 nm and lengths from 5 to 50 μm. X-ray diffraction (XRD) measurements, not shown, confirm the tetragonal rutile crystal structure of the nanobelts with lattice constants of a = b = 0.473 nm and c = 0.318 nm.17

Figure 2. (a) Photoluminescence (PL) spectrum (IR-VI and UV region) of the investigated SnO2 nanobelt sample at 4 K, showing a strong and broad orange and green luminescence, at 2.07 and 2.57 eV, respectively. (b) UV PL spectra of SnO2 nanobelts taken at 4 K, showing the eA°, D0x, and DAP transitions, and LO replicas of the DAP transition.

3. PHOTOLUMINESCENCE EXPERIMENTS Low-temperature photoluminescence experiments were carried out to study the optical transitions in the SnO2 nanobelts and to support the photoconductivity studies that will be presented. A 325 nm line of a HeCd laser (Kimmon instruments) was used to perform the photoluminescence (PL) experiments in a bundle of SnO2 nanobelts. The PL emission was analyzed with a PI-Acton SpectraPro-2558i spectrometer equipped with a UV-enhanced liquid nitrogen cooled Si charge couple device (CCD) detector. PL was measured at 4 K in a coldfinger Janis ST-100 cryostat, with standard UV grade optical view ports. Figure 2 shows the PL spectrum of the SnO2 nanobelts at 4 K. In the sub-UV region (infrared (IR), visible (VI)) of the spectrum (Figure 2(a)), six peaks were observed at: 2.57, 2.07, 1.47, 1.40, 1.32, and 0.90 eV. The strong and broad orange (2.07 eV) and green (2.57 eV) gap luminescence is assigned to defect states due to oxygen vacancies in the tin oxide, as those recently published by Kar et al.18,19 Oxygen vacancies in three charge states have been noted in SnO2, neutral Vφ, single Vo+, and double ionized Vo+2. The peak at 2.07 eV is believed to be the Vo+ state, and the peak at 2.57 eV is believed to be assigned to isolated Vo+ centers, which lies at higher energy than the complex Vo+ center band at 2.07 eV.18,19 The IR peaks, 0.90, 1.32, 1.40, and 1.47 eV, are assumed to be transitions from trapped electrons, forming donor states (ED1, ED2, and ED3), to

the midgap oxygen vacancies states, (Vo+) and (Vo+)iso, that were defined by the broad and strong orange/green peaks. In the UV region of the spectrum (Figure 2(b)), six peaks were observed at: 3.16, 3.20, 3.24, 3.28, 3.33, and 3.36 eV. The peaks at 3.36 and 3.33 eV are assigned to the recombination of electrons from the conduction band to acceptor levels (eA°) and to excitons bound to neutral (D0x),20−22 respectively. The peaks at 3.28, 3.24, 3.20, and 3.16 eV are associated with donor−acceptor pair (DAP) recombination, and the DAP-1 to DAP-3 are the longitudinal phonon (LO) replicas. All PL peaks of the SnO2 nanobelts observed in Figure 2 are in agreement with previous reports.18−20,22 Considering the accepted value of 3.60 eV for the energy gap of the SnO2 nanobelts,23,24 the ionization energy of the acceptor and the donor level involved in the optical transitions eA° and DAP could also be determined to be EA = 240 meV and ED1 = 80 meV. The donor level ED1 could also be defined by the IR-UV peaks, 2.07 and 1.47 eV, since both are broad peaks and the value of ED1 obtained was 60 meV; so to corroborate with the other transition the donor states were defined to be in the range of ED1 = 60−80 meV. By transitions between midgap Vo+ states, another two donor levels could also be determined as ED2 = 130 meV and ED3 = 210 meV. Finally the schematic energy diagram of these transitions is shown in Figure 3, such that down arrows are the 7845

dx.doi.org/10.1021/jp312191c | J. Phys. Chem. C 2013, 117, 7844−7849

The Journal of Physical Chemistry C

Article

optical transitions; Vo+ and (Vo+)iso are the oxygen-vacancy states, and the double arrows are the donor/acceptor levels.

Figure 5 shows the PC measurement results at temperatures of 300, 350, and 400 K, acquired with applied voltage Vds = 100

Figure 5. Time response of the induced photocurrent In for the SnO2 nanobelts, as a function of temperature.

Figure 3. Schematic energy band diagram and the optical transitions in the SnO2 nanobelts.

mV and a UV-light intensity of 20.56 mW/cm2. Upon exposure to UV illumination, a fast increase in the photocurrent In was observed. We can also observe that the generation rate of In does not change significantly with temperature, but the maximum induced photocurrent (In)max increases as temperature decreases. When the UV light is turned off, the current slowly decays, as shown in Figure 5, and continues decaying for hours and even at temperatures as high as 400 K. This is characteristic of the phenomenon named persistent photoconductivity (PPC).

4. DEVICE FABRICATION To measure the photoconductivity (PC) of a single SnO2 nanobelt, devices with two contacts were fabricated by photolithography using a LaserWriter (UV-Laser, λ = 405 nm) and standard lift-off process. Electrodes, 2−3 μm apart, consisting of 10 nm of Cr followed by 150 nm of Au were defined on top of a 300 nm thick SiO2 deposited on a highly doped p-type Si substrate. The SnO2 nanobelts are surrounded by an amorphous high resistivity oxide layer, so a H2 plasma treatment was carried out to improve the quality of the contacts. A schematic of the device is shown in Figure 4.

Figure 6. Normalized decay photocurrent IPPC(t)/IPPC(0) curves at temperatures 250, 300, and 350 K. The solid lines are the respective fittings using eq 1.

The decay of In(t), photocurrent as a function of time, shown in Figure 6 can be well described by the stretched-exponential function25

Figure 4. Schematic two-contact device used in this work to measure the photoconductivity of individual nanowires.

IPPC(t ) = IPPC(0) exp( −(t /τ )β )

(1)

where τ and β represent the lifetime of the induced photocarriers and the decay exponent, respectively. Both parameters, τ and β, were determined by fitting the IPPC curves with eq 1, for each temperature measured, as shown in Figure 6. This stretched-exponential function is usually used in the barrier model25−27 to explain the PPC effect caused by traps such as AX centers in GaN26 and DX27 and EL228 centers in AlGaAs. In processes in which trapped electrons or holes are being thermally freed from traps to the conduction or to the valence

5. PHOTOCONDUCTIVITY MEASUREMENTS For the PC measurements the SnO2 device was placed in a coldfinger Oxford cryostat in He-7 atm, with a precise temperature controller, and a Keithley 237 source meter was used to measure the current Ids passing through the nanobelts as a function of time t, with a constant applied voltage Vds. An UV light-emitting diode (LED) emitting at 403 nm (3.08 eV) was placed near the sample and used as an excitation source in the PC experiments. 7846

dx.doi.org/10.1021/jp312191c | J. Phys. Chem. C 2013, 117, 7844−7849

The Journal of Physical Chemistry C

Article

indicate that the presence of oxygen in the atmosphere determines the change of the photoconductivity of the SnO2 nanobelt.

band, the temperature dependence of the photocarrier lifetime can be expressed as29 ⎛ −ΔEtrap ⎞ ⎛1⎞ ⎜ ⎟ = N δ exp⎜ ⎟ c n ⎝τ⎠ ⎝ kBT ⎠

6. DISCUSSION It is believed that the nature of the oxygen surface species depends on the temperature. Oxygen adsorbs on the SnO2 surface in a nondissociatively, i.e., in molecular, form, either as neutral O2(ads) or charged O2−(ads) at low temperatures (T < 500 K). At higher temperatures (T > 500 K), O2(ads) dissociates into atomic form, as neutral O(ads) or in charged O−(ads).18 The molecular oxygen O2(g) traps electrons according to Adsorption:

(2)

where Nc is the effective density of states in the conduction band; δn is the capture coefficient of the traps for electrons; ΔEtrap is the estimated trap depth or ionization energy; and kB is the Boltzmann constant. Figure 7 shows the temperature dependence of the parameters τ and β in the range of 200−400 K. A strong

O2 (g) + e− → O−2 (ad)

(3)

O−2 (ad)

where O2(g) and indicate oxygen in its free and adsorbed states, respectively. The reverse process, namely, the desorption of oxygen from the surface, requires a photogenerated hole h+ Desorption: h+ + O−2 (ad) → O2 (g)

(4)

Considering these two processes and the schematic energy band obtained from the PL measurements, we want to explain the PPC phenomena in the SnO2 nanobelts. Once the UV light is turned on, oxygen molecules (that are supposed to be naturally adsorbed at the range of temperatures measured) are desorbed from the surface, and photoexcited holes become available, thereby reducing the surface potential and the depletion width, until a steady state with photocurrent (In)max is reached, as presented in Figure 9.

Figure 7. Temperature dependence of the PPC lifetime τ (a) and of the decay exponent β (b). The solid line corresponds to the fitting of the experimental data with eq 2.

decrease of τ and a slow increase of β are observed as the temperature increases. The activation energy of τ, ΔEtrap = 230 meV, was determined by fitting the temperature dependence of τ with eq 2. To explain the PPC behavior in the SnO2 nanobelt and the high value achieved for ΔEtrap, photocurrent experiments were carried out in different atmospheres: air, helium, and vacuum. The photoresponses were measured at fixed temperature of 300 K and with the same UV light intensity defined previously. Figure 8 shows that the maximum induced photocurrent (In)max is already two times higher in vacuum than in helium and at least four times higher than in air. Besides, the decay time in vacuum (τvacuum = 1.20 × 105 s) is almost 20 times longer than in helium (τhelium = 7.35 × 103 s) and one thousand times longer than in air (τair = 1.20 × 102 s). These results

Figure 9. UV photogenerated hole promotes desorption of the negatively charged adsorbed oxygen at the surface, reducing the surface potential Φ and the depletion width, increasing the photocurrent.

As presented in Figure 5 and Figure 8, oxygen desorbs in vacuum, air, or helium (at temperatures between 200 and 400 K) at approximately the same rate. However, the steady-state value, (In)max, is different due to the available concentration of molecular oxygen in each case. As the illumination is turned off and at temperatures near the room temperature, the adsorbed oxygen molecules trap the electrons thermally activated, gradually bend the bands, and

Figure 8. Time response of the induced photocurrent In for the SnO2 nanobelts for different atmospheres, air, helium, and vacuum, taken at 300 K. 7847

dx.doi.org/10.1021/jp312191c | J. Phys. Chem. C 2013, 117, 7844−7849

The Journal of Physical Chemistry C

Article

raise the surface potential barrier and the internal field, as presented in Figure 10. This reduces the chance of new

So the source of the electrons in the process of oxygen O2(g) adsorption (eq 2) is not from the oxygen vacancies. Since a LED emitting at 403 nm (energy of 3.08 eV) is used as the source of the UV light, we are pumping the nanobelt with light that has energy below the band gap (3.60 eV). This way, we are excluding the formation of electron−hole pairs in the conduction/valence bands, which is why holes in the process of oxygen O2−(ads) desorption (eq 3) have to be thermally promoted from the acceptor energy level EA, in the absence of UV light, giving rise to PPC. Besides the effects of temperature in the photoconductivity of SnO2 nanobelts here presented, surface passivation can be also used to tune the photoconductivity of these nanostructures. This approach has already been carried out by covering nanostructures with polymers14,30 or even with thin metal nanolayers,31 which strongly modifies the band bending at the surface14,30 and consequentially the PPC effect.30 However, further studies are necessary to determine how the PCC effect is influenced by the surface passivation of SnO2 nanobelts.

Figure 10. Schematic oxygen absorption at the surface process by an electron trapping, increasing the surface potential Φ and the depletion width.

7. CONCLUSIONS In this work we have studied the photoconductivity (PC) of SnO2 nanobelts as a function of temperature and atmosphere (air, helium, or vacuum), as well as low-temperature photoluminescence (PL) experiments. In the PL experiment the optical transitions were studied in the visible-infrared region (VI-IR) and in the ultraviolet (UV) region. Peaks at the VI-IR region are associated to transitions from the oxygen-vacancy states to three donor levels, ED1 = (60−80) meV, ED2 = 130 meV, and ED3 = 210 meV, and peaks in the UV region to transitions from a donor level or from the conduction band to an acceptor level, EA = 230 meV. Not surprisingly the donor level ED1 was determined simultaneously from VI-IR and UV transitions, reinforcing the PL results presented. In the PC measurements we observed that the maximum UV-induced photocurrent (In)max increases with a decrease in temperature or in oxygen concentration, in the atmosphere around the sample. A persistent photoconductivity (PPC) effect was also observed, lasting for hours, even at temperatures as high as 400 K. The kinetics of the PPC decay curves was found to be well explained by the stretched-exponential function, with an increase in the decay times as temperature decreases. Besides that a reduction of the available oxygen in the atmosphere around the sample (vacuum or helium) also modifies drastically the PPC decay time. The photoconductivity and the PPC results were explained in terms of adsorption and desorption of molecular oxygen at the surface of the SnO2 nanobelts, which promotes the creation of a band bending at the nanobelt surface, consequentially spatial charge separation and a decrease in the free carriers’ recombination time. The temperature dependence of the PPC decay time was found to exhibit activation energy around 230 meV that corresponds to the energy necessary for thermal ionization of free holes from acceptor levels to the valence band, in agreement with the photoluminescence results presented. So, holes ionized from acceptor levels are responsible for the PPC effect in SnO2, when we are using a light source with energy below the band gap, that prohibits band-to-band transitions. We claim that the experiments carried out in SnO2 can be extrapolated for other metal oxide semiconductors, especially

electrons being trapped at the surface, thereby reducing the rate of oxygen adsorption, which promotes a partial separation between the electrons and the photogenerated holes, resulting in an enhancement in their recombination time.3 This explains the PPC effect and the observed difference between the PPC in air, helium, and vacuum. So, the changes in the surface potential and in the depletion width are determined by the interplay between oxygen adsorption (eq 3) and the net desorption rate (eq 4) and consequently determine the photoconductivity behavior. But how is the temperature dependence explained? The activation energy ΔEtrap associated with the decay time was found to be about 230 meV. This value is very close to the ionization energy of the acceptor level EA = 240 meV found in the PL experiments, as can be seen in Figure 3. The increase of free hole density due to ionization thermally activated will further decrease the band bending at the surface (Figure 10), which in turn reduces the spatial separation and recombination time of free electrons and holes when the illumination is turned off, in agreement with our observations. This way, we understand that the temperature dependence of the PPC decay time is due to the thermal ionization of holes from the acceptor level. The high value of the activation energy obtained explains the observation of PPC even at temperatures as high as 400 K. This is the main contribution of the present paper since many authors have proposed that the mechanisms behind PPC on metal oxide semiconductors are related to oxygen adsorption and desorption but give no support for the temperature dependence.3,12,14 Undoped SnO2 typically shows n-type conductivity. To prove that, I−V curves (not shown) as a function of gate bias, and carried out before illumination and also after the illumination is turned off, confirm that the conductivity in our SnO2 nanobelts is n-type. By the results presented in Figure 3, the n-type behavior of the SnO2 nanobelts at near room temperature can be explained by the shallow donor state ED1 = 60−80 meV since part of the carriers can be thermally activated (ED1 ≤ 3kBT) from ED1 to the conduction band. The n-type behavior may not be explained from the donor state ED2, ED3 that is too deep to be thermally activated, or from the oxygen vacancy surface levels, (Vo+) and (Vo+)iso, that are even more. 7848

dx.doi.org/10.1021/jp312191c | J. Phys. Chem. C 2013, 117, 7844−7849

The Journal of Physical Chemistry C

Article

(22) Liu, B.; Cheng, C. W.; Chen, R.; Shen, Z. X.; Fan, H. J.; Sun., H. D. J. Phys. Chem. C 2010, 114, 3407−3410. (23) Zhou, X. T.; Heigl, F.; Murphy, M. W.; Sham, T. K.; Regier, T.; Coulthard, I.; Blyth, R. I. R. Appl. Phys. Lett. 2006, 89, 213109. (24) Mondal, S. P.; Ray, S. K.; Ravvichandran, J.; Manna, I. Bull. Mater. Sci. 2010, 33, 357−364. (25) Chang, D. C.; Chang, C. Y. J. Phys.: Condens. Matter 1995, 7, 4525−4532. (26) Li, J. Z.; Lin, J. Y.; Jiang, H. X.; Salvador, A.; Botchkarev, A.; Morkoç, H. Appl. Phys. Lett. 1996, 69, 1474. (27) Lin, J. Y.; Dissanayake, A.; Brown, G.; Jiang, H. X. Phys. Rev. B 1990, 42, 5855−5858. (28) Rubinger, R. M.; Bezerra, J. C.; Chagas, E. F.; Gonzalez, J. C.; Rodrigues, W. N.; Ribeiro, G. M.; Moreira, M. V. B.; de Oliveira, A. G. J. Appl. Phys. 1999, 84, 3764. (29) Bube, R. H. Photoelectronic Properties of Semiconductors; Cambridge University Press: Cambridge, U.K., 1995. (30) Huh, J.; Joo, M.-K.; Jang, D.; Lee, J.-H.; Kim, G. T.. J. Mater. Chem. 2012, 22, 24012. (31) Liu, K. W.; Liu, B.; Wang, S. J.; Wei, Z. P.; Wu, T.; Cong, C. X.; Shen, Z. X.; Sun, X. W.; Sun., H. D. J. Appl. Phys. 2009, 106, 083110.

for ZnO nanowires/nanobelts since both materials are naturally n-type semiconductors, having similar crystal structure and band gaps. Within the results presented the authors want to exclude the common sense of the literature, that the oxygen vacancies are the striking forces on metal oxide nanostructures in all aspects; in this case, it is not!



AUTHOR INFORMATION

Corresponding Author

*Tel.: +55(31)3409.5625. Fax: +55(31)3409.5637. E-mail: emilsonfi[email protected] (E. R. Viana). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge the financial support of CNPq, CAPES, and FAPEMIG, Brazilian funding agencies, and the technical support of the Microscopy Center of the Universidade Federal de Minas Gerais.



REFERENCES

(1) Lopez-Richard, V.; González, J. C.; Matinaga, F. M.; TralleroGiner, C.; Ribeiro, E.; Rebello Sousa Dias, M.; Villegas-Lelovsky, L.; Marques., G. E. Nano Lett. 2009, 9, 3129−3136. (2) González, J. C.; da Silva, M. I. N.; Lozano, X. S.; Zanchet, D.; Ugarte, D.; Ribeiro, E.; Gutiérrez, H. R.; Cotta, M. A. J. of Nanosci. Nanotechnol. 2006, 6, 2182−2186. (3) Jiming, B.; Ilan, S.; Zhihua, S.; Ron, G.; Federico, C.; Xiaowei, W.; Zhifeng, R.. Nanoscale Res. Lett. 2011, 6, 404. (4) Yin, Z. G.; Zhang, X. W.; Fu, Z.; Yang, X. L.; Wu, J. L.; Wu, G. S.; Gong, L.; Paul, K. Phys. Status Solid RRL 2012, 6, 117. (5) Falcão, B. P.; Leitão, J. P.; González, J. C.; Correia, M. R.; ZayasBazán, K. G.; Matinaga, F. M.; Moreira, M. B.; Leite, C. F.; de Oliveira, A. G. J. Mater. Sci. 2013, 48, 1794−1798. (6) Dattoli, E. N.; Wan, Q.; Guo, W.; Chen, Y.; Pan, X.; Wei, L. Nano Lett. 2007, 7, 2463−2469. (7) Sun, J.; Tang, Q.; Lu, A.; Jiang, X.; Wan, Q. Nanotechnology 2009, 20, 255202. (8) Hullavarad, S.; Hullavarad, N.; Look, D.; Claffin, B. Nanoscale Res. Lett. 2009, 4, 1421−1427. (9) Viana, E. R.; González, J. C.; Ribeiro, G. M.; de Oliveira., A. G. Phys. Status Solidi RRL 2012, 6, 262−264. (10) Schlenker, E.; Bakin, A.; Weimann, T.; Hinze, P.; Weber, D. H.; Golzhauser, A.; Wehmann, H-H; Wang, A. Nanotechnology 2008, 19, 365707. (11) Wan, Q.; Li, Q. H.; Chen, Y. J.; Wang, T. H.; He, X. L. Appl. Phys. Lett. 2004, 84, 3654. (12) Lin, C.-H.; Chen, R.-S.; Chen, T.-t.; Chen, H.-Y.; Chen, Y.-F.; Chen, K.-H.; Chen, L.-C. Appl. Phys. Lett. 2008, 93, 112115. (13) Prades, J. D.; Hernandez-Ramirez, F.; Jimenez-Diaz, R.; Manzanares, M.; Andreu, T.; Cicera, A.; Romano-Rodriguez, A.; Morante., J. R. Nanotechnology 2008, 19, 465501. (14) Liu, K. W.; Chen, R.; Xing, G. Z.; Wu, T.; Sun., H. D. Appl. Phys. Lett. 2010, 96, 023111. (15) Huang, K.; Qing, Z. Nanoscale Res. Lett. 2011, 6, 52. (16) Wagner, R. S.; Ellis, W. C. Appl. Phys. Lett. 1964, 4, 89. (17) Tamanaka; Kurashima, R.; Mimaki, J. Z. Kristallogr. 2000, 215, 424. (18) Kar, A.; Stroscio, M. A.; Meyyappan, M.; Gosztola, D. J.; Wiederrecht, G. P.; Dutta, M. Nanotechnology 2011, 22, 285709. (19) Kar, A.; Stroscio, M. A.; Dutta, M.; Jyoti, K.; Meyyappan, M. Semicond. Sci. Technol. 2010, 25, 024012. (20) Ayan, K.; Stroscio, M. A.; Dutta, M.; Kumari, J.; Meyyappan, M. Appl. Phys. Lett. 2009, 94, 101905. (21) Chen, R.; Xing, G. Z.; Gao, J.; Zhang, Z.; Wu, T.; Sun, H. D. Appl. Phys. Lett. 2009, 95, 061908. 7849

dx.doi.org/10.1021/jp312191c | J. Phys. Chem. C 2013, 117, 7844−7849