Identification of oxygen vacancy types from Raman spectra of SnO 2 nanocrystals

Similar documents
Preparation and characterization of Co BaTiO 3 nano-composite films by the pulsed laser deposition

Supporting Online Material for

Supporting Information. for Efficient Low-Frequency Microwave Absorption

EFFECT OF GROWTH TEMPERATURE ON THE CATALYST-FREE GROWTH OF LONG SILICON NANOWIRES USING RADIO FREQUENCY MAGNETRON SPUTTERING

This journal is The Royal Society of Chemistry S 1

Preparation and Characterization of Copper Oxide -Water Based Nanofluids by One Step Method for Heat Transfer Applications

STRUCTURAL AND LUMINENSENCE CHARACTERISTICS OF NANOCRYSTALLINE SnO 2 DOPED WITH Co 2+

Supporting Information

X-ray Diffraction and Vibrational Spectroscopy of Catalysts for Exhaust Aftertreatment

Fabrication and characterization of photocatalyst coatings by heat treatment in carbon powder for TiC coatings

Simple Synthesis of Single-crystalline Nanoplates of Magnesium Oxide

Long-Range Lattice engineering of MoTe 2 by 2D Electride

Formation of HCP Rhodium as a Size Effect

Rapid degradation of methylene blue in a novel

Mater. Res. Soc. Symp. Proc. Vol Materials Research Society

Supporting Information. Solution-Processed 2D PbS Nanoplates with Residual Cu 2 S. Exhibiting Low Resistivity and High Infrared Responsivity

Experimental Study on Nitrogen-Doped Nano-Scale TiO 2 Prepared by Microwave-Assisted Process at Low Temperature

Defense Technical Information Center Compilation Part Notice

Urchin-like V 2 O 3 /C Hollow Nanospheres Hybrid for High-Capacity and Long-Cycle-Life Lithium Storage

Solid-Phase Synthesis of Mg2Si Thin Film on Sapphire substrate

Electronic Supplementary Information (ESI) Self-assembly of Polyoxometalate / Reduced Graphene Oxide

CHINESE JOURNAL OF CHEMICAL PHYSICS VOLUME 21, NUMBER 2 APRIL 27, 2008

Microstructure, morphology and their annealing behaviors of alumina films synthesized by ion beam assisted deposition

Visible-Light-Driven Photocatalysis and NO 2 Gas Sensing

Supplementary Information

Influence of stress on Raman spectra in Ba 1 x Sr x TiO 3 thin films

Supplementary Figure S1. High-resolution XPS spectra in the Cu 2p region and Cu LMM spectra are shown in (A) and (B) respectively for CdSe NCs

Graphene/Fe 3 O Quaternary Nanocomposites: Synthesis and Excellent Electromagnetic Absorption Properties

High pressure Raman spectrum study of Ga2O3 Abstract 1. Introduction

Supplementary Materials for

Supplementary Figure 1 Heating E-chip. (A) The preparation of heating E-chip with

The Effects of Nonstoichiometry on Optical Properties of Oxide Nanopowders

SYNTHESIS OF SnO 2 NANORODS BY HYDROTHERMAL METHOD FOR GAS SENSOR APPLICATION

Supplementary Figures

quantum dots by chemical precipitation assisted by ultrasound

Microstructure and Magnetic Properties of Iron Oxide Nanoparticles Prepared by Wet Chemical Method

Advances in Engineering Research (AER), volume 102 Second International Conference on Mechanics, Materials and Structural Engineering (ICMMSE 2017)

Supporting Information. Experimental and Theoretical Investigation of Mesoporous MnO 2

Disket-Nanorings of K 2 Ti 6 O 13 Formed by Self-Spiraling of a Nanobelt

Three-dimensional graphene-based hierarchically porous carbon. composites prepared by a dual-template strategy for capacitive

Growth of Hexagonal Phase Sodium Rare Earth Tetrafluorides Induced by Heterogeneous Cubic Phase Core

Carbon Black At-line Characterization. Using a Portable Raman Spectrometer

Deposited by Sputtering of Sn and SnO 2

Synthesis of beta gallium oxide nano-ribbons from gallium arsenide by plasma immersion ion implantation and rapid thermal annealing

Performance at Wafer-Scale

Chemically Tunable Full Spectrum Optical Properties of 2D Silicon Telluride Nanoplates

Microstructural Characterization of Materials

Nanjing University, Nanjing , P.R.China, 2. Max-Planck-Institut für Mikrostrukturphysik, Weinberg 2, D Halle, Germany

Supplementary Figures

Supporting Information

Highly thermally conductive and electrically insulating polymer nanocomposites with boron nitride nanosheet/ionic liquid complexes

6th International Conference on Advanced Design and Manufacturing Engineering (ICADME 2016)

Crystallographic Characterization of GaN Nanowires by Raman Spectral Image Mapping

Tin Oxide Nanowires, Nanoribbons, and Nanotubes

MgAl 2 O 4 nanoparticles: A new low-density additive for accelerated thermal decomposition of ammonium perchlorate

Formation of crystallized titania nanotubes and their transformation into nanowires

Enhanced Light Trapping in Periodic Aluminum Nanorod Arrays as Cavity Resonator

Synthesis and Thermoelectric Properties of Bi 2 S 3 Nanobeads

University of Groningen

CHAPTER 8 CONCLUSIONS AND SCOPE FOR FUTURE WORK

Department of Chemistry, University of California, Davis, California 95616, USA 2

Optimized by Disorder Engineering on Iron doped Ni 3 S 2. nanosheets for Oxygen Evolution Reaction

Pelotas, , Pelotas, RS, Brazil 2 Instituto de Física, Universidade Federal do Rio Grande do Sul C.P ,

Characterization of carbon nitride thin films deposited by microwave plasma chemical vapor deposition

X-ray Studies of Magnetic Nanoparticle Assemblies

Single-crystal nanorings of b-gallium oxide synthesized using a microwave plasma

Bi-functional ZnO/RGO/Au substrate : photocatalysts for degrading. pollutant and SERS substrates for real-time monitoring

De-ionized water. Nickel target. Supplementary Figure S1. A schematic illustration of the experimental setup.

Supplementary Figure 1. Energy-dispersive X-ray spectroscopy (EDS) of a wide-field of a) 2 nm, b) 4 nm and c) 6 nm Cu 2 Se nanocrystals (NCs),

Optical, microstructural and electrical studies on sol gel derived TiO 2 thin films

Formation mechanism of alumina nanotube array

Ultra-stretchable, sensitive and durable strain sensors based on. polydopamine encapsulated carbon nanotubes/elastic bands

Supplementary Figure 1 TEM of external salt byproducts. TEM image of some salt byproducts precipitated out separately from the Si network, with

Reviewers' Comments: Reviewer #1 (Remarks to the Author)

Supplimentary Information. Large-Scale Synthesis and Functionalization of Hexagonal Boron Nitride. Nanosheets

Why does polycrystalline natural diamond turn black after annealing?

Doping and Temperature Dependence of Raman Scattering of

Bimetallic Nanoparticle Oxidation in Three Dimensions by Chemically Sensitive Electron Tomography and In-Situ Transmission Electron Microscopy

STRUCTURAL INFORMATION FROM THE RAMAN SPECTRA OF ACTIVATED CARBON MATERIALS

CHAPTER 5. HIGH PRESSURE MICRO RAMAN STUDIES ON SINGLE CRYSTAL Yb 2 O 3

Variation of fundamental and higher-order Raman spectra of ZrO 2 nanograins with annealing temperature

Synthesis and Characterization of nanocrystalline Ni-Co-Zn ferrite by Sol-gel Auto-Combustion method.

Supporting information

Heteroepitaxy of Monolayer MoS 2 and WS 2

Structure and optical properties of M/ZnO (M=Au, Cu, Pt) nanocomposites

Mechanism of the Oxidation of Iron. Yiqian Wang 1,e

Supporting Information. on Degradation of Dye. Chengsi Pan and Yongfa Zhu* Department of Chemistry, Tsinghua University, Beijing, , China

An Analysis of Structural and Optical Properties Undoped ZnS and Doped (with Mn, Ni) ZnS Nano Particles

Electronic Supplementary Information

Electronic Supplementary Material

Fast Growth of Strain-Free AlN on Graphene-Buffered Sapphire

Pressure effects on the absorption of α-phase Nb-H system

INVESTIGATION OF NANOCRYSTALS USING TEM MICROGRAPHS AND ELECTRON DIFFRACTION TECHNIQUE

Microstructural Evolution of Ti-Mo-Ni-C Powder by Mechanical Alloying

Supporting Information. Selective Metallization Induced by Laser Activation: Fabricating

Synthesis and characterization of strontium chloroborate whiskers

Supporting Information for

Evolution of tetragonal phase in the FeSe wire fabricated by a novel

PREPARATION OF NEODYMIUM HYDROXIDE NANORODS AND NEODYMIUM OXIDE NANORODS BY A HYDROTHERMAL METHOD

respectively. A plot of the Raman shift position as a function of layer number in Figure S1(c)

Transcription:

Research article Received: 21 December 2011 Revised: 27 February 2012 Accepted: 9 March 2012 Published online in Wiley Online Library: 20 July 2012 (wileyonlinelibrary.com) DOI 10.1002/jrs.4078 Identification of oxygen vacancy types from Raman spectra of SnO 2 nanocrystals L. Z. Liu, a T. H. Li, a,b X. L. Wu, a *J.C.Shen a and P. K. Chu c Raman spectra acquired from spherical SnO 2 nanocrystals prepared by pulsed laser ablation and hydrothermal synthesis exhibit three oxygen-vacancy-related Raman modes at 234, 573, and 618 cm 1. The peak location and intensity vary with annealing temperature under O 2 finally approaching those of bulk materials. Density functional calculation discloses that the three Raman modes stem from subbridging, in-plane, and bridging oxygen vacancies, respectively. Raman spectra can thus be used to discern different types of oxygen vacancies in SnO 2 nanocrystals. Copyright 2012 John Wiley & Sons, Ltd. Supporting Information may be found in the online version of this article. Keywords: SnO 2 nanocrystals; Raman spectroscopy; oxygen vacancies; annealing processing; density functional calculation Introduction Tin oxide (SnO 2 ) is one of the attractive functional materials because of potential applications to biophysics, gas sensing, catalysis, and batteries. [1 6] SnO 2 nanostructures with different morphologies such as nanowires (nanorods), [7] nanoribbons (nanobelts), [8,9] and nanocrystals (NCs) [10,11] exhibit some novel characteristics. Because oxygen vacancies (OVs) play a critical role in many of the new physical phenomena, [9,12,13] more accurate identification and better understanding of the different OV structures are of fundamental and technical interest. SnO 2 has a tetragonal rutile structure (space groupd 14 4h, P42/mnm) with two molecules per unit cell. The irreducible representation of optic modes is Γ = A 1g + A 2g + B 1g + B 2g + E g + A 2u +2B 1u +3E u. [14] The A 2g and 2B 1u modes are neither Raman active nor infrared (IR) active. The symmetrical A 2u and E u modes are IR active, whereas the remaining A 1g,B 1g,B 2g, and E g modes are Raman active. Raman scattering is a useful tool to identify microstructural changes, and our previous investigations revealed that the Raman mode at ~574 cm 1 was closely associated with the inplane OVs in SnO 2 NCs. [13] The findings suggest that it is possible to correlate the different Raman modes with OV types (bridging and subbridging) and that Raman scattering can be employed to identify the types of OVs and to explore the physical mechanism. In this work, the Raman spectra acquired from SnO 2 NCs fabricated by pulsed laser ablation are studied systematically. The IR active mode at ~234 cm 1 is found to transform into a Raman active mode because of relaxation of the k = 0 selection rule induced by subbridging OV. Meanwhile, the existence of bridging OVs downshifts the A 1g mode at ~635 to ~618 cm 1. To investigate the underlying mechanism, the density functional theory (DFT) is adopted to derive the Raman spectra of SnO 2 NCs with different OV types, and the results indicate that the 234, 573, and 618 cm 1 Raman modes can be used as a fingerprint to identify the types of OVs in SnO 2 nanostructures. Samples and experiments Pulsed laser ablation was conducted in water with tin (99.99%) being the target to prepare the SnO 2 NC samples. The detailed experimental setup has been described previously. [15] After drying, the white powders were annealed at temperatures T a of 400, 600, 800, and 1000 C for 6 h under both oxygen and vacuum. The preparation conditions of the samples for both pulsed laser ablation and hydrothermal synthesis are listed in Table S1 (Supporting Information) for comparison. The samples were examined by high-resolution transmission electron microscopy (HR-TEM, JEOL-2100) and X-ray diffraction (XRD) (Philips X Per Pro X-ray diffractometer). The Raman scattering spectra were obtained on a T64000 triple monochromator Raman system with the 514.5-nm line of an argon ion laser as the excitation source without the polarization configuration. The resolution of the spectrometer was 0.5 cm 1. The diameter of the beam spot was 5 mm, and the power on the samples was less than 4 mw to avoid sample degradation due to laser heating. Results and discussion The HR-TEM image in Fig. 1(a) depicts the morphology of the SnO 2 NCs in the sample without any post-treatment. They have * Correspondence to: Wu, XL, Nanjing National Laboratory of Microstructures and Department of Physics, Nanjing University, Nanjing 210093, China. E-mail: hkxlwu@nju.edu.cn a Nanjing National Laboratory of Microstructures and Department of Physics, Nanjing University, Nanjing 210093, China b College of Electronic Engineering, Guangxi Normal University, Guilin 541004, China c Department of Physics and Materials Science, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong, China 1423 J. Raman Spectrosc. 2012, 43, 1423 1426 Copyright 2012 John Wiley & Sons, Ltd.

L. Z. Liu et al. Figure 1. (a) HR-TEM image of the SnO 2 NC sample fabricated by pulsed laser ablation without any post-processing. (b) Size distribution of the SnO 2 NCs calculated from HR-TEM results. (c) SAED pattern of the SnO 2 NCs. an almost spherical morphology to attain minimum surface free energy. The lattice fringes of 0.343 and 0.265 nm correspond to the (110) and (101) planes of crystalline SnO 2. The Gaussian fit of the SnO 2 NC size distribution is shown in Fig. 1(b) with the probable size and half-width being 3.45 and 0.66 nm. The selected-area electron diffraction (SAED) pattern is displayed in Fig. 1(c), and the three diffraction rings (from inner to outer) are associated with the (110), (101), and (211) planes of rutile SnO 2. The XRD spectra of the SnO 2 NC samples annealed in O 2 at different T a are shown in Fig. 2(a). The scan rate of the XRD pattern is 0.25 /s. All the diffraction peaks [(110), (101), and (211)] can be indexed to the crystalline SnO 2 growth planes. As T a is increased, the diffraction peaks become sharper and stronger indicating that the particles are become bigger and the crystal quality is improved. The Scherrer formula is employed to calculate the average size of the SnO 2 NCs. As T a is increased, the average sizes increase from 3.5 (the as-made) to 5.5 (400 C), 6.4 (600 C), 6.9 (800 C), and 10.4 nm (1000 C). Because the XRD patterns acquired from the samples annealed in vacuum are similar to those from the samples annealed in O 2, we infer that the SnO 2 NCs with similar sizes can be obtained in the samples with annealing treatment in vacuum. The Raman spectrum acquired from a SnO 2 NC sample fabricated hydrothermally is also shown in Fig. 2(b) for comparison, and the average size of the NCs is about 3.7 nm as determined by TEM. In addition to the E u LO mode (S 1 ) at 353.5 cm 1,E g mode (S 2 ) at 430.5 cm 1, and A 1g mode (S 4 ) at 630.0 cm 1, which are related to the small size effect according to Matossi force constant model, [6,16] a strong mode appears at 573.5 cm 1 (S3). This mode has been identified to arise from in-plane OVs. [13] It is known that the types of OVs can be controlled by adopting different fabrication methods and conditions. Here, pulsed laser ablation is utilized to fabricate SnO 2 nanostructures that possibly contain subbridging and bridging OVs. Figure 3(a) shows the representative Raman spectra acquired from the SnO 2 NC samples annealed under oxygen at different T a. Three weak peaks at 475.5 (f 5 ), 540.5 (f 3 ), and 498.5 cm 1 (f 2 ) correspond to the B 1u(2),B 1u(3), and E g modes of crystal SnO 2, respectively, [14] and the slight shifts can be attributed to the OV-induced phonon confinement effect. At T a =400 C, the f 4 mode (A 1g ) wavenumber (at ~619.5 cm 1 ) and intensity only change slightly compared with those in the unannealed sample, but the intensity of the f 1 mode (E u TO at ~234.5 cm 1 ) decreases rapidly. Factor-group analysis indicates that this mode is an IR active mode and it should not appear in the Raman spectrum. [14] However, Abello et al. [17] have suggested relaxation of the k = 0 selection rule because the perfect crystal is destroyed by the existence of OVs leading the IR to Raman transformation. Die guez et al. [18] have also reported that the SnO 2 nanoparticle surface region is usually covered by a disorder layer containing OVs, and this provides a reasonable explanation of the origin of this mode. At T a = 600 and 800 C, the Raman spectra of mode f 1 are only slightly modified, but the mode f 4 wavenumber f 1 (a) f 4 (b) Oxygen f 2 f 3 Vacuum 1424 Figure 2. (a) XRD patterns acquired from the SnO 2 NC samples annealed at T a = 0, 400, 600, 800, and 1000 C for 6 h under oxygen. (b) Raman spectrum acquired from the spherical SnO 2 NCs with a diameter of 3.7 nm produced hydrothermally. The inset displays the structural schematic diagram used in the DFT calculation for the SnO 2 (110) surface containing different types of OVs P: without OV; P 1 : bridging OV; P 2 : in-plane OV; P 3 : subbridging OV. Big and small balls represent Sn and O atoms, respectively. Raman Intensity 0 o C 0 o C 400 o C 400 o C 600 o C 600 o C 800 o C f 5 800 o C 1000 o C 1000 o C 260 390 520 650 260 390 520 650 Figure 3. Raman spectra acquired from SnO 2 NC samples annealed at T a = 0, 400, 600, 800, and 1000 C under (a) oxygen and (b) vacuum. wileyonlinelibrary.com/journal/jrs Copyright 2012 John Wiley & Sons, Ltd. J. Raman Spectrosc. 2012, 43, 1423 1426

Identifying oxygen vacancy types from Raman spectra of SnO 2 nanocrystals up-shifts significantly to 631.0 cm 1 (bulk) and the intensity is also enhanced. Disappearance of the weak peak (f 1 )att a =1000 C implies that the OVs are finally removed by high-temperature annealing. The results provide evidence that the f 1 and f 4 modes are closely related to OVs. Campbell and Fauchet s phonon confinement model is used to explain the f 4 mode wavenumber shift. [19,20] When the NC size increases from 5.5 to 10.3 nm, our calculation shows that the wavenumber only shifts from 628.1 to 632.3 cm 1. Therefore, it can be inferred that the size change is not the main reason for the f 4 mode shift. On the basis of the origin of mode f 1, mode f 4 should also be related to the OVs in the NCs. However, an obvious shift in the f 4 mode can be observed at T a = 600 C, implying that the f 4 mode depends on other types of OVs compared with the f 1 mode. According to the atom displacements model diagram, [14] the physical origins of the f 1 and f 4 modes can be ascribed to the subbridging and bridging OV types, respectively. For confirmation, the samples are also annealed in vacuum, and the corresponding Raman spectra are shown in Fig. 3(b). The Raman spectra are mostly unchanged and the small intensity change can be ascribed to better crystallization. After annealing in vacuum, the sizes of SnO 2 NCs become larger and their surfaces are reconstructed. However, the OV content hardly changes with their morphology transformation. [21] This verifies that OVs play a crucial role in the f 1 and f 4 modes. In addition, a broad shoulder type band appears at ~440 450 cm 1 in most of the Raman spectra. The shoulder band may arise from deeper OVs inside SnO 2 NCs. When the sample is annealed in O 2, the NCs grow and the surface OVs vanish easily, but the deeper internal OVs still exist. When the annealing temperature reaches 1000 C, OVs decrease largely. This leads to the appearance of the shoulder band. [22] For annealing treatment in vacuum, it only makes the SnO 2 NCs grow but does not largely eliminate OVs (especially internal OVs). As a result, the shoulder band hardly changes. To further explore the origins of the f 1 and f 4 modes, DFT is adopted to calculate the Raman characteristics of SnO 2 structures with different types of OVs along the (110) plane. With the use of the Broyden Fletcher Goldfarb Shanno minimizer in the CASTEP package with default convergence tolerances of 1.0 10 5 ev for energy and 0.3 ev/nm for maximum displacement, [9,13,23] the geometric structures of all the supercells are optimized to obtain stable structures as shown in the inset of Fig. 2. In our calculation, the norm-conserving pseudo-potential method is chosen together with the gradient correction and Perdew Burke Ernzerhof potential function. [24] The calculated Raman spectra of the SnO 2 structures containing bridging, subbridging, and in-plane OVsaswellaswithoutOVs(P 1,P 3,P 2, and P, respectively in the inset of Fig. 2) are presented in Fig. 4(a). In the presence of inplane OVs, the S 3 mode at ~573.5 cm 1 (line b) can be clearly observed. The existence of a bridging OV type (line c) causes the wavenumber of the f 4 mode to downshift, and a weak f 1 mode appears at the same time. Compared with the result of no OV, the subbridging OVs hardly affect the position of mode f 4 but slightly change its intensity. Evidently, the behavior of the f 1,S 3,andf 4 modes is different from that of the sample withoutov(linea),anditdependsstronglyontheovtypes. According to the calculated Raman spectra shown in Fig. 4(b), the experimental results can be explained. Firstly, the combined action of subbridging and bridging OVs not only effectively enhances the f 1 mode intensity but also downshifts the wavenumber of mode f 4 (line g), as consistent with the experimental Raman Intensity (a) f 1 260 390 520 650 results obtained from the T a =0 C sample in Fig. 3(a). Annealing at T a =400 C eliminates the subbridging OVs on the NC surface. As a result, the intensity of mode f 1 diminishes abruptly, but the wavenumber only up-shifts slightly, whereas the intensity and wavenumber of mode f 4 do not vary. This implies that bridging OVs still exist as verified by the calculated result (line f). Annealing at T a =600 C removes the bridging OVs. Mode f 4, which is affected by the bridging OV, returns to the position of the bulk mode. The residual weak peak at ~245.6 cm 1 (f 1 ) (T a = 600 and 800 C) may be related to OVs inside the NC. They can be completely eliminated at T a =1000 C and approach those of the calculated Raman spectrum without OVs (line e). The experimental and theoretical results indicate that the subbridging, bridging, and in-plane OVs are responsible for the f 1,f 4,andS 3 modes, respectively. The small deviation between the calculated and experimental data may be due to the complex nature of the samples. In order to read conveniently the peak positionsin Figs2 4, we have listed various peak wavenumbers in Tables S2 and S3 (Supporting Information) for comparison. Conclusion d: Subbridging c: Bridging S 1 b: In-Plane S 2 a: No OV In summary, the Raman spectra acquired from spherical SnO 2 NCs produced by pulsed laser ablation show that as the annealing temperature T a increases, the 234.5 and 618 cm 1 modes exhibit reduced intensity and location shift, respectively. Our DFT calculation reveals that the two modes depend strongly on the subbridging and bridging OVs and the 573.5 cm 1 mode arises from the in-plane OVs. [13] Our results demonstrate that Raman scattering can be used to identify the types of OVs in SnO 2 nanostructures. Acknowledgements f 3 f f 4 5 S 3 S 4 g: Subbridging and Bridging f: Bridging e: No OV 260 390 520 650 Figure 4. (a) Calculated Raman spectra with different OV types. (b) Calculated Raman spectra for comparison with the corresponding experimental Raman results. This work was jointly supported by Grants (Nos. 2011CB922102 and 60976063) from the National Natural Science Foundation and Basic Research Programs of China, and Postdoctoral Science Foundation of China and Jiangsu Province (Nos. 2011M500889 and 1102001B). Partial support was also from PAPD and Hong Kong Research Grants Council (RGC) General Research Fund (GRF) No. CityU 112510. (b) f 5 1425 J. Raman Spectrosc. 2012, 43, 1423 1426 Copyright 2012 John Wiley & Sons, Ltd. wileyonlinelibrary.com/journal/jrs

L. Z. Liu et al. Supporting information Supporting Information may be found in the online version of this article. References [1] S. Brovelli, A. Chiodini, A. Lauria, F. Meinardi, A. Paleari, Phys. Rev. B 2006, 73, 073406. [2] C. Killic, A. Zunger, Phys. Rev. Lett. 2002, 88, 095501. [3] S. O. Kucheyev, T. F. Baumann, P. A. Sterne, Y. M. Wang, T. Buuren, A. V. Hamza, L. J. Terminello, T. M. Willey, Phys. Rev. B 2005, 72, 035404. [4] Y. Idota, T. Kubota, A. Matsufuji, Y. Maekawa, T. Miyasaka, Science 1997, 276, 1395. [5] R. Asahi, T. Morijawa, T. Ohwaki, K. Aoki, Y. Taga, Science, 2001, 293, 269. [6] Y. J. Chen, L. Nie, X. Y. Xue, Y. G. Wang, T. H. Wang, Appl. Phys. Lett. 2006, 88, 083105. [7] L. Z. Liu, X. X. Li, X. L. Wu, X. T. Chen, P. K. Chu, Appl. Phys. Lett. 2011, 98, 133102. [8] J. Q. Hu, Y. Bando, Q. L. Liu, D. Golberg, Adv. Funct. Mater. 2003, 13,493. [9] H. T. Chen, S. J. Xiong, X. L. Wu, J. Zhu, J. C. Shen, P. K. Chu, Nano Lett. 2009, 1926. [10] E. J. H. Lee, C. Ribeiro, T. R. Giraldi, E. Longo, E. R. Leite, J. A. Varela, Appl. Phys. Lett. 2004, 84, 1745. [11] L. H. Jiang, G. Q. Sun, Z. H. Zhou, S. G. Sun, Q. Wang, S. Y. Yan, H. Q. Li, J. Tian, J. S. Guo, B. Zhou, Q. Xin, J. Phys. Chem. B 2005, 109, 8774. [12] M. A. Mäki-Jaskari, T. T. Rantal, Phys. Rev. B 2002, 65, 245428. [13] L. Z. Liu, X. L. Wu, F. Gao, J. C. Shen, T. H. Li, P. K. Chu, Solid State Commun. 2011, 151, 811. [14] T. Hirata, K. Ishioka, M. Kitajima, H. Doi, Phys. Rev. B 1996, 53, 8442. [15] T. H. Li, L. Z. Liu, X. L. Wu, J. C. Shen, F. Gao, P. K. Chu, Appl. Phys. Lett. 2010, 97, 121901. [16] J. Zou, C. Y. Xu, X. M. Liu, C. S. Wang, C. Y. Wang, Y. Hu, Y. T. Qian, J. Appl. Phys. 1994, 75, 1835. [17] L. Abello, B. Bochu, A. Gaskov, S. Koudryavtseva, G. Lucazeau, M. Roumyantseva, J. Solid State Chem. 1998, 135, 78. [18] A. Diéguez, A. Romano-Rodríguez, A. Vilà, J. R. Morante, J. Appl. Phys. 2001, 90, 1550. [19] L. Z. Liu, X. L. Wu, Z. Y. Zhang, T. H. Li, P. K. Chu, Appl. Phys. Lett. 2009, 95, 093109. [20] I. H. Campbell, P. M. Fauchtet, Solid State Commun. 1986, 58, 739. [21] T. H. Li, L. Z. Liu, X. X. Li, X. L. Wu, H. T. Chen, P. K. Chu, Opt. Lett. 2011, 36 4296. [22] K. N. Yu, Y. H. Xiong, Y. L. Liu, C. S. Xiong, Phys. Rev. B 1997, 55, 2666. [23] L. Z. Liu, X. L. Wu, J. C. Shen, T. H. Li, F. Gao, P. K. Chu, Chem. Comm. 2010, 46, 5539. [24] B. G. Pfrommer, M. Cote, S. G. Louie, M. L. J. Cohen, J. Comput. Phys. 1997, 131, 233. 1426 wileyonlinelibrary.com/journal/jrs Copyright 2012 John Wiley & Sons, Ltd. J. Raman Spectrosc. 2012, 43, 1423 1426

Supporting Information Identification of Oxygen Vacancy Types from Raman Spectra of SnO 2 Nanocrystals L. Z. Liu, T. H. Li, X. L. Wu, J. C. Shen, and Paul K. Chu Table S1. The comparison for preparation conditions between pulsed laser ablation and hydrothermal synthesis methods Method Hydrothermal synthesis Pulsed laser ablation Preparation condition Reactant: NH 3 H 2 O and SnCl 4 5H 2 O Condition: ph=7 Temperature: 200 o C Reactant: Sn target and H 2 O Laser power: 250 mj/pulse Laser wavelength: 248 nm Table S2. Peak positions in the experimental Raman spectra of Figure 2 and comparison with the calculated results of Figure 4 Shift (cm -1 ) Peak Mode Experimental Calculated S 1 E u 353.5 337.7 S 2 E g 430.5 S 3 (In-Plane OV) 573.5 573.5 S 4 A 1g 630.0 627.2 The sample is prepared by hydrothermal synthesis method, and peak S 3 originates from in-plane oxygen vacancies.

Table S3. Peak positions in the experimental Raman spectra of Figure 3 at different annealing temperatures, T a = 0, 400, 600, 800, 1000 o C, and comparison with the calculated results of Figure 4 Shift (cm -1 ) T a =0 o C T a =400 o C T a =600 o C T a = 800 o C T a = 1000 o C Calculated Peak Mode Oxygen Vacuum Oxygen Vacuum Oxygen Vacuum Oxygen Vacuum Oxygen Vacuum OV no OV f 1 E u 234.5 234.5 245.6 234.5 245.6 234.5 245.6 234.5 234.5 232.8 f 2 E g 498.5 498.5 498.5 498.5 499.3 498.5 499.3 498.5 499.3 498.5 f 3 B 1u(3) 540.5 540.5 540.5 540.5 540.5 540.5 540.5 540.5 540.5 540.5 543.5 543.5 f 4 A 1g 619.5 619.5 619.8 619.5 630.4 619.5 631.2 619.5 633.2 619.5 602.9 627.2 f 5 B 1u(2) 475.5 475.5 475.5 475.5 475.5 The samples are prepared by pulsed laser ablation, and peaks f 1 and f 4 are affected by subbridging and bridging oxygen vacancies, respectively.