Interfacial effect on Mn-doped TiO2 nanoparticles: from paramagnetism to ferromagnetism

Le Zhangab, Liping Zhu*ab, Liang Huab, Yaguang Liab, Hui Songab and Zhizhen Yeab
aState Key Laboratory of Silicon Materials, Department of Materials Science and Engineering, Zhejiang University, Hangzhou 310027, People's Republic of China. E-mail: zlp1@zju.edu.cn
bCyrus Tang Center for Sensor Materials and Applications, Zhejiang University, Hangzhou 310027, People's Republic of China

Received 12th March 2016 , Accepted 7th June 2016

First published on 10th June 2016


Abstract

Manganese-doped TiO2 nanoparticles with different crystalline structures, namely, anatase, rutile, mixed-phase of anatase and rutile, have been synthesized by a template method. We found that different structures lead to different defect concentrations and, as a result, strongly influence the magnetic properties. The pure anatase and rutile Mn-doped TiO2 are paramagnetic while the mixed-phase exhibits robust room temperature ferromagnetism. Extended X-ray absorption fine structure data reveal increasing oxygen vacancies when the mixed-phase formed. The key factor for activating ferromagnetism is found to be the interfacial defects created during phase-transformation, which is proposed to be oxygen vacancies. The defect effective Bohr radii are sufficiently large to overlap dopant ions, causing a net alignment of the dopant spins and then leading to a long-range ferromagnetic order. This discovery demonstrates that controlling the interfaces is an effective route to regulate magnetic properties and establish a prominent example of emergent phenomena at oxide interfaces.


1. Introduction

The appearance and manipulation of magnetism in traditional dilute magnetic semiconductors (DMSs) are nowadays some of the most active branches of material science. Over the past decade, research has been mainly focused on transition metal ions (Mn-, Co-, Cu-) doped oxides because of their potential application in spintronic devices.1 Recent experimental results indicate that the observed ferromagnetism of ZnO/TiO2 is attributed to the presence of impurities, defects, grain boundaries and surface/interface effects, but the origin is still a matter of controversy.2 Ever since the formation of a metallic layer was observed at the interface between the two insulators LaAlO3 and SrTiO3 around a decade ago, the oxide interface has been suggested to be a critical component of interesting phenomena, including magnetism and superconductivity.3 Prompted by the eagerness to understand the abnormal magnetic properties, considerable work has been carried out. García et al. experimentally showed that the interface formed at Zn diffusion front into Mn oxide resulted in room temperature ferromagnetism (RTFM) in Mn–Zn–O system.4 Kopnov et al. found an evidential relation between the surface roughness of silicon/silicon oxide interface and the measured magnetism by chemical etching.5 Serrano et al. confirmed a surface reduction of octahedral Co3+ to Co2+ in the mixtures of anatase-TiO2 and Co3O4 was related to the observed ferromagnetism.6 Recently, interface-based ferromagnetism was observed in CuO/Cu2O microspheres, and this observation confirmed the interface effect on engineering architecture with enhanced magnetic property.7 What's more, the electronic and magnetic properties of transition metal perovskites based interfaces have been widely studied due to manipulation of magnetism.8 Tunnelling experiments revealed that ultrathin BaTiO3 films simultaneously possessed a magnetization and a polarization with Fe or Co at the interface.9 Using LaAlO3/SrTiO3 as model, Lee et al. demonstrated an unexpected property with no bulk analogue in the constituent materials.10 Most recently, high-quality devices have been demonstrated by simply evaporating patterned top gates on the surface of a LaAlO3/SrTiO3 interface.11

These observations of interface-related ferromagnetism and recent technical advances in atomic-scale characterization have pioneered a new avenue for explorations of magnetic oxides-based interface systems, accelerating the discovery of new magnetic materials for spintronics application. Transition-metal (TM) doped TiO2 is a fantastic material with numerous applications in room temperature semiconductor spintronics,12 photocatalysis,13 sensitized solar cells,14 etc. Following the discovery of RTFM in Co-doped TiO2 by Matsumoto et al.,15 the research on the peculiar magnetic properties of these materials was focused on both anatase and rutile TM-doped TiO2 (e.g. Co2+, Mn2+, Ni2+) thin films systems.16 The most investigated DMS systems are characterized by super-exchange between magnetic ions via electrons trapped in oxygen vacancies, indicating that the doped TM ions and defects play a crucial role in these systems.17,18 The dd double exchange of multiple-valence transition metal d states can also induce local ferromagnetic-like behavior, reduced rutile TiO2−δ nanoparticles being one possible example.19 Although particular attention was devoted to Mn-, Co-doped anatase/rutile TiO2, recent work released enhanced ferromagnetism in Mn-doped TiO2 films during phase transition from anatase to rutile.20 Yıldırım et al. found different TM-implanted TiO2 structures strongly influence the magnetic properties.21,22 In this work, Mn-doped TiO2 nanoparticles with different magnetic properties have been prepared by adjusting their crystalline phase. Only in mixed-phase robust RTFM can be observed; however, pure anatase or rutile shows paramagnetism. This discrepancy is attributed to an increasing number of active magnetic Mn2+ ions and interfacial defects created during the phase transformation. These results demonstrate new possibility for effectively manipulating the ferromagnetic behavior in oxide semiconductors.

2. Experiment

The carbon spheres were firstly synthesized by hydrothermal method. A moderate amount of tetrabutyl titanate and manganese nitrate were dissolved in 100.0 mL anhydrous ethanol, followed by adding 1.0 g carbon spheres for templating. After ultrasonication 30 minutes, the as-obtained solution was vigorously stirred for 4 hours to get Ti4+, Mn2+ ions absorbed homogeneously. The obtained products were collected by centrifugation and rinsed with ethanol. The crystalline Mn-doped TiO2 powder was obtained by annealing the as-prepared products in air at 400 °C, 500 °C, 600 °C for 6 hours, respectively. XRD patterns were recorded by using a Panalytical X'PERT PRO MPD diffractometer with Cu Kα monochromatic radiation (λ = 1.54056 Å). SEM (SU-70) and HR-TEM (FEI TECNAI G2 F20) and corresponding energy-dispersive spectroscopy (EDS) mapping analyses were carried out to identify the morphology and nanostructure. Raman spectra were performed at room temperature by using a Thermo Fisher DXR SmartRaman spectrograph with 532 nm laser. X-ray photoelectron spectra (XPS) were acquired on an ESCALAB 250Xi with 0.45 eV resolution. Magnetic properties were measured on a Quantum Design magnetic property measurement system (MPMS3). Mn concentration was determined by inductively coupled optical emission spectrometry (ICP-OES, Agilent Technologies model 710). Electronic paramagnetic resonance (EPR) measurement was performed using a Bruker ESRA-300 spectrometer at 9.87 GHz (X-band). The Mn K-edge extended X-ray absorption fine structure (EXAFS) spectra were measured at the beamline 20-ID of the Advanced Photon Source (APS), Argonne National Laboratory.

3. Results and discussion

Determined by ICP-OES, Mn concentration is 1.85 at%, which is close to the nominal one (2.0 at%). Fig. 1 shows the XRD patterns of Mn-doped TiO2 annealing at different temperature. When the sample was annealed at 400 °C, intensive peaks at 2θ = 25.24°, 38.01°, 48.07°, 54.58° and 62.95° are observed, which correspond to the indices of (101), (004), (200), (211) and (213) planes of anatase phase, respectively.23 When the sample was annealed at 500 °C, besides typical anatase-like peaks, peaks at 2θ = 27.39° and 36.08° occur. These peaks represent the indices of (110) and (101) planes of rutile phase, respectively, indicating that a fraction of anatase phase has transformed to rutile phase after annealed at 500 °C. The rutile phase component is estimated to be 17% by using the following equation:24
 
image file: c6ra06606b-t1.tif(1)
where WR is rutile weight fraction, Aana and Arut are the X-ray integrated intensities for anatase (101) and rutile (110) diffraction peaks, respectively. When the sample was annealed at 600 °C, the diffraction peaks of anatase phase completely vanish. All the X-ray diffraction peak corresponds to the rutile phase, indicating the completeness of phase transformation. It is noted that no secondary phases of crystalline Mn2O3 or Mn3O4 are detected by XRD.

image file: c6ra06606b-f1.tif
Fig. 1 XRD pattern of Mn-doped TiO2 after annealing at 400 °C, 500 °C, 600 °C.

The template method maintains the hollow spheres morphology as observed by scanning electron microscopy (SEM) with a diameter from 150 to 200 nm (Fig. 2a). A transmission electron microscopy (TEM) image (Fig. 2b) demonstrates a thin shell with a thickness of about 10 nm. Scanning transmission electron microscopes (STEM) mode EDS mapping confirms the homogeneous distribution of Ti, O, Mn elements. High resolution TEM (HRTEM) analysis (Fig. 3) was applied to investigate the microstructure changes during temperature-variable annealing. It's clear to observe that lattice spacing corresponding to anatase (d = 3.54 Å, d = 2.42 Å) changes to that of rutile (d = 3.25 Å, d = 2.51 Å). Combined with TEM measurement, Penn et al.25 proposed that structural elements common to rutile can be produced at a subset of anatase interface, and these might serve as rutile nucleation sites. Lee et al.26 found that the nucleation of rutile occur at the amorphous interface of anatase particles due to strain and the disordered structure. On the basis of UV Raman, visible Raman and XRD results, Zhang et al.27 suggested that the phase transformation actually starts from the interfaces between the agglomerated anatase grains. Here, we propose that interfaces formed during phase transformation are likely to give rise to defects such as oxygen vacancy.28,29


image file: c6ra06606b-f2.tif
Fig. 2 (a) SEM image. (b) Magnified TEM image. (c) STEM image and corresponding elemental mapping of (d) Ti, (e) O and (f) Mn elements.

image file: c6ra06606b-f3.tif
Fig. 3 HRTEM images of Mn-doped TiO2. (a) Pure anatase. (b) Mixed-phase. (c) Pure rutile.

The Raman spectra collected from out-phase Mn-doped TiO2 in the region of 100–900 cm−1 are shown in Fig. 4a, and for convenience of comparison, undoped anatase TiO2 Raman spectra are also shown. According to previous report, anatase TiO2 (space group D4h, Z = 2) shows strong Raman peaks at 143 (Eg), 399 (B1g), 516 (A1g) and 638 cm−1 (Eg).30 When the sample was annealed at 400 °C, Eg Raman scattering mode shift to higher wavenumbers due to the distortion of Ti–O band by Mn incorporation.31 When the sample was annealed at 500 °C, the Raman scattering peaks become broader due to the increase of grain size.32 When the sample was annealed at 600 °C, Raman scattering mode at 143 cm−1 (Eg) diminishes completely. At the same time, new bands at 425 and 618 cm−1 appear, which are assigned to Eg and A1g modes of rutile phase respectively.24 Meanwhile, the intensity of the bands decrease due to the disorder of oxygen lattice induced by Mn incorporation.31 The phase transformation revealed by Raman spectra is consistent with the results of XRD and HRTEM well.


image file: c6ra06606b-f4.tif
Fig. 4 (a) Raman spectrum. (b) Ti 2p, (c) O 1s, (d) Mn 2p core level spectra of Mn-doped TiO2 with various annealing temperature.

To get a better insight into the local chemical environments, all these samples were further investigated by means of XPS. All the spectra were corrected for the C 1s peak at 284.6 eV. The X-ray photoelectron spectrum of Ti 2p core electron (Ti 2p3/2, binding energy 458.8 eV; Ti 2p1/2, binding energy 464.5 eV) for phase-pure anatase clearly matches previous reports, which are typical BE values of Ti4+ in anatase TiO2.33 For phase-pure rutile, these peak positions of Ti 2p3/2 and Ti 2p1/2 are at 458.5 eV and 464.3 eV, respectively, shifting slightly toward lower BE values. The Ti 2p3/2 spectrum from mixed-phase sample is found to be at 458.6 eV binding energy and fitted through a Gaussian–Lorentzian function according to phase composite.34 The peak area ratio of anatase to rutile is 4[thin space (1/6-em)]:[thin space (1/6-em)]1, which is consistent with the observed compositional ratio of anatase to rutile phase by XRD.

Further, the symmetric shape of Ti 2p lines suggest that any magnetic species of titanium such as Ti3+ can be ruled out.19 The binding energy value of the O 1s is found to be around 529.5 eV which is attributed to O–Ti4+ bond.31 And the O 1s spectrum shows an identical trend to Ti 2p spectrum, indicating the phase transformation. The O 1s peaks situated at 531.5 and 533.3 eV correspond to chemisorbed oxygen in hydroxyl group, with less than 10% area of the whole O 1s peaks.35 Hence the Ti 2p peak is supposed to contain only a small contribution from hydroxylated Ti environment.

In order to identify chemical state of Mn in TiO2 lattice, a slow scan of Mn 2p core level spectrum is recorded. The 12.2 eV-spaced Mn 2p3/2 and 2p1/2 peaks located at 641.1 eV and 653.3 eV indicate the presence of Mn2+ or Mn3+. Since the binding energy of Mn2+ and Mn3+ are both around ∼641 eV, it's difficult to distinguish Mn2+ from Mn3+ completely. While Mn2+ is discernible and Mn3+ is usually invisible in X-band EPR,36 on the basis of EPR result (discuss later), thus we can conclude that Mn ions are present as divalent state substituting Ti4+ in TiO2 crystal lattice curves of out-phase Mn-doped TiO2 nanoparticles. The pure anatase and rutile ones are paramagnetic due to the presence of magnetic Mn2+ ions. The mixed-phase sample shows unsaturated magnetization. The actual saturation magnetization possibly occurs at low magnetic field and the linear magnetization occurs at a higher magnetic field. The ferromagnetism may be due to ferromagnetic exchange between the Mn2+ ions and oxygen vacancies. Therefore, we can attribute the total magnetization of the sample to the sum of ferromagnetic and paramagnetic part.37 For comparison, the MH curve of undoped TiO2 at 300 K is also shown, which exhibits diamagnetism. The large Hc proves a strong ferromagnetic coupling rather than any contamination.38 Fig. 5b shows zero field cooling and field cooling (ZFC–FC) curves for the mixed-phase sample, obtained with a 500 Oe applied field. There is a peak closed to 60 K, which may be attributed to antiferromagnetic oxygen species.38 This particular peak has been widely reported in various O2-contaminated samples.39


image file: c6ra06606b-f5.tif
Fig. 5 (a) MH curves of Mn-doped TiO2 at room temperature. The inset shows MH curve of undoped TiO2. (b) Temperature dependence of ZFC and FC magnetization (MT curves) of mixed-phase sample at a fixed field (500 Oe). The inset shows enlarged view of coercivity.

In order to explore the origin of magnetism of these out-phase Mn-doped TiO2 nanoparticles, we measured their EPR spectra. Mn3+ (S = 2) is usually silent in X-band EPR due to a large zero-field splitting,36,40 here we can exclude the presence of Mn3+ ions. The Mn2+–Mn2+ antiferromagnetic super exchange interaction is known to be silent in X-band EPR41 and the isolated Mn2+ (S = 5/2) ions exhibit hyperfine splitting sextet lines in EPR spectrum due to allowed (ΔmI = 0) and forbidden (ΔmI = ±1) hyperfine transitions between the Zeeman sublevels mS = ±1/2, where mI and mS are nuclear spin and electron spin quantum numbers, respectively.42 The broad mainly symmetrical lines with a line width of 92 mT is probably due to higher concentration Mn2+ ions with enhanced dipolar interaction.43 The g factor in the Zeeman interaction is found centered at ∼2, which is close to previous report, indicating that the incorporated Mn2+ ions on Ti4+ lattice sites.44 The EPR integrated intensity demonstrates higher paramagnetic Mn2+ concentration in mixed-phase (Fig. 6).


image file: c6ra06606b-f6.tif
Fig. 6 (a) EPR spectra. (b) Mn K-edge Fourier transform EXAFS spectra.

To locally probe the fine structure, the Mn K-edge EXAFS spectra were taken for all samples. Curve-fitting was performed by ab initio FEFF calculation, to extract quantitative structural parameters.45 The structure model used was a anatase/rutile structure with the core Ti atom being replaced with Mn.46 This structural model provides a good fit to the experimental spectrum, and the Mn–O bond lengths are observed to be 1.94, 1.98 and 2.04 Å, respectively, excluding the presence the manganese-based oxides (Mn–O bond length is 2.20 Å of manganese-based oxides). It can be also confirm that Mn2+ ions substitute Ti4+ lattice sites, consistent with the EPR result. When Ti4+ is replaced with Mn2+, electroneutrality should decrease the O neighbors, namely, creating some oxygen vacancies. The coordination number is observed to be 4, 3.3 and 4, respectively, indicating more oxygen vacancies formed during phase transformation.

Finally, we briefly discuss the physical origin of the observed ferromagnetism of out-phase Mn-doped TiO2 nanostructure. Donor type defects, e.g. oxygen vacancies, have been proposed to be vital elements necessary for RTFM in TM-doped oxides.17,18 According to previous report,17 hydrogenic orbital radius of TiO2 is 0.48 nm, which is far smaller than grain size. So we can exclude the possibility of carriers localization stem from the spatial confinement effect.41,47 Pure anatase and rutile Mn[thin space (1/6-em)]:[thin space (1/6-em)]TiO2 nanoparticles are paramagnetic, lacking sufficient oxygen vacancies to couple with the magnetic Mn2+ ions to form bound magnetic polaron (BMP).37,48,49 The phase-transformation process creates interfacial defects, whose effective Bohr radii are large enough to overlap the concentrated Mn2+ ions. Following the BMP mechanism,17 a net alignment of the dopant spins occurs, leading to a long-range ferromagnetic order. These results additionally provide new microscopic information of interfacial defects. Finally, this finding suggests approaches to manipulating DMS ferromagnetism by tailoring the interfaces that will help guide experiments aimed at introducing appropriate defects into DMS and can be further applied to thin film-based spintronics application.

4. Conclusions

In summary, we have experimentally demonstrated the feasibility of achieving ferromagnetic interaction in Mn-doped TiO2 nanostructure, by phase-transformation strategy. Combined a detailed structural analysis, we confirm the phase-transformation process. The pure anatase and rutile Mn-doped TiO2 are paramagnetic while mixed-phase exhibits robust room temperature ferromagnetism, with saturation magnetization of 0.28 emu g−1. Based on the EPR and EXAFS experiment, we consider that an increasing concentration of interfacial defects forms BMP with the dopant magnetic ions, making the magnetic moment of mixed-phase observable. This finding provides a versatile system to induce and manipulate magnetic moments in non-magnetic materials, which have potential applications in spintronics.

Acknowledgements

This work was supported by National Natural Science Foundation of China 51372224, 51072181 and 51572239, Program for Innovative Research Team in University of Ministry of Education of China (IRT13037) and National Science and Technology Support Program (2012BAC08B08). We would especially like to thank Dr Wenge Yang and Dr Bin Chen at Center for High Pressure Science & Technology Advanced Research (HPSTAR) for arranging EXAFS measurement at Advanced Photon Source (APS), Argonne National Laboratory.

References

  1. F. Pan, C. Song, X. J. Liu, Y. C. Yang and F. Zeng, Mater. Sci. Eng., R, 2008, 62, 1–35 CrossRef.
  2. S. B. Ogale, Adv. Mater., 2010, 22, 3125–3155 CrossRef CAS PubMed.
  3. A. Ohtomo and H. Y. Hwang, Nature, 2004, 427, 423–426 CrossRef CAS PubMed.
  4. M. A. García, M. L. Ruiz-González, A. Quesada, J. L. Costa-Krämer, J. F. Fernández, S. J. Khatib, A. Wennberg, A. C. Caballero, M. S. Martín-González, M. Villegas, F. Briones, J. M. González-Calbet and A. Hernando, Phys. Rev. Lett., 2005, 94, 217206 CrossRef PubMed.
  5. G. Kopnov, Z. Vager and R. Naaman, Adv. Mater., 2007, 19, 925–928 CrossRef CAS.
  6. A. Serrano, E. F. Pinel, A. Quesada, I. Lorite, M. Plaza, L. Pérez, F. Jiménez-Villacorta, J. de la Venta, M. S. Martín-González, J. L. Costa-Krämer, J. F. Fernandez, J. Llopis and M. A. García, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79, 144405 CrossRef.
  7. D. Gao, Z. zhang, Q. Xu, J. Zhang, Z. Yan, J. Yao and D. Xue, Appl. Phys. Lett., 2014, 104, 022406 CrossRef.
  8. F. Matsukura, Y. Tokura and H. Ohno, Nat. Nanotechnol., 2015, 10, 209–220 CrossRef CAS PubMed.
  9. S. Valencia, A. Crassous, L. Bocher, V. Garcia, X. Moya, R. O. Cherifi, C. Deranlot, K. Bouzehouane, S. Fusil, A. Zobelli, A. Gloter, N. D. Mathur, A. Gaupp, R. Abrudan, F. Radu, A. Barthelemy and M. Bibes, Nat. Mater., 2011, 10, 753–758 CrossRef CAS PubMed.
  10. J. S. Lee, Y. W. Xie, H. K. Sato, C. Bell, Y. Hikita, H. Y. Hwang and C. C. Kao, Nat. Mater., 2013, 12, 703–706 CrossRef CAS PubMed.
  11. S. Goswami, E. Mulazimoglu, L. M. K. Vandersypen and A. D. Caviglia, Nano Lett., 2015, 15, 2627–2632 CrossRef CAS PubMed.
  12. Y. Yamada, K. Ueno, T. Fukumura, H. T. Yuan, H. Shimotani, Y. Iwasa, L. Gu, S. Tsukimoto, Y. Ikuhara and M. Kawasaki, Science, 2011, 332, 1065–1067 CrossRef CAS PubMed.
  13. Y. Ma, X. Wang, Y. Jia, X. Chen, H. Han and C. Li, Chem. Rev., 2014, 114, 9987–10043 CrossRef CAS PubMed.
  14. P. K. Santra and P. V. Kamat, J. Am. Chem. Soc., 2012, 134, 2508–2511 CrossRef CAS PubMed.
  15. Y. Matsumoto, Science, 2001, 291, 854–856 CrossRef CAS PubMed.
  16. F. L. Deepak, M. Banobre-Lopez, E. Carbo-Argibay, M. F. Cerqueira, Y. Pineiro-Redondo, J. Rivas, C. M. Thompson, S. Kamali, C. Rodriguez-Abreu, K. Kovnir and Y. V. Kolen'ko, J. Phys. Chem. C, 2015, 119, 11947–11957 CAS.
  17. J. M. Coey, M. Venkatesan and C. B. Fitzgerald, Nat. Mater., 2005, 4, 173–179 CrossRef CAS PubMed.
  18. T. Zhao, S. R. Shinde, S. B. Ogale, H. Zheng, T. Venkatesan, R. Ramesh and S. Das Sarma, Phys. Rev. Lett., 2005, 94, 126601 CrossRef CAS PubMed.
  19. M. Parras, Á. Varela, R. Cortés-Gil, K. Boulahya, A. Hernando and J. M. González- Calbet, J. Phys. Chem. Lett., 2013, 4, 2171 CrossRef CAS.
  20. J. P. Xu, Y. B. Lin, Z. H. Lu, X. C. Liu, Z. L. Lu, J. F. Wang, W. Q. Zou, L. Y. Lv, F. M. Zhang and Y. W. Du, Solid State Commun., 2006, 140, 514–518 CrossRef CAS.
  21. O. Yıldırım, S. Cornelius, M. Butterling, W. Anwand, A. Wagner, A. Smekhova, J. Fiedler, R. Böttger, C. Bähtz and K. Potzger, Appl. Phys. Lett., 2015, 107, 242405 CrossRef.
  22. O. Yildirim, S. Cornelius, A. Smekhova, G. Zykov, E. A. Gan'shina, A. B. Granovsky, R. Hübner, C. Bähtz and K. Potzger, J. Appl. Phys., 2015, 117, 183901 CrossRef.
  23. J. Zhang, M. Li, Z. Feng, J. Chen and C. Li, J. Phys. Chem. B, 2006, 110, 927–935 CrossRef CAS PubMed.
  24. W. Su, J. Zhang, Z. Feng, T. Chen, P. Ying and C. Li, J. Phys. Chem. C, 2008, 112, 7710–7716 CAS.
  25. R. L. Penn and J. F. Banfield, Am. Mineral., 1999, 84, 871–876 CrossRef CAS.
  26. G. H. Lee and J.-M. Zuo, J. Am. Ceram. Soc., 2004, 87, 473–479 CrossRef CAS.
  27. K. N. P. Kumar, K. Keizer, A. J. Burggraaf, T. Okubo, H. Nagamoto and S. Morooka, Nature, 1992, 358, 48–51 CrossRef CAS.
  28. J. D. Bryan, S. M. Heald, S. A. Chambers and D. R. Gamelin, J. Am. Chem. Soc., 2004, 126, 11640–11647 CrossRef CAS PubMed.
  29. J. D. Bryan, S. A. Santangelo, S. C. Keveren and D. R. Gamelin, J. Am. Chem. Soc., 2005, 127, 15568–15574 CrossRef CAS PubMed.
  30. T. Ohsaka, F. Izumi and Y. Fujiki, J. Raman Spectrosc., 1978, 7, 321–324 CrossRef.
  31. S. Sharma, S. Chaudhary, S. C. Kashyap and S. K. Sharma, J. Appl. Phys., 2011, 109, 083905 CrossRef.
  32. X. Xue, W. Ji, Z. Mao, H. Mao, Y. Wang, X. Wang, W. Ruan, B. Zhao and J. R. Lombardi, J. Phys. Chem. C, 2012, 116, 8792–8797 CAS.
  33. X. Y. Ma, Z. G. Chen, S. B. Hartono, H. B. Jiang, J. Zou, S. Z. Qiao and H. G. Yang, Chem. Commun., 2010, 46, 6608–6610 RSC.
  34. D. O. Scanlon, C. W. Dunnill, J. Buckeridge, S. A. Shevlin, A. J. Logsdail, S. M. Woodley, C. R. Catlow, M. J. Powell, R. G. Palgrave, I. P. Parkin, G. W. Watson, T. W. Keal, P. Sherwood, A. Walsh and A. A. Sokol, Nat. Mater., 2013, 12, 798–801 CrossRef CAS PubMed.
  35. H. Huang and D. Y. C. Leung, ACS Catal., 2011, 1, 348–354 CrossRef CAS.
  36. S. Bhattacharyya, A. Pucci, D. Zitoun and A. Gedanken, Nanotechnology, 2008, 19, 495711 CrossRef PubMed.
  37. B. Choudhury and A. Choudhury, Curr. Appl. Phys., 2013, 13, 1025–1031 CrossRef.
  38. L. Hu, J. Huang, H. He, L. Zhu, S. Liu, Y. Jin, L. Sun and Z. Ye, Nanoscale, 2013, 5, 3918–3930 RSC.
  39. T. Dubroca, J. Hack and R. Hummel, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 74, 026403 CrossRef.
  40. F. Iacomi, N. Apetroaei, G. Calin, G. Zodieriu, M. M. Cazacu, C. Scarlat, V. Goian, D. Menzel, I. Jursic and J. Schoenes, Thin Solid Films, 2007, 515, 6402–6406 CrossRef CAS.
  41. S. T. Ochsenbein, Y. Feng, K. M. Whitaker, E. Badaeva, W. K. Liu, X. Li and D. R. Gamelin, Nat. Nanotechnol., 2009, 4, 681–687 CrossRef CAS PubMed.
  42. P. Szirmai, E. Horváth, B. Náfrádi, Z. Micković, R. Smajda, D. M. Djokić, K. Schenk, L. Forró and A. Magrez, J. Phys. Chem. C, 2013, 117, 697–702 CAS.
  43. Y. Yang, Y. Li, L. Zhu, H. He, L. Hu, J. Huang, F. Hu, B. He and Z. Ye, Nanoscale, 2013, 5, 10461–10471 RSC.
  44. P. Umek, C. Bittencourt, P. Guttmann, A. Gloter, S. D. Škapin and D. Arčon, J. Phys. Chem. C, 2014, 118, 21250–21257 CAS.
  45. Z. Sun, W. Yan, T. Yao, Q. Liu, Y. Xie and S. Wei, Dalton Trans., 2013, 42, 13779–13801 RSC.
  46. J. Resasco, N. P. Dasgupta, J. R. Rosell, J. Guo and P. Yang, J. Am. Chem. Soc., 2014, 136, 10521–10526 CrossRef CAS PubMed.
  47. D. A. Bussian, S. A. Crooker, M. Yin, M. Brynda, A. L. Efros and V. I. Klimov, Nat. Mater., 2009, 8, 35–40 CrossRef CAS PubMed.
  48. B. Santara, P. K. Giri, S. Dhara, K. Imakita and M. Fujii, J. Phys. D: Appl. Phys., 2014, 47, 235304 CrossRef.
  49. B. Choudhury, R. Verma and A. Choudhury, RSC Adv., 2014, 4, 29314 RSC.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.