Elsevier

Chemical Physics Letters

Volume 643, January 2016, Pages 126-130
Chemical Physics Letters

Identification of different tin species in SnO2 nanosheets with 119Sn solid-state NMR spectroscopy

https://doi.org/10.1016/j.cplett.2015.11.035Get rights and content

Highlights

  • Sn ions in 1st/2nd layer, and ‘bulk’ in SnO2 nanosheets can be resolved with NMR.

  • NMR spectroscopy provides fine resolution that no other spectroscopy currently shows.

  • This approach based on NMR should be able to be extended to other diamagnetic metal oxides.

Abstract

119Sn solid-state nuclear magnetic resonance (NMR) spectroscopy was employed to investigate the structure of hydroxylated SnO2 nanosheets. Three 119Sn resonances can be observed and assigned to Sn ions in the first layer, the bulk and the second layer from high to low frequencies with the help of density functional theory (DFT) calculations. The results suggest that 119Sn NMR spectroscopy can be a sensitive method to monitor the structure of SnO2 based nanomaterials and extension of this approach to other diamagnetic metal oxides.

Graphical abstract

Tin ions in the first layer, second layer and ‘bulk-like’ environment of ultrathin SnO2 nanosheets can be distinguished by using 119Sn solid-state NMR spectroscopy.

  1. Download : Download high-res image (116KB)
  2. Download : Download full-size image

Introduction

Nanostructured materials show different, often better properties from their bulk counterparts, which arise from their higher surface areas and sometimes electronic properties, and thus are used or have potential applications in energy storage, catalysis, biosciences and environmental management. In order to rationally design nanomaterials with improved performances, it is critical to understand their structure details, especially surface structure, and develop structure and property relationships [1], [2], [3], [4]. Although electron microscopy is the most important tool to determine the surface structure, the small volumes investigated under such methods may not be representative of the entire sample. Other characterization techniques such as Raman [5], infrared (IR) [6], [7], [8] and ultraviolet (UV) spectroscopy [9], X-ray photoelectron spectroscopy (XPS) [10], [11] and low energy electron diffraction (LEED) [12] can also be used to extract the structural information of nanomaterials, however, much lower resolution is achieved.

Recently we showed that 17O NMR spectroscopy, in combination with DFT calculations, can be an extremely sensitive probe to detect and resolve oxygen ions in oxide nanostructures [13]. For example, the 1st, 2nd and 3rd layer of ceria nanoparticles, oxygen ions in the bulk, as well as surface hydroxyl species. However, this approach requires surface selective isotopic labeling due to the low natural abundance of 17O. In principle, the NMR parameters (e.g., chemical shift) of any nucleus should be sensitive to its local structure and thus NMR spectroscopy of metal ions should also be a powerful method for investigations of surface structure of nanosized oxides. Since nanostructured tin dioxide (SnO2) have remarkable electronic, electrochemical and optical properties which lead to a variety of applications, such as catalytic carrier [14], [15], lithium-ion batteries [16], [17], [18], solar cells [19] and gas sensors [20], [21], [22], [23], and 119Sn has relatively high gyromagnetic ratio (γ(119Sn) = 0.374γ(1H)) and natural abundance (8.58%), we attempted to develop the new characterization method on the basis of 119Sn solid-state NMR spectroscopy on SnO2 nanomaterials. Specifically, hydroxylated SnO2 nanosheets were prepared and investigated because the well-defined nanostructures. We are able to demonstrate here that 119Sn NMR spectroscopy can also be a sensitive technique to monitor the surface structure of nanosized SnO2. We find that the 119Sn chemical shifts can distinguish Sn ions in the first, second layer and ‘bulk’ environment in the nanosheets, thus can be used to follow the structure evolution of SnO2 based materials in related applications..

Section snippets

Synthesis of SnO2 nanosheets

In a typical synthesis process, 374 mg of the precursor SnCl2·2H2O was added into 100 mL mixture of ethanol and deionized water (volume ratio = 1:1). After magnetically stirred for 5 min, ammonia was slowly (0.5 mL/min) introduced into the mixture until the pH value reached 11. After vigorous stirring for 30 min, the obtained white turbid suspension was transferred into a 100 mL Teflon-lined autoclave, heated with magnetic stirring at 393 K for different time in an oil bath pan before it was rapidly

Results and discussion

The XRD pattern of as-obtained SnO2 sample (Figure S2) can be indexed to rutile phase with a tetragonal structure (JCPDS No. 41-1445). The broad peak widths indicate the presence of nanosized crystallites. The TEM and HRTEM images of SnO2 nanostructures are shown in Figure S3and Figure 1A, in which light and dark regions can be observed. The light regions, which are the majority of the sample, suggest planar thin sheets lying on the substrate, while dark regions imply that nanosheets may either

Conclusions

119Sn solid-state NMR spectroscopy, in combination with the DFT calculations, has been demonstrated to identify different Sn species in SnO2 nanosheets. The resonances at −585, −604 and −618 ppm can be assigned to Sn ions in the 1st layer, ‘bulk-like’ environment and 2nd layer in the hydroxylated nanosheets, respectively. The relatively large chemical shift differences found in different species suggest 119Sn NMR chemical shift is a sensitive structural probe. The results also suggest that NMR

Acknowledgements

This work is supported by National Basic Research Program of China (2013CB934800), National Natural Science Foundation of China (NSFC) (21222302, 20903056 and 21322307), NSFC – Royal Society Joint Program (21111130201), Program for New Century Excellent Talents in University (NCET-10-0483), Fundamental Research Funds for the Central Universities (1124020512), National Science Fund for Talent Training in Basic Science (J1103310) and a Project Funded by the Priority Academic Program Development

References (33)

  • A. Ayeshamariam et al.

    Spectrochim. Acta A

    (2014)
  • P. Chetri et al.

    J. Alloys Compd.

    (2015)
  • I. Hannus et al.

    J. Mol. Struct.

    (1995)
  • J.H. Kwak et al.

    Science

    (2009)
  • Q. Fu et al.

    Science

    (2010)
  • H.G. Yang et al.

    Nature

    (2008)
  • G.V. Lier et al.

    Chem. Phys. Lett.

    (2000)
  • R.G. Pavelko et al.

    Phys. Chem. Chem. Phys.

    (2010)
  • K. Nejati

    Cryst. Res. Technol.

    (2012)
  • S. Shah et al.

    J. Mater. Chem.

    (2012)
  • C.W. Cheng et al.

    ACS Nano

    (2009)
  • M. Kwokaa et al.

    Thin Solid Films

    (2005)
  • S.Y. Ho et al.

    Cryst. Growth Des.

    (2009)
  • M.S. Hettenbach et al.

    Surf. Sci.

    (2001)
  • M. Wang et al.

    Sci. Adv.

    (2015)
  • A. Kowal1 et al.

    Nat. Mater.

    (2009)
  • Cited by (15)

    • Hyperpolarization transfer pathways in inorganic materials

      2021, Journal of Magnetic Resonance
      Citation Excerpt :

      The surface spectrum of SnO2 is characterized by three notable peaks corresponding to different sites. These signals can be observed in hydroxylated SnO2 nanosheets and have been assigned to different atomic layers [33]. The bulk-like 119Sn is at −604 ppm (B) and the first and second atomic layer, respectively, were assigned chemical shifts of −585 ppm (S1) and −618 ppm (S2).

    • Sol-gel synthesis of sno2 nanoparticles doped with zr+4/cd+2 and co-doped with f: Their structural, optical and electrical properties

      2021, Materials Today: Proceedings
      Citation Excerpt :

      EDS analysis has shown incorporation of dopant ions into SnO2 network Fig. 3. Solid state 119Sn MAS NMR spectroscopy is an important tool for study the different chemical environment around Sn+4 ions in SnO6 polyhedra of casitarite structure of SnO2 [28–30]. To analyze presence of different chemical environment around tin in SnO2, 119Sn MAS NMR spectra of one of the doped samples (4) along with pure SnO2 synthesized with similar route have been recorded and are shown in Fig. 4.

    • Relationship between tin environment of SnO<inf>2</inf> nanoparticles and their electrochemical behaviour in a lithium ion battery

      2021, Materials Chemistry and Physics
      Citation Excerpt :

      The shoulder line is observed for solvo and hydrothermal materials at about −610 ppm. Chen et al. reported that this peak is derived from surface Sn species that are close to absorbed water molecules and hydroxyl groups [38]. This is consistent with the observation of the Sn–OH bond identified by infrared for solvo and hydrothermal materials.

    • Highly efficient removal of U(VI) by the photoreduction of SnO<inf>2</inf>/CdCO<inf>3</inf>/CdS nanocomposite under visible light irradiation

      2020, Applied Catalysis B: Environmental
      Citation Excerpt :

      Fig. 1B shows the FTIR spectra of the samples. For SnO2, the broad absorption bands appearing at 567 and 679 cm−1 are assigned to SnO antisymmetric vibrations and the absorption band at 1037 cm−1 is consistent with the previously reported overtone and combination bands of bulk cassiterite [27,28]. The bands appearing at ∼1630 cm−1 in all curves are related to the adsorbed H2O molecules.

    • Efficient solid acid catalysts based on sulfated tin oxides for liquid phase esterification of levulinic acid with ethanol

      2018, Applied Catalysis A: General
      Citation Excerpt :

      Thus we suggest that the calcination of the precursor SnO2.xH2O gel at 500 °C resulted in formation of the inactive β-form which has less bound water [43]. The progressive increase of crystallite size with increasing temperature is also documented in the literature [46]. The 119Sn NMR spectra of samples SO42−/SnO2(100) and SO42−/SnO2(140) synthesized by sulfating of SnO2 precursors obtained without calcination step (SnO2(100) and SnO2(140)) show that the sulfation procedure resulted in disappearance of the signal at −601 ppm and in the appearance of new peaks at −769 ppm and −786 ppm (Fig. 7).

    View all citing articles on Scopus
    1

    These authors contributed equally.

    View full text