Hostname: page-component-8448b6f56d-jr42d Total loading time: 0 Render date: 2024-04-23T21:21:54.942Z Has data issue: false hasContentIssue false

NMR chemical shift of single-wall carbon nanotubes

Published online by Cambridge University Press:  15 March 2011

Sylvain Latil
Affiliation:
Groupe de Dynamique des Phases Condensées, CNRS-Université de Montpellier 2, France
Christopher Goze
Affiliation:
Groupe de Dynamique des Phases Condensées, CNRS-Université de Montpellier 2, France
Goze Bac
Affiliation:
Groupe de Dynamique des Phases Condensées, CNRS-Université de Montpellier 2, France
Patrick Bernier
Affiliation:
Groupe de Dynamique des Phases Condensées, CNRS-Université de Montpellier 2, France
Luc Henrard
Affiliation:
Laboratoire de Physique du Solide, Facultés Universitaries Notre-Dame de la Paix, Namur, Belgium
Angel Rubio
Affiliation:
Donostia International Physics Center (DIPC) and Centro Mixto CSIC-UPV, 20018 San Sebastián/Donostia, Basque Country, Spain and Departamento de Física Teórica, Universidad de Valladolid, Spain
Get access

Abstract

We report calculation of the NMR chemical shift anisotropy (CSA) tensor →μ of single-wall carbon nanotubes, within the London approximation (ring currents contribution). Our results indicate that the isotropic line as measured by high resoultion experiments is splited about 11ppm between metallic and semiconductor nanotubes. We carefully check that this result remains vaild and observable when the bundle packing is taken into account. The resulting broadening is aroune 20ppm, but reduces onto a sharp lorentzian (<1ppm) when averaging by high resoultion NMR measurements.

Type
Research Article
Copyright
Copyright © Materials Research Society 2001

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

1. Tang, X.-p. et al. , Scinece 288, 492 (2000).Google Scholar
2. Bac, C. Goze et al. , Phys. Rev. B. 63 100302 (2001) Ibid, 13 CNMR study of functionalized nanotubes, in this conference.Google Scholar
3. Mintmire, J., Dunlap, B., and White, C., Phys. Rev. Lett. 68, 631 (1992).Google Scholar
4. Lu, J.-P. Phys. Rev. Lett. 74, 1123 (1995). the susceptibility of incinite and isolated tubes when Ho∥z, is diamagentic for semiconducting tubes and paramagnetic for metallic tubes. When Ho ⊥, it is diamagetic in any case.Google Scholar
5. Satio, R., Dresselhaus, G., and Dresselhaus, M., Physical properties of carbon nanotubes (Imperial College Press, London, 1998).Google Scholar
6. Mintmire, J. and White, C., Phys. Rev. Lett. 80, 2506 (1998).Google Scholar
7. Pople, J., J. Chem. Phys. 37, 53 (1962).Google Scholar
8. Pasquarello, A., Sclüter, M., and Haddon, R.C., Phys. Rev. A47, 1783 (1993).Google Scholar
9. Rols, S. et al. , Eur. Phys. J. 10, 263 (1999).Google Scholar
10. For the present work this approximation is reasonable. However, the given results are not realtive to tetrmatethysilane. In our case, a positive (negative) shift correspond to a diamagnetic (paramagnetic) shift, For an absolute chemical shift computation, we need to use first principles ab initio techniques: See Mauri, F. et al. Phys. Rev. Lett. 79, 2340 (1999) and refernces therein for the method applied to carbon materials.Google Scholar
11. Memory, J.D.. Quantum theory of magnetic resonance parameters (McGraw-Hill, New york, 1968).Google Scholar
12. Salem, L.The molecular orbital theroy of conjugaged systems (W.A) Benjamin inc, Reading 1966).Google Scholar
13. In this report, the value of the first neighbour hopping integral is 2.66 eV (Benzene).Google Scholar
14. Mehring, M.. Hign resolution NMR in solids (Springer-Verlag, Heidelberg, 1976), see chapter 2.4.Google Scholar