Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-20T03:08:21.741Z Has data issue: false hasContentIssue false

Computational Studies of Novel Thermoelectric Materials

Published online by Cambridge University Press:  15 February 2011

D.J. Singh
Affiliation:
Code 6691, Naval Research Laboratory, Washington, DC 20375
I.I. Mazin
Affiliation:
CSI, George Mason University and Code 6691 NRL, Washington, DC 20375
S.G. KIM
Affiliation:
UES Inc. and Code 6691 NRL, Washington, DC 20375
L. Nordstrom
Affiliation:
Physics Department, Uppsala University, Box 530, S-75121 Uppsala, Sweden
Get access

Abstract

The thermoelectric properties of La-filled skutterudites and β-Zn4Sb3 are discussed from the point of view of their electronic structures. These are calculated from first principles within the local density approximation. The electronic structures are in turn used determine transport related quantities. β-Zn4Sb3 is found to be metallic with a complex Fermi surface topology, which yields a non-trivial dependence of the Hall concentration on the band filling. Calculations of the variation with band filling are used to extract the carrier concentration from the experimental Hall number. At this band filling, which corresponds to 0.1 electrons per 22 atom unit cell, we calculate a Seebeck coefficient and temperature dependence in good agreement with the experimental value. The high Seebeck coefficients in a metallic material are remarkable, and arise because of the strong energy dependence of the Fermiology near the experimental band filling. Virtual crystal calculations for La(Fe,Co)4Sb12 show that the system obeys almost perfect rigid band behavior with varying Co concentration, and has a substantial band gap at a position corresponding to the composition LaFe3CoSb12. The valence band maximum occurs at the Γ point and is due to a singly degenerate dispersive (Fe,Co)-Sb band, which by itself would not be favorable for TE. However, very flat transition metal derived bands occur in close proximity and become active as the doping level is increased, giving a non-trivial dependence of the properties on carrier concentration and explaining the favorable TE properties.

Type
Research Article
Copyright
Copyright © Materials Research Society 1997

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. Mahan, G., Sales, B. and Sharp, J., Physics Today 50 #3, 42 (1997).Google Scholar
2. Caillat, T., Fleurial, J.P. and Borshchevsky, A. in Proceedings ICT'96 edited by Caillat, T., Borshchevsky, A. and Fleurial, J.P. (IEEE Press, Piscataway, 1996), pp. 151154.Google Scholar
3. Sales, B.C., Mandrus, D. and Williams, R.K., Science 272, 1325 (1996).Google Scholar
4. Fleurial, J.P., Borshchevsky, A., Caillat, T., Morelli, D.T. and Meisner, G.P. in Proceedings ICT'96 edited by Caillat, T., Borshchevsky, A. and Fleurial, J.P. (IEEE Press, Piscataway, 1996), pp. 9195.Google Scholar
5. Morelli, D.T. and Meisner, G.P., J. Appl. Phys. 77, 3777 (1995).Google Scholar
6. Nordstrom, L. and Singh, D.J., Phys. Rev. B 53, 1103 (1996).Google Scholar
7. Singh, D.J. and Pickett, W.E., Phys. Rev. B 50, 11235 (1994).Google Scholar
8. Feldman, J.L. and Singh, D.J., Phys. Rev. B 53, 6273 (1996); 54, 712 (1996).Google Scholar
9. Singh, D.J., Planewaves, Pseudopotential and the LAPW Method, Kluwer, Boston, 1994.Google Scholar
10. Ziman, J.M., Principles of the Theory of Solids, Cambridge University Press, Cambridge, 1972.Google Scholar
11. Hurd, C.M., The Hall effect in Metals and Alloys, Plenum, New York, 1972.Google Scholar
12. Schulz, W.W., Allen, P.B. and Trivedi, N., Phys. Rev. B 45, 10886 (1992).Google Scholar
13. Karakozov, A.E., Mazin, I.I. and Uspenski, Y.A., Sov. Phys, Doklady 277, 848 (1984).Google Scholar
14. Slack, G.A. in Solid State Physics edited by Seitz, F. and Turnbull, D. (Academic, New York, 1979), Vol.34, p. 1 Google Scholar
15. Cahill, D.G., Watson, S.K. and Pohl, R.O., Phys. Rev. B 46, 6131 (1992).Google Scholar
16. Caillat, T., Kulleck, J., Borshchevsky, A. and Fleurial, J.P., J. Appl. Phys. 79, 8419 (1996).Google Scholar
17. Nolas, G.S., Slack, G.A., Morelli, D.T., Tritt, T.M. and Ehrlich, A.C., J. Appl. Phys. 79, 4002 (1996).Google Scholar
18. Morelli, D.T., Caillat, T., Fleurial, J.P., Borshchevsky, A., Vandersande, J., Chen, B. and Uher, C., Phys. Rev. B 51, 9622 (1995).Google Scholar
19. Mandrus, D., Migliori, A., Darling, T.W., Hundley, M.F., Peterson, E.J. and Thompson, J.D., Phys. Rev. B 52, 4926 (1995).Google Scholar
20. Caillat, T., Borshchevsky, A. and Fleurial, J.P., J. Appl. Phys. 80, 4442 (1996).Google Scholar
21. The inverse effective mass tensor is not fully isotropic especially for the heavier band. This is symmetry allowed because of the missing mirror planes in the spacegroup.Google Scholar
22. In metals convergence is usually obtained with an energy range of ±5kT, but because of the strong variation of σ due to the band onsets we used ±10kT in the present calculations.Google Scholar
23. The light band is derived from the Sb p states directed perpendicular to the plane of the Sb4 rings. These hybridize strongly with Co 4p like states that mix with the La 4f resonance. This resonance, which is above the gap, pushes the light band down.Google Scholar