Thermoelectric Properties of SrTiO3
Thermoelectric Properties of SrTiO3
Thermoelectric Properties of SrTiO3
6, 2015
DOI: 10.1007/s11664-014-3560-0
2014 The Minerals, Metals & Materials Society
INTRODUCTION
With the rapid depletion of fossil fuels and the
increasing demand for energy, renewable sources of
energy and ways to increase energy efficiency are of
growing importance. For over 50 years, thermoelectric generators (TEGs) have been employed
which are based upon the Seebeck effect,1 whereby
a potential difference is developed across a material
when its ends are held at different temperatures.
TEGs can be used to generate electrical power from
environments where there are temperature gradients,2 and can be utilised to convert waste heat
energy into usable electrical energy. The conversion
efficiency (g) is dependent on the temperature difference (between Tcold and Thot) and a dimensionless
term, ZT, defined as the figure of merit for the
material, given by Eqs. 1 and 2 below3:
p
1 ZT 1
(1)
g conversion g carnot p
1 ZT TTcold
hot
efficiency
efficiency
(Received July 15, 2014; accepted November 24, 2014;
published online December 18, 2014)
ZT
S2 r
T
j
(2)
1804
High Temperature Thermoelectric Properties of (1 x) SrTiO3 (x) La1/3NbO3 Ceramic Solid Solution
1805
Fig. 1. X-ray diffraction spectra for compositions in the Sr1xLax/3Ti1xNbxO3 solid solution.
Fig. 2. Backscattered electron SEM images of samples sintered in air: (a) STO, (b) L1, (c) L2, (d) L3. Note lower magnification used in (d), to
image the grain distribution and second phase.
1806
Fig. 3. (a) HAADF (Zone axis: h1 0 0i) image for L3 showing both A-site and B-site atomic columns. The variation in the optical intensity along
the rows indicates random distribution of Sr, La and vacancies in A-site and Ti and Nb in B-sites; (b) detailed intensity scan along A-sites;
(c) detailed intensity scan along B-sites.
High Temperature Thermoelectric Properties of (1 x) SrTiO3 (x) La1/3NbO3 Ceramic Solid Solution
1807
Fig. 4. Thermoelectric properties of samples in the (1 x) SrTiO3 (x) (La1/3NbO3) system sintered in air. Temperature dependence of (1 x)
SrTiO3 (x) (La1/3NbO3) is shown as (a) resistivity, (b) Seebeck coefficient, (c) thermal conductivity and (d) figure of merit, ZT. Compositions
STO, L1, L2, L3 corresponds to x = 0.0, 0.1, 0.2, 0.3, respectively.
centres for phonon propagation at high temperatures. Moreover, the clustering of vacancies and the
uneven distribution of vacancies or La/Nb ions
(Fig. 3a) creates disorder in the crystal structure,
assisting phonon scattering across a wider spectrum. The development of second phase segregation
in sample L3 could also make a significant contribution to reducing lattice thermal conductivity,
yielding the lowest thermal conductivity of 2.5 W/
mK in sample L3 at 1000 K. The mean free path for
phonon conduction in STO-based compositions is
reported to be in the range of 0.210 nm which
accounts for most of the heat conduction in bulk
samples.26,27 It should be possible to achieve further
reduction in the overall thermal conductivity by
nanostructuring the grain structure and by developing nanoinclusions at grain boundaries or nanoprecipitates within the grains.
From the thermoelectric parameters determined
at 5501050 K, the ZT values were determined; the
highest ZT value of 0.004 was obtained for L2
sample as seen in Fig. 4d. The value is very low due
to fact that no excess charge carrier had been added.
However, it defines the best composition in studied
solid solution to investigate further by controlled
nanostructuring and generating excess charge carriers within the material. For example, a pilot study
of L2 sintered in reducing conditions gave a ZT of
1808
REFERENCES
1. T.J. Seebeck, Ueber die magnetische Polarisation der Metalle und Erze durch Temperatur-Differnz. Berlin: gedruckt
in der Druckerei der Konigl. Akademie der Wissenschaften.
26 cm (1825).
2. T.M. Tritt, Annu. Rev. Mater. Res. 41, 433 (2011).
3. D.M. Rowe and C. Chemical Rubber, CRC Handbook of
Thermoelectrics [electronic resource]. Boca Raton, FL: CRC
Press, 1995. 1 online resource (701).
4. G.S. Nolas, in Materials Research Society Symposium Proceedings, November 2530, 2012, Boston, MA, ed. by
Y. Grin, D.C. Johnson, and A.J. Thompson (Warrendale:
Materials Research Society, 2013), pp. xiv, 245.
5. Y. Pei, et al., Energy Environ. Sci. 4, 2085 (2011).
6. T. Caillat, J.P. Fleurial, and A. Borshchevsky, J. Phys.
Chem. Solids 58, 1119 (1997).
7. G.S. Nolas, et al., J. Appl. Phys. 79, 4002 (1996).
8. J.W. Fergus, J. Eur. Ceram. Soc. 32, 525 (2012).
9. K. Koumoto, et al., J. Am. Ceram. Soc. 96, 1 (2013).
10. I. Terasaki, Phys. B 328, 63 (2003).
11. K. Koumoto, I. Terasaki, and R. Funahashi, MRS Bull. 31,
206 (2006).
12. K. Koumoto, et al., Annu. Rev. Mater. Res. 40, 363 (2010).
13. T. Okuda et al., Phys. Rev. B 63 (2001).
14. S. Ohta, et al., J. Appl. Phys. 97, 034106 (2005).
15. N. Wang, et al., J. Ceram. Soc. Jpn. 118, 1098 (2010).
16. O. Madelung, U. Rossler, and M. Schulz (ed.), SrTiO3 crystal
structure, lattice parameters.
17. A.N. Salak, et al., Appl. Phys. Lett. 93, 162903 (2008).
18. M.I. Mendelson, J. Am. Ceram. Soc. 52, 443 (1969).
19. W.J. Parker, et al., J. Appl. Phys. 32, 1679 (1961).
20. R.D. Cowan, J. Appl. Phys. 34, 926 (1963).
21. A.R. Denton and N.W. Ashcroft, Phys. Rev. A 43, 3161 (1991).
22. R.D. Shannon, Acta Crystallogr. Sect. A 32, 751 (1976).
23. W. Wunderlich, H. Ohta, and K. Koumoto, Phys. B 404, 2202
(2009).
24. R.R. Heikes and R.W. Ure, Thermoelectricity: Science and
Engineering (New York: Interscience Publishers, 1961),
p. 576.
25. C. Yu, et al., Appl. Phys. Lett. 92, 092118 (2008).
26. W. Yifeng, et al., Appl. Phys. Express 3, 031101 (2010).
27. G.G. Yadav, et al., Nanoscale 3, 4078 (2011).