Hostname: page-component-77c89778f8-rkxrd Total loading time: 0 Render date: 2024-07-20T06:20:25.508Z Has data issue: false hasContentIssue false

Bulk response and grain boundary microelectrical activity of high TC BaTiO3–(Bi1/2K1/2)TiO3-based positive temperature coefficient of resistance ceramics

Published online by Cambridge University Press:  23 October 2013

Mohammad A. Zubair
Affiliation:
Department of Metallurgy and Ceramic Science, Graduate School of Science and Engineering, Tokyo Institute of Technology, Meguro, Tokyo 152-8552, Japan; and Materials Science Center, School of Materials, University of Manchester, Manchester M1 7HS, United Kingdom
Hiroaki Takeda*
Affiliation:
Department of Metallurgy and Ceramic Science, Graduate School of Science and Engineering, Tokyo Institute of Technology, Meguro, Tokyo 152-8552, Japan
Colin Leach
Affiliation:
Materials Science Center, School of Materials, University of Manchester, Manchester M1 7HS, United Kingdom
Robert Freer
Affiliation:
Materials Science Center, School of Materials, University of Manchester, Manchester M1 7HS, United Kingdom
Takuya Hoshina
Affiliation:
Department of Metallurgy and Ceramic Science, Graduate School of Science and Engineering, Tokyo Institute of Technology, Meguro, Tokyo 152-8552, Japan
Takaaki Tsurumi
Affiliation:
Department of Metallurgy and Ceramic Science, Graduate School of Science and Engineering, Tokyo Institute of Technology, Meguro, Tokyo 152-8552, Japan
*
a)Address all correspondence to this author. e-mail: htakeda@ceram.titech.ac.jp
Get access

Abstract

Lead-free positive temperature coefficient of resistance (PTC) thermistors were synthesized from (1 − x/100)BaTiO3–(x/100)(Bi1/2K1/2)TiO3-based solid solutions, using a conventional mixed-oxide fabrication route, and sintered in N2 followed by air annealing. A maximum TC of 205 °C was achieved for x = 20. An increase in x from 0 to 20 decreased the grain size by more than 92% and increased room temperature resistivity (ρRT) by 7 orders of magnitude. For x ≤ 10, PTC ratio (ρmaxmin) ≈ 104.5 and temperature coefficient of resistivity (α) > 10.3%/°C were achieved using Mn and Al2O3:SiO2:TiO2 (AST) additions. For x > 10, ρmaxmin > 103 and α > 8%/°C were only obtained in samples sintered in N2 without subsequent air annealing. Complex impedance analysis revealed three relaxation processes, attributed to a semiconducting grain core, a PTC active grain boundary interface, and a grain boundary insulating layer. Local electrical activity was investigated by hot-stage conductive mode microscopy. The existence of symmetrical grain boundary electron beam-induced current and β-conductivity contrast at the grain boundaries, consistent with the presence of an electron trapping two-dimensional grain boundary plane, compensated by positive space charge layers and a low conductivity vacancy-rich layer, was revealed for the first time within this system.

Type
Articles
Copyright
Copyright © Materials Research Society 2013 

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

REFERENCES

Heywang, W.: Resistivity anomaly in doped barium titanate. J. Am. Ceram. Soc. 47(10), 484 (1964).CrossRefGoogle Scholar
Jonker, G.H.: Some aspects of semiconducting barium titanate. Solid State Electron. 7, 895 (1964).CrossRefGoogle Scholar
Japan Electronic Manufactures Association Standard EMAS-8202, 1988.Google Scholar
Huybrechts, B., Ishikazi, K., and Takata, M.: Review: The positive temperature coefficient of resistivity. J. Mater. Sci. 30, 2463 (1995).CrossRefGoogle Scholar
Kuwabara, M. and Kumamoto, K.: Ptcr characteristics in barium titanate ceramics with Curie point between 60 and 360 degrees centigrade. J. Am. Ceram. Soc. 66, C-214 (1983).CrossRefGoogle Scholar
Directive 2002/95/EC of the European Parliament and of the Council. Restriction of the use of certain hazardous substances in electrical and electronic equipment. Off J L 037, 00190023 (2003).Google Scholar
Takeda, H., Shimada, T., Katsuyama, Y., and Shiosaki, T.: Fabrication and operation limit of lead-free PTCR ceramics using BaTiO3-(Bi1/2Na1/2)TiO3 system. J. Electroceram. 22, 263 (2009).CrossRefGoogle Scholar
Shimada, T., Touji, K., Katsuyama, Y., Takeda, H., and Shiosaki, T.: Lead free PTCR ceramics and its electrical properties. J. Eur. Ceram. Soc. 27, 3877 (2007).CrossRefGoogle Scholar
Xiang, P-H., Takeda, H., and Shiosaki, T.: High TC lead free BaTiO3-(Bi1/2Na1/2)TiO3 positive temperature coefficient of resistivity ceramics with electrically heterogeneous structure. Appl. Phys. Lett. 91, 162904 (2007).CrossRefGoogle Scholar
Ihring, H.: PTC effect in BaTiO3 as a function of doping with 3d Elements. J. Am. Ceram. Soc. 64, 617 (1981).CrossRefGoogle Scholar
Kuwabara, M., Morimo, K., and Matsunaga, T.: Single gain-boundary in PTC resistors. J. Am. Ceram. Soc. 79, 997 (1996).CrossRefGoogle Scholar
Seaton, J. and Leach, C.: Conductive mode imaging of thermistor grain boundaries. J. Eur. Ceram. Soc. 24, 1191 (2004).CrossRefGoogle Scholar
Hayashi, K., Yamamoto, T., and Sakuma, T.: Grain orientation dependence of the PTCR effect in niobium-doped barium titanate. J. Am. Ceram. Soc. 79, 1669 (1996).CrossRefGoogle Scholar
Russell, J.D. and Leach, C.: Problems associated with imaging resistive barriers in BaTiO3 ceramics using the SEM conductive mode. J. Eur. Ceram. Soc. 15, 617 (1995).CrossRefGoogle Scholar
Russell, J.D. and Leach, C.: β-Conductivity contrast at barium titanate thermistor grain boundaries. J. Eur. Ceram. Soc. 16, 10351039 (1996).CrossRefGoogle Scholar
Leach, C.: The effect of voltage bias on the EBIC contrast present at varistor grain boundaries. Interface Sci. 8, 375 (2000).CrossRefGoogle Scholar
Buhrer, C.F.: Some properties of bismuth perovskites. J. Chem. Phys. 36, 798 (1962).CrossRefGoogle Scholar
Nakahara, M. and Murakami, T.: Electronic states of Mn ions in Ba0.97Sr0.03TiO3 single crystals. J. Appl. Phys. 45, 3795 (1974).CrossRefGoogle Scholar
Illingsworth, J., Al-Allak, H.M., Brinkman, A.W., and Woods, J.: The influence of Mn on the grain-boundary potential barrier characteristics of donor-doped BaTiO3 ceramics. J. Appl. Phys. 67(4), 2088 (1989).CrossRefGoogle Scholar
Cheng, H-F.: Effect of sintering aids on the electrical properties of positive temperature coefficient of resistivity BaTiO3 ceramics. J. Appl. Phys. 66(3), 1382 (1989).CrossRefGoogle Scholar
Allak, H.M., Parry, T.V., Russell, G.J., and Woods, J.: Effect of aluminum on the electrical and mechanical properties of PTCR BaTiO3 ceramics as a function of the sintering temperature. J. Mater. Sci. 23, 10831089 (1988).CrossRefGoogle Scholar
Xiang, P-H., Takeda, H., and Shiosaki, T.: Characterization of manganese doped BaTiO3-(Bi1/2Na1/2)TiO3 positive temperature coefficient of resistivity ceramics using impedance spectroscopy. J. Appl. Phys. 103, 064102 (2008).CrossRefGoogle Scholar
Moriwaka, H.: First-principle calculation of formation energy of neutral point defects in perovskite type BaTiO3. Int. J. Quantum Chem. 99, 7556 (2004).Google Scholar
Chan, N-H. and Smyth, D.M.: Defect chemistry on donor-doped BaTiO3. J. Am. Ceram. Soc. 67(4), 285 (1984).CrossRefGoogle Scholar
Leach, C. and Zubair, M.A.: Microstructural and electrical property evolution in an acceptor-dopant free positive temperature coefficient thermistor. Mater. Sci. Semicond. Process. 15(1), 4751 (2012).CrossRefGoogle Scholar
Ting, C-J., Peng, C-J., Lu, H-Y., and Wu, S-T.: Lanthanum-magnesium and lanthanum-manganese donor-acceptor codoped semiconducting barium titanate. J. Am. Ceram. Soc. 73, 329 (1990).CrossRefGoogle Scholar
Sinclair, D.C. and West, A.R.: Impedance and modulus spectroscopy of semiconducting BaTiO3 showing positive temperature coefficient of resistance. J. Appl. Phys. 66, 3850 (1989).CrossRefGoogle Scholar
Zubair, M.A. and Leach, C.: Modeling the resistance: Temperature characteristics of a positive temperature coefficient thermistor, using the experimentally determined permittivity data. Appl. Phys. Lett. 91, 082105 (2007).CrossRefGoogle Scholar
Palm, J.: Local investigation of recombination at grain boundaries of silicon by grain boundary-electron beam induced current. J. Appl. Phys. 74, 1169 (1993).CrossRefGoogle Scholar
Zubair, M.A. and Leach, C.: Modeling the effect of SiO2 additions and cooling rate on the electrical behavior of donor-acceptor co-doped positive temperature coefficient thermistors. J. Appl. Phys. 103, 123713 (2008).CrossRefGoogle Scholar
Daniels, J. and Wernicke, R.: New aspects of an improved PTC model. Philips Res. Rep. 31, 594 (1976).Google Scholar
Leach, C. and Seaton, J.: Local variability in PTC thermistor grain boundary structures. Scanning 30, 339 (2008).CrossRefGoogle ScholarPubMed
Hou, J., Zhang, Z., Press, W., sitte, W., and Dehm, G.: Electrical properties and structure of grain boundaries in n-conducting BaTiO3 ceramics. J. Eur. Ceram. Soc. 31, 763 (2011).CrossRefGoogle Scholar
Zubair, M.A. and Leach, C.: The effect of SiO2 addition on the development of low-Σ grain boundaries in PTC thermistors. J. Eur. Ceram. Soc. 30, 107 (2010).CrossRefGoogle Scholar
Press, W. and sitte, W.: Modeling of transport properties of interfacially controlled electroceramics: Application to n-conducting barium titanate. J. Electroceram. 27, 83 (2011).CrossRefGoogle Scholar
Palm, J., Steinbach, D., and Alexander, H.: Local investigation of the electrical properties of the grain boundaries. Mater. Sci. Eng., B 24, 54 (1994).CrossRefGoogle Scholar