Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-16T15:16:29.544Z Has data issue: false hasContentIssue false

Structure and temperature-dependent phase transitions of lead-free Bi1/2Na1/2TiO3–Bi1/2K1/2TiO3–K0.5Na0.5NbO3 piezoceramics

Published online by Cambridge University Press:  09 July 2012

Eva-Maria Anton*
Affiliation:
Institute of Materials Science, Technische Universität Darmstadt, 64287 Darmstadt, Germany
Ljubomira Ana Schmitt
Affiliation:
Institute of Applied Geosciences, Technische Universität Darmstadt, 64287 Darmstadt, Germany
Manuel Hinterstein
Affiliation:
Institut für Werkstoffwissenschaft, Technische Universität Dresden, 01069 Dresden, Germany
Joe Trodahl
Affiliation:
MacDiarmid Institute of Advanced Materials and Nanotechnology, Victoria University, Wellington, New Zealand
Ben Kowalski
Affiliation:
Department of Materials Science & Engineering, University of Florida, Gainesville, Florida 32611
Wook Jo
Affiliation:
Institute of Materials Science, Technische Universität Darmstadt, 64287 Darmstadt, Germany
Hans-Joachim Kleebe
Affiliation:
Institute of Applied Geosciences, Technische Universität Darmstadt, 64287 Darmstadt, Germany
Jürgen Rödel
Affiliation:
Institute of Materials Science, Technische Universität Darmstadt, 64287 Darmstadt, Germany
Jacob L. Jones
Affiliation:
Department of Materials Science & Engineering, University of Florida, Gainesville, Florida 32611
*
a)Address all correspondence to this author. e-mail: anton@ceramics.tu-darmstadt.de
Get access

Abstract

Structure and phase transitions of (1 − y)((1 − x)Bi1/2Na1/2TiO3xBi1/2K1/2TiO3)–yK0.5Na0.5NbO3 (x; y) piezoceramics (0.1 ≤ x ≤ 0.4; 0 ≤ y ≤ 0.05) were investigated by transmission electron microscopy, neutron diffraction, temperature-dependent x-ray diffraction, and Raman spectroscopy. The local crystallographic structure at room temperature (RT) does not change by adding K0.5Na0.5NbO3 to Bi1/2Na1/2TiO3xBi1/2K1/2TiO3 for x = 0.2 and 0.4. The average crystal structure and microstructure on the other hand develop from mainly long-range polar order with ferroelectric domains to short-range order with polar nanoregions displaying a more pronounced relaxor character. The (0.1; 0) and (0.1; 0.02) compositions exhibit monoclinic Cc space group symmetry, which transform into Cc + P4bm at 185 and 130 °C, respectively. This high temperature phase is stable at RT for the morphotropic phase boundary compositions of (0.1; 0.05) and all compositions with x = 0.2. For the compositions of (0.1; 0) and (0.1; 0.02), local structural changes on heating are evidenced by Raman; for all other compositions, changes in the long-range average crystal structure were observed.

Type
Articles
Copyright
Copyright © Materials Research Society 2012

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

Rödel, J., Jo, W., Seifert, K.T.P., Anton, E-M., Granzow, T., and Damjanovic, D.: Perspective on the development of lead-free piezoceramics. J. Am. Ceram. Soc. 92, 1153 (2009).CrossRefGoogle Scholar
Takenaka, T., Nagata, H., and Hiruma, Y.: Current developments and prospective of lead-free piezoelectric ceramics. Jpn. J. Appl. Phys. 47, 3787 (2008).CrossRefGoogle Scholar
Aksel, E. and Jones, J.L.: Advances in lead-free piezoelectric materials for sensors and actuators. Sensors 10, 1935 (2010).CrossRefGoogle ScholarPubMed
Jaffe, B., Cook, W.R., and Jaffe, H., editors: Piezoelectric Ceramics (Academic Press, London, 1971).Google Scholar
Kounga, A.B., Zhang, S-T., Jo, W., Granzow, T., and Rödel, J.: Morphotropic phase boundary in (1-x)Bi0.5Na0.5TiO3-xK0.5Na0.5NbO3 lead-free piezoceramics. Appl. Phys. Lett. 92, 222902 (2008).CrossRefGoogle Scholar
Takenaka, T., Maruyama, K., and Sakata, K.: (Bi1/2Na1/2)TiO3-BaTiO3 system for lead-free piezoelectric ceramics. Jpn. J. Appl. Phys., Part 1 30, 2236 (1991).CrossRefGoogle Scholar
Elkechai, O., Manier, M., and Mercurio, J.P.: Na0.5Bi0.5TiO3-K0.5Bi0.5TiO3 (NBT-KBT) system: A structural and electrical study. Phys. Status Solidi A 157, 499 (1996).CrossRefGoogle Scholar
Sasaki, A., Chiba, T., Mamiya, Y., and Otsuki, E.: Dielectric and piezoelectric properties of (Bi1/2Na1/2)TiO3-(Bi0.5K0.5)TiO3 systems. Jpn. J. Appl. Phys., Part 1 38, 5564 (1999).CrossRefGoogle Scholar
Nagata, H., Yoshida, M., Makiuchi, Y., and Takenaka, T.: Large piezoelectric constant and high Curie temperature of lead-free piezoelectric ceramic ternary system based on bismuth sodium titanate-bismuth potassium titanate-barium titanate near the morphotropic phase boundary. Jpn. J. Appl. Phys., Part 1 42, 7401 (2003).CrossRefGoogle Scholar
Zhang, S-T., Kounga, A.B., Aulbach, E., Ehrenberg, H., and Rödel, J.: Giant strain in lead-free piezoceramics Bi0.5Na0.5TiO3-BaTiO3-K0.5Na0.5NbO3 system. Appl. Phys. Lett. 91, 112906 (2007).CrossRefGoogle Scholar
Seifert, K.T.P., Jo, W., and Rödel, J.: Temperature-insensitive large strain of (Bi1/2Na1/2)TiO3-(Bi1/2K1/2)TiO3-(K0.5Na0.5)NbO3 lead-free piezoceramics. J. Am. Ceram. Soc. 93, 1392 (2010).CrossRefGoogle Scholar
Singh, A. and Chatterjee, R.: Structural, electrical, and strain properties of stoichiometric 1-x - y(Bi0.5Na0.5)TiO3 - x(Bi0.5K0.5TiO3) - y(Na0.5K0.5)NbO3 solid solutions. J. Appl. Phys. 109, 024105 (2011).CrossRefGoogle Scholar
Anton, E-M., Jo, W., Trodahl, J., Damjanovic, D., and Rödel, J.: Effect of K0.5Na0.5NbO3 on properties at and off the morphotropic phase boundary in Bi1/2Na1/2TiO3-Bi1/2K1/2TiO3 ceramics. Jpn. J. Appl. Phys. 50, 055802 (2011).CrossRefGoogle Scholar
Jones, G.O., Kreisel, J., and Thomas, P.A.: A structural study of the (Na1-xKx)0.5Bi0.5TiO3 perovskite series as a function of substitution (x) and temperature. Powder Diffr. 17, 301 (2002).CrossRefGoogle Scholar
Glazer, A.M.: Classification of tilted octahedra in perovskites. Acta Crystallogr., Sect. B: Struct. Sci. 28, 3384 (1972).CrossRefGoogle Scholar
Zhao, W., Zhou, H.P., Yan, Y.K., and Liu, D.: Morphotropic phase boundary study of the BNT-BKT lead-free piezoelectric ceramics. Key Eng. Mater. 368372, 1908 (2008).CrossRefGoogle Scholar
Yang, Z., Liu, B., Wei, L., and Hou, Y.: Structure and electrical properties of (1-x)Bi0.5Na0.5TiO3-xBi0.5K0.5TiO3 ceramics near morphotropic phase boundary. Mater. Res. Bull. 43, 81 (2008).CrossRefGoogle Scholar
Otonicar, M., Skapin, S.D., Spreitzer, M., and Suvorov, D.: Compositional range and electrical properties of the morphotropic phase boundary in the Na0.5Bi0.5TiO3-K0.5Bi0.5TiO3 system. J. Eur. Ceram. Soc. 30, 971 (2010).CrossRefGoogle Scholar
Tai, C.W., Choy, S.H., and Chan, H.L.W.: Ferroelectric domain morphology evolution and octahedral tilting in lead-free (Bi1/2Na1/2)TiO3-(Bi1/2K1/2)TiO3-(Bi1/2Li1/2)TiO3-BaTiO3 ceramics at different temperatures. J. Am. Ceram. Soc. 91, 3335 (2008).CrossRefGoogle Scholar
Tai, C.W. and Lereah, Y.: Nanoscale oxygen octahedral tilting in 0.90(Bi1/2Na1/2)TiO3-0.05(Bi1/2K1/2)TiO3-0.05BaTiO3 lead-free perovskite piezoelectric ceramics. Appl. Phys. Lett. 95, 062901 (2009).CrossRefGoogle Scholar
Otonicar, M., Skapin, S.D., and Jancar, B.: TEM analyses of the local crystal and domain structures in (Na1-xKx)0.5Bi0.5TiO3 perovskite ceramics. IEEE Trans. Ultrason., Ferroelectr. Freq. Control 58, 1928 (2011).CrossRefGoogle ScholarPubMed
Woodward, D.I. and Reaney, I.M.: Electron diffraction of tilted perovskites. Acta Crystallogr., Sect. B: Struct. Sci. B61, 387 (2005).CrossRefGoogle Scholar
Gorfman, S. and Thomas, P.A.: Evidence for a non-rhombohedral average structure in the lead-free piezoelectric material Na0.5Bi0.5TiO3. J. Appl. Crystallogr. 43, 1409 (2010).CrossRefGoogle Scholar
Aksel, E., Forrester, J.S., Jones, J.L., Thomas, P.A., Page, K., and Suchomel, M.R.: Monoclinic crystal structure of polycrystalline Na0.5Bi0.5TiO3. Appl. Phys. Lett. 98, 152901 (2011).CrossRefGoogle Scholar
Levin, I., Reaney, I.M., Anton, E.-M., Jo, W., Rödel, J., Pokorny, J., Schmitt, L.A., Kleebe, H.-J., Hinterstein, M., Trodahl, J., and Jones, J.L.: Local Structure, Pseudo-Symmetry, and Phase Transitions in Na1/2Bi1/2TiO3-K1/2Bi1/2TiO3 Ceramics. Phys. Rev. B (2012, submitted).Google Scholar
Yao, Z.H., Liu, H.X., Chen, L., and Cao, M.H.: Morphotropic phase boundary and piezoelectric properties of (Bi1/2Na1/2)1-x(Bi1/2K1/2)xTiO3-0.03(Na0.5K0.5)NbO3 ferroelectric ceramics. Mater. Lett. 63, 547 (2009).CrossRefGoogle Scholar
Jo, W., Daniels, J.E., Jones, J.L., Tan, X., Thomas, P.A., Damjanovic, D., and Rödel, J.: Evolving morphotropic phase boundary in lead-free (Bi1/2Na1/2)TiO3–BaTiO3 piezoceramics. J. Appl. Phys. 109, 014110 (2011).CrossRefGoogle Scholar
Wylie-van Eerd, B., Damjanovic, D., Klein, N., Setter, N., and Trodahl, J.: Structural complexity of (Na0.5Bi0.5)TiO3-BaTiO3 as revealed by Raman spectroscopy. Phys. Rev. B 82, 104112 (2010).CrossRefGoogle Scholar
Miehe, G.: Program for Interpreting Electron Diffraction Patterns (PIEP). Version 7.12 (Institute for Materials Science, Technische Universität Darmstadt, Germany, 2002).Google Scholar
Roisnel, T. and Rodriguez-Carvajal, J.: WinPLOTR: A windows tool for powder diffraction pattern analysis. Mater. Sci. Forum 378381, 118 (2001).CrossRefGoogle Scholar
Fousek, J. and Janovec, V.: The orientation of domain walls in twinned ferroelectric crystals. Phys. Rev. B 40, 135 (1969).Google Scholar
Dai, X.H., Xu, Z., Li, J.F., and Viehland, D.: Effects of lanthanum modification on rhombohedral Pb(Zr1-xTix)O3 ceramics .1. Transformation from normal to relaxor ferroelectric behaviors. J. Mater. Res. 11, 618 (1996).CrossRefGoogle Scholar
Honjo, G., Kodera, S., and Kitamura, N.: Diffuse streak diffraction patterns from single crystals. I. General discussion and aspects of electron diffraction diffuse streak patterns. J. Phys. Soc. Jpn. 19, 351 (1964).CrossRefGoogle Scholar
Welberry, T.R.: Diffuse X-Ray Scattering and Models of Disorder (Oxford University Press, New York, 2004).Google Scholar
Daniels, J.E., Jo, W., Rodel, J., Rytz, D., and Donner, W.: Structural origins of relaxor behavior in a 0.96(Bi1/2Na1/2)TiO3-0.04BaTiO3 single crystal under electric field. Appl. Phys. Lett. 98, 252904 (2011).CrossRefGoogle Scholar
Jeong, I., Park, C.Y., Kim, D.J., Kim, S-h., Moon, B.K., Kim, I.W., and Ahn, C.W.: Neutron total scattering studies on A-site disorder in lead-free ferroelectric Bi0.5(Na1–xKx)0.5TiO3. Z. Kristallogr. 226, 150 (2011).CrossRefGoogle Scholar
Aksel, E., Forrester, J.S., Kowalski, B., Jones, J.L., and Thomas, P.A.: Phase transition sequence in sodium bismuth titanate observed using high-resolution x-ray diffraction. Appl. Phys. Lett. 99, 222901 (2011).CrossRefGoogle Scholar
Xie, H.D., Jin, L., Shen, D.Z., Wang, X.Q., and Shen, G.Q.: Morphotropic phase boundary, segregation effect and crystal growth in the NBT-KBT system. J. Cryst. Growth 311, 3626 (2009).CrossRefGoogle Scholar
Aksel, E., Forrester, J.S., Kowalski, B., Deluca, M., Damjanovic, D., and Jones, J.L.: Structure and properties of Fe-modified Na0.5Bi0.5TiO3 at ambient and elevated temperature. Phys. Rev. B 85, (2012).CrossRefGoogle Scholar
Anton, E-M., Jo, W., Damjanovic, D., and Rödel, J.: Determination of depolarization temperature of (Bi1/2Na1/2)TiO3-based lead-free piezoceramics. J. Appl. Phys. 110, 094108 (2011).CrossRefGoogle Scholar
Davies, M., Aksel, E., and Jones, J.L.: Enhanced high-temperature piezoelectric coefficients and thermal stability of Fe- and Mn-substituted Na0.5Bi0.5TiO3 ceramics. J. Am. Ceram. Soc. 94, 1314 (2011).CrossRefGoogle Scholar
Frantti, J., Ivanov, S., Eriksson, S., Rundlöf, H., Lantto, V., Lappalainen, J., and Kakihana, M.: Phase transitions of Pb(ZrxTi1-x)O3 ceramics. Phys. Rev. B 66, 064108 (2002).CrossRefGoogle Scholar
Hinterstein, M., Knapp, M., Holzel, M., Jo, W., Cervellino, A., Ehrenberg, H., and Fuess, H.: Field-induced phase transition in Bi1/2Na1/2TiO3-based lead-free piezoelectric ceramics. J. Appl. Crystallogr. 43, 1314 (2010).CrossRefGoogle Scholar
Schmitt, L.A., Hinterstein, M., Kleebe, H.J., and Fuess, H.: Comparative study of two lead-free piezoceramics using diffraction techniques. J. Appl. Crystallogr. 43, 805 (2010).CrossRefGoogle Scholar
Peng, J. and Bursill, L.A.: Polar and chemical domain structures of lead scandium tantalate (PST). Mod. Phys. Lett. B 7, 609 (1993).CrossRefGoogle Scholar
Zhou, D.H., Hoatson, G.L., Vold, R.L., and Fayon, F.: Local structure in perovskite relaxor ferroelectrics by 207Pb NMR. Phys. Rev. B 69, 134104 (2004).CrossRefGoogle Scholar
Maier, B.J., Angel, R.J., Marshall, W.G., Mihailova, B., Paulmann, C., Engel, J.M., Gospodinov, M., Welsch, A.M., Petrova, D., and Bismayer, U.: Octahedral tilting in Pb-based relaxor ferroelectrics at high pressure. Acta Crystallogr., Sect. B: Struct. Sci. 66, 280 (2010).CrossRefGoogle ScholarPubMed
Jones, G.O. and Thomas, P.A.: Investigation of the structure and phase transitions in the novel A-site substituted distorted perovskite compound Na1/2Bi1/2TiO3. Acta Crystallogr., Sect. B: Struct. Sci. 58, 168 (2002).CrossRefGoogle Scholar
Kreisel, J., Bouvier, P., Dkhil, B., Thomas, P.A., Glazer, A.M., Welberry, T.R., Chaabane, B., and Mezouar, M.: High-pressure x-ray scattering of oxides with a nanoscale local structure: Application to Na1/2Bi1/2TiO3. Phys. Rev. B 68, 014113 (2003).CrossRefGoogle Scholar
Said, S. and Mercurio, J.P.: Relaxor behaviour of low lead and lead free ferroelectric ceramics of the Na1/2Bi1/2TiO3-PbTiO3 and Na1/2Bi1/2TiO3-K0.5Bi0.5TiO3 systems. J. Eur. Ceram. Soc. 21, 1333 (2001).CrossRefGoogle Scholar
Jo, W., Schaab, S., Sapper, E., Schmitt, L.A., Kleebe, H-J., Bell, A.J., and Rodel, J.: On the phase identity and its thermal evolution of lead free (Bi1/2Na1/2)TiO3-6 mol% BaTiO3. J. Appl. Phys. 110, 074106 (2011).CrossRefGoogle Scholar
Supplementary material: File

Anton Supplementary Material

Tables

Download Anton Supplementary Material(File)
File 724.5 KB