Hostname: page-component-8448b6f56d-sxzjt Total loading time: 0 Render date: 2024-04-20T01:08:29.675Z Has data issue: false hasContentIssue false

Phase transitions, relaxor behavior, and electrical properties in (1−x)(Bi0.5Na0.5)TiO3x(K0.5Na0.5)NbO3 lead-free piezoceramics

Published online by Cambridge University Press:  17 October 2012

Jigong Hao
Affiliation:
Functional Materials Research Laboratory, Tongji University, Shanghai 200092, China
Wangfeng Bai
Affiliation:
Functional Materials Research Laboratory, Tongji University, Shanghai 200092, China
Wei Li
Affiliation:
Functional Materials Research Laboratory, Tongji University, Shanghai 200092, China
Bo Shen
Affiliation:
Functional Materials Research Laboratory, Tongji University, Shanghai 200092, China
Jiwei Zhai*
Affiliation:
Functional Materials Research Laboratory, Tongji University, Shanghai 200092, China
*
a)Address all correspondence to this author. e-mail: apzhai@tongji.edu.cn
Get access

Abstract

Phase structures and electrical properties of lead-free piezoelectric (1−x)(Bi0.5Na0.5)TiO3x(K0.5Na0.5)NbO3 (BNT–xKNN) ceramics with 0.08 ≤ x ≤ 0.19 were systematically investigated. Results showed that a phase transition from a tetragonal to a pseudocubic phase occurred in this system, as KNN content increases. The addition of KNN shifted both the depolarization temperature Td and rhombohedral–tetragonal phase transition temperature TR-T to lower temperatures and tended to enhance the relaxor behavior of the ceramics, which was well explained by the microdomain–macrodomain transition theory with calculating criterion K. At x = 0.10–0.11, Td reached room temperature (RT), which accordingly induced an enhancement of the unipolar strain that peaks at a value of 0.22% was obtained. Furthermore, as the compositions (x = 0.12–0.15) have Td below RT, samples exhibited high electrostrictive response with large electrostrictive coefficient Q33 (0.017–0.019 m4/C2) and good thermostability comparable with that of traditional Pb-based electrostrictors.

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

Takenaka, T. and Nagata, H.: Current status and prospects of lead-free piezoelectric ceramics. J. Eur. Ceram. Soc. 25, 2693 (2005).CrossRefGoogle Scholar
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
Sasaki, A., Chiba, T., Mamiya, Y., and Otsuki, E.: Dielectric and piezoelectric properties of (Bi0.5Na0.5)TiO3–(Bi0.5K0.5)TiO3 systems. Jpn. J. Appl. Phys. 38, 5564 (1999).CrossRefGoogle Scholar
Liu, L.Y., Zhu, M.K., Hou, Y.D., Yan, H., and Liu, R.P.: Abnormal piezoelectric and dielectric behavior of 0.92Na0.5Bi0.5TiO3-0.08BaTiO3 induced by La doping. J. Mater. Res. 22, 1188 (2007).CrossRefGoogle Scholar
Kounga, A.B., Zhang, S.T., Jo, W., Granzow, T., and Rödel, J.: Morphotropic phase boundary in (1-x)Bi0.5Na0.5TiO3xK0.5Na0.5NbO3 lead-free piezoceramics. Appl. Phys. Lett. 92, 222902 (2008).CrossRefGoogle Scholar
Hiruma, Y., Yoshii, K., Nagata, H., and Takenaka, T.: Phase transition temperature and electrical properties of (Bi1/2Na1/2)TiO3-(Bi1/2A1/2)TiO3 (A = Li and K) lead-free ferroelectric ceramics. J. Appl. Phys. 103, 084121084127 (2008).CrossRefGoogle Scholar
Shieh, J., Wu, K.C., and Chen, C.S.: Switching characteristics of MPB compositions of (Bi0.5Na0.5)TiO3-BaTiO3-(Bi0.5K0.5)TiO3 lead-free ferroelectric ceramics. Acta Mater. 55, 30813087 (2007).CrossRefGoogle Scholar
Zhang, S., Shrout, T.R., Nagata, H., Hiruma, Y., and Takenaka, T.: Piezoelectric properties in (K0.5Bi0.5)TiO3-(Na0.5Bi0.5)TiO3-BaTiO3 lead-free ceramics. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 54, 910 (2007).CrossRefGoogle ScholarPubMed
Shrout, T.R. and Zhang, S.J.: Lead-free piezoelectric ceramics: Alternatives for PZT? J. Electroceram. 19, 111 (2007).CrossRefGoogle Scholar
Hiruma, Y., Nagata, H., and Takenaka, T.: Phase diagrams and electrical properties of (Bi1/2Na1/2)TiO3-based solid solutions. J. Appl. Phys. 104, 124106 (2008).CrossRefGoogle Scholar
Lee, S-H., Yoon, C-B., Lee, S-M., Kim, H-E., and Lee, K-W.: Piezoelectric properties of lead-free (Na0.5Bi0.5)TiO3–(Na0.5K0.5)NbO3–BaTiO3 ceramics. J. Mater. Res. 23, 115 (2008).CrossRefGoogle Scholar
Wang, K., Hussain, A., Jo, W., and Rödel, J.: Temperature-dependent properties of (Bi1/2Na1/2)TiO3–(Bi1/2K1/2)TiO3–SrTiO3 lead-free piezoceramics. J. Am. Ceram. Soc. 95, 2241 (2012).CrossRefGoogle Scholar
Guo, Y.P., Gu, M.Y., Luo, H.S., Liu, Y., and Withers, R.L.: Composition-induced antiferroelectric phase and giant strain in lead-free (Nay, Biz)Ti1−xO3(1−x)-xBaTiO3 ceramics. Phys. Rev. B 83, 054118 (2011).CrossRefGoogle Scholar
Hiruma, Y., Nagata, H., and Takenaka, T.: Detection of morphotropic phase boundary of (Bi1/2Na1/2)TiO3–Ba(Al1/2Sb1/2)O3 solid-solution ceramics. Appl. Phys. Lett. 95, 052903 (2009).CrossRefGoogle Scholar
Zhang, S-T., Kounga, A.B., Aulbach, E., Ehrenberg, H., and Rödel, J.: Giant strain in lead-free piezoceramics. 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
Wang, F.F., Xu, M., Tang, Y.X., Wang, T., Shi, W.Z., and Leung, C.M.: Large strain response in the ternary Bi0.5Na0.5TiO3–BaTiO3–SrTiO3 solid solutions. J. Am. Ceram. Soc. 95, 1955 (2012).CrossRefGoogle Scholar
Jarupoom, P., Patterson, E., Gibbons, B., Rujijanagul, G., Yimnirun, R., and Cann, D.: Lead-free ternary perovskite compounds with large electromechanical strains. Appl. Phys. Lett. 99, 152901 (2011).CrossRefGoogle Scholar
Ullah, A., Ahn, C.W., Hussain, A., Lee, S.Y., and Kim, I.W.: Phase transition, electrical properties, and temperature-insensitive large strain in BiAlO3-modified Bi0.5(Na0.75K0.25)0.5TiO3 lead-free piezoelectric ceramics. J. Am. Ceram. Soc. 94, 3915 (2011).CrossRefGoogle Scholar
Jo, W., Granzow, T., Aulbach, E., Rödel, J., and Damjanovic, D.: Origin of the large strain response in (K0.5Na0.5)NbO3-modified (Bi0.5Na0.5)TiO3–BaTiO3 lead-free piezoceramics. J. Appl. Phys. 105, 094102 (2009).CrossRefGoogle Scholar
Jo, W., Schaab, S., Sapper, E., Schmitt, L.A., Kleebe, H-J., Bell, A.J., and Rödel, 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
Bobnar, V., Malič, B., Holc, J., Kosec, M., Steinhausen, R., and Beige, H.: Electrostrictive effect in lead-free relaxor K0.5Na0.5NbO3–SrTiO3 ceramic system. J. Appl. Phys. 98, 024113 (2005).CrossRefGoogle Scholar
Ang, C. and Yu, Z.: High, purely electrostrictive strain in lead-free dielectrics. Adv. Mater. 18, 103 (2006).CrossRefGoogle Scholar
Zhang, S.T., Kounga, A.B., Jo, W., Jamin, C., Seifert, K., Granzow, T., Rödel, J., and Damjanovic, D.: High-strain lead-free antiferroelectric electrostrictors. Adv. Mater. 21, 4716 (2009).CrossRefGoogle Scholar
Zhang, S.T., Yan, F., Yang, B., and Cao, W.: Phase diagram and electrostrictive properties of Bi0.5Na0.5TiO3–BaTiO3–K0.5Na0.5NbO3 ceramics. Appl. Phys. Lett. 97, 122901 (2010).CrossRefGoogle Scholar
Ngoc, V.D., Han, T.H-S., Yoo, C-H., Lee, J-S., Jo, W., and Rödel, J.: Lead-free electrostrictive bismuth perovskite ceramics with thermally stable field-induced strains. Mater. Lett. 65, 2607 (2011).Google Scholar
Li, J.M., Wang, F.F., Qin, X.M., Xu, M., and Shi, W.Z.: Large electrostrictive strain in lead-free Bi0.5Na0.5TiO3–BaTiO3–KNbO3 ceramics. Appl. Phys. A 104, 117 (2011).CrossRefGoogle Scholar
Haertling, G.H.: Ferroelectric ceramics: History and technology. J. Am. Ceram. Soc. 82, 797 (1999).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 and temperature. Powder Diffr. 17, 301 (2002).CrossRefGoogle Scholar
Ito, K., Tezuka, K., and Hinatsu, Y.: Preparation, magnetic susceptibility, and specific heat on interlanthanide perovskites ABO3 (A = La-Nd, B = Dy-Lu). J. Solid State Chem. 157, 173 (2001).CrossRefGoogle Scholar
Lee, W-C., Huang, C-Y., Tsao, L-K., and Wu, Y-C.: Chemical composition and tolerance factor at the morphotropic phase boundary in (Bi0.5Na0.5)TiO3-based piezoelectric ceramics. J. Eur. Ceram. Soc. 29, 1443 (2009).CrossRefGoogle Scholar
Sapper, E., Schaab, S., Jo, W., Granzow, T., and Rödel, J.: Influence of electric fields on the depolarization temperature of Mn-doped (1-x)(Bi1/2Na1/2)TiO3-xBaTiO3. J. Appl. Phys. 111, 014105 (2012).CrossRefGoogle Scholar
Hiruma, Y., Nagata, H., and Takenaka, T.: Thermal depoling process and piezoelectric properties of bismuth sodium titanate ceramics. J. Appl. Phys. 105, 084112 (2009).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
Tan, X., Aulbach, E., Jo, W., Granzow, T., Kling, J., Marsilius, M., Kleebe, H.J., and Rödel, J.: Effect of uniaxial stress on ferroelectric behavior of (Bi1/2Na1/2)TiO3-based lead-free piezoelectric ceramics. J. Appl. Phys. 106, 044107 (2009).CrossRefGoogle Scholar
Dittmer, R., Jo, W., Damjanovic, D., and Rödel, J.: Lead-free high-temperature dielectrics with wide operational range. J. Appl. Phys. 109, 034107 (2011).CrossRefGoogle Scholar
Hinterstein, M., Knapp, M., Hölzel, M., Jo, W., Cervellino, A., and Ehrenberg, H.: Field-induced phase transition in Bi1/2Na1/2TiO3-based lead-free piezoelectric ceramics. J. Appl. Crystallogr. 43, 1314 (2010).CrossRefGoogle Scholar
Webber, K.G., Zhang, Y., Jo, W., Daniels, J.E., and Rödel, J.: High temperature stress-induced “double loop-like” phase transitions in Bi-based perovskites. J. Appl. Phys. 108, 014101 (2010).CrossRefGoogle Scholar
Laoratanakul, P., Yimnirun, R., and Wongsaenmai, S.: Phase formation and dielectric properties of bismuth sodium titanate potassium sodium niobate ceramics. Curr. Appl. Phys. 11, S161 (2011).CrossRefGoogle Scholar
Patterson, E.A., Cann, D.P., Pokorny, J., and Reaney, I.M.: Electromechanical strain in Bi(Zn1/2Ti1/2)O3–(Bi1/2Na1/2)TiO3–(Bi1/2K1/2)TiO3 solid solutions. J. Appl. Phys. 111, 094105 (2012).CrossRefGoogle Scholar
Zuo, R.Z., Fang, X.S., and Ye, C.: Phase structures and electrical properties of new lead-free (Na0.5K0.5)NbO3–(Bi0.5Na0.5)TiO3 ceramics. Appl. Phys. Lett. 90, 092904 (2007).CrossRefGoogle Scholar
Uchino, K. and Nomura, S.: Critical exponents of the dielectric constants in diffused phase transition crystals. Ferroelectr. Lett. Sect. 44, 55 (1982).CrossRefGoogle Scholar
Zhang, D.J. and Yao, X.: Dynamics on microdomain–macrodomain transition of relaxor ferroelectrics. Acta Phys. Chim. Sin. 20, 712 (2004).Google Scholar
Gao, F., Liu, L.L., Xu, B., Hu, G.X., Cao, X., Hong, R.Z., and Tian, C.S.: Texture development and dielectric relaxor behavior of 0.80Na0.5Bi0.5TiO3–0.20K0.5Bi0.5TiO3 ceramics templated by plate-like NaNbO3 particles. J. Eur. Ceram. Soc. 31, 2987 (2011).CrossRefGoogle Scholar
Hirum, Y., Nagata, H., and Takenaka, T.: Formation of morphotropic phase boundary and electrical properties of (Bi1/2Na1/2)TiO3–Ba(Al1/2Nb1/2)O3 solid solution ceramics. Jpn. J. Appl. Phys. 48, 09KC08 (2009).Google Scholar
Petzelt, J., Kamba, S., Fabry, J., Noujni, D., Porokhonskyy, V., Pashkin, A., Franke, I., Roleder, K., Suchanicz, J., Klein, R., and Kugel, G.E.: Infrared, Raman and high-frequency dielectric spectroscopy and the phase transition in Na1/2Bi1/2TiO3. J. Phys. Condens. Matter 16, 2719 (2004).CrossRefGoogle Scholar
Zhang, S.T., Kounga, A.B., Aulbach, E., and Deng, Y.: Temperature-dependent electrical properties of 0.94Bi0.5Na0.5TiO3–0.06BaTiO3 ceramics. J. Am. Ceram. Soc. 91, 3950 (2008).CrossRefGoogle Scholar