Hostname: page-component-7c8c6479df-p566r Total loading time: 0 Render date: 2024-03-27T20:21:41.427Z Has data issue: false hasContentIssue false

Piezoelectric and mechanical properties of structured PZT–epoxy composites

Published online by Cambridge University Press:  01 February 2013

Nijesh Kunnamkuzhakkal James*
Affiliation:
Novel Aerospace Materials Group, Faculty of Aerospace Engineering, Department of Structures and Materials, Delft University of Technology, 2629 HS Delft, The Netherlands
Daan van den Ende
Affiliation:
Novel Aerospace Materials Group, Faculty of Aerospace Engineering, Delft University of Technology, 2629 HS Delft, The Netherlands; and Holst Centre, TNO, 5605 KN Eindhoven, The Netherlands
Ugo Lafont
Affiliation:
Novel Aerospace Materials Group, Faculty of Aerospace Engineering, Department of Aerospace Materials and Structures, Delft University of Technology, 2629 HS Delft, The Netherlands
Sybrand van der Zwaag
Affiliation:
Novel Aerospace Materials Group, Faculty of Aerospace Engineering, Department of Aerospace Materials and Structures, Delft University of Technology, 2629 HS Delft, The Netherlands
Wilhelm A. Groen
Affiliation:
Novel Aerospace Materials Group, Faculty of Aerospace Engineering, Department of Aerospace Structures and Materials, Delft University of Technology, 2629 HS Delft, The Netherlands; and Holst Centre, TNO, 5605 KN Eindhoven, The Netherlands
*
a)Address all correspondence to this author. e-mail: n.kunnamkuzhakkaljames@tudelft.nl
Get access

Abstract

Structured lead zirconium titanate (PZT)–epoxy composites are prepared by dielectrophoresis. The piezoelectric and dielectric properties of the composites as a function of PZT volume fraction are investigated and compared with the corresponding unstructured composites. The effect of poling voltage on piezoelectric properties of the composites is studied for various volume fractions of PZT composites. The experimentally observed piezoelectric and dielectric properties have been compared with theoretical models. Dielectrophoretically structured composites exhibit higher piezoelectric voltage coefficients compared to 0–3 composites. Structured composites with 0.1 volume fraction of PZT have the highest piezoelectric voltage coefficient. The flexural strength and bending modulus of the structured and random composites were analyzed using three-point bending tests.

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

Gururaja, T.R., Schulze, W.A., Cross, L.E., Newnham, R.E., Auld, B.A., and Wang, Y.J.: Piezoelectric composite materials for ultrasonic transducer applications. Part I: Resonant modes of vibration of PZT rod-polymer composites. IEEE Trans. Son. Ultrason. 32, 481 (1985).CrossRefGoogle Scholar
Akdogan, E.K., Allahverdi, M., and Safari, A.: Piezoelectric composites for sensor and actuator applications. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 52, 746 (2005).CrossRefGoogle ScholarPubMed
Newnham, R.E., Bowen, L.J., Klicker, K.A., and Cross, L.E.: Composite piezoelectric transducers. J. Mater. Eng. 2, 93 (1980).Google Scholar
Newnham, R.E., Skinner, D.P., and Cross, L.E.: Connectivity and piezoelectric and pyroelectric composite. Mater. Res. Bull. 13, 525 (1978).CrossRefGoogle Scholar
McNulty, T.F., Janas, V.F., and Safari, A.: Novel processing of 1-3 ceramic/polymer composites for transducer applications. J. Am. Ceram. Soc. 78, 2913 (1995).CrossRefGoogle Scholar
Taunaumang, H., Guy, I.L., and Chan, H.L.W.: Electromechanical properties of 1-3 piezoelectric ceramic/polymer composites. J. Appl. Phys. 76, 484 (1994).CrossRefGoogle Scholar
Panda, R.K., Janas, V.F., and Safari, A.: Fabrication and properties of fine scale 1-3 piezoelectric ceramic/polymer composites for ultrasonic transducer applications. In Proceedings of the 10th IEEE International Symposium Applications of Ferroelectrics, 1996; J.W. Silwa, Jr., S. Ayter, J.P. Mohr, ed., IEEE, Piscataway, NJ, 1996. p. 551.Google Scholar
Dias, C.J. and Das Gupta, K.D.: Inorganic ceramic/polymer ferroelectric composite electrets. IEEE Trans. Dielectr. Electr. Insul. 39, 706 (1996).CrossRefGoogle Scholar
Klicker, K.A., Biggers, J.V., and Newnham, R.E.: Piezoelectric composites of PZT and epoxy for hydrostatic transducer applications. J. Am. Ceram. Soc. 64, 5 (1981).CrossRefGoogle Scholar
Silwa, J.W., Ayter, J.S., and Mohr, III: Method for making piezoelectric composite. U.S. Patent No. 5 239 736, 1993.Google Scholar
Bast, U., Kaarmann, H., Lubitz, K., Vogt, M., Wersing, W., and Carmer, D.: Composite ultrasound transducer for manufacturing a structured component therefor of piezoelectric ceramic. U.S. Patent No. 5 164 920, 1992.Google Scholar
Livneh, S., Janas, V., and Safari, A.: Development of fine scale PZT ceramic fiber/polymer shell composite transducer. J. Am. Ceram. Soc. 78, 1900 (1995).CrossRefGoogle Scholar
Safari, A., Allahverdi, M., and Akdogan, E.K.: Solid freeform fabrication of piezoelectric sensors and actuators. J. Mater. Sci. 41, 177 (2006).CrossRefGoogle Scholar
Jans, V.F. and Safari, A.: Overview of fine-scale piezoelectric ceramic/polymer composite processing. J. Am. Ceram. Soc. 78, 2945 (1995).CrossRefGoogle Scholar
Bowen, C.P., Shrout, T.R., Newnham, R.E., and Randall, C.A.: Tunable electric field processing of composite materials. J. Intell. Mater. Syst. Struct. 6, 159 (1995).CrossRefGoogle Scholar
van den Ende, D.A., Bory, B.F., Groen, W.A., and van der Zwaag, S.: Improving the d33 and g33 properties of 0–3 piezoelectric composites by dielectrophoresis. J. Appl. Phys. 107, 024107 (2010).CrossRefGoogle Scholar
Randall, C.A., Miller, D.V., Adair, J.H., and Bhalla, A.S.: Processing of electroceramic-polymer composites using the electrorheological effect. J. Mater. Res. 8, 899 (1993).CrossRefGoogle Scholar
Wilson, S.A., Maistros, G.M., and Whatmore, R.W.: Structure modification of 0-3 piezoelectric ceramic/polymer composites through dielectrophoresis. J. Phys. D: Appl. Phys. 38, 175 (2005).CrossRefGoogle Scholar
19.Park, C. and Robertson, R.E.: Aligned microstructure of some particulate polymer composites obtained with an electric field. J. Mater. Sci. 33, 3541 (1998).CrossRefGoogle Scholar
Yamada, T., Ueda, T., and Kitayama, T.: Piezoelectricity of a high-content lead zirconate titanate/polymer composite. J. Appl. Phys. 53, 4328 (1982).CrossRefGoogle Scholar
21.Furukawa, T., Fujino, K., and Fukada, E.: Electromechanical properties in the composites of epoxy-resin and PZT ceramics. Jpn. J. Appl. Phys. 15, 2119 (1976).CrossRefGoogle Scholar
Furukawa, T., Ishida, T.K., and Fukada, E.: Piezoelectric properties in the composite systems of polymers and PZT ceramics. J. Appl. Phys. 50, 4904 (1979).CrossRefGoogle Scholar
23.Bowen, C.P., Newnham, R.E., and Randall, C.A.: Dielectric properties of dielectrophoretically assembled particulate-polymer composites. J. Mater. Res. 13, 205 (1998).CrossRefGoogle Scholar
Zakari, T., Laurent, J.P., and Vaculin, M.: Theoretical evidence for ‘Lichtenecker’s mixture formulae’ based on the effective medium theory. J. Phys. D: Appl. Phys. 31, 1589 (1998).CrossRefGoogle Scholar
Lichtenecker, K.: Dielectric constant of natural and synthetic mixtures. Phys. Z. 27, 115 (1926).Google Scholar
Dowling, N.E.: Mechanical Behaviour of Materials: Engineering Materials for Deformation, Fracture, and Fatigue (Prentice-Hall Inc., Englewood Cliffs, NJ, 1993).Google Scholar
van den Ende, D.A., De Almeida, P., and van der Zwaag, S.: Piezoelectric and mechanical properties of novel composites of PZT and a liquid crystalline thermosetting resin. J. Mater. Sci. 42, 6417 (2007).CrossRefGoogle Scholar
Babu, I., van den Ende, D.A., and De With, G.: Processing and characterization of piezoelectric 0-3 PZT/LCT/PA composites. J. Phys. D: Appl. Phys. 43, 425402 (2010).CrossRefGoogle Scholar
Hiremath, B.V., Kingon, A., and Biggers, J.V.: Reaction sequence in the formation of lead zirconate-lead titanate solid solution: Role of raw materials. J. Am. Ceram. Soc. 66(11), 790793 (1983).CrossRefGoogle Scholar
van den Ende, D.A., Groen, W.A., and van der Zwaag, S.: The effect of calcining temperature on the properties of 0-3 piezoelectric composites of PZT and a liquid crystalline thermosetting polymer. J. Electroceram. 27, 13 (2011).CrossRefGoogle Scholar
Lee, M.H., Halliyal, A., and Newnham, R.E.: Poling of coprecipitated lead titanate-epoxy 0–3 composites. J. Am. Ceram. Soc. 72, 986 (1989).CrossRefGoogle Scholar
Rasband, W.S.: IMAGEJ (U.S. National Institutes of Health, Bethesda, MD, 2007).Google Scholar
Rashid, E.S.A., Akil, H.M., Ariffin, K., and Choong Kooi, C.: The mechanical and thermal properties of alumina filled epoxy. In Proceedings of 31st International Conference on Electronics Manufacturing and Technology, (IEEE Publications, New Brunswick, NJ, 2006); p. 282.Google Scholar