Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Home Search Collections Journals About Contact us My IOPscience

Piezoelectric and mechanical properties of fatigue resistant, self-healing PZT–ionomer

composites

This content has been downloaded from IOPscience. Please scroll down to see the full text.

2014 Smart Mater. Struct. 23 055001

(http://iopscience.iop.org/0964-1726/23/5/055001)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 131.204.172.32
This content was downloaded on 31/08/2014 at 08:06

Please note that terms and conditions apply.


Smart Materials and Structures
Smart Mater. Struct. 23 (2014) 055001 (9pp) doi:10.1088/0964-1726/23/5/055001

Piezoelectric and mechanical properties of


fatigue resistant, self-healing PZT–ionomer
composites
N K James1 , U Lafont1 , S van der Zwaag1 and W A Groen1,2
1
Novel Aerospace Materials Group, Faculty of Aerospace Engineering, Delft University of Technology,
Kluyverweg 1, 2629 HS Delft, The Netherlands
2
Holst Centre, TNO, High Tech Campus 31, 5605 KN Eindhoven, The Netherlands

E-mail: n.kunnamkuzhakkaljames@tudelft.nl

Received 21 November 2013, revised 7 February 2014


Accepted for publication 7 February 2014
Published 18 March 2014

Abstract
Piezoelectric ceramic–polymer composites with 0–3 connectivity were fabricated using lead
zirconium titanate (PZT) powder dispersed in an ionomer (Zn ionomer) and its reference
ethylene methacrylic acid copolymer (EMAA) polymer matrix. The PZT–Zn ionomer and
PZT–EMAA composites were prepared by melt extrusion followed by hot pressing. The
effects of poling conditions such as temperature, time and electric field on the piezoelectric
properties of the composites were investigated. The experimentally observed piezoelectric
charge coefficient and dielectric constant of the composites were compared with theoretical
models. The results show that PZT–Zn ionomer composites have better piezoelectric
properties compared to PZT–EMAA composites. The static and fatigue properties of the
composites were investigated. The PZT–Zn ionomer composites were found to have excellent
fatigue resistance even at strain levels of 4%. Due to the self-healing capabilities of the
ionomer matrix, the loss of piezoelectric properties after high strain tensile cyclic loading
could be partially recovered by thermal healing.

Keywords: piezoelectric composite, ionomer, mechanical properties, self-healing


(Some figures may appear in colour only in the online journal)

1. Introduction are embedded in a polymeric matrix. These composites bring


together the required properties such as a high piezoelectric
Lead zirconium titanate (PZT) ceramics have good piezoelec- voltage constant, a low acoustic impedance and mechanical
tric and pyroelectric properties and are widely used in trans- flexibility [7, 8].
ducers, actuators and sensors. However, their poor mechanical The mechanical and electrical properties of the composite
properties, limited freedom in creating sensors with complex materials can be tailored by the selection of the constituents,
curved surfaces and the acoustic impedance mismatch between the volume fractions of constituent phases and the way in
ceramics and transmitting media limit their use in many which the individual phases are connected. Newnham et al
applications [1–6]. Moreover, piezoelectric ceramics have a described ten possible connectivity patterns between particles
relatively low hydrostatic piezoelectric coefficient and piezo- in diphasic composite systems [3]. The inter particle connectiv-
electric voltage coefficient (g33 ) because of their negative d31 ity has a large influence on the final electrical and mechanical
and high dielectric constant respectively. In order to overcome properties of such composites. Composites with 0–3 connec-
the aforementioned limitations, piezoelectric composite mate- tivity are obtained by distributing the ceramic particles in a
rials were developed in which piezoelectric ceramic particles polymer matrix in such a way that the particles are not in

0964-1726/14/055001+09$33.00 1 c 2014 IOP Publishing Ltd


Printed in the UK
Smart Mater. Struct. 23 (2014) 055001 N K James et al

contact with one another while the polymer phase is connected piezoelectric ceramic particles with a low dielectric constant.
in all three dimensions. The 0–3 composites are considered to Different approaches have been taken to increase the poling
be the least complicated in terms of manufacturing and ease of efficiency, for example the use of piezoelectrically active
molding in complex, predesigned shapes. Although there have polymers, the use of various conductive additives such as
been many reports about piezoelectric–polymer composites carbon black or graphite and the use of functionalized PZT
with 0–3 connectivity, only a handful of investigations have ceramics by a silane coupling agent to enhance the binding
been carried out on the effect of poling conditions on the piezo- between PZT and the polymer matrix [13, 19–21].
electric properties of the composites for polymer matrices with Yamada et al proposed a model to predict the piezoelectric
different electrical conductivities [9–15]. charge constant and dielectric constant of a composite by
In the present study an ionomer has been selected as the assuming a perfect degree of poling of individual ellipsoidal
polymer matrix because of its high flexibility, good mechanical piezoelectric particles which are dispersed in a continuous low
properties, excellent adhesion to the ceramic phase and self- dielectric polymer matrix [18]. The final equations are given
healing potential [16, 17]. The effective poling conditions for below.
nφ(εc − εp )
 
PZT–Zn ionomer composites were investigated and the results ε = εp 1 + (3)
were compared with the reference PZT–EMAA (ethylene nεp + (εc − εp )(1 − φ)
methacrylic acid copolymer) composites and PZT ceramics. nε
The experimentally observed dielectric and piezoelectric coef- d33 = φ d33c (4)
(nε + (εc − ε))
ficient are compared with Yamada’s model [18]. Furthermore,
where n is a parameter reflecting the aspect ratio of the ferro-
to investigate the suitability of potential applications of such
electric particle and d33 and d33c are the piezoelectric charge
composites, the tensile properties and high cycle fatigue of the
constant of the composite and ceramic phase respectively [19].
composites for large strain levels has been studied. Finally, for
The piezoelectric voltage constant of the composite, g33 , can
the first time ever it is demonstrated that for a well chosen
be calculated according to the following equation:
self-healing polymer matrix the loss of sensorial functionality
of the composites after high cyclic tensile fatigue can be d33
partially recovered by thermal healing. g33 = (5)
εo ε

2. Theory where d33 is the piezoelectric charge constant in pC N−1 and


ε0 is the permittivity of vacuum.
One of the major factors influencing the macroscopic piezo-
electric properties of a polymer–piezoelectric ceramic com- 3. Experimental details
posite is the degree of poling of the ceramic phase. A fine
grained piezoelectric particle in a polymeric matrix requires The piezoelectric ceramic powder used for manufacturing
high poling fields to achieve sufficient poling efficiency. The the composites was a soft doped Pb(Zr, Ti)O3 (PZT507)
electric field E 1 acting on an isolated spherical grain in a (Morgan Electro Ceramics, Ruabon, UK). In order to improve
polymer matrix is given by (equation (1)): its piezoelectric properties the as-received powder was cal-
cined at 1200 ◦ C for 2 h in a closed alumina crucible in
3εp
E1 = E0 (1) a Nabertherm furnace with a heating rate of 5 K min−1 .
εc + 3εp The calcined powder was ball milled using yttria stabilized
zirconia balls of 5 mm diameter in cyclohexane for 4 h.
where εc and εp are the dielectric constants of the spherical
Subsequently, the powder was dried at 150 ◦ C and stored in
piezoelectric grains and polymer matrix respectively, and E 0
a closed container to avoid moisture absorption. The average
is the applied electric field [12]. The effective electric field
particle size (d50 ) of the milled powder is 4.26 µm as measured
experienced by the ceramic phase is generally low due to the
by laser scattering with a Horiba LA950. The particle size
low dielectric constant of the polymer phase. While applying a
measured with SEM is about 3–5 µm, which is in good agree-
DC poling electric field to the sample for a time longer than the
ment particle size measured by laser scattering. The phase
sample relaxation time, the field distribution is also controlled
purity and compositional homogeneity are studied by x-ray
by σp /σc , the ratio of the electrical conductivity of the polymer
powder diffraction using a Bruker-AXS D5005 diffractometer
to that of the ceramics. During the poling process, the mobility
[11, 15]. The calcined powder crystallizes in a perovskite
of charges in the polymer matrix plays an important role. It
structure with the coexistence of rhombohedral and tetragonal
has been theoretically shown that the poling time should be
phases.
sufficiently longer than the relaxation time τ , which is defined
Zn-based ionomer (Surlyn 9320 DuPont) and ethylene
as [19]:
methacrylic acid (EMAA) copolymer were used as the poly-
φ3εp + (1 − φ)(εc + 2εp ) mer matrix materials. In EMAA-based ionomers, the ethylene
τ= (2) methacrylic acid copolymer (see figure 1(a)) is neutralized
φ3σp + (1 − φ)(σc + 2σp )
such that a certain percentage of metallic cations such as
where φ is the volume fraction of the ceramic particles. zinc (Zn), sodium (Na) or lithium (Li) are retained along the
The poling process can be accelerated and optimized by polymer back bone (figure 1(b)) [16]. Self-healing behavior
using a polymer matrix with high electrical conductivity and of ionomers has been well reported [16, 17, 22], in which the

2
Smart Mater. Struct. 23 (2014) 055001 N K James et al

(a) (b)

Figure 1. Molecular structure of (a) EMAA and (b) Zn ionomer.

self-healing phenomena was either initiated by high energy of both the as-produced composites and cyclically loaded
impact or a puncture test, such as bullet penetration or by composites were measured. The specimens were fatigue tested
more controlled quasi-static penetration experiments [23] or under different strain levels, by properly selecting the stress
by mild thermal treatments and looking at the healing of ratio defined as the ratio of the minimum stress to the maximum
surface scratches [24]. As shown elsewhere [17], the Zn-based stress. The shape of the loading waveform was sinusoidal and
ionomers show the highest degree of self-healing (i.e. the the employed frequency was restricted to 1 Hz to avoid sample
ability to close cracks or other local mechanical damage upon heating.
application of a modest temperature well below the melting Finally, to measure the functional healing, the degraded
point of the polymer). An additional desirable characteristic of tensile fatigue loaded composites were annealed at 70 ◦ C in a
the ionomeric polymer is the relaxation of the ionic species at conventional circulating hot air oven while the piezoelectric
elevated temperatures which increases the electrical conduc- properties were measured at intermediate times.
tivity significantly, which should be beneficial to the poling
efficiency.
4. Results and discussion
The composites were prepared by mixing a defined
ratio of polymeric matrix and PZT powder in a labscale 4.1. Microstructure
counter rotating twin screw extruder (DSM Xplore Research
Netherlands). The processing temperature was kept at 170 ◦ C The as-produced composites are in the form of opaque whitish
for both the PZT–Zn ionomer and the PZT–EMAA composites. flexible sheets of about 1 mm thickness. An SEM micrograph
The rotating force was set to 4.5 kN. In order to get a of the cross-sectional view of a 30 vol.% filled PZT–Zn
homogeneous distribution of PZT ceramic particles in the ionomer composite is shown in figure 2(a). The micrograph
polymer matrix, the residence time was kept for 5 min, which confirms (figure 2(b)) that fully dense PZT–Zn ionomer
ensured multiple passes of the material through the mixing composites with a spatially uniform distribution of the PZT
section of the extruder. The composites were finally extruded particles were created. The PZT particles are well distributed,
through a 2 mm diameter outlet. In order to get planar sheets show limited agglomeration (hence a 0–3 configuration) and
of composites, the extruded wire-shaped materials were hot adhere well to the polymer matrix which is a key requirement
pressed at temperature of 150 ◦ C with a pressure of 1 MPa for for optimal piezoelectric and mechanical properties. A higher
5 min, yielding flexible planar sheets with a thickness of about magnification SEM micrograph showing the primary particles
1 mm. Electrodes on the samples were made by sputtering inside the composite is given in figure 2(c).
of gold. The samples were poled in an organic oil (rape
seed oil) bath under different poling conditions (electric field 4.2. Poling behavior
strength, temperature and time). The dielectric constants of the
composites were measured using the parallel plate capacitor The effect of poling temperature on the d33 of 30 vol.% filled
method using an Agilent 4263B LCR meter at 1V AC signal PZT–Zn ionomer and PZT–EMAA composites is shown in
and 1 kHz. The d33 of the composites was measured using a figure 3 for a poling field of 15 kV mm−1 and a time of
Berlincourt type d33 meter (KCF technology PM3001). The 60 min. The results are also compared with those for bulk
static tensile properties of the composite were investigated PZT sintered ceramics. It is evident that for the PZT–polymer
using a Zwick Roell 20 kN tensile bench with a 1 kN load cell composites the poling temperature has a marked effect on
and a cross-head speed of 0.5 mm min−1 at room temperature. the d33 , especially around 50–80 ◦ C, whereas the d33 of
Rectangular specimens with dimensions 50 mm × 5 mm × the bulk PZT ceramics remains more or less constant. The
1 mm were tested. The average value over four samples are dielectric breakdown of the composites prevents higher poling
reported. Fatigue tests were conducted at room temperature temperatures being applied. It is clear that PZT–Zn ionomer
on an MTS 831 Elastomer test system. Rectangular specimens composites show significantly better piezoelectric properties
with dimensions 50 mm × 5 mm × 1 mm were prepared and than the corresponding PZT–EMAA composites. The increase
within these samples two circular regions were metalized and of d33 with poling temperature observed for the PZT–Zn
poled at 15 kV mm−1 prior to the fatigue test. The d33 values ionomer composite can be attributed to the relaxation of the

3
Smart Mater. Struct. 23 (2014) 055001 N K James et al

Figure 2. (a) Photograph of 30 vol.% filled ionomer composite, (b) SEM micrograph of cross-sectional view of 30 vol.% filled PZT–Zn
ionomer composite and (c) SEM micrograph showing the primary particles.

800 The effect of poling time on the d33 of 30 vol.% filled


PZT- Zn ionomer composite PZT–Zn ionomer and PZT–EMAA composites is shown in
780 PZT- EMAA composite figure 4 for a poling field of 15 kVmm−1 and at a poling
PZT ceramic
temperature of 60 ◦ C. It is clear that the d33 of the compos-
760
ites increases considerably with poling time during the first
740
60 min and reaches a stable value thereafter. This increase
and subsequent saturation is attributed to two separate effects.
d33 (pC.N )
-1

In a polymer–ceramic composite the effective electric field


5 experienced by the ceramic phase is controlled by the electrical
4 conductivity and dielectric constants of both the polymer and
3 the ceramic phase. The effective electric field is always much
2 lower than the actual applied electric field [22, 25]. This
1 results in a longer poling time requirement for composites
0 as compared to bulk PZT ceramics. In addition to this, while
20 30 40 50 60 70 80 the reversal of 180◦ ferroelectric domains takes place mainly at
the initial stage of poling, 90◦ domain rotation is accompanied
o
Poling temperature ( C)
by local stress and strain which requires a longer amount of
Figure 3. Variation of the piezoelectric charge coefficient (d33 ) as a time. Therefore, the degree of poling in composites can be
function of poling temperature with an electric field of 15 kV mm−1 improved by increasing the poling time, but after a poling time
for 1 h. exceeding 60 min the composite d33 reaches its maximum
value indicating the reorientation of ferroelectric domains
ion clusters of the polymer chain at elevated temperatures to be near to completion. It is observed that the d33 of the
which increases the electrical conductivity of the polymer PZT–Zn ionomer composites increases more rapidly than the
matrix, which in turn increases the effective electric field PZT–EMAA composites which again can be attributed to the
experienced by the ceramic phase. This promotes an easier Zn ion cluster relaxation and the higher dielectric constant of
polarization of the ceramic phase in the ionomer matrix. The the ionomer polymer matrix.
temperature at which the poling efficiency increases for the The dependence of the piezoelectric charge coefficient
ionomer (between 50 and 60 ◦ C) is in agreement with the on the poling electric field is shown in figure 5 for a poling
de-clustering temperature of this ionomer [16, 17]. temperature of 60 ◦ C and a time of 60 min. As expected, the

4
Smart Mater. Struct. 23 (2014) 055001 N K James et al

800
PZT ceramic
780 PZT- Zn ionomer composite
PZT- EMAA composite
760

740
d33 (pC.N )
-1

0
0 20 40 60 80 100 120 140 160 180 200
Poling time (min)

Figure 4. Variation of the piezoelectric charge coefficient as a


function of poling time with an electric field of 15 kV mm−1 and
poling temperature of 60 ◦ C.

770
PZT ceramic
765 PZT - Zn ionomer composite
PZT - EMAA composite
760

755
d33 (pC.N )
-1

750

5
4
3
2
1
0 Figure 6. (a) The dielectric constant and (b) loss tangent of 30 vol.%
4 6 8 10 12 14 16 18 20 22 PZT–EMAA and PZT–ionomer composites as a function of
-1
Poling field (kV.mm ) frequency and temperature (the number inside the brackets in the
legend indicates the temperature in ◦ C).
Figure 5. Variation of the piezoelectric charge coefficient (d33 ) as a
function of poling field. 30 vol.% PZT–EMAA and PZT–ionomer composites. For all
conditions the dielectric constant shows an almost frequency
piezoelectric charge coefficient of the composites increases independent behavior with a small decrease in the low fre-
with the poling field for both PZT–Zn ionomer and PZT– quency region. The loss tangent spectra are characterized by a
EMAA composites and reaches a constant level at about peak appearing in the low frequency region of PZT–ionomer
17 kV mm−1 . In PZT ceramics, the maximum is already composites especially at a temperature of 60 ◦ C (i.e. at the
observed at 4 kV mm−1 . The poling field is the driving force selected poling temperature). The strength and frequency of
for the reversal of domains. The higher the field, the faster and relaxation depend on the characteristics of molecular and dipo-
more complete the reorientation of the domains in the direction lar relaxation [26]. In PZT–ionomer composites, the relaxation
of the poling field. In a 0–3 polymer–ceramic composite, and mobility of ionic clusters is evident by the peak shift
a large part of the applied electric field during poling is towards the higher frequency side. The relatively high mobility
obscured by the polymer phase. According to equation (1), at higher temperatures enhances the transport properties which
the effective electric field experienced by the ceramic phase in turn increase the loss tangent of PZT–ionomer composites.
is of the order of 1.7–28 V mm−1 for the range of applied In figure 7 the measured dielectric constant and piezoelec-
electric fields of 1–15 kV mm−1 . Hence, a polymer–ceramic tric charge coefficient of the composites are compared to the
composite requires a high electric field for effective poling of values predicted using Yamada’s model. This model predicts
the piezoelectric phase. Based on these results, the optimum that these properties mainly depend on the volume fraction
poling conditions for the described composites are an electric of PZT and the aspect ratio of the particles. The dielectric
field of 17 kV mm−1 to be applied at a temperature of 60 ◦ C constant and piezoelectric charge coefficient increases signifi-
for 60 min. cantly with increasing PZT volume fraction. The theoretically
Figures 6(a) and (b) show the temperature and frequency predicted values match well with the experimentally observed
dependence of the dielectric constant and loss tangent of data for the composites. The d33 and dielectric constant of the

5
Smart Mater. Struct. 23 (2014) 055001 N K James et al

12 PZT- EMMA experimental 60


PZT- Zn ionomer experimental PZT- Zn Ionomer composite
PZT- Zn ionomer Yamada model PZT- EMAA composite
10 PZT- EMMA Yamada model 50
Dielectric constant

8 40

g33 (mV.m.N )
-1
6 30

4 20

2 10

0 0
0 10 20 30 0 10 20 30
Volume fraction of PZT loading (%) volume fraction of PZT loading (%)
(a)
Figure 8. Variation of piezoelectric voltage coefficient (g33 ) of the
composites as a function of PZT volume percentage.
and dynamic mechanical properties of such composites were
studied in more detail.

4.3.1. Tensile properties. Changes in Young’s modulus, elon-


gation and strength at break as a function of PZT loading are
shown in figures 9(a) and (b). The results indicate that as the
PZT filler loading content increases, the Young’s modulus of
the composite increases which is a common trend for any 0–3
polymer–ceramic composites. The 30 vol.% filled PZT–Zn
ionomer composite exhibits a relatively low Young’s modulus
of 90 MPa, a tensile strength of 5 MPa and a elongation
at break of 150%. Few 0–3 composites are reported with
a comparable low modulus and good piezoelectric voltage
coefficient [11, 25]. It has been reported that the filler particle
b size, morphology, volume fraction and adhesion between the
particles and polymer can have a detrimental effect on the
Figure 7. (a) The dielectric constant and (b) piezoelectric charge mechanical properties of the composite [26]. The good adhe-
coefficient of the composites as a function of PZT volume sion between PZT ceramic particles and the ionomer matrix
percentage.
leads to high strength and elongation at break. Both tensile
strength and elongation at break of the composite decreases as
composite follow the same trend as reported for other 0–3 hot
the PZT filler loading increases in the polymer matrix, which
pressed composites [8, 11, 18]. is a normal behavior for any polymer ceramic composite.
The variation of piezoelectric voltage coefficient of the
composites as a function of PZT volume fraction is shown in 4.3.2. Fatigue life. The effect of single and cyclic tensile
figure 8. The piezoelectric voltage coefficient of ionomer–PZT loading to specific strain values on the piezoelectric prop-
composites increases with increasing PZT volume fraction in erties of the 30 vol.% filled PZT–Zn ionomer composite
the composite. A maximum value for g33 of 52 mV m N−1 is shown in figures 10(a) and (b) respectively. It is clear
is obtained for a 30% volume fraction of PZT–Zn ionomer that a significant degradation of piezoelectric properties only
composite. Compared to the ionomer-based composites, the occurs upon straining the composites above 80%. Just before
voltage coefficient of EMAA–PZT composites increases lin- failure, at 125% of elongation, the d33 value has decreased
early up to 20 vol.% of PZT and reaches a maximum the plateau with 20% as compared to the initial d33 value for the un-
at around 18 mV m N−1 . The voltage sensitivity obtained for stressed composites. Figure 10(b) shows that no degradation
the PZT–Zn ionomer composites is reasonably good compared of piezoelectric properties is observed for strain levels of
to other 0–3 composites [15, 27, 28]. 4% up to 105 cycles. At higher strain levels, a decrease of
piezoelectric properties is observed. The d33 of the PZT–Zn
4.3. Mechanical properties ionomer composites was measured to be 5.2 pC N−1 before
the fatigue test, and it dropped by 57% after 103 cycles for a
The results described above indicate that 30 vol.% of PZT strain of 8%.
filled ionomer composites attain a reasonably good voltage At higher strain levels, and larger number of cycles a
coefficient at the optimized poling conditions. Hence, the static decrease in sensorial properties is observed. It is most likely

6
Smart Mater. Struct. 23 (2014) 055001 N K James et al

b
Figure 9. Variation of (a) Young’s modulus and (b) elongation and
strength at break as a function of PZT volume percentage in a Figure 10. Piezoelectric charge coefficient of the 30 vol.% filled
Zn-ionomer-based composite. PZT–Zn ionomer composite (a) as a function of the applied strain
during single pass loading and (b) after cyclic loading to three
different maximal strain values.
that such a drop in d33 is caused by delamination between
PZT particles and the polymer matrix as there are no intrinsic
changes in the properties of the constituent phases are to be
expected at room temperature testing. The assumed debonding
is also evident from a SEM micrographs (figure 12(a)) of the
fracture surface of the PZT–Zn ionomer composite.

4.3.3. Self-healing properties. The piezoelectric charge coef-


ficient recovery of the degraded composites which endured
tensile cyclic loading at 6 and 8% strain levels for 50 000 and
15 000 cycles, upon annealing at 70 ◦ C is shown in figure 11
as function of healing time. The results indicate that the 6%
strain level fatigued composites recovered their properties
much faster than the composites strained at 8%. This is due
to the fact that fewer and smaller voids are formed in the
composites strained at 6% than at 8%, which in turn will also
require a shorter time to be healed. A high magnification SEM
micrograph (figure 12(b)) of the composites healed at 70 ◦ C Figure 11. Self-healing of the PZT–Zn ionomer composite
piezoelectric charge coefficient at 70 ◦ C.
for 1 h reveals that the debonded interfaces areas healed,
contributing to the recovery of the piezoelectric properties.
The fact that the d33 properties were not fully restored to 5. Conclusion
their original properties suggest that for the loading condition
explored the healing was imperfect and not all interphases A highly flexible PZT–Zn ionomer piezoelectric composite has
were reformed. been developed which can be used for sensor or transducer

7
Smart Mater. Struct. 23 (2014) 055001 N K James et al

‘Smart systems based on integrated piezo’. The authors are


grateful to Morgan Electro Ceramics for providing the PZT507
powder used in this research.

References

[1] Jaffe B, Cook W R and Jaffe H 1971 Piezoelectric Ceramics


(New York/London: Academic)
[2] Gururaja T R 1994 Piezoelectrics for medical ultrasonic
imaging Am. Ceram. Soc. Bull. 73 50–5
[3] Newnham R E, Skinner D P and Cross L E 1978 Connectivity
and piezoelectric–pyroelectric composites Mater. Res. Bull.
13 525–36
[4] Fukada E 2000 History and recent progress in piezoelectric
polymers IEEE Trans. Ultrason. Ferroelectr. Freq. Control
47 1277–90
[5] Qiu X 2010 Patterned piezo-, pyro-, and ferroelectricity of
poled polymer electrets J. Appl. Phys. 108 011101
[6] Furukawa T, Ishida K and Fukada E 1979 Piezoelectric
properties in the composite J. Appl. Phys. 50 4904–12
[7] Dias C J and Das-Gupta D K 1996 Inorganic ceramic/polymer
ferroelectric composite electrets IEEE Trans. Dielectr.
Electr. Insul. 3 706–34
[8] Venkatragavaraj E, Satish B, Vinod P R and Vijaya M S 2001
Piezoelectric properties of ferroelectric PZT-polymer
composites J. Phys. D: Appl. Phys. 34 487–92
[9] Lam K H and Chan H L W 2005 Piezoelectric and
pyroelectric properties of 65PMN-35PT/P(VDF-TrFE) 0–3
composites Compos. Sci. Technol. 65 1107–11
[10] Wegener M and Arlt K 2008 PZT/P(VDF-HFP) 0–3
composites as solvent-cast films: preparation, structure and
piezoelectric properties J. Phys. D: Appl. Phys. 41 165409
[11] van den Ende D A, Groen W A and van der Zwaag S 2011
The effect of calcining temperature on the properties of 0–3
piezoelectric properties of PZT and a liquid crystalline
Figure 12. SEM micrographs of 8% strained and fatigued PZT–Zn
thermosetting polymer J. Electroceram. 27 13–9
ionomer composite (a) before and (b) after healing. [12] Gong G S, Safari A, Jang S J and Newnham R E 1986 Poling
of flexible piezoelectric composites Ferroelectr. Lett.
5 131–42
applications in high strain applications. The manufacturing [13] Lau S T, Kwok K W and Shin F G 2007 A poling study of
process of such composites is straightforward and can be lead zirconate titanate/polyeurathane 0–3 composites
easily up-scaled. At optimized poling conditions, the 30 vol.% J. Appl. Phys. 102 044104
filled PZT–Zn ionomer composites exhibit a d33 and g33 of [14] Lee M H, Halliyal A and Newnham R E 1988 Poling of
5.2 pC N−1 and 52 mV m N−1 respectively. The measured coprecipitated lead titanate-epoxy 0–3 piezoelectric
dielectric and piezoelectric properties are in good agreement composites Ferroelectrics 81 71–80
[15] James N K, van den Ende D A, Lafont U, van der Zwaag S
with theoretical models. The tensile fatigue loading has shown
and Groen W A 2013 Piezoelectric and mechanical
that PZT–Zn ionomer composites exhibit minimal degradation
properties of structured PZT-epoxy composites J. Mater.
in performance at lower strain levels. For an imposed maximal Res. 28 635–41
strain of 4% the PZT–Zn ionomer composites have a useful [16] Rees R W and Vaughan D J 1965 Physical structure of
life of more than 105 cycles. The self-healing experiments ionomers ACS Polymer Preprints 6 287–96
demonstrated that the degraded piezoelectric properties after [17] Tadano K, Hirasawa E, Yamamoto H and Yano S 1988
high strain tensile cyclic loading can be recovered to a Order–disorder transition of ionic clusters in ionomers
significant degree due to the thermal healing capability of J. Macromol. 22 226–33
Zn-based ionomer matrix. [18] Yamada T, Ueda T and Kitayama T 1982 Piezoelectricity of a
high content lead zirconate titanate/polymer composite
J. Appl. Phys. 53 4328–32
Acknowledgments [19] Wong C K, Wong Y W and Shin F G 2002 Effect of interfacial
charge on polarization switching of lead zirconate particles
This work was financially supported by the Smartmix funding in lead zirconate titanate/polyurethane composites
program (Grant No. SMVA06071), as part of the program J. Appl. Phys. 92 3974–8

8
Smart Mater. Struct. 23 (2014) 055001 N K James et al

[20] Sakamoto W K, De Souza E and Das-Gupta D K 2001 [24] van der Zwaag S 2007 Self Healing Materials: An Alternative
Electroactive properties of flexible piezoelectric composites Approach to 20 Centuries of Materials Science (Dordrecht:
Mater. Res. 4 201–4 Springer) ISBN 978-1-4020-6250-6 pp 95–114
[21] Capsal J F, Dantras E, Laffont L, Dandurand J and [25] Chure M C, Chen P C, Wu L, Chen B H and Wu K K 2011
Lacabanne C 2010 Nanotexture influence of BaTiO3 Influence of poling conditions on the characteristics of PZT
particles on piezoelectric behavior of PA 11/BaTiO3 ceramics Adv. Mater. Res. 284 1375–80
[26] Kremer F and Schonhals A 2003 Broadband Dielectric
nanocomposites J. Non-Cryst. Solids 356 629–34
Spectroscopy (Berlin: Springer)
[22] Mendiola J, Alemany C, Jimenez B and Maurer E 1985 Poling [27] Carlsson L, Crane R L and Uchino K 2000 Comprehensive
of (Pb, La)(Zr, Ti)O3 J. Mater. Sci. Lett. 4 1383–6 Composite Materials vol 5 (New York: Elsevier) pp 523–32
[23] Varley R J and van der Zwaag S 2010 Development of a [28] Tavman I H 1996 Thermal and mechanical properties of
quasi-static test method to ionomer Polym. Int. aluminum powder filled high density polyethylene
59 1031–8 composites J. Appl. Polym. Sci. 62 2161–7

You might also like