You are on page 1of 7

Materials Research Bulletin 48 (2013) 4395–4401

Contents lists available at SciVerse ScienceDirect

Materials Research Bulletin


journal homepage: www.elsevier.com/locate/matresbu

Short communication

Structure and ferroelectric studies of (Ba0.85Ca0.15)(Ti0.9Zr0.1)O3


piezoelectric ceramics
Venkata Ramana E. a,*, A. Mahajan b, M.P.F. Graça a, S.K. Mendiratta a, J.M. Monteiro a,
M.A. Valente a
a
I3N-Aveiro, Department of Physics, University of Aveiro, Aveiro 3810 193, Portugal
b
Department of Materials and Ceramics Engineering, University of Aveiro, Aveiro 3810 193, Portugal

A R T I C L E I N F O A B S T R A C T

Article history: We have synthesized and studied the structural and ferroelectric properties of lead-free 0.5(Ba0.7Ca0.3)-
Received 21 December 2012 TiO3–0.5Ba(Zr0.2Ti0.8)O3 ceramics in the temperature region of its ferroelectric transition. The
Received in revised form 26 April 2013 synthesized material showed high dielectric constant, low loss and good pyroelectric figure of merit.
Accepted 27 May 2013
From the temperature dependent X-ray diffraction measurements, we determined the tricritical point to
Available online 10 June 2013
be in the temperature range of 303–400 K. The dielectric measurements indicate a diffuse ferroelectric
phase transition (DPT) around 360 K in agreement with the X-ray measurements. We studied the
Keywords:
evolution of Raman spectra with temperature to understand the nature of phase transition in BaTiO3
A. Ceramics
C. Atomic force microscopy
(BTO) and the BCTZO. The results indicates that the transition of ferroelectric–paraelectric state is not
C. X-ray diffraction sharp as in the case of BTO and the polar state persists through the paraelectric state. In general, our
D. Crystal structure study indicates that there are ferroelectric domains of nanometer size beyond the commonly defined
D. Ferroelectricity transition temperature. The observation of local piezoelectric hysteresis loop indicated the existence of
intrinsic ferroelectric property of the ceramic at the nanoscale. The ceramics exhibited electric field
tunable dielectric properties with a tunability of 82% at an applied DC field of 40 kV cm1, low dielectric
loss of 0.001 and room temperature pyroelectric coefficient of 6  108 C cm2 K1 and the detectivity of
1.9  108 C cm1 J1; larger than those reported for other BaTiO3-based materials. Overall, our results
indicate that BCTZO ceramics with coexistence of rhombohedral–tetragonal phases is a promising
candidate for lead-free ferroelectric applications.
ß 2013 Elsevier Ltd. All rights reserved.

1. Introduction with MPB exhibited significantly improved piezoelectric proper-


ties, albeit inferior to the lead-based ones. Recently, Liu et al. [7]
For the past few decades, Pb-based perovskite ferroelectrics reported large piezoelectric properties in 0.5[Ba(Zr0.2Ti0.8)O3]–
such as Pb(Zr,Ti)O3 (PZT) and Pb(MgNb)O3 (PMN) have been 0.5[(Ba0.7Ca0.3)O3] (BZT–BCT) ceramic with a d33  600 pC/N,
studied extensively as potential technological materials in view of higher than in PZT, and with a tricritical point (TCP) near room
their excellent piezoelectric properties [1–3]. However, in view of temperature. Enhancement of the piezoelectric properties in BZT–
Pb toxicity the search has been started to find suitable lead-free BCT has been achieved due to the existence of MPB near room
materials that have piezoelectric properties comparable to temperature, with a triple point of paraelectric cubic (C),
archtypical Pb compounds, PZT and PMN. In this context, ferroelectric rhombohedral (R) and tetragonal (T) phases. This
Na0.5Bi0.5TiO3 (NBT) and its solid solutions with BaTiO3 [4], triple point leads to an anisotropically flattened energy profile so
K0.5Bi0.5TiO3 [5], (K0.5,Na0.5)NbO3 [6] have been studied. Strong that the polarization can easily be rotated between h1 1 1iR and
piezoelectricity of PZT and PMN materials is believed to result from h0 0 1iT states by external stress or electric field [6].
the coexistence of two phases and the composition driven phase By means of transmission electron microscopy combined with
transition between the two ferroelectric phases (morphotropic the electron diffraction studies, Gao et al. [8], showed that the
phase boundary, MPB). On the other hand, lead-free compounds strong piezoelectric character near MPB is due to easy polarization
rotation between the coexisting nano-scale tetragonal and
rhombohedral domains. In a recent study, Xue et al. [9] reported
* Corresponding author. Tel.: +351 911825261.
enhanced electromechanical coupling factor, KP = 65%, piezoelec-
E-mail addresses: venkataramanaesk@rediffmail.com, tric voltage coefficient g33 = 15.3  103 Vm/N and elastic constant
ramana_evr@rediffmail.com (V.R. E.). SE33 ¼ 19:7  1012 m2 =N in a BCT–BZT system; the values are

0025-5408/$ – see front matter ß 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.materresbull.2013.05.108
4396 V.R. E. et al. / Materials Research Bulletin 48 (2013) 4395–4401

superior to that reported for the PZT. Yao et al. [10] reported the (Agilent 4294A). DC electric field dependent dielectric measure-
enhanced pyroelectric properties in this ceramic with a room ments were carried at 10 kHz using the Keithley 2410 high voltage
temperature pyroelectric coefficient of 5.84  104 C m2 K1. source. Microwave dielectric measurements were carried (reflec-
Piorra et al. [11] have successfully grown polycrystalline BCT– tance method) at room temperature using the HP Network
BZT thin films by pulsed laser deposition and achieved a room analyzer (8753D) with Agilent 85070E dielectric probe, in the
temperature dielectric constant of 1010 and clamped piezoelectric frequency range 500 MHz to 6 GHz. Pyroelectric measurements
response of 80 pm/V similar to the results of PZT films. were carried out by the Byer–Roundy method by measuring the
In the present work, we synthetized and studied the depolarization current in Kiethley 617 programmable electrome-
characteristics of this MPB material by means of temperature ter from 200 to 400 K at a heating rate (dT/dt) of 4 K/min [12].
dependent X-ray diffraction and Raman spectroscopy, piezo- Piezoresponse measurements were carried out on the thermally
response force microscopy along with the dielectric, ferroelectric etched samples (at 1300 8C) using a commercial PFM setup
and pyroelectric properties. Our aim is primarily to verify whether (Multimode, NanoScope IIIA, Veeco). Piezoresponse was recorded
temperature dependent Raman measurements show coexisting with a voltage of amplitude (Vac) 15 V at 160 kHz with a Pt coated
ferroelectric phases. Si tip-cantilever (spring constant of 5.7 N/m).

2. Experimental 3. Results and discussion

Polycrystalline ceramics of (Ba0.85Ca0.15)(Ti0.9Zr0.1)O3 (BCTZO) Fig. 1 shows the room temperature as well as the temperature
were synthesized by the conventional solid state reaction method. dependent X-ray diffraction patterns of the BCTZO ceramics
High purity (99.9%) raw chemicals, BaCO3, CaCO3, TiO2 and ZrO2 sintered at 1480 8C. The ceramic was formed with a pure
were mixed in stoichiometric ratio, and milled for 12 h in ethanol perovskite phase without any trace of impurities. All the diffraction
media. The resulting slurry was dried and calcined twice at 1350 8C peaks, at room temperature, can be indexed based on the standard
for 6 h with intermediate grindings. The powder obtained was X-ray pattern of polycrystalline tetragonal BaTiO3 with P4mm
again milled for 24 h and pressed into cylindrical discs with space group (JCPDS card no. 05-0626). The characteristic tetrago-
diameter of 10 and 25 mm. Final sintering was carried out in the nal reflection (1 1 0) was observed at 2u = 31.38; peak splitting was
temperature range 1450–1500 8C for 3 h. observed around 2u = 458 which is assigned to (0 0 2)/(2 0 0)
The crystal structure of the specimens was analyzed using the splitting of the tetragonal features and is related to the MPB
X-ray diffraction (XRD). Prior to the measurements, ceramic discs (rhombohedral/tetragonal) composition of 0.5BCT–0.5BZT solid
were crushed into powders and annealed at 400 8C for 1 h to solutions [7]. Fig. 1b shows this tetragonal peak splitting more
release strains. u–2u scans in the range 10–808 were done using the clearly. For comparison, the room temperature XRD of BaTiO3
Philips diffractometer with Cu Ka radiation from room tempera- (BTO) is also shown. The room temperature lattice parameters,
ture to 473 K. Surface morphology was evaluated using the calculated using the standard procedure, were found to be
scanning electron microscopy (SEM, Hitachi S-4100). Raman a = 4.0055 Å, c = 4.0165 Å and volume = 64.44 Å, which are slightly
spectroscopic studies were carried out in a microscope coupled larger than that of pure BTO (a = 3.991 Å, c = 4.032 Å) synthesized
to a spectrometer SP300I (Acton Research) and a sensitive cooled under the same conditions. Partial substitution of Ca2+ with ionic
CCD detector. Laser of wavelength 532 nm was used and the radius (IR) = 1.0 Å for Ba2+ (IR = 1.35 Å) and Zr4+ (IR = 0.72 Å) for
spectra were recorded in the temperature range 296–500 K using a Ti4+ (IR = 0.68 Å), respectively are expected to be primarily
specially designed furnace (Linkham, UK). For electrical measure- responsible for the altered unit cell dimensions. The tetragonal
ments gold electrodes were sputtered on polished ceramic strain (c/a) at room temperature is 1.003 which is in good
specimens. Electrical poling was done in the range RT-313 K in a agreement with the results of Liu and Ren [7]. With increasing
silicone oil bath under the d.c. electric field of 40–60 kV cm1. temperature, the peak corresponding to the (1 1 0) reflection shifts
Piezoelectric charge coefficient (d33) was measured using the to a lower 2u value indicating volume expansion at higher
piezometer (Piezotest-PM100). Polarization–electric field hyster- temperatures. An important observation from the figure is that the
esis loops were measured at different temperatures using a (0 0 2)/(2 0 0) peak merges into a single peak around 360 K. This
ferroelectric loop tracer (Radient Technologies). Dielectric mea- feature is similar to that observed by Wu et al. [13]. Above 380 K,
surements were carried out in the frequency range 0.1–500 kHz in the XRD pattern has the features of a cubic BTO phase.
the temperature range 200–400 K using the impedance analyzer Temperature dependence of lattice parameters plotted in Fig. 2a

Fig. 1. (a) XRD patterns of BCTZO measured at a temperature in the range 303–463 K and (b) enlarged data in the 2u range 44.5–468 (0 0 2/2 0 0 plane). The room temperature
tetragonal splitting of the BTO is also shown.
V.R. E. et al. / Materials Research Bulletin 48 (2013) 4395–4401 4397

Fig. 3. Temperature dependence of (a) dielectric permittivity and (b) dielectric loss
for BCTZO. Inset: log((1/e)  (1/emax)) vs. log(T  TC).

results of Shi et al. [14]. Dielectric loss, shown in Fig. 3b, also
Fig. 2. (a) Temperature dependent unit cell parameters for BCTZO and (b) surface exhibits a maximum at 360 K corresponding to the phase
morphology. transition. From Fig. 3a it can be seen that the phase transition
in BCTZO is not classical but represents a diffused nature. Such a
shows that the lattice parameter a gradually increases with transition can be represented by a modified Curie-Weiss law
temperature while c contracts merging with a in the range 360– expressed as ð1=eÞ  ð1=emax Þ ¼ ðT  T m Þg =C where Tm is the
380 K. This temperature is considered as the tetragonal $ cubic temperature corresponding to the dielectric maxima and g is
phase transition temperature. the fitting constant. The limiting values g = 1 and g = 2 reduce the
Fig. 2b shows the surface morphology of the BCTZO ceramics equation to Curie–Weiss law, valid for the case of normal
sintered at 1480 8C. A dense microstructure with clear grain to ferroelectrics and to the quadratic, valid for the relaxor ferroelec-
grain contact can be seen from the SEM image. The average grain tric, respectively. The plot derived from logð1=eÞ  ð1=emax Þ vs. log
size estimated from the intercept line method is 3 mm. The grain (T  TC) for T > TC at frequencies 0.1, 1 and 10 kHz exhibits a linear
size distribution determined from the atomic force microscopy relationship (inset of Fig. 3a) confirming a power law behaviour.
(AFM), Fig. 8, is in good agreement with this value. The measured The values of g = 1.7, 1.5 and 1.1 for 0.1, 1 and 10 kHz, respectively,
density of the sintered ceramic was 5.4 g cm3, which is 94% of the are indicative of diffuse phase transition (DPT) observed in BCTZO
theoretical density. and may be attributed to the presence of heterovalent ions at A and
Fig. 3a shows the temperature dependence of real part of B-sites of the perovskite structure.
dielectric permittivity for the BCTZO ceramic at 0.1, 1, 5, 10, 50, 100 The broadening of the phase transition is a common phenome-
and 500 kHz. The ceramic exhibits two dielectric anomalies: a non in solid solutions and other disordered structures [15,16]. The
broad hump around 310 K and a dielectric maximum around broad maximum observed in the present system is a result of
360 K. The observation of hump like behaviour is due to the overlap of several ferroelectric and non-ferroelectric regions. The
coexistence of two phases, rhombohedral (R) as well as tetragonal ferroelectric–paraelectric phase transition in this system is not
(T) at room temperature and its polymorphic phase transition from abrupt but they are gradual diffuse transitions occurring over a
R–T. The dielectric maximum corresponding to the tetragonal– temperature range called Curie range. It is in this temperature
cubic (T–C) phase transition (TC) is within the range of temperature range that the defects or impurity-defect complexes are still
obtained from the XRD results (see Fig. 2a) and agrees well with frozen. This behaviour is attributed to the structural disorder and
recent other reports on the 50BCT–50BZT ceramics [10,13]. The the compositional fluctuations in the crystalline systems [17].
present ceramics possess low dielectric loss of 0.001 and high Fig. 4a shows the dependence of dielectric permittivity on the
dielectric constant of 3152 (at 100 kHz) at room temperature, applied electric field at 10 kHz for BCTZO and BTO ceramics. We
respectively. The values are comparable to those obtained by observed a decreasing trend of er with the increase in electric field
4398 V.R. E. et al. / Materials Research Bulletin 48 (2013) 4395–4401

Fig. 5. Room temperature Raman spectra of BCTZO and BTO compositions. The
positions of Raman modes are superimposed onto the spectra.

is in good agreement with that (570) measured using the resonant


cavity method by Li et al. [22]. With the increasing frequency
dielectric constant of both the ceramics showed a decreasing trend
in the range 4.5–6 GHz. The lower values of dielectric constant at
microwave region compared to the low frequencies is due to
suppression of domain wall motion. BCTZO exhibits low dielectric
loss of (0.009) and high quality factor (Q = 1/tan d) which is
comparable to BST ceramics reported by Li et al. [22]. In this
context, the present ceramic with high tunability and low
dielectric loss is a promising material for dielectric tunable
applications.
In order to study the effect of phase transitions on the
vibrational character of the structure of this compound with
TCP, we used the in situ micro Raman spectroscopy where the size
of the spot of the exciting laser (532 nm) on the sample was less
than 50 micron; for comparison we also recorded the phonon
modes of BTO ceramic having about the same grain size.
Ferroelectric BTO in tetragonal structure allows 3A1 modes and
3E modes arising from the zone centre infrared active phonons
(F1u) and one silent mode (E + B1) comes from F2u [23,24]. Due to
Fig. 4. (a) Electric field dependence of dielectric permittivity for BTO and BCTZO
ceramics. Inset: variation of tunability with the applied field. (b) Dielectric the long-range electrostatic force associated with the lattice
permittivity and loss for BTO and BCTZO at microwave frequencies. iconicity, each of the E and A1 modes further splits into two optical
modes, transverse (TO) and longitudinal (LO).
Fig. 5 shows the results of room temperature Raman spectra
up to 40 kV cm1. Figure shows the butterfly shaped curve with corresponding to BTO and BCTZO ceramics. The vibrational modes
hysteresis for both the ceramics. The tunability is determined from are indicated in the diagram. BCTZO ceramic exhibited several
the change in dielectric constant at zero electric field compared to weak modes A1(TO1), E(TO)/E(LO)/A1(LO), B1 and/or E(TO + LO) at
field dependent values following the relation ððeð0Þ  eðEÞÞ=eð0ÞÞ  148 cm1, 182 cm1, 275 cm1, respectively. The onset of asym-
100 where e(0) and e(E) are dielectric constant under zero and metry in the E(TO), A1(TO) mode, broad A1(LO), E(LO) bands are
external electric fields, respectively. The change in tunability is visible at 519 cm1 and 723 cm1, respectively. The presence of
shown in the inset of Fig. 4a. The observed value of tunability for observed Raman modes match well with those of polycrystalline
BTO, 76%, is in good agreement with the earlier report [18]. It was BTO ceramics reported in literature [25–27]. The Raman modes of
observed that BCTZO ceramics exhibited a maximum tunability of undoped BTO were observed at 174, 262, 300, 513 and 718 cm1.
82% compared to that of BTO (76%) under an electric field of The peak at 300 cm1, E(TO3) mode, is sharper than the
40 kV cm1.This value is in good agreement with the tunability corresponding peak of BCTZO, agreeing with the higher tetra-
obtained for BCTZO thin films (83%) [11]. The present ceramics gonality of BTO seen in XRD results. The dip around 180 cm1 for
exhibit higher tunability compared to other BTO–based ferro- BTO is attributed to the interference effect between a narrow and a
electrics such as Ca and Zr containing BTO (60%)[19], Zr-doped broad A1(TO) mode indicating the signature of tetragonal
Ba0.6Sr0.4TiO3 (57%) [20], Sn-modified BTO (74%)[18], Zr-doped ferroelectric phase. Broadening of all the peaks in the mixed
BTO (37%) [21]. The existence of tricritical point close to room system suggests the presence of higher disorder compared to BTO.
temperature could be the possible reason for the larger tunability. This can be interpreted as the signature of coexistence of R and T
Dielectric constant measured at microwave frequency. The phases. In BCTZO it can also be seen that the intense peaks at 300
dielectric permittivity was also measured at microwave frequency and 262 cm1 are shifted towards lower wave number. The peaks
using the reflectance method. Fig. 4b shows the variation of e0r with around 513 and 718 cm1 correspond to phonon vibrations of the
frequency. BCTZO shows a dielectric constant of 710 at 3 GHz Ba-O bonds while the peaks in the range 180–300 correspond to
which is significantly lower compared to the value of 3152 at 100 the phonon vibrations of Ti–O bonds. A weak Raman active
KHz. The value of e0r at microwave frequencies for BTO is 540 which asymmetric breathing mode (A1g) was seen at 816 cm1 for BCTZO
V.R. E. et al. / Materials Research Bulletin 48 (2013) 4395–4401 4399

Fig. 7. (a) Polarization–electric field hysteresis loops measured at selected


temperatures in the range RT-400 K. Inset: variation of Pr at different electric
fields. (b) Temperature dependence of pyroelectric coefficient and spontaneous
polarization for BCTZO ceramic.

in Fig. 6a and b. In BTO the presence of sharp E + B1 mode at


300 cm1 and the high frequency LO modes around 718 cm1 mark
the existence of long-range ferroelectric tetragonal phase. With the
increase in temperature these modes become asymmetric and
broader. For T > 392 K, the mode around 718 cm1 diminished
completely indicating the transition to paraelectric cubic phase.
However, BCTZO showed slightly different behaviour (Fig. 6c) in
which the intensity of peak around 718 cm1 is diminished and
resolved into weak reflections even after T > 396 K. It is generally
agreed that the cubic phase is characterized by two second order
broadened peaks around 225 and 520 cm1 (Pasha et al. [27]).
From the result of Fig. 2 (XRD), we know that BCTZO attains the
cubic phase around 360 K; however, the sharpness of the transition
Fig. 6. Temperature dependence of Raman spectra of BTO (a) and BCTZO (c). is not evident in the Raman data. Similarly, the mode at 148 cm1
Variation of Raman shifts around 515 cm1 and 715 cm1 with temperature are loses its intensity, becomes broad and weak showing the signature
shown in (b) and (d) for BTO and BCTZO, respectively. Verticle lines indicate of the phase transition from tetragonal to cubic. The presence of
structural phase transitions. this peak at high temperatures (T > 360 K) indicates that the
transition is not complete. Thus we are led to conclude that the
while this mode is absent for pure BTO. This high frequency broad phase transition of BCTZO is unlike that of BTO and the polar state
peak was also reported in other components of BCT–BZT series of may persist through TC and there exists regions retaining
compounds and is due to the presence of several dissimilar atoms tetragonal symmetry into the cubic phase. Raman shift vs
at A- and B-sites forming a complex perovskite solid solution temperature graph shown in Fig. 6d clearly shows that the phase
system [26]. Shifting of peaks in BCTZO to higher frequency transition is of diffuse nature, as evidenced by the dielectric
compared to BTO, especially at 519 and 718 cm1, is expected due measurements. Some features of the phase transition observed in
to the different ionic radius between Ba2+ and Ca2+ as well as Ti4+ the present samples are similar to the relaxor type BTO-based
and Zr4+ which results in the distortion of the lattice and results in materials such as Ba(Zr1xTix)O3[23] and Ba(Ti,Sn)O3 [28]. The
energy band widening. present ceramic with the presence of dissimilar ions at A and B-
The evolution of Raman spectra with temperature for BTO and sites of the lattice exhibited diffused phase transition, which is
BCTZO are shown in Fig. 6. Results for the pure BTO are presented suspected primarily to arise from the short range cation disorder
4400 V.R. E. et al. / Materials Research Bulletin 48 (2013) 4395–4401

between Ti4+ and Zr4+ of the polar nano regions (PNR). In a recent Shi et al. [14] suggested that the optimized condition to pole the
study, Damjanovic et al. [29], detected residual piezoelectricity for sample for its maximum piezoelectric properties should be at a
temperatures 30 8C above TC (dielectric) and showed that the field equal to two times the EC and the temperature should be in
macroscopic symmetry of the BCTZO remains polar above Curie the range 303–413 K. We measured the piezoelectric charge
temperature. This result is consistent with our Raman measure- coefficient, d33, of samples poled electrically at 50 kV cm1 at
ments of Fig. 6. Such a polar nature in this region are expected due 313 K. The obtained d33 of 470 pC/N is smaller than the value of
to the PNR, precursor order or defects in the material. 620 pC/N reported by Liu and Ren [7] but comparable to the result
Ferroelectric hysteresis loops measured from room tempera- (480 pC/N) of Yao et al. [10].
ture to 130 8C are shown in Fig. 7a. A well saturated, ‘‘soft’’ P–E Fig. 7b shows the temperature dependence of the pyroelectric
loops with a remanent polarization (Pr) of 12.2 mC cm2 and a coefficient for BCTZO ceramic poled at 50 kV cm1. The pyroelec-
coercivity (EC) = 2.8 kV cm1 was obtained at room temperature. tric coefficient (p) was calculated from the measured current (IP)
From the inset of Fig. 7a it can be seen that the Pr increases with the using the relation p ¼ I p =ðA  ðdT=dtÞÞ where A is the sample’s
applied electric field and saturates around 23 kV cm1. The electrode area and dT/dt is the rate of heating [12]. The value of ‘p’
obtained polarization limits are in good agreement with those increases with temperature and exhibits a peak at 367 K. A small
reported by others; Shi et al. (Pr = 13 mC cm2, EC = 2.7 kV cm1) hump observed around 310 K is similar to the result of dielectric
[14], Wu et al. (Pr = 10 mC cm2, EC = 2.35 kV cm1) [13] and anomaly at this temperature and can be ascribed to the
slightly lower than the results of Liu and Ren [7] (Pr = 15 mC cm2, rhombohedral–tetragonal phase transition. Here, it can be
EC = 1.7 kV cm1). With the increasing temperature, hysteresis observed that the peak corresponding to the TC appeared at
loops become slimmer with a drop in Pr values. Above 358 K the slightly higher temperature. The existence of pyroelectric current
loops are almost closed (paraelectric) and the ferroelectric is expected from the existence of permanent dipoles oriented
behaviour almost disappears, confirming the phase transition randomly in different regions beyond TC. This feature can be
results of Figs. 2 and 3. correlated to the observed peak broadening in the dielectric

Fig. 8. Piezoelectric force microscopy images of BCTZO ceramic. (a) Topogaphy, (b) and (d) out-of-plane piezoresponse amplitude before and after poling, (b) and (e)out-of-
plane piezoresponse phase before and after poling the area under circle with 75 V and (f) local piezoresponse hysteresis loop.
V.R. E. et al. / Materials Research Bulletin 48 (2013) 4395–4401 4401

behaviour, which has been also attributed to compositional transition is in good agreement the results of X-ray studies. It
heterogeneities [12]. The room temperature value of p in the was observed that BCTZO ceramics exhibit a dielectric tunability
present study is 6  108 C cm2 K1 which is comparable to that of 82% under an electric field of 40 kV cm1 and good quality factor
reported by Yao et al. (5.84  108 C cm2 K1) [10]. The sponta- of 1 1 1 at microwave frequencies. By using the Raman spectroscopy
R
neous polarization estimated from the relation P S ¼ pdT is we observed that the nature of ferroelectric–paraelectric transition
shown in Fig. 7b. The value of PS at room temperature is for BCTZO is not sharp as observed in the case of BTO and the polar
15 mC cm2. This spontaneous polarization is normally larger state persists through the range of phase transition temperatures
than Pr as this is P (E = 0). However, we observed no saturation in detected by the dielectric measurements. The sample exhibited well
the PS at low temperatures similar to the study of Benabdallah et al. saturated P–E loops with a remanent polarization of 12.2 mC cm2
[30], who attributed it to the fast change of ferroelectric domains. and a coercivity (EC) = 2.8 kV cm1. Pyroelectric measurements
The value of PS approaches zero at the maximum of p indicating the exhibited a room temperature coefficient of 6  108 C cm2 K1at
paraelectric state. The pyroelectric response of a material is room temperature and a detectivity of 1.9  108 C cm1 J1, higher
characterized by its figure of merit, detectivity, given by F D ¼
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi than the BaTiO3-based ceramics. By means of Piezoresponse force
p= C V  e0r  tan d where p is the pyroelectric coefficient, CV is the microscopy measurements we observed that the direction of
volume specific heat, e0r and tan d is the dissipation factor [12]. The polarization is mostly oriented in out-of-the-plane direction. A
calculated value of FD for BCTZO is 1.9  108 C cm1 J1 which is clear hysteresis observed in local piezoresponse indicated switch-
more than one order larger than the one observed for the BTO- able ferroelectricity at the nanoscale.
based pyroelectric arrays reported by Noh et al. [31] and indicates
that BCTZO is a useful material for infrared detection. Acknowledgement
We studied the ferroelectricity of BCTZO ceramic by domain
imaging in piezoforce microscopy. In PFM image, the local Authors would like to acknowledge the financial support from
vibration amplitude of the test sample is proportional to the FCT, Portugal (SFRH/BPD/75582/2010).
effective piezoelectric coefficient while the phase is related to the
orientation of ferroelectric domains. From the changes in piezo-
response contrast, domains with oppositely oriented polarization References
can be detected. Fig. 8a shows the topography of the BCTZO [1] B. Noheda, D.E. Cox, G. Shirane, J. Gao, Z.G. Ye, Phys. Rev. B 66 (2002) 054104–
indicating polycrystalline rough grains/islands (rms roughness is 54113.
100 nm). We observed a poor correlation between the topography [2] D.E. Cox, B. Noheda, G. Shrine, Y. Uesu, K. Fujishiro, Y. Yamada, Appl. Phys. Lett. 79
(2011) 400–403.
and in-plane amplitude (not shown) indicating that most of the
[3] N. Setter, D. Damjanovic, L. Eng, G. Fox, S. Gevorgian, S. Hong, A. Kingon, H.
domains are oriented towards the sample surface. Fig. 8b and c Kohlstedt, N.Y. Park, G.B. Stephenson, I. Stolitchnov, A.K. Taganstev, D.V. Taylor, T.
shows the out-of-plane (vertical) amplitude and phase images Yamada, S. Streiffer, J. Appl. Phys. 100 (2006) 051606–51651.
with dark and bright regions corresponding to weak and strong [4] Q. Zhang, Y. Zhang, F. Wang, Y. Wang, D. Lin, X. Zhao, H. Luo, W. Ge, D. Viehland,
Appl. Phys. Lett. 95 (2009) 102942–102945.
piezoresponse (PR) of the topographic image. The PR seems to be [5] A. Sasaki, T. Chiba, Y. Mamiya, E. Otsuki, Jpn. J. Appl. Phys. 38 (1999) 5564–5567.
suppressed around the boundary of islands. From the phase image, [6] Y. Saito, H. Takao, T. Tani, T. Nonoyama, K. Takatori, T. Homma, T.M. Nagaya,
it can be seen that the domains exhibit polarization in out-of-plane Nakamura Nat. 432 (2004) 84–87.
[7] W. Liu, X. Ren, Phys. Rev. Lett. 103 (2009) 257602–257606.
direction. In ceramics of the present class the large fraction of [8] J. Gao, D. Xue, Y. Wang, D. Wang, L. Zhang, H. Wu, S. Guo, H. Bao, C. Zhou, W. Liu, S.
grains showing piezoresponse are coupled in clusters of several Hou, G. Xiao, X. Ren, Appl. Phys. Lett. 99 (2011) 092901–92903.
grains and the intragranular interactions are responsible in the [9] D. Xue, Y. Zhou, H. Bao, C. Zhou, J. Gao, X. Ren, J. Appl. Phys. 109 (2011) 054110–
54115.
formation of domains. To study the ferroelectric behaviour by [10] S. Yao, W. Ren, H. Ji, X. Wu, P. Shi, D. Xue, X. Ren, Z.G. Ye, J. Phys. D: Appl. Phys. 45
means of domain switching we applied a dc voltage of 75 V with a (2012) 195301–195305.
pulse duration of 0.1 s between the AFM tip, at the location [11] A. Piorra, A. Petraru, H. Kohlstedt, M. Wuttig, E. Quandt, J. Appl. Phys. 109 (2011)
104101–104104.
encircled in Fig. 8b, and the bottom electrode of the sample. The [12] E.V. Ramana, V.V. Kiran, T. Bhima Sankaram, J. Alloys Compd. 456 (2008) 271–276.
brightened area in the circle of Fig. 8d indicates that the switched [13] J. Wu, D. Xiao, W. Wu, Q. Chen, J. Zhu, Z. Yang, J. Wang, J. Eur. Ceram. Soc. 32 (2012)
domain extends to the whole region and the induced polarization 891–898.
[14] S. Shi, R. Zuo, S. Lu, Z. Xu, X. Wang, L. Li, Curr. Appl. Phys. 11 (2011) S120–S123.
is quite stable. A similar response can be seen in the phase image of
[15] G.A. Smolenskii, J. Phys. Soc. Jpn. Suppl. 28 (1970) 26–37.
Fig. 8e. To further prove the ferroelectricity of the sample, we [16] M.E. Lines, A.M. Glass, Principles and applications of ferroelectrics and related
measured local hysteresis in piezoresponse inside individual materials, Clarendon Press, Oxford, 1977p. 285.
grains by keeping the AFM tip fixed above the surface and [17] G.A. Smolenskii, Ferroelectrics 53 (1984) 129–135.
[18] N. Horchidana, A.C. Ianculescu, L.P. Curecheriu, F. Tudorache, V. Musteata, S.
applying a sequence of voltage pulses. Fig. 8f clearly shows a Stoleriu, N. Dragan, D. Crisan, S. Tascu, L. Mitoseriu, J. Alloys Compd. 509 (2011)
switchable ferroelectric response from the circled area in Fig. 8b 4731–4737.
proving the control of ferroelectric domains at a nanoscale. It was [19] X.G. Tang, K.-H. Chew, J. Wang, H.L.W. Chan, Appl. Phys. Lett. 85 (2004) 991–993.
[20] X. Wang, R. Huang, Y. Zhao, Y. Zhao, H. Zhou, Z. Jia, J. Alloys Compd. 533 (2012)
observed that the ferroelectric hysteresis can be obtained from the 25–28.
grains which initially showed weak PR. This could be due to the [21] T. Maiti, R. Guo, A.S. Bhalla, Appl. Phys. Lett. 89 (2006) 122909–122911.
ferroelectric grains which might have their polarization parallel to [22] L. Li, Y. Fang, X.M. Chen, J. Am. Ceram. Soc. 95 (2012) 982–985.
[23] U.D. Venkateswaran, V.M. Naik, R. Naik, Phys. Rev. B 58 (1998) 14256–14260.
the surface and then are rotated perpendicular to the surface by the [24] N.K. Karan, R.S. Katiyar, T. Maiti, R. Guo, A.S. Bhalla, J. Raman Spectrosc 40 (2009)
applied electric field. In short, the PFM measurements clearly 370–375.
demonstrated the domain switchability of BCTZO ceramic. [25] Y. Shiratori, C. Pithan, J. Dornseiffer, R. Waser, J. Raman Spectrosc. 38 (2007)
1300–1306.
[26] V.S. Puli, A. Kumar, D.B. Chrisey, M. Tomozawa, J.F. Scott, R.S. Katiyar, J. Phys. D:
4. Conclusions Appl. Phys. 44 (2011) 395403–395412.
[27] U.M. Pasha, H. Zheng, O.P. Thakur, A. Feteira, K.R. Whittle, D.C. Sinclair, I.M. Reany,
Appl. Phys. Lett. 91 (2007) 062908–62913.
Ferroelectric (Ba0.85Ca0.15)(Ti0.9Zr0.1)O3 ceramics were synthe-
[28] M. Deluca, L. Stoleriu, L.P. Curecheriu, N. Horchidan, A.C. Ianculescu, C. Galassi, L.
sized by the conventional ceramic rout. By using the XRD Mitoseriu, J. Appl. Phys. 111 (2012) 084102–84111.
measurements the phase tetragonal–cubic transition associated [29] D. Damjanovic, A. Biancoli, L. Batooli, A. Vahabzadeh, J. Trodahl, Appl. Phys. Lett.
with the tricritical point, has been demonstrated. The dielectric 100 (2012) 192907–192911.
[30] F. Benabdallah, A. Simon, H. Khemakhem, C. Elissalde, M. Maglione, J. Appl. Phys.
measurements exhibited a diffused transition behaviour in 109 (2011) 124116–124126.
ferroelectric–paraelectric transition around 360 K and this [31] H.J. Noh, S.G. Lee, S.P. Nam, Y.H. Lee, Mater. Res. Bull. 45 (2010) 339–342.

You might also like