ferroelectric relaxor behavior in hafnium doped barium-titanate ceramic

6
Ferroelectric relaxor behavior in hafnium doped barium-titanate ceramic Shahid Anwar, P.R. Sagdeo, N.P. Lalla * UGC-DAE Consortium for Scientific Research, University Campus, Khandwa Road, Indore 452017, India Received 22 June 2005; received in revised form 7 March 2006; accepted 18 March 2006 by B.-F. Zhu Available online 4 April 2006 Abstract Temperature and frequency dependence of the real (3 0 ) and imaginary (3 00 ) parts of the dielectric permitivity of cubic Ba(Ti 0.7 Hf 0.3 )O 3 ceramic has been studied in the temperature range of 100 K to 350 K at the frequencies 0.1 kHz, 1 kHz, 10 kHz, 100 kHz for the first time. Diffuse phase transition and frequency dispersion is observed in the permittivity-vs-temperature plots. This has been attributed to the occurrence of relaxor ferroelectric behavior. The observed relaxor behavior has been quantitatively characterized based on phenomenological parameters. A comparison with the Zr doped BaTiO 3 has also been presented. For Hf doped samples transmission electron microscopy (TEM) characterization do show the presence of highly disordered microstructure at length scales of few tens of nano-meters. q 2006 Elsevier Ltd. All rights reserved. PACS: 77.; 77,22.Ch Keywords: A. Ferroelectric; C. Structure; D. Dielectric 1. Introduction Perovskite-based ferroelectrics materials attract consider- able interest owing to rich diversity of their physical properties and possible applications in various technologies like memory storage devices [1], micro-electromechanical systems [2], multilayer ceramic capacitors [3], and recently in the area of opto-electronic devices [4]. These useful properties have most often been observed in lead based perovskite compounds, such as PMN, PST, PLZT [5–7]. The enhanced properties of these compounds are attributed to their relaxor behavior, observed in doped (mixed) perovskites. However these compositions have obvious disadvantages of volatility and toxicity of PbO. Therefore much effort has been carried out towards investi- gating environmental friendly ‘Pb-free’ ceramic materials. Specifically, BaTiO 3 and its isovalent substituted materials are the promising candidates for microwave and opto-electronic applications. The effect of substitution on dielectric relaxation, ferro- electric phase transition and electrical properties of BaTiO 3 has been extensively studied [8,9]. On partial substitution of dopants like Ca, Sr, Zr [10–12], the variation of 3 0 around T c gets broadened out in ceramics and single crystal samples both. Broadening increases with increasing concentration of the dopant, as also does the deviation from Curie–Weiss behavior at temperatures above the peak temperature (T m ) of the 3 0 KT variation. The observed broadening in 3 0 KT variation has generally been attributed to the presence of nano-regions resulting from local composition variation over length scale of 100–1000 A ˚ . Different nano-regions in a macroscopic sample transform at different temperatures giving rise a range of transformation temperatures, the so-called ‘Curie range’. Thus the compositional fluctuation [6,11] in an otherwise composi- tionally homogenous system leads to diffuse phase transition (DPT). In compositionally homogenous systems quenched random disorder breaks the long range polar order at unit cell level, leading to broad 3 0 KT response [7]. Such materials exhibit slow enough relaxation dynamics and hence have been termed as ferroelectric relaxors [5–7]. A series of impurity doped BaTiO 3 system such as Sn, Ce, Zr etc have shown ferroelectric relaxor behavior. Among these the Zr-substituted BaTiO 3 ceramics have received renewed attention due to its enhanced properties both in single crystals and ceramics [12]. In the present investigation, we have studied the ferro- electric relaxor behavior in high concentration Hf substituted BaTiO 3 , i.e. Ba(Ti 0.7 Hf 0.3 )O 3 ceramics, by monitoring the variation of its dielectric permittivity with temperature in the range of 90–350 K and in the frequency range of 0.1–100 KHz. Till date only limited amount of work has been carried out for Hf doping on Ti site in BaTiO 3 [13,14] ceramics like its effect Solid State Communications 138 (2006) 331–336 www.elsevier.com/locate/ssc 0038-1098/$ - see front matter q 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.ssc.2006.03.018 * Corresponding author. Tel.: C91 731 2463913; fax: C91 731 2462294. E-mail address: [email protected] (N.P. Lalla).

Upload: shahid-anwar

Post on 29-Jun-2016

216 views

Category:

Documents


0 download

TRANSCRIPT

Ferroelectric relaxor behavior in hafnium doped barium-titanate ceramic

Shahid Anwar, P.R. Sagdeo, N.P. Lalla *

UGC-DAE Consortium for Scientific Research, University Campus, Khandwa Road, Indore 452017, India

Received 22 June 2005; received in revised form 7 March 2006; accepted 18 March 2006 by B.-F. Zhu

Available online 4 April 2006

Abstract

Temperature and frequency dependence of the real (3 0) and imaginary (3 00) parts of the dielectric permitivity of cubic Ba(Ti0.7Hf0.3)O3 ceramic

has been studied in the temperature range of 100 K to 350 K at the frequencies 0.1 kHz, 1 kHz, 10 kHz, 100 kHz for the first time. Diffuse phase

transition and frequency dispersion is observed in the permittivity-vs-temperature plots. This has been attributed to the occurrence of relaxor

ferroelectric behavior. The observed relaxor behavior has been quantitatively characterized based on phenomenological parameters.

A comparison with the Zr doped BaTiO3 has also been presented. For Hf doped samples transmission electron microscopy (TEM)

characterization do show the presence of highly disordered microstructure at length scales of few tens of nano-meters.

q 2006 Elsevier Ltd. All rights reserved.

PACS: 77.; 77,22.Ch

Keywords: A. Ferroelectric; C. Structure; D. Dielectric

1. Introduction

Perovskite-based ferroelectrics materials attract consider-

able interest owing to rich diversity of their physical properties

and possible applications in various technologies like memory

storage devices [1], micro-electromechanical systems [2],

multilayer ceramic capacitors [3], and recently in the area of

opto-electronic devices [4]. These useful properties have most

often been observed in lead based perovskite compounds, such

as PMN, PST, PLZT [5–7]. The enhanced properties of these

compounds are attributed to their relaxor behavior, observed in

doped (mixed) perovskites. However these compositions have

obvious disadvantages of volatility and toxicity of PbO.

Therefore much effort has been carried out towards investi-

gating environmental friendly ‘Pb-free’ ceramic materials.

Specifically, BaTiO3 and its isovalent substituted materials are

the promising candidates for microwave and opto-electronic

applications.

The effect of substitution on dielectric relaxation, ferro-

electric phase transition and electrical properties of BaTiO3 has

been extensively studied [8,9].On partial substitution of dopants

like Ca, Sr, Zr [10–12], the variation of 3 0 around Tc gets

0038-1098/$ - see front matter q 2006 Elsevier Ltd. All rights reserved.

doi:10.1016/j.ssc.2006.03.018

* Corresponding author. Tel.: C91 731 2463913; fax: C91 731 2462294.

E-mail address: [email protected] (N.P. Lalla).

broadened out in ceramics and single crystal samples both.

Broadening increases with increasing concentration of the

dopant, as also does the deviation fromCurie–Weiss behavior at

temperatures above the peak temperature (Tm) of the 3 0KT

variation. The observed broadening in 3 0KT variation has

generally been attributed to the presence of nano-regions

resulting from local composition variation over length scale of

100–1000 A. Different nano-regions in a macroscopic sample

transform at different temperatures giving rise a range of

transformation temperatures, the so-called ‘Curie range’. Thus

the compositional fluctuation [6,11] in an otherwise composi-

tionally homogenous system leads to diffuse phase transition

(DPT). In compositionally homogenous systems quenched

random disorder breaks the long range polar order at unit cell

level, leading to broad 3 0KT response [7]. Such materials

exhibit slow enough relaxation dynamics and hence have been

termed as ferroelectric relaxors [5–7]. A series of impurity

doped BaTiO3 system such as Sn, Ce, Zr etc have shown

ferroelectric relaxor behavior. Among these the Zr-substituted

BaTiO3 ceramics have received renewed attention due to its

enhanced properties both in single crystals and ceramics [12].

In the present investigation, we have studied the ferro-

electric relaxor behavior in high concentration Hf substituted

BaTiO3, i.e. Ba(Ti0.7Hf0.3)O3 ceramics, by monitoring the

variation of its dielectric permittivity with temperature in the

range of 90–350 K and in the frequency range of 0.1–100 KHz.

Till date only limited amount of work has been carried out for

Hf doping on Ti site in BaTiO3 [13,14] ceramics like its effect

Solid State Communications 138 (2006) 331–336

www.elsevier.com/locate/ssc

20 40 60 80 100 120

BaTiO3

Ba(Ti0.7Hf0.3)O3

43.5 44.0 44.5 45.0 45.5 46.073 74 75 76

Inte

nsity

(a.

u)

Fig. 1. Room temperature XRD pattern of BaTiO3 (grey circles) and

Ba(Ti0.7Hf0.3)O3 (black circles), Fig. also shows the fitted pattern, Braggs

peaks and difference between observed and calculated pattern. Inset show the

expansion around the (200) and (310) peaks.

Fig. 2. Secondary electron micrographs corresponding to (a) BaTiO3 and (b)

Ba(Ti0.7Hf0.3)O3 respectively.

S. Anwar et al. / Solid State Communications 138 (2006) 331–336332

on dielectric and structural properties. Detailed structural and

dielectric studies have been carried out. The data has been

quantitatively analyzed in terms of parameters characterizing

the relaxor behavior.

2. Experimental

BaTiO3 and Ba(Ti0.7Hf0.3)O3 ceramics were prepared

following the conventional solid-state reaction technique.

High purity (99.99%) powders of BaCO3, TiO2 and HfO2

were weighed in stoichiometric proportions and wet-mixed,

taking acetone as mixing medium. After mixing and drying, the

powder was calcined at 1100 8C for 6 h. The calcined powder

was remixed, dried, and then palletized at 100 kN/cm2 pressure

into a 15 mm dia. pallet using polyvinyl alcohol as binder. The

pallets were then sintered at 1250 8C for 12 h and cooled

naturally to room temperature while furnace power was off.

The as prepared pallets of BaTiO3 and Ba(Ti0.7Hf0.3)O3were

subjected to structural and phase purity characterization

employing powder X-ray diffraction (XRD), scanning electron

microscopy (SEM-JEOL 5600) and transmission electron

microscopy (TEM). The XRD analysis was carried out using

CuKaX-ray employing a Rigakumake goniometer and rotating

anode X-ray generator working at 10 kW output power. Sample

for TEM studies were prepared using ion-beam polishing at

well-optimized kV and angle settings (3.3 kV and 38). TEM

characterization of these samples was carried out in diffraction

and imaging modes using FEI-TECNAI G2K20 working at

200 kV. Electroding of the pallets for dielectric measurement

was then done by painting its flat surfaces with high temperature

silver-paint and then firing it at 500 8C for 15 min. Dielectric

characterization of pallets, as a function of temperature, was

carried out employing temperature-controlled measurement of

its dielectric permitivity (3 0) and dielectric loss (tan d) in the

temperature range of 90–350 K. For these measurements a

computer interfaced and programmed LCRmeter (Hioki-3538)

and a Lakeshore temperature controller with a Pt-100,

connected in 4-probe mode, was used for temperature

measurement. The peak-to-peak ac signal of 1.0 V in the

frequency range of 102–105 Hz and a temperature step of 0.5 K

with set-point stability better than 20 mK were used. Thus the

measurement process ensures that no anomaly remains

unobserved in the entire studied temperature range.

3. Results and discussions

3.1. Structural studies

The results of X-ray diffraction characterization of the as

prepared Ba(Ti0.7Hf0.3)O3 and BaTiO3 samples are shown in

Fig. 1. The insets of Fig. 1 show two sets of (200) & (002) and

(301) & (310) diffraction peaks which is a characteristics of the

tetragonal BaTiO3 phase at room temperature. In the case of

Ba(Ti0.7Hf0.3)O3 the two peaks merge into single peak, i.e.

(200) and (310), which is a characteristic of tetragonal to cubic

phase transition. The XRD data of Ba(Ti0.7Hf0.3)O3 and

BaTiO3 samples were subjected to Reitveld refinement with

space-group of Pm3m (cubic) and P4 mm (tetragonal)

respectively. The goodness of fit parameter were found to be

w2.44. The continuous line in Fig. 1 shows the Reitveld fit to

the XRD data of Ba(Ti0.7Hf0.3)O3. The quality of fit and the

absence of any unaccounted diffraction peak indicate that the

synthesized Ba(Ti0.7Hf0.3)O3 ceramic is cubic single phase

material. Fig. 2(a) and (b) depicts the secondary electron

Fig. 3. Transmission electron micrographs corresponding to BaTiO3 (a, c) and

Ba(Ti0.7Hf0.3)O3 (b, d) respectively.

S. Anwar et al. / Solid State Communications 138 (2006) 331–336 333

micrographs of BaTiO3 and Ba(Ti0.7Hf0.3)O3 phases respect-

ively. Well faceted grains of sizes ranging from 2–8 mm can be

seen.

Fig. 3(a)–(b) shows TEM micrographs of BaTiO3 and

Ba(Ti0.7Hf0.3)O3. The occurrence of large grains of sizes

3–5 mm of BaTiO3 and Ba(Ti0.7Hf0.3)O3 phases can be seen in

Fig. 3(a) and (b), respectively. This is in conformity with the

SEM results. A relatively high-magnification micrograph of

BaTiO3 is shown in Fig. 3(c), which depicts the occurrence of

ferroelectric twin-domains of the tetragonal BaTiO3 phase.

Fig. 3(d) shows a high-magnification micrograph of Ba(Ti0.7Hf0.3)O3 phase. One can see that unlike BaTiO3 phase no

contrast showing the twin-domains are present in its

microstructure but it is rich in nano contrast regions of

4–10 nm apparent sizes. The detailed features of these nano

contrasts are typical to the presence of strained regions [16]. To

confirm that these strains are not due to some secondary phase

inclusions, we carried out selected area diffraction (SAD) from

these areas as shown in the inset of Fig. 3(d). It is obvious from

this SAD pattern that no extra spots are present other than those

from the basic perovskite structure. This confirms that these

contrasts are not due to any secondary phase inclusions. Since

these regions were found to change contrast simultaneously

during tilt, these might originate from strain fields due to some

defect, like tiny dislocation loops extending just to few lattice

sites with their burger vectors parallel. The occurrence of these

defect features in Ba(Ti0.7Hf0.3)O3 will lower its structural

correlation length as compared to BaTiO3 in which no such

features were found. The ‘structural correlation length’ is

basically a measure of the effective extent to which the long-

range order of an atomic arrangement gets limited as a result of

the cumulative effect of all types of defect features, which

some how either interrupt the chemical order or produce strain

in the lattice. Beyond this extent the atoms don’t scatter

coherently and contribute to the width of the diffraction

maxima. Thus the structural correlation length may be used as

a measure of comparison of defects presence in two similar

types of structures. Thus to compare the presence of defects in

the bulk samples of Ba(Ti0.7Hf0.3)O3 and BaTiO3, structural

correlation lengths were calculated from the half-widths of the

(111) XRD peak, after applying instrumental broadening and

Ka2 corrections. These were found to be w980 A and

w1120 A respectively for Ba(Ti0.7Hf0.3)O3 and BaTiO3. It

should be noted that the size of even the smallest grain, as seen

through SEM, is at least an order of magnitude larger than the

structural correlation length. Nearly 10% smaller structural

correlation length was found for Ba(Ti0.7Hf0.3)O3 as compared

to that of the BaTiO3, indicating the presence of more defects

in it. This tells that the defect feature observed through TEM is

a characteristic of the whole bulk and is not only limited to few

grains.

3.2. Dielectric studies

The temperature dependence of real (3 0) and imaginary (3 00)

parts of the dielectric permitivity of Ba(Ti0.7Hf0.3)O3 ceramic

are shown in Fig. 4. Unlike BaTiO3 the transition is quite

diffuse. The paraelectric to ferroelectric phase transition

temperature (Tc) as compared to that of the BaTiO3, have

decreased. The three phase transitions which are observed in

BaTiO3 have got pinched and merged into one round peak in

3 0KT variation. The results obtained can be described as in the

following:

100 150 200 250 300 350

0.0003

0.0006

0.0009

0.0012

0.0015

Tdev

1/ε'

Temperature (K)

Fig. 5. The inverse dielectric constant (1/3 0) as a function of temperature at

10 KHz for ceramics Ba(Ti0.7Hf0.3)O3. The Symbol represents experimental

data points and solid line shows fitting to the Curie–Weiss law.

100 150 200 250 300 350

500

1000

1500

2000

2500

3000

3500

100 KHz

100 Hz

100 KHz

100 Hz

0.1KHz1KHz10KHz100KHz

ε'

100 150 200 250 300 350

0

50

100

150

200

250

300

ε''

Temperature(K)

Fig. 4. Temperature dependence of 3 0 and 300 of Ba(Ti0.7Hf0.3)O3 ceramic at 0.1,

1, 10 and 100 kHz.

–5.5

S. Anwar et al. / Solid State Communications 138 (2006) 331–336334

(1) There is a broad peak around TmZ200 K in the 3 0KT

curve. With increasing frequency Tm increases, while the

magnitude of the peak decreases.

(2) There is a strong dielectric dispersion in radio frequency

region around and below Tm in the 3 0KT.

(3) The value of T 0m (dielectric absorption maxima tempera-

ture), is much less than Tm and around and above T 0m the

dielectric absorption (3 00) exhibits a strong frequency

dependence. With increasing frequency T 0m shifts to

higher temperature with increasing dielectric absorption.

The above described features of (30KT) and (300KT) variations

shown in Fig. 4 are very much similar to the observations by

Cross, Lu, Cheng, Chen and other workers [5,6,10,12,15] for

various lead based and lead free ferroelectric relaxormaterials. In

order to further confirm the relaxor behavior, the quantitative

characterizations as described in the following have been done.

0.25 0.50 0.75 1.00

–7.0

–6.5

–6.0

log(

1/ε'-

1/ε' m

)

log(T-Tε'm)

Fig. 6. ln(1/30K1/3 0m)-vs-ln(TKT3 0m) plot for Ba(Ti0.7Hf0.3)O3 ceramics at

10 kHz Symbol represents experimental data and solid line shows fitting.

3.3. Permittivity variation in the high temperature side

It is known that dielectric permittivity of a normal

ferroelectric above Curie temperature follows the Curie–

Weiss law described by

1=30 Z ðTKTcÞ=C ðTOTcÞ; (1)

Where Tc is the Curie temperature and C is the Curie–Weiss

constant. Fig. 5 shows the inverse of 30 as a function of

temperature at 10 kHz and its fit to the experimental data by

Curie–Weiss law. A deviation from Curie–Weiss law starting at

Tdev can be clearly seen. The parameterDTm, which is often usedto show the degree of deviation from the Curie-Weiss law is

defined as

DTm Z TdevKTm (2)

The Tdev as determined from the Curie–Weiss fit, isTdevZ288 K,

andDTm is thus found to beZ132 K at 10 kHz. For such relaxor

behavior a modified Curie–Weiss law has been proposed by

Uchino and Nomura, [17] to describe the diffuseness of the phase

transition. This is defined as in the following.

1=30K1=30m Z ðTKT30mÞg=C1 (3)

where g and C1 are modified constants, with 1ZgZ2. The

parameter gives the information on the character of the phase

transition. Its limiting values aregZ1 and gZ2 in expression (3)

of the Curie–Weiss law, gZ1 is for the case of a normal

ferroelectric and the quadratic dependence is valid for an ideal

ferroelectric relaxor respectively [12,18]. Thus the value ofg can

also characterize the relaxor behavior. Theplot of log(1/30K1/30m)

as a function of log(TKT3 0m) is shown in the Fig. 6 by fitting

S. Anwar et al. / Solid State Communications 138 (2006) 331–336 335

with Eq. (3), the exponent g, determining the degree of the

diffuseness of the phase transition, is obtained from the slope of

log(1/3 0K1/3 0m)-vs- log(TKT3 0m) plot. We obtained the value

of the parameterg to be 1.88,which is very close to 2, suggestingthat the prepared ceramic is a relaxor ferroelectric [17,18]. Yet

another parameter, which is used to characterize the degree of

relaxation behavior in the frequency range of 100 Hz to

100 kHz, is described [18] as

DTrelax Z T30mð100 KHzÞKT30mð100 HzÞ (4)

The value of DTrelax was determined to be 23 K for the present

sample. The above characterization done on the basis of Curie–

Weiss law and the value of empirical parameters like DTm, g,and DTrelax suggest that the permittivity of Ba(Ti0.7Hf0.3)O3

ceramic follows Curie-weiss law only at temperatures much

higher than Tm. Thus the large deviation from the Curie–Weiss

behavior, large relaxation temperature g Trelax, and gZ1.88,

suggests that Ba(Ti0.7Hf0.3)O3 is a relaxor ferroelectric.

The frequency dependence of Tm is shown in Fig. 7 as

ln f-vsK1000/Tm. The observed frequency dependence of Tmwas empirically evaluated using Vogel–Fulcher’s relationship

given as

f Z f0expfKEa=kBðTmKTfÞg (5)

where Ea is the activation energy, Tf the freezing temperature

of polarization–fluctuation, and f0 is the pre-exponential factor.

The value of Ea, Tf and f0 for Ba(Ti0.7Hf0.3)O3 are found to be

0.14 eV, 78.5 K, 1.53!1013 Hz respectively. Similar values

have been reported for Ba(Ti0.7Zr0.3)O3 relaxor ferroelectric

system too [12,19].

The occurance of relaxor in Zr substituted barium titanate

[12,20,21] has been attributed to the existence of nano-polar

region due to Zr doping. The replacements of Ti4C by Zr4C

ions is known in the classical ferroelectric [19,20]. The

increasing substitution decreases the Curie temperature, which

is related to the ionic radius of the dopand. In the case of

BaZr0.3Ti0.7O3 (BZT) [12,19,21] where the ionic radius of

Zr4C(0.72 A) is higher than Ti4C(0.605 A) by 0.095 A the

TcZ280 K is lower than that of the pure BaTiO3 (TcZ395 K).

6.0 6.2 6.4 6.6 6.8 7.0 7.24

6

8

10

12

Ln(f

req)

1000/T

Fig. 7. Frequency dependence of Tm for Ba(Ti0.7Hf0.3)O3 ceramics. The

symbols and solid line indicate data points and fit to Vogel–Fulcher

relationship, respectively.

For BaHf0.3Ti0.7O3 (BHT), in the present studies, TcZ156 K,

which is much lower than that of the Zr doped one. Whereas

the ionic radius of Hf4C(0.71 A) is not much different than the

ionic radius of Zr4C. This appears to be due to lower activation

energy of BHT (0.14 eV) as compared to BZT (0.21 eV) [12].

It can be seen from the comparative structural study of

BaHf0.3Ti0.7O3 and BaTiO3 using XRD and TEM, that nano

scale defect features are present in BaHf0.3Ti0.7O3. It appears

that substitution of Hf at Ti site is causing nano-scale

compositional heterogeneity due to presence of defects

[19–21] creating nano polar domains. It has been suggested

that polar nano-domains are responsible for the relaxational

behavior in PMN and PLZT [10] systems. The substitution of

Hf ions tends to make the distance between off center Ti

dipoles larger and thus weakening the correlation between

these dipoles. The mismatch in the size of Ti and Hf ions will

cause substitutional distortion of the oxygen octahedra, giving

rise to the local electric-field and strain-field. Presence of such

strain-field contrast localized in nano-meter range, has been

very clearly observed for BaHf0.3Ti0.7O3 in the present

investigation, see Fig. 3(d). Present TEM investigations reveal

that these local fields and strain fields occur randomly but are

coherent in nature and tend to break the large ferroelectric

domains into polar nano-domain [18]. Also the ferroelectric

behavior of Ba(Ti0.7Hf0.3)O3 depends on the competitions

between long-range ordering owing to strong correlation of the

off center Ti dipoles and the random fields induced by Hf

doping, thus the existence of these fields leads to the

destruction of long-range ferroelectric ordering and nano-

polar coherent domains get formed giving rise to the relaxor

behavior.

4. Conclusions

Based on the X-ray diffraction and dielectric studies of

Ba(Ti0.7Hf0.3)O3 ceramic it can be concluded that Hf

substitution in the place of Ti causes a tetragonal to cubic

transformation at room temperature. The occurrence of diffuse

phase-transition and strong frequency dispersion of maxima in

the permitivity versus temperature, strongly indicate the relaxor

behavior for this Ba(Ti0.7Hf0.3)O3 ceramic. The quantitative

characterization and comparison of the relaxor behavior based

on empirical parameters (DTm, g, gTdif andDTrelax) confirms its

relaxor behavior. The TEM observations suggest that the

observed relaxor behavior is due to the coherent strained nano-

regions, which form on Hf substitution on Ti site.

Acknowledgements

Authors would like to acknowledge Dr P. Chaddah, the

Director UGC-DAE CSR, Prof V.N. Bhoraskar, the Ex-Director

UGC-DAECSR and Prof AjayGupta, the Center Director UGC-

DAE CSR-Indore for their encouragement and interest in the

work. Authors sincerely acknowledge Prof L.E. Cross for

providing some basic references and Dr Archna Jaiswal, for

helping at various stages. Partial support from DST is also

sincerely acknowledged.

S. Anwar et al. / Solid State Communications 138 (2006) 331–336336

References

[1] A.I. Kingon, S.K. Streiffier, C. Basceri, S.R. Summerfelt, MRS Bull. 21

(1996) 46.

[2] D.L. Polla, L.F. Francis, MRS Bull. 21 (1996) 59.

[3] Y. Sakabe, T. Takagi, K. Wakino, D.M. Smith, in: J.B. Blum,

W.R. Cannon (Eds.), Advances in Ceramics, Vol. 19, Am. Ceram. Soc.

Inc.,, Ohio, 1986, p. 103.

[4] F.J. Walker, A. McKee, Nanostruct. Mater 7 (1996) 221.

[5] Z.G. Lu, G. Calvarin, Phys. Rev. B 51 (1995) 2694.

[6] (a) L.E. Cross, Ferroelectric 76 (1987) 241;

(b) L.E. Cross, Ferroelectric 151 (1994) 305–320.

[7] N. Settler, L.E. Cross, J. App. Phy. 51 (8) (1980).

[8] M. Maglione, M. Belkaoumi, Phys. Rev. B 45 (1992) 2029.

[9] J.N. Lin, B. Wu, J. Appl. Phy. 68 (1990) 985.

[10] Victor P., Ranjith R., Krupanidhi S.B. J. App. Phy. 94 (12) 7702.

[11] (a) V.S. Tiwari, N. Singh, D. Pandey, J. Phys. Cond.Matter 7 (1995) 1441;

(b) Neelam Singh and Dhananjai Pandey. 8, 4269 (1996).

(c) Neelam Singh et al. 8, 7813–7827 (1996).

[12] Yu Zhi, Chen Ang, Ruyan Guo, A.S. Bhalla, J. App. Phy. 92 (2002)

2655.

[13] W.H. Payne, V.J. Tennery, J. AM. Ceram. Soc. 48 (1965) 413.

[14] V. Tura, L. Mitoseriu, Euro phys Letters 50 (6) (2000) 810.

[15] Z.Y. Cheng, R.S. Katiyar, X. Yao, Aqiang Guo, Phys. Rev. B 55 (1997)

8165.

[16] P.B. Hirsch, A. Howie, R.B. Nicholson, D.W. Pashley, J. Whelan,

Electron microscopy of thin crystals, Butterworths, London, 1969. pp.

327–337.

[17] K. Uchino, S Nomura, Integr. Ferroelectr. 44 (1982) 55.

[18] Chen Ang, Zhi Jing, Zhi Yu, J. Phys. Cond. Matter 14 (2002) 8901.

[19] J. Ravez, A Simon, Eur. J. Solid State Inorg. Chem. 37 (1997) 1199.

[20] J. Ravez, M. Pouchard, P. Hagenmuller, Eur. J. Solid State Inorg. Chem.

28 (1991) 1107.

[21] J. Ravez, A. Simon, Solid state Sci. 2 (2000) 525–529.