Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics

Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics

Materials Chemistry and Physics xxx (2013) 1e8 Contents lists available at SciVerse ScienceDirect Materials Chemistry and Physics journal homepage: ...

1MB Sizes 0 Downloads 26 Views

Materials Chemistry and Physics xxx (2013) 1e8

Contents lists available at SciVerse ScienceDirect

Materials Chemistry and Physics journal homepage: www.elsevier.com/locate/matchemphys

Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics Dhananjay K. Sharma a, *, Nawnit Kumar b, Seema Sharma a, Radheshyam Rai c a

Ferroelectric Research Laboratory, Department of Physics, A N College, Patna 800013, India Department of Physics and Meteorology, Indian Institute of Technology, Kharagpur 721302, India c Department of Ceramics and Glass Engineering and CICECO, University of Aveiro, 3810-193 Aveiro, Portugal b

h i g h l i g h t s  Detailed analysis and interpretation of impedance data of BT doped NBT reported.  Complex impedance analyses of compounds reveal grain and grain boundary effect.  Magnitude of Z0 decreases at lower frequencies with temperature showing NTCR effect.  Decrease in relaxation time with increasing temperature represents semiconducting behaviour.  Semiconducting behaviour of the compounds is believed to be due to the presence of oxygen defects.

a r t i c l e i n f o

a b s t r a c t

Article history: Received 4 December 2012 Received in revised form 28 March 2013 Accepted 26 April 2013

The near morphotropic phase boundary (MPB) compositions of lead-free piezoelectric ceramics based on sodium bismuth titanate (Na0.50Bi0.50TiO3: NBT) and barium titanate (BaTiO3: BT) were carefully investigated by conventional high temperature mixed-oxide method. All the ceramics exhibit single phase rhombohedral symmetry. The frequency (100 Hz to 1 MHz) and temperature (Room temperaturee500  C) dependence of impedance spectroscopy of (1  x)Na0.50Bi0.50TiO3exBaTiO3 (x ¼ 0.0, 0.06, 0.07 and 0.08) ceramics were investigated by impedance analyzer. The frequency explicit plots of Z00 versus frequency at various temperatures show peaks in the higher temperature range (>400  C). The compounds show dielectric relaxation, which is found to be of non-Debye type and the relaxation frequency shifted to higher side with increase in temperature. The activation energy values obtained for different BT content suggest that the electrical conduction in NBT is mainly due to the mobility of the ionized oxygen defects. Ó 2013 Elsevier B.V. All rights reserved.

Keywords: D. Crystal structure C. Piezoelectric C. Electron microscopy (SEM) D. Electrical properties

1. Introduction Piezoelectric materials play an important role for electronic devices such as actuators, accelerators, piezoelectric motors, transducers, filters and resonators, and microelectromechanical systems (MEMS). The most widely used piezoelectric materials are PbZrO3ePbTiO3 (PZT)-based multicomponent systems [1e4] because of their excellent piezoelectric properties. However, it is recently desired to use lead-free materials for environmental protection during the waste disposal of products. Therefore, lead-free piezoelectric materials have been attracting attention worldwide

* Corresponding author. Tel.: þ91 9199017908. E-mail addresses: [email protected] (D.K. Sharma), seema_sharma26@ yahoo.com (S. Sharma).

[5e8] as new materials in place of PZT-based piezoelectric ceramics. Lead-free piezoelectric materials, such as piezoelectric single crystals, e.g. langa-site [9], and ferroelectric ceramics with a perovskite structure [10e18], a tungsten bronze structure [19,20], and bismuth layer-structured ferroelectric [21e24], have been extensively reported. Recently, various perovskite-structured ferroelectrics such as BaTiO3 (Na0.5Bi0.5)TiO3, (Bi0.5K0.5)TiO3, KNbO3, (K,Na)NbO3, and their solid solutions have been actively studied as candidates for lead-free piezoelectric ceramics. Sodium Bismuth Titanate, (Na0.5Bi0.5)TiO3, is a perovskitestructured ferroelectric with rhombohedral symmetry (R3C) at room temperature (RT) and its phase transition is complicated. The phase transition temperatures, from rhombohedral to tetragonal and from tetragonal to cubic are approximately 340  C and 540  C on heating, respectively, for NBT single crystals. NBT ceramic shows the strong ferroelectric properties of a large remanent polarization

0254-0584/$ e see front matter Ó 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.matchemphys.2013.04.038

Please cite this article in press as: D.K. Sharma, et al., Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.04.038

and relatively high piezoelectric properties compared with other lead-free piezoelectric ceramics. Therefore, the NBT is considered to be an excellent candidate as a key material of lead-free piezoelectric ceramics. However, NBT ceramic is difficult to pole due to a large coercive field [25]. In the recent 2 decades, NBT-based solid solutions [17,19e32] and A-site substituted NBT [25e28] that can be poled easily were extensively studied. Among these NBT-based systems, (1  x) (Na0.5Bi0.5)TiO3exBaTiO3 is more focused. Takenaka et al. [29] pointed out that at room temperature it has a rhombohedral (Fa)etetragonal (Fb) Morphotropic Phase Boundary (MPB) at x ¼ 0.06e0.07 where the system shows outstanding piezoelectric and dielectric properties of all the lead free ceramic resources. In the present work, the 1  x(Na0.5Bi0.5TiO3)ex(BaTiO3) system with x ¼ 0.0, 0.06, 0.07, and 0.08, (NBTeBT) near MPB were prepared by the conventional solid state reaction method and their structural, microstructural and electrical properties were studied systematically. 2. Experimental procedure Polycrystalline samples of (1  x)Na0.50Bi0.50TiO3exBaTiO3 (NBTeBT), where (x ¼ 0.0, 0.06, 0.07, and 0.08), were synthesized from high purity oxides of Bi2O3 (99.9%, Aldrich Chem. Co), Na2CO3 (99.5%, Aldrich Chem. Co), BaCO3 (99.9%, Aldrich Chem. Co.) and TiO2 (99.9%, Aldrich Chem. Co.) using high temperature solid state reaction technique in an ambient atmosphere. The constituent compounds in suitable stoichiometry were thoroughly mixed in a ball milling unit for 24 h. The calcined fine powder was cold pressed into cylindrical pellets of 10 mm in diameter and 1e2 mm in thickness using a hydraulic press at a pressure of 6  107 kg m2. These pellets were sintered between 1000 and 1050  C for 2 h in air. The formation and quality of compounds were verified by X-ray diffraction (XRD) technique. The XRD pattern of the compounds was recorded at room temperature using X-ray powder diffractometer (Rigaku Minifiex Japan) with Cu Ka radiation (¼1.5418  A) in a wide range of Bragg angles 2q (20  2q  80 ) at a scanning rate of 2 min1. The microstructural examination was carried out by scanning electron microscopy (SEM) using Jeol 6300 and Philips XL30, equipped with energy dispersive spectrometer (EDS) for chemical analysis. For electrical measurements, the polished surfaces of the sintered pellets were electroded with air drying silver paste. Impedance spectroscopy of the compounds was investigated using an Impedance Analyzer (Newtons 4th Limited) as a function of frequency at room temperature (RT) and temperature (RT to 500  C) at different frequencies (100 Hz to 1 MHz). 3. Result and discussion Fig. 1 shows the X-ray diffraction patterns of NBTeBT ceramics. The spectra examined are single phase with perovskite type structure without any secondary phases. At room temperature, the symmetry of NBT is rhombohedral and BT is tetragonal. Our solid solutions have rhombohedral morphotropic phase boundary. The narrow and symmetric X-ray diffraction peaks of the NBTeBT compounds indicate homogeneity and good crystallization of the samples. All the reflection peaks were indexed using observed interplanar spacing d, and lattice parameters of compound were determined using a least-square refinement method. A good agreement between calculated and observed d values of all diffraction lines (reflections) of NBTeBT system with different x content suggests that the basic crystal structure is rhombohedral. It is also noted that the diffraction peaks shift slightly towards low diffraction angles as x increases means lattice parameter increases with increase in x. This may be attributed to the larger ionic radii of Ba2þ than those of (Na0.50B0.50)þ and Ti4þ. In addition, it was seen

Intensity (arb. unit)

D.K. Sharma et al. / Materials Chemistry and Physics xxx (2013) 1e8

Intensity (arb. unit)

2

(110)

(111)

(d)(100)

(200)

(211) (210)

2θ (degrees) (220) (221) (310)

(c) (b) (a) 20

30

40

50

60

70

80

2 (degrees) Fig. 1. Room temperature XRD pattern of (1  x)NBTexBT, (a) x ¼ 0.00, (b) x ¼ 0.06, (c) x ¼ 0.07 and (d) x ¼ 0.08.

that the cell volumes of the ceramics increased gradually with increasing x, which can explain why the diffraction peaks shifted slightly towards low diffraction angles as x increased. Speculation regarding the A-site substitution is made based on the random solid solution model, ionic radii, and coordination number of the introduced guest ion in respect to the A- and B-sites. Since the perovskite is a closed-packed structure, the possible interstitial sites are bounded by charged ions (negative or positive ions). This implies that there is a smaller chance of having a high concentration of interstitial sites. Therefore the cation lattice sites (A or B) are locations with the highest possibility of occupation by the guest ion. The ionic radii of B-site ion Ti4þ with a coordination number (CN of 6) is 0.60  A respectively. The ionic radii of A-site ions such as Naþ and Biþ (CN of 6) are 1.02  A and 1.03  A respectively. Thus due to the close proximity of the ionic radii of Naþ and Biþ3, it is postulated  that the Ba2þ (CN: 6 and rþ Ba: 1.35 A) substitutes for the potassium in the A-site of perovskite lattice. Fig. 2 exhibits the microstructure (SEM) of the compounds with different x values. The fracture surfaces show homogeneous grains and dense structure which is in good agreement with the density (5.65 g cm3), determined experimentally by Archimedes method. The introduction of BT in NBTeBT ceramics results in the reduction of grain size of the parent compound. It is evident from the figure that with BaTiO3 content above 6%, the grain size is well smaller than 1 mm. For PZT- and PLZT-based ceramics, the substitution of Ba2þ for Pb2þ generally leads to a great inhibition of grain growth. Similar to Ba-modified lead-based ceramics, in the NBTeBT ceramics, grain growth is inhibited after the introduction of BaTiO3 into (Na0.5Bi0.5)TiO3. Fig. 3 represents the dielectric spectrum 3 0 (u) of (1  x)NBTexBT (x ¼ 0, 0.06, 0.07 and 0.08) at different temperatures. The observed variation of 3 0 with frequency can be attributed to frequency relaxation in the material. A high degree of dispersion of the permittivity is identified at high temperature (from 300  C) and at low frequencies (<1 kHz). This behaviour is found for dielectric materials, in which a mechanism conduction of the hopping type is present [30]. Inset of Fig. 3 shows the 3 00 (u) as a function of frequency at several temperatures for all the compositions. All the curves show an intense increase of the loss magnitude at increased temperature compared to their room temperature values. At high frequencies, the losses are much lower than one occurring at low frequencies.

Please cite this article in press as: D.K. Sharma, et al., Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.04.038

D.K. Sharma et al. / Materials Chemistry and Physics xxx (2013) 1e8

3

Fig. 2. Scanning electron micrographs of (1  x)NBTexBT, (a) x ¼ 0.00, (b) x ¼ 0.06, (c) x ¼ 0.07 and (d) x ¼ 0.08.

3

35

2

0 0.1

1

10

100

3 2

48

0

1000

0.1

100

RT 100 200 300 400 500

ε ''× 10

3

75

ε '× 103

4

50

18 C C C C C

25 0

2

0.1

1

10

100

1000

10

100

1000

1 10 100 Frequency (kHz)

(d)

240

3

(c)

0.1

1000

3

1 10 100 Frequency (kHz)

160

12

ε '×10

0.1

1

Frequency (Hz)

1

0

C C C C C

24

Frequency (kHz)

1

6

RT 100 200 300 400 500

72

ε ''× 10

3

ε ''× 10

3

70

(b)

4

ε ''× 10

ε '×10

3

4

RT 100 C 200 C 300 C 400 C 500 C

105

3

(a)

ε '× 10

5

RT 100 C 200 C 300 C 400 C 500 C

80

0 0.1

6

1000

1

10

100

1000

Frequency (kHz)

Frequency (kHz)

0

0 0.1

1 10 100 Frequency (Hz)

1000

0.1

1 10 100 Frequency (kHz)

1000

Fig. 3. Variation of 3 0 with frequency and in inset variation of 3 00 with frequency at selective temperatures of (1  x)NBTexBT, (a) x ¼ 0.00, (b) x ¼ 0.06, (c) x ¼ 0.07 and (d) x ¼ 0.08.

Please cite this article in press as: D.K. Sharma, et al., Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.04.038

4

D.K. Sharma et al. / Materials Chemistry and Physics xxx (2013) 1e8

This kind of dependence of the dielectric loss with frequency is associated with losses due to the conduction mechanism which happens at high temperatures for ceramics. Similar behaviour has also been reported for Bi-modified lead lanthanum zirconium titanate material at high temperature [31]. Fig. 4 shows temperature dependent spectra (Nyquist plot) of NBTeBT material. The impedance spectrum is featured by semicircular arcs. The nature of variation of the arcs with temperature and frequency provides various clues of the materials. The impedance spectra are characterized by the appearance of a single semicircular arc and the intercept of the semicircular arc with the real axis (Z0 ) gives us an estimate of the bulk resistance (Rb) of the material. It has been observed that the bulk resistance of the material decreases with increase in temperature showing a typical semiconducting property, i.e. negative temperature coefficient of resistance (NTCR) behaviour. It is observed that with the increase in temperature the slope of the lines decrease and the lines bend towards real (Z0 ) axis and at 450  C; a semicircle could be traced, indicating the increase in conductivity of the sample. It can also be observed that the peak maxima of the plots decrease and the frequency for the maximum shifts to higher values with the increase in temperature. It can be noticed that the complex impedance plots are not represented by full semicircle, rather the semicircular arc is depressed and the centre of the arc lies below the real (Z0 ) axis suggesting the relaxation to be of polydispersive non-Debye type in samples. This may be due to the presence of distributed elements in the material electrode system [32,33]. The correlation among the Debye relaxators may start developing via formation of nanopolar clusters of NaeTiO3 and BieTiO3 [34,35]. Since the relaxation times of the relaxators within polar clusters are distributed over a wide spectrum at higher temperatures, their response to external field is in a different time domain. This results in the deviation from Colee

(a)

(c)

Cole plots. It is clear from figure that with the increase of measuring temperature ColeeCole plots become stretched and/or splitted into two discrete semicircles, inferring the possible average profile of various ColeeCole semicircles. The split as well as stretched semicircles may be due to secondary elements like interfacial capacitance or defects. According to Debye’s model, a material having single relaxation time gives rise to an ideal semicircle centred on the real axis. The first semicircle (at higher frequencies) attributed to transport phenomena in the bulk/grain and another at lower frequencies, is related to transport phenomena at grain boundary [36,37]. The low frequency second dispersion curve has been assigned to the grain boundary (blocking core) conduction. It can be seen from Fig. 5 that both grain and grain boundary resistance decreases with increase in temperature which indicate the NTCR behaviour. It also indicates that the grain boundary effect has assisted in lowering the barrier to the motion of charge carriers paving the way for increased electrical transport with rise in temperature. An equivalent circuit is being used to provide a complete picture of the system and establish the structural property relationship of the materials. Comparison of complex impedance plots (symbols) with fitted data (lines) using commercially available software ZSimp WIN Version 2 has been given in the figure. To Model the non-Debye response, constant phase element (CPE) is used in addition to resistors and capacitors. It has also been clearly observed from the Nyquist plots that the influence of grain size on the intergrain resistivity increases with decreasing grain size. This plot indicates the changes in grain boundary resistance at elevated temperatures representing the role of grain boundaries in electrical conduction process of the material. The decrease in grain boundary resistance with rise in temperature may be due to the lowering of barrier favouring the increase of mobility of charge carriers that adds to the conduction process. The conduction mechanism for

(b)

(d)

Fig. 4. Nyquist plot of (1  x)NBTexBT at selective temperatures, (a) x ¼ 0.00, (b) x ¼ 0.06, (c) x ¼ 0.07 and (d) x ¼ 0.08.

Please cite this article in press as: D.K. Sharma, et al., Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.04.038

D.K. Sharma et al. / Materials Chemistry and Physics xxx (2013) 1e8

Z'' (k )

80

270

C C C C C

Z' (k )

Z' (k )

200

400 425 450 475 500

40

0 0.1

100

1

10

100

1000

Frequency (kHz)

180

70

C C C C C

35

0 0.1

90

0

1

10

100

1000

Frequency (kHz)

0

0.1

1

10

100

1000

0.1

1

Frequency (kHz)

(c)

160

400 425 450 475 500

80 40

200

270

C C C C C

Z' (k )

Z'' (k )

120

300

0 0.1

1

10

100

1000

Frequency (kHz)

10

100

120

(d)

180

400 425 450 475 500

80

Z'' (k )

400

Z' (k )

400 425 450 475 500

105

(b)

Z'' (k )

120

(a)

300

5

C C C C C

40

0 0.1

90

1000

1

10

100

1000

Frequency (kHz)

Frequency (kHz)

100

0

0 0.1

1

10

100

1000

0.1

1

10

100

1000

Frequency (kHz)

Frequency (kHz)

Fig. 5. Variation of Z0 with frequency and in inset variation of Z00 with frequency at selective temperatures of (1  x)NBTexBT, (a) x ¼ 0.00, (b) x ¼ 0.06, (c) x ¼ 0.07, and (d) x ¼ 0.08.

400 425 450 475 500

2

400 425 450 475 500

C C C C C

C C C C C

2

7

0 0.1

1

10

100

1000

Frequency (kHz)

1

400 425 450 475 500

400 425 450 475 500

8

(b) M'' (10-3)

M'' (10-3)

3

M'' (10-4)

(a)

M'' (10-4)

14

4

C C C C C

C C C C C

4

0 0.1

1

10

100

1000

Frequency (kHz)

1

1

2

3

0 0

4

1

M' (10-3 )

400 425 450 475 500

C C C C C

6

C C C C C

5

(d)

2 0 0.1

1

10

100 1000

Frequency (kHz)

3 2

400 425 450 475 500

20

4

4

M'' (10)-3

M'' (10-3)

3

2

400 425 450 475 500

8

(c) M'' (10-4)

4

2

M' (10 -3)

400 425 450 475 500

C C C C C

M'' (10-4)

0 0

C C C C C

10

0 0.1

1

10

100 1000

Frequency (kHz)

1

1 0 0

1

2

M' (10 -3)

3

4

0 0

1

2

3

M' (10)-3

4

5

Fig. 6. M0 vs M00 plot and in inset variation of M00 with frequency at selective temperatures of (1  x)NBTexBT, (a) x ¼ 0.00, (b) x ¼ 0.06, (c) x ¼ 0.07, and (d) x ¼ 0.08.

Please cite this article in press as: D.K. Sharma, et al., Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.04.038

6

D.K. Sharma et al. / Materials Chemistry and Physics xxx (2013) 1e8

grains and grain boundaries are different. A hopping mechanism through various defect sites contributes the intergrain conduction whereas the interface barrier potential plays the major role for intragrain boundary condition. In both cases, resistance of grain and grain boundary decreases with temperature showing the negative temperature coefficient of resistance (NTCR) behaviour like that of a semiconductor. Fig. 5 shows the variation of real part of impedance (Z0 ) as a function of frequency (100 Hz to 1 MHz) at different temperatures for various x values. The pattern shows a sigmoidal variation as a function of frequency in the low frequency region followed by a saturation region in the high frequency region. This suggests the presence of mixed nature of polarization behaviour in the material. The pattern also shows a very steep Z0 dispersion in the low frequency region of the spectrum followed by a plateau region in the high frequency region. The extent of steepness in the low frequency region is observed to have a very strong dependence of Z0 on the composition of BaTiO3 irrespective of temperature. Further, the impedance value is observed to decrease at higher temperatures. A decreasing trend of Z0 with rise in temperature suggests the presence of negative temperature coefficient of resistance (NTCR) in the material in the low frequency region but tends to merge in the high frequency region at almost all temperatures. These results indicate a possibility of increase in a.c. conductivity with rise in temperature in the high frequency region (possibly) due to the release of space charge, and lowering in the barrier properties of the material. Inset of Fig. 5 exhibits frequency and temperature dependence of imaginary part of impedance (Z00 ) for NBT and BT solid-solutions. In low temperature/frequency range, Z00 has very high value for all the compositions which decreases with rise in frequency and

attains low values. This is an interesting (unusual) trend in the variation of Z00 with frequency/temperature for NBTeBT composite. The effect of BaTiO3 substitution on the electrical behaviour of the sample can clearly be seen with the appearance of peaks in the impedance spectrum at and above 400  C for x ¼ 0.00e0.08. This observation clearly suggests the presence of electronic, dipolar and space charge polarizations in the material, and the appearance of the peak (s) being a result of dipolar contribution. The appearance and nature of peaks at a characteristic angular frequency umax (¼2pfmax) provides information about the type and strength of the dielectric relaxation phenomenon occurring in the material. At low temperatures (400  C), the spectrum shows a monotonous decrease having dispersive nature of pattern in the low frequency region followed by a plateau at the higher frequencies indicating weaker relaxations. The width of the peaks suggests a spread of relaxation times, which may involve more than two equilibrium positions. The spread factor is represented by the half-width of the peaks. The substitution of Baþ for (NaBi)þ2 may alter the long-range polar order with TiO6 octahedra and this may result in a disturbance in the local polarizations present in the lattice. The samples contain defects or impurityedefect complexes and these would have a corresponding effect on relaxation. At lower temperatures, the complexes do not have an orientation effect and no peaks are observed. The relaxation species may not relax at lower temperatures as there is a polarization field in the lattice. Electrical response of the materials can also be analyzed through complex electric modulus formalism, which provides an alternative approach based on polarization analysis. Complex electric formalism gives the inhomogeneous nature of the polycrystalline ceramic, which can be probed into bulk and grain

Fig. 7. Variation of M0 with frequency at selective temperatures of (1  x)NBTexBT, (a) x ¼ 0.00, (b) x ¼ 0.06, (c) x ¼ 0.07 and (d) x ¼ 0.08.

Please cite this article in press as: D.K. Sharma, et al., Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.04.038

D.K. Sharma et al. / Materials Chemistry and Physics xxx (2013) 1e8

boundary effects. The complex electric modulus (M*) has been calculated from the complex impedance data using the following relations: complex modulus, M* (u) ¼ M þ jM ¼ juC0Z*, where M (real part) ¼ uC0Z and M (imaginary part) ¼ uC0Z (u ¼ angular frequency i.e., 2pf/C0 ¼ geometrical capacitance ¼ 3 0A/t, 3 0 ¼ permittivity of free space, A ¼ area of the electrode surface and t ¼ thickness). The complex electric modulus spectrum M0 versus M00 is shown in Fig. 6 for all the samples at different temperatures. The patterns are characterized by the presence of little asymmetric and depressed semicircular arcs whose centre does not lie on M0 axis. The behaviour of electric modulus spectrum is suggestive of the temperature dependent hopping type of mechanism for electric conduction (charge transport) in the system and non-Debye type dielectric relaxation. In a relaxation system, one can determine the probable relaxation time (s) from the position of the loss peak in the Z00 as well as M00 vs. log f plots according to the relation: s ¼ 1/ u ¼ 1/2pf (f is the relaxation frequency). It may be noted from inset of Fig. 6 that the position of the peak 00 Mmax shifts to higher frequencies as the temperature is increased. The frequency region below peak maximum M00 determines the range in which charge carriers are mobile on long distances. At frequency above peak maximum, the carriers are confined to potential wells, being mobile on short distances. The peaks are asymmetric and broader than the ideal Debye curve. The frequency range where the peaks occur is indicative of transition from long range to short range mobility [37,38]. Fig. 7 displays the frequency (angular) dependence of M0 (u) for NBTeBT as a function of temperature. The variation of M0 with frequency for NBTeBT samples shows a dispersion tending towards MN (the asymptotic value of M0 at higher frequencies) and it (dispersion) shifts towards higher frequency side as temperature increases. The asymmetric plot of M0 is because of the stretched exponential character of relaxation time of the material. Monotonous dispersion on increasing frequency at lower temperatures may be caused by short range mobility of charge carriers. Such results may possibly be related to a lack of restoring force governing the mobility of the charge carriers under the action of an induced electric field. The value of M0 decreases with rise in temperature in the observed frequency range. The nature of variation of relaxation time s with temperature of the compounds for impedance plot has been shown in Fig. 8. The

-3

4. Conclusions NBTeBT ceramics have been fabricated by solid state reaction technique and its structure, microstructure and impedance properties have been studied systematically. The results of XRD measurement reveal that BT has diffused into the NBT lattice to form a solid solution in single phase perovskite structure with rhombohedral symmetry. Displacements of the XRD peaks upon doping suggest that Ba enters the A site of the perovskite structure. SEM micrographs show spherical morphology with uniform particle distribution. The ceramics have a dense fine microstructure when sintered at 1020  C for 2 h. Impedance analyses indicated the presence of grain and grain boundary effect in NBTeBT samples. The compounds showed dielectric relaxation, which is found to be of non-Debye type and the relaxation frequency shifted to higher side with the increase of temperature. The Nyquist plot and conductivity studies showed the NTCR character of samples. The activation energy values obtained for different BT content suggest that the electrical conduction in NBT is mainly due to the mobility of the ionized oxygen defects.

One of the authors, Nawnit Kumar would like to thank Council of Scientific and Industrial Research, New Delhi for research fellowship.

1.8

Ea (eV)

00 frequency u (corresponding to Zmax ) gives the most probable relaxation time s from the condition us ¼ 1. All the curves find to follow the Arrehenius relation: s ¼ so exp(Ea/kT) where so is pre exponential factor, Ea is the activation energy, k is Boltzmaan constant and T is the absolute temperature. The s-value of the NBTeBT compounds was found to be decreasing on increasing temperature, which is a typical behaviour of a semiconductor. The semiconducting nature of the grains in ceramics is believed to be due to the loss of oxygen during high temperature sintering process. The activation energy (Ea) was evaluated from the slope of log s against 103 /T curve. The Ea values (inset Fig. 8) show that the addition of BaTiO3 in NBT decreases the activation energy and was minimum for BT ¼ 0.07 which further increased abruptly for BT ¼ 0.08. The % error in the evaluation of the Ea for x ¼ 0.00, 0.06, 0.07 and 0.08 are 0.1, 0.3, 0.4 and 0.2 respectively. The values of Ea show that the dielectric relaxation processes in the studied samples are closely related to the oxygen vacancies, which have been reported as the most mobile ionic defects in perovskites [39]. According to Steinsvik et al. [40], the activation energy in the ABO3 perovskite structure decreases with increasing oxygen vacancy content.

Acknowledgements

1.9

10

7

1.7 1.6

References

τ (sec)

1.5

10

10

-4

0.00

0.02

0.04

0.06

0.08

x

x = 0.00 x = 0.06 x = 0.07 x = 0.08 Linear Fit

-5

1.25

1.30

1.35 1.40 3 -1 10 /T (K )

1.45

Fig. 8. Variation of relaxation time with inverse of temperature and in inset variation of activation energy (Ea) with x value for (1  x)NBTexBT.

[1] E. Sawaguchi, J. Phys. Soc. Jpn. 8 (1953) 615. [2] B. Jaffe, R.S. Roth, S. Marzullo, J. Appl. Phys. 25 (1954) 809. [3] B. Jaffe, W.R. Cook, H. Jaffe, Piezoelectric Ceramics, Academic, New York, 1971, p. 142. [4] T. Yamamoto, Jpn. J. Appl. Phys. 35 (1996) 5104. [5] T. Takenaka, Piezoelectric properties of some lead-free ferroelectric ceramics, Ferroelectrics 230 (1999) 87. [6] Y. Saito, H. Takao, T. Tani, T. Nonoyama, K. Takatori, T. Homma, T. Nagaya, M. Nakamura, Lead free piezoceramics, Nature 42 (2004) 84. [7] T. Takenaka, H. Nagata, J. Eur. Ceram. Soc. 25 (2005) 2693. [8] T. Takenaka, H. Nagata, Y. Hiruma, Jpn. J. Appl. Phys. 47 (2008) 3787. [9] K. Shimamura, H. Takeda, T. Kohno, T. Fukuda, J. Cryst. Growth 163 (1996) 388. [10] H. Ishii, H. Nagata, T. Takenaka, Jpn. J. Appl. Phys. 40 (2001) 5660. [11] H. Nagata, M. Yoshida, Y. Makiuchi, T. Takenaka, Jpn. J. Appl. Phys. 42 (2003) 7401. [12] Y. Hiruma, R. Aoyagi, H. Nagata, T. Takenka, Jpn. J. Appl. Phys. 43 (2004) 7556. [13] Y. Makiuchi, R. Aoyagi, Y. Hiruma, H. Nagata, T. Takenaka, Jpn. J. Appl. Phys. 44 (2005) 4350. [14] Y. Hiruma, R. Aoyagi, H. Nagata, T. Takenaka, Jpn. J. Appl. Phys. 44 (2005) 5040.

Please cite this article in press as: D.K. Sharma, et al., Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.04.038

8

D.K. Sharma et al. / Materials Chemistry and Physics xxx (2013) 1e8

[15] H. Nagata, T. Shinya, Y. Hiruma, T. Takenaka, I. Sakaguchi, H. Haneda, Ceram. Trans. 167 (2005) 213. [16] K. Yoshii, Y. Hiruma, H. Nagata, T. Takenaka, Jpn. J. Appl. Phys. 45 (2006) 4493. [17] S. Zhang, T.R. Shrout, H. Nagata, Y. Hiruma, T. Takenaka, IEEE Trans. Ultrason. Ferroelectr. Freq. Control 54 (2007) 910. [18] Y. Hiruma, H. Nagata, T. Takenaka, IEEE Trans. Ultrason. Ferroelectr. Freq. Control 54 (2007) 2493. [19] M. Kimura, T. Minamikawa, A. Ando, Y. Sakabe, Jpn. J. Appl. Phys. 36 (1997) 6051. [20] Y. Doshida, S. Kishimoto, K. Ishii, H. Kishi, H. Tamura, Y. Tomikawa, S. Hirose, Jpn. J. Appl. Phys. 46 (2007) 4921. [21] T. Takenaka, K. Sakata, J. Appl. Phys. 55 (1984) 1092. [22] T. Takenaka, K. Sakata, Ferroelectrics 94 (1989) 175. [23] T. Takenaka, J. Ceram Soc. Jpn. 110 (4) (2002) 215. [24] Zuo-Guang Ye (Ed.), Handbook of Advanced Dielectric, Piezoelectric and Ferroelectric Materials (Part VII Novel Processing and New Materials, No. 27 “Grain Orientation and Electrical Properties of Bismuth Layer-structure Ferroelectrics”), Woodhead Publishing Limited, CRC Press LLC, 2008, p. 818. [25] J.K. Yi, J.-K. Lee, K.-S. Hong, J. Am. Ceram. Soc. 85 (2003) 3004. [26] A. Herabut, A. Safari, J. Am. Ceram. Soc. 80 (1997) 2954.

[27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40]

Y. Watanabe, Y. Hiruma, H. Nagata, T. Takenaka, Ceram. Int. 34 (2008) 761. S.S. Sundari, Binay Kumar, R. Dhanasekaran, Ceram. Int. 39 (2013) 555. T. Takenaka, K.I. Maruyama, K. Sakata, Jpn. J. Appl. Phys. 30 (1991) 2236. A.K. Jonscher, J. Mater. Sci. 16 (1981) 2037. S. Dutta, R.N.P. Choudhary, P.K. Sinha, Eur. Phys. J. Appl. Phys. 36 (2006) 141. J.R. MacDonald (Ed.), Impedance Spectroscopy Emphasizing Solid Materials and Systems, Wiley, New York, 1987. K.S. Rao, B. Tilak, K.C.V. Rajulu, A. Swathi, H. Workineh, J. Alloy. Compd. 509 (2011) 7121. K. Prasad, K. Kumari, K.P. Chandra, K.L. Yadav, S. Sen, Mater. Sci. Poland 27 (2009) 2. P. Bonneau, O. Garnier, G. Calvarin, E. Husson, J.R. Gavarri, A.W. Hewat, A. Morrel, J. Solid State Chem. 91 (1991) 350. S. Zha, C. Xia, G. Meng, J. Power Sources 115 (2003) 44. A. Shukla, R.N.P. Choudhary, Phys. B: Condens. Matter 406 (2011) 2492. A. Shukla, R.N.P. Choudhary, A.K. Thakur, J. Phys. Chem. Solids 70 (2009) 1401. S. Poykko, D.J. Chadi, Appl. Phys. Lett. 76 (2000) 499. S. Steinsvik, R. Bugge, J. Gjonnes, J. Tafto, T. Norby, J. Phys. Chem. Solids 58 (1997) 969.

Please cite this article in press as: D.K. Sharma, et al., Effect of BaTiO3 additive on the electrical properties of Na0.50Bi0.50TiO3 lead free ceramics, Materials Chemistry and Physics (2013), http://dx.doi.org/10.1016/j.matchemphys.2013.04.038