Top Banner
University of Wollongong Research Online University of Wollongong esis Collection University of Wollongong esis Collections 2013 Lead-free piezoelectric materials Abolfazl Jalalian University of Wollongong Research Online is the open access institutional repository for the University of Wollongong. For further information contact the UOW Library: [email protected] Recommended Citation Jalalian, Abolfazl, Lead-free piezoelectric materials, Doctor of Philosophy thesis, Institute for Superconducting and Electronic Materials, University of Wollongong, 2013. hp://ro.uow.edu.au/theses/4037
129

Lead-free piezoelectric materials - Research Online

May 06, 2023

Download

Documents

Khang Minh
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Lead-free piezoelectric materials - Research Online

University of WollongongResearch Online

University of Wollongong Thesis Collection University of Wollongong Thesis Collections

2013

Lead-free piezoelectric materialsAbolfazl JalalianUniversity of Wollongong

Research Online is the open access institutional repository for theUniversity of Wollongong. For further information contact the UOWLibrary: [email protected]

Recommended CitationJalalian, Abolfazl, Lead-free piezoelectric materials, Doctor of Philosophy thesis, Institute for Superconducting and ElectronicMaterials, University of Wollongong, 2013. http://ro.uow.edu.au/theses/4037

Page 2: Lead-free piezoelectric materials - Research Online
Page 3: Lead-free piezoelectric materials - Research Online

Department of

Institute for Superconducting and Electronic Materials

Lead-Free Piezoelectric Materials

ABOLFAZL JALALIAN

"This thesis is presented as part of the requirements for the award of the Degree of Doctor of Philosophy

of the University of Wollongong"

Sep. / 2013

Page 4: Lead-free piezoelectric materials - Research Online

i

DECLARATION

I, Abolfazl Jalalian, declare that this thesis, submitted in partial fulfilment of the

requirements for the award of Doctor of Philosophy, in the Institute for

Superconducting & Electronic Materials, University of Wollongong, is entirely my

own work unless otherwise referenced or acknowledged. The document has not been

submitted for qualifications at any other academic institution.

Abolfazl Jalalian

September, 2013

Page 5: Lead-free piezoelectric materials - Research Online

ii

ABSTRACT

Quasi-one-dimensional (1D) materials, including nanofibers, nanotubes, nanowires,

and nanobelts, have been exploited widely in nanogenerators, sensors, transducers,

microelectromechanical systems (MEMS) devices, and other applications, such as

microwave varactors and ferroelectric field effect transistors. Currently available

nanostructured piezoelectric materials show a low piezoelectric coefficient d33 of

merely 100 pm V-1, with Pb(Zr, Ti)O3 (PZT)-based materials at the high end. The

health impact of lead poisoning is well known, however, and intensive efforts have

begun to discover new lead-free piezoelectric compounds which possess comparable

piezoelectric performances to those of the lead-based piezoelectric materials.

Recently, a lead-free (1-x)Ba(Ti0.80Zr0.20)O3-x(Ba0.70Ca0.30)TiO3 ((1-x)BTZ-xBCT)

piezoelectric system with optimal composition of x = 0.5 was reported to show

superior room temperature piezoelectricity, with the piezoelectric coefficient d33 =

620 pC N-1, the piezoelectric voltage constant g33 = 15.38 × 10-3 Vm N-1, and the

electromechanical (converse piezoelectric) response as high as 1140 pm V-1. These

superior piezoelectric properties are comparable to or higher than those of state-of-

the-art PZT or other lead-free piezoelectric compounds, due to the low polarization

anisotropy and low energy barrier for lattice distortions in the morphotropic phase

boundary (MPB) region.

The main work presented in this dissertation is focused on the synthesis and

characterization of Ba (Ti0.80Zr0.20) O3-(Ba0.70Ca0.30) TiO3 (BTZ-BCT) in different

forms including ceramics, thin films, and nanofibers.

Different structural analysis techniques, including X-ray diffraction (XRD), Raman

spectroscopy, and transmission electron microscopy (TEM), have been employed to

aj393
Highlight
aj393
Highlight
Page 6: Lead-free piezoelectric materials - Research Online

iii

investigate the evolution of the crystal structure and phase content in the samples.

The coexistence of two ferroelectric phases, tetragonal and rhombohedral, and

crystallization of the ceramics, the fibers, and the thin films in the vicinity of the

MPB region have been demonstrated. The lattice constants have been defined using

the Rietveld method. The impact of lattice parameter variations on the ferroelectric

properties of the (1-x)BTZ-xBCT ceramics has been investigated.

Different scanning probe microscopy techniques, including piezoresponse force

microscopy (PFM), scanning capacitance microscopy (SCM), and scanning

spreading resistance microscopy (SSRM), have been employed to study the

piezoelectric, ferroelectric domain switching, and the electrical properties of the

nanofibers and thin films. Very large piezoelectricity in low-dimensional BTZ-BCT

sintered as thin films (d33 = 141 pm V-1) and nanofibers (d33 = 180 pm V-1) has been

achieved. These values are comparable to those of PZT films and nanofibers.

Observations of ferroelectric nanodomains with high spatial resolution using SCM

and PFM techniques are also presented. The influences of lateral size, geometry, and

the clamping effect on the piezoelectric performance are investigated for both the

thin films and the nanofibers. The current distribution and resistivity have been

studied by SSRM. The results show a uniform distribution of resistance and very

high resistance of 1010 ohms in the BTZ-BCT nanofibers. Combining a high

piezoelectric coefficient with environmental benefits, the BTZ-BCT nanostructures

provide the superior functions that are in demand for highly efficient piezoelectric

devices and electromechanical systems.

In the last chapter of the thesis, the synthesis and characterization of biocompatible

and piezoelectric (Na,K)NbO3 (NKN) nanofibers are presented. The X-ray

diffraction pattern of the nanofibers reveals a pure single phase with polar structure

aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
Page 7: Lead-free piezoelectric materials - Research Online

iv

after annealing at 700°C. TEM images and electron diffraction patterns show the

growth of NKN single crystals in the form of nanofibers. The ferroelectric domain

switching and piezoelectric response of the nanofibers have been investigated using

PFM. A higher piezoelectric response is achieved in NKN nanofibers (d33 = 58 pm

V-1) than in its thin films (d33 = 40 pm V-1). Owing to the existence of permanently

charged regions in the NKN nanofibers known as ferroelectric domains, electrical

signals can be generated in them via the piezoelectric effect to provide a new

opportunity for construction of a smart biocompatible scaffold that can be used for

repair, engineering, and regeneration of damaged tissues.

Page 8: Lead-free piezoelectric materials - Research Online

v

ACKNOWLEDGEMENTS

Firstly, I would like to express my sincere gratitude to my supervisor, Prof. S.X.

Dou, and my co-supervisor, Prof. X.L. Wang, for their continuous academic

guidance and support during my PhD study at the Institute for Superconducting and

Electronic Materials (ISEM), University of Wollongong.

It is my pleasure to pay tribute to Prof. A.M. Grishin, head of the Department of

Condensed Matter Physics, at the Royal Institute of Technology (KTH), Sweden, for

giving me the opportunity to work in his laboratory as a visiting PhD student and

providing me with the opportunity to attend very worthwhile courses and workshops

in scanning probe microscopy techniques. Furthermore, I would also like to thank

Dr. S. Khartsev for his technical assistance and Prof. A. Hallen and Dr. A. Srinivasan

for the use of an atomic force microscope with a scanning capacitance module.

I acknowledge Dr. Tania Silver for her help in proofreading and correction the

English in the thesis and Dr. Z.X. Cheng for his review and comments on my thesis.

I am grateful to Dr. D. Wexler, Mr. D. Attard, and Dr. M. Higgins for their

invaluable guidance and training courses in microscopy techniques in ISEM. I am

also thankful to the administrative assistants, Ms. Roberta Lynch and Mrs. Crystal

Mahfouz. I thank all my friends for the memorable times that we had together.

I would especially like to convey my gratitude to my fiancée, Ms. Nioosha Nasseh,

and all my family for their warmth, enthusiasm, love, and boundless moral support

over my entire PhD journey. Their support and care helped me to stay focused on my

work.

Finally, I would like to acknowledge the Australian Research Council (ARC) for

Discovery grant DP0879070, and the University of Wollongong and ISEM for

Page 9: Lead-free piezoelectric materials - Research Online

vi

providing me with a Matching Scholarship and International Postgraduate Tuition

Award during my PhD study.

Page 10: Lead-free piezoelectric materials - Research Online

vii

TABLE OF CONTENTS

ABSTRACT ……………………………………………………………………………... ii ACKNOWLEDGEMENTS ………………………………………………………...…... v TABLE OF CONTENTS ………………………………………………..............……... vii LIST OF FIGURES ..…………...………………………………………………….….... x LIST OF TABLES …………………………………………………………..………….. vii

Chapter 1. Literature Review ………………..…………………………………........... 1

1.1 Piezoelectricity ………………….……………………………………..... 1

1.1.1 Pyroelec.ricity ………………………………………………....... 4

1.1.2 Ferroelectricity ………………………………………………...... 7

1.1.3 Other important piezoelectric parameters …………………….... 7

1.2 Lead-based piezoelectric materials …………………………………....... 9

1.2.1 Lead titanate (PbTiO3) ………………………………………..... 9

1.2.2 Lead zirconate titanate ( Pb(Zr, Ti)O3) ………………………..... 10

1.2.3 Other lead-based materials …………………………………….... 12

1.2.4 Environmental issues ………………………………………….... 12

1.3 Lead-free piezoelectric materials ……………………………..………..... 12

1.3.1 Bismuth based layered perovskite structures ……….…………... 13

1.3.2 Potassium sodium niobates ………………………….………...... 15

1.3.3 Barium titanates …………………………………….………….... 16

1.3.4 Other lead-free piezoelectric materials ……..…………………... 22

1.4 Low-dimensional piezoelectric properties measurements ………............. 23

1.5 Research motivation …………………………………………………....... 25

Page 11: Lead-free piezoelectric materials - Research Online

viii

1.6 References…………………………………………………………….... 28

Chapter 2. BTZ-BCT Ceramics ……………………………………………….…...... 35

2.1 Introduction ………………………………………..………………...... 35

2.2 Experimental procedure ……………………………..……………....... 36

2.3 Results and discussions …………………….………………………..... 36

2.3.1 Crystalline structure and vibration modes …………………...... 36

2.3.2 Ferroelectric and piezoelectric properties ……………………... 42

2.4 Conclusions ………………………………………………………......... 46

2.5 References ……………………………………………………………... 47

Chapter 3. BTZ- BCT Nanofibers ………………………………………..................... 49

3.1 Introduction …………………………………………………………..... 49

3.2 Experimental procedure ………………………………………….......... 50

3.3 Results and discussion …………………………………………............ 52

3.3.1 Microstructure and crystalline phase evolution ……………...... 52

3.3.2 Vibration modes ……………………………………………….. 56

3.3.3 Crystalline phase observations by TEM ……………………...... 58

3.3.4 Piezoelectric and ferroelectric measurements by PFM ……….... 59

3.3.5 Ferroelectric nanodomain imaging by SCM ………………….... 65

3.3.6 Electric resistance measurements by SSRM ………………….... 67

3.4 Conclusions ……………………………………………………..................... 69

3.5 References …………………………………………………………….. 70

aj393
Highlight
aj393
Highlight
Page 12: Lead-free piezoelectric materials - Research Online

ix

Chapter 4. BTZ -BCT Thin Films ……………………………………….................... 74

4.1 Introduction ………………………………………………………........ 74

4.2 Experimental procedure …………………………………................. 75

4.3 Results and discussion ……………………………………………....... 76

4.3.1 Morphology studies using FE-SEM and AFM ……………...... 76

4.3.2 Phase content investigation using XRD and Raman spectroscopy 77

4.3.3 Piezoelectric and ferroelectric properties investigation by PFM 78

4.4 Conclusions ………………………………………………………........ 84

4.5 References ……………….………………………………………......... 84

Chapter 5. Biocompatible Piezoelectric Nanofibers …………………………........... 87

5.1 Introduction ………………………………………………………........ 87

5.2 Experimental procedure ……………………………………………..... 89

5.3 Results and discussions ……………………………………………..... 90

5.3.1 Morphology and Crystalline phase evolution ………………... 90

5.3.2 Nanostructure studies by TEM ……………………...………... 94

5.3.3 Piezoelectric and ferroelectric properties ………………..….... 95

5.4 Conclusions …………………………………………………...…….... 99

5.5 References ………………………………………………………...….. 100

CONCLUSIONS RECOMMENDATIONS..................................................................... 103

PUBLICATIONS ……………………………………………………………………... 107

aj393
Highlight
Page 13: Lead-free piezoelectric materials - Research Online

x

LIST OF FIGURES

Figure 1.1 Symmetry hierarchy for piezoelectricity according to the

crystallographic point groups.

2

Figure 1.2 Schematic representation of the direct piezoelectric effect. 3

Figure 1.3 Schematic representation of the inverse piezoelectric effect. 3

Figure 1.4 Directions in a piezoelectric unit cell according to the Cartesian

coordinates system.

4

Figure 1.5 (a) Hexagonal structure of β-quartz and its coordination vectors. (b)

Schematic illustration of the piezoelectric phenomenon in quartz as

a non-polar piezoelectric material.

6

Figure 1.6 Phase diagram of lead zirconate titanate. 10

Figure 1.7 Schematic illustration of possible domain orientations in (a)

tetragonal structure: 6 directions and (b) rhombohedral structure: 8

directions.

11

Figure 1.8 Bismuth layered structure in BLSF [27]. 13

F igure 1.9 (a) Effect of Nb5+ doping on the resistivity and (b) effect of V5+ on

the dielectric permittivity of the Bi4Ti3O13 .

14

Figure 1.10 Equilibrium phase diagram of the KNbO3-NaNbO3 system. 15

Figure 1.11 (a) Barium titanate crystalline structures at different temperatures.

(b) Dielectric constant of BaTiO3 as a function of temperature.

17

Figure 1.12 (a) Tetragonal perovskite structure of BaTiO3. (b) Piezoelectric

effects in a BaTiO3 unit cell due to Ti displacement under an

external electric field E .

18

Figure 1.13 (a) Equilibrium phase diagram of the BaTiO3-CaTiO3 system [45]. (b)

Piezoelectric constants of <100>c oriented (Ba,Ca)TiO3 single

crystals and zone-melt biphasic polycrystals.

19

Figure 1.14 (a) Hysteresis loops of 40BCT, 40BCT,and 60BCT. (b1) Saturation

polarization, Pm, (b2) remnant polarization, Pr, (b3) coercive field,

Ec, (b4) permittivity, (b5) piezoelectric coefficient, d33, and (b6)

converse piezoelectric coefficient d = S/E. Values of various PZTs

are also shown as a reference. (c) Comparison of d33 among BZT-

50BCT and other non-Pb piezoelectrics, and yjr PZT family. (d)

20

Page 14: Lead-free piezoelectric materials - Research Online

xi

Comparison of the electrostrain versus electric field among BZT-

50BCT and various PZTs. (e1) Temperature dependence of the direct

piezoelectric coefficient in BZT-50BCT. (e2) Temperature

dependence of the direct piezoelectric coefficient in BZT-45BCT

(f) Phase diagram of pseudo-binary ferroelectric system

Ba(Zr0.20Ti0.80) O3-(Ba0.70Ca0.30)TiO3, abbreviated as BZT-BCT.

Figure 1.15 (a) Schematic illustration of the tilted MPB developed in the BZT-

xBCT system. Point b represents the TCP and c, d, and e identify

three different molar ratios x at different temperatures along the

MPB. (b),(c),(d), and (e) represent the 1D free energy barrier plot

against the polarization rotation from rhombohedral (PR) to

tetragonal (PT) and vice versa. (f) Dielectric constant vs.

temperature change in Ba(Zr0.15Ti0.85)O3 (rhombohedral phase) and

(Ba0.80Ca0.20)TiO3 (tetragonal phase) .

21

Figure 1.16 ZnO unit cell in the wurtzite structure. 22

Figure 1.17 (a) Schematic illustration of the PFM experimental set up. (b)

Surface displacement of a piezoelectric sample due to the inverse

piezoelectric effect exhibited by the ferroelectric domains under

applied electric field.

23

Figure 1.18 (a) Schematic illustration of the a-domains and c-domains in a

crystal. (b) Possible movements of the cantilever due to a force

developed by the interaction of different domains with the applied

AC signal, in which Fdefl results in deflection, Fbuck leads to buckling,

and Ftor creates torsion in the cantilever. (c) Side- and top-view of

the cantilever movement. (d) Possible movements of the laser spot

on the photodetector. Fdefl and Fbuck result in a vertical signal, while

only Ftor results in a lateral signal.

25

Figure 2.1 XRD patterns (left) of the (1-x)BTZ-xBCT ceramics sintered at

1450°C for 2 h in air and enlarged peaks located at 2θ ≈ 31.5º fitted

by the Lorentzian function for different x values. Coexistence of two

polar phases, both tetragonal (T) and rhombohedral (R), confirms

the crystallization of BTZ-BCT ceramics in yjr vicinity of the MPB

region. The peak shifts observed in different samples can be

37

aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
Page 15: Lead-free piezoelectric materials - Research Online

xii

attributed to the internal strain developed by the substitution of Ca2+

in the Ba+2 sites and Ti+4 in the Zr4+ sites in the BTZ ceramics.

Figure 2.2 Refinement patterns of the (a) BTZ, (b) 0.9BTZ-0.1BCT, (c)

0.8BTZ-0.2BCT, (d) 0.7BTZ-0.3BCT, and (e) 0.6BTZ-0.4BCT

ceramics sintered at 1450°C for 2 h in air. The symbols mark the

experimental results for the XRD pattern that is fitted in red. The

short blue vertical lines mark the line positions of the standard, and

the green spectrum at the bottom is the difference spectrum between

the fit and the experimental results.

38

Figure 2.3 (a) FE-SEM image of the particle size distribution and morphology

of the grains in the BTZ-BCT ceramic sintered at 1450°C for 2 h in

the air. (b) Higher magnification image of the area indicated by the

square in (a) that shows the growth steps in individual grains.

41

Figure 2.4 Different vibration modes in the (1-x)BTZ-xBCT system were

investigated by Raman spectroscopy at room temperature.

41

Figure 2.5 Typical polarization versus electric field hysteresis loop (P-E loop)

in a ferroelectric material.

43

Figure 2.6 (a) Maximum polarization Ps is increased by introducing BCT into

the BTZ structure, and the maximum Ps= 13.32 μC cm-2 at x = 0.5 in

the (1-x)BTZ-xBCT system. (b) Correlation of the polarization

saturation enhancement with the decrease in the unit cell volume in

the (1-x)BTZ-xBCT ceramics. (c) The BTZ-BCT ceramic achieved

the maximum spontaneous polarization Ps = 13.86 μC cm-2 under

an electric field E = 40 kV cm-1 at room temperature. (d)

Piezoelectric response enhancement in the BTZ-BCT ceramic

compared to other compositions in this system.

45

Figure 3.1 FE-SEM images of the BTZ-BCT nanofibers (a) calcined at 500°C,

(b) at 600°C (c) at 700°C. And (d) at 800°C for 1 hour. Insets:

highly magnified views of annealed BTZ-BCT nanofibers. All

images were collected under 0.5 kV acceleration voltage and 3.7

mm working distance without conductive coating.

53

Figure 3.2 Energy dispersive X-ray spectroscopy (EDS) of the BTZ-BCT

nanofibres annealed at 700°C. The inset table shows the

54

aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
Page 16: Lead-free piezoelectric materials - Research Online

xiii

concentrations of the different elements.

Figure 3.3 (a) XRD patterns of the BTZ-BCT nanofibers annealed at different

temperatures with corresponding Miller indices of BaTiO3.

Enlarged scans of Bragg diffraction reflections (b) for {110} at 2θ =

31.7 ° and (c) {200} at 2θ = 45.5 ° for the BTZ-BCT nanofibers

annealed at 700 °C. The peaks are fitted by Lorentzian functions.

Coexistence of rhombohedral and tetragonal phases proves that

crystallization of the nanofibers has occurred in the vicinity of the

morphotropic phase boundary.

55

Figure 3.4 Raman scattering spectra of the BTZ-BCT nanofibers annealed at

500, 600, 700, and 800 °C for 1 hour in air. Since the [A1(TO)] peak

at 164 cm-1 exists only in the rhombohedral phase of BaTiO3

nanocrystals, it indicates the coexistence of rhombohedral (R) and

tetragonal (T) phases in the MPB region at temperatures above

700°C.

56

Figure 3.5 TEM results obtained from sample heat-treated at 700° C: (a) low

magnification image of a nanofibre containing larger tetragonal and

smaller rhombohedral particles, with inset selected area electron

diffraction pattern; (b) indexing of the regions indicated in red and

green in (a) of {110} group reflections according to the indicated

rhombohedral (R) and tetragonal (T) phase reflections; (c) a second

selected area diffraction pattern with rhombohedral and tetragonal

reflections as indicated; (d) high magnification image with region

containing fine twins indicated.

58

Figure 3.6 . Force-distance curve recorded from the surface of a silicon

substrate using the cantilever employed in our PFM measurements.

This plot was used for the optical sensitivity calibration of the

cantilever.

60

Figure 3.7 (a) Local PFM phase hysteresis loops and (b) piezoelectric butterfly

loops of the BTZ-BCT nanofiber obtained by switching the dc-bias

voltage from -5 to +5 volts. 180° domain switching in the phase

hysteresis loop shows the switching of spontaneous electrical

61

aj393
Highlight
aj393
Highlight
Page 17: Lead-free piezoelectric materials - Research Online

xiv

dipoles built up in the BTZ-BCT fiber. (c) d33 hysteresis loop

calculated by using the converse piezoelectric equation, Δz = d33V,

which indicates a significant enhancement of d33 = 180 pmV-1 in the

BTZ-BCT nanofiber. (d) 3D topographic atomic force microscope

(AFM) image of the nanofiber used for the PFM measurements.

Figure 3.8 Comparison of piezoelectric coefficient d33 in nanostructured

components used to build piezoelectric generators with d33 in our

BTZ-BCT films and fibers.

63

Figure 3.9 (a) Topographical (AFM), (b) SCM images, obtained in contact

mode, of a single BTZ-BCT nanofiber annealed at 700°C on a

Si/SiO2/Ti/Ir conductive substrate. The two distinct types of regions

in black and white represent opposite out-of-plane electric domains,

averaging 24 nm in size and (c) schematic illustration of the SCM

measurement. The sign and magnitude of the C-V slope in the SCM

image is recorded while the AC bias is applied to a local capacitor

constructed by the nanofiber and conductive tip, and the substrate as

top and bottom electrodes, respectively. The ferroelectric domain

polarities and their configurations are developed in the dC/dV image

in the SCM as a result of different trends in the dC/dV with respect

to the polarization states.

66

Figure 3.10 (a) Topographical (AFM), (b) SSRM images, obtained in contact

mode, of a single BTZ-BCT nanofiber annealed at 700°C on a

Si/SiO2/Ti/Ir conductive substrate. The two distinct types of regions

in dark and bright colours represent the current distribution in the

conducting substrate and ferroelectric nanofiber. (c) Section analysis

of the SSRM image, and (d) reference transfer curve of the SSRM

logarithmic current amplifier in different biases.

68

Figure 4.1 (a) FE-SEM image of BTZ-BCT thin film deposited on the

Si/SiO2/Ti/Ir substrate and annealed at 700°C for 1 hour in air. Inset:

magnified view of the thin film surface. (b) FE-SEM image of cross-

section of BTZ-BCT thin film about 200 nm in thickness with an

average particle size of 33 nm. (c) 3D AFM image of topography of

BTZ-BCT thin film annealed at 700°C for 1 hour, showing 2 nm

76

aj393
Highlight
Page 18: Lead-free piezoelectric materials - Research Online

xv

rms surface roughness.

Figure 4.2 (a) X-ray diffraction pattern of the BTZ-BCT thin film annealed at

700°C on the Si substrate. (b) Raman spectrum of BTZ-BCT 200

nm thick BTZ-BCT film spin-coated on Si/SiO2/Ti/Ir substrate

annealed at 700°C for 1 hour in air. The spectrum of the substrate

alone is included for reference. All features are characteristic of the

polar structure of BaTiO3 and indicate successful evolution of the

perovskite structure in all samples. The existence of a peak at 164

cm-1 [A1(TO)] is only observed in the rhombohedral symmetry,

while the other peaks, which appear in both tetragonal and

rhombohedral symmetries, indicate the crystallization in

morphotropic phase boundary (MPB) region.

78

Figure 4.3 3D PFM phase (a) and amplitude (b) images of the surface of 200

nm thick BTZ-BCT film. Ferroelectric domain patterns in the phase

image (a) display the opposite polarity of ferroelectric domains 20 to

40 nm in size. These domains are revealed by 180° phase difference

contrast. The amplitude image (b) reproduces the domain shape.

Sharp contrast around domains visualizes domain boundaries. In (c),

phase and amplitude profiles recorded along the marked lines in (a)

and (b) are overlaid on an AFM topographical image. All three

images were collected simultaneously under 1.0 V ac modulation

voltage.

79

Figure 4.4 (a) Local PFM phase hysteresis loops and (b) piezoelectric

butterfly loops of the BTZ-BCT thin film obtained by switching the

bias voltage from -5 to +5 volts (±300 kV cm-1 electric field). 180°

domain switching in the phase hysteresis loop shows the switching

phenomenon of spontaneous electrical dipoles built up in the BTZ-

BCT thin film. (c) d33 hysteresis loop calculated by using the

converse piezoelectric equation, Δz = d33V, which indicates the

piezoelectric coefficient d33 = 141 pmV-1 in the BTZ-BCT thin film.

(d) Schematic illustration of lateral size and contact area in the thin

film and nanofiber.

80

Figure 4.5 Polarization versus electric field (P-E) loop obtained at 1 kHz of the 83

aj393
Highlight
aj393
Highlight
aj393
Highlight
Page 19: Lead-free piezoelectric materials - Research Online

xvi

BTZ-BCT thin film with about 200 nm thickness deposited on the

Si/SiO2/Ti/Ir substrate.

Figure 5. 1 FE-SEM images of (Na,K)NbO3 nanofibers mat (a) annealed at 600

°C, (b) 700 °C and (c) 800 °C in air. (d) Thermogravimetric analysis

of NKN nanofibers from room temperature up to 800 °C in air. 62

wt% of the NKN precursor evaporates or burns out during the

annealing process at different stages.

90

Figure 5.2 Energy dispersive x-ray spectroscopy (EDS) spectrum of the NKN

nanofibers annealed at 800°C. Inset: quantitative analysis of the

EDS that approves the existence of the constructive elements in the

(Na, K) NbO3 nanofibers which follow the stoichiometric ratios with

reasonable accuracy.

91

Figure 5.3 XRD patterns of the (Na, K) NbO3 nanofibers annealed at 700 and

800°C reveal the crystallization of the nanofibers in a monoclinic

structure. All peaks are indexed according to the PDF-card number

77-0038 corresponds to the (Na0.35K0.65) NbO3.

92

Figure 5.4 Experimental Raman spectrum of the NKN-nanofibers annealed at

800 °C for 1 hour in air.

93

Figure 5.5 TEM results obtained from sample heat treated at 800° C; (a) low

magnification image of a nanofiber containing single crystals. (b)

Selected area electron diffraction pattern which reveals the single

crystal. (c) High resolution TEM image of the NKN nanofiber. (d)

Schematic of a monoclinic structure.

95

Figure 5.6 3D-PFM (a) topography, (b) phase and (c) amplitude images of the

NKN nanofiber. Ferroelectric domain pattern in the phase image (b)

displays the opposite polarity of ferroelectric domains. These

domains are revealed by 180° phase difference contrast. The

amplitude image (c) represents the piezoelectric response of the

NKN nanofiber generated by the movement of the nanofiber under

alternate electric field developed between the tip and conducting

substrate. The PFM-phase and amplitude images are overlaid on the

topography image and collected simultaneously under 1.2 V ac

modulation voltage. (d) Displays the interaction of ferroelectric

98

aj393
Highlight
Page 20: Lead-free piezoelectric materials - Research Online

xvii

domains with essential elements for the tissue growth and

osteoconduction including fibrins and calcium ions respectively.

Figure 5.7 (a) Local PFM phase hysteresis loop and (b) piezoelectric butterfly

loop of the NKN nanofiber obtained by switching the bias voltage

from -5 to +5 volts at 25 Hz. 180° domain switching in the phase

hysteresis loop shows the switching phenomenon of spontaneous

electrical dipoles built up in the NKN nanofiber. (c) d33 hysteresis

loop calculated by using the converse piezoelectric equation, Δz =

d33V, indicates a significant enhancement of d33 = 58 pmV-1 in the

NKN nanofiber. (d) 3D topographical image of the NKN nanofiber

with 200 nm width and 100 nm height, used in PFM measurements.

99

Page 21: Lead-free piezoelectric materials - Research Online

xviii

LIST OF TABLES

Table 1.1 Centrosymmetric and non-centrosymmetric point groups in crystals. 5

Table 1.2 Comparison of piezoelectric properties in different BLSF

compositions.

14

Table 1.3 Piezoelectric coefficients of ZnO, AlN, and GaN in the forms of thin

films and single crystal.

22

Table 2.1 Lattice parameters and unit cell volumes of the (1-x)BTZ-xBCT

ceramics extracted from the Rietveld refinement results on the

ceramics.

40

Table 3.1 Comparison of the piezoelectric coefficient d33 of the BTZ-BCT

nanofibres with lead-based and lead-free piezoelectric nanofibres.

62

Table 3.2 Calculated effective stress induced by electric field in the nanofiber-

substrate interface.

65

Table 4.1 Comparison of the piezoelectric coefficient d33 in some Pb-based and

Pb-free piezoelectric thin films.

81

Table 4.2 Calculated effective stress induced by electric field in the thin film −

substrate interface.

82

aj393
Highlight
aj393
Highlight
Page 22: Lead-free piezoelectric materials - Research Online

xix

Page 23: Lead-free piezoelectric materials - Research Online

1

Chapter 1. Literature Review

1.1 Piezoelectricity

Piezoelectricity was discovered by the Curie brothers, Jacques and Pierre Curie, in

1880[1]. They observed electrical charges generated on the surfaces of some crystals,

including quartz (SiO2), the boron silicate crystal mineral group (tourmaline), topaz

(Al2SiO4(F,OH)2), cane sugar (C12H22O11), and Rochelle salt (KNaC4H4O6·4H2O),

when an appropriate load (stress) was applied to them. The phenomenon, which

involves the generation of electricity by mechanical deformation (strain) in some

materials, was called piezoelectricity, and the behaviour was considered as a direct

piezoelectric response. In 1881, Lippmann mathematically predicted an inverse

piezoelectric phenomenon, in which an electric field leads to a mechanical

deformation. He established the fundamental thermodynamic principles to represent

the converse piezoelectricity. The Curie brothers immediately confirmed the

converse piezoelectric phenomenon experimentally in 1881.

Just after the discovery of piezoelectricity, much more work was done to define the

crystallographic principles of this effect. In 1910, the piezoelectric effect was defined

with respect to crystallographic point groups in the first textbook published on

physical crystallography by Voigt[1]. Among the 32 crystallographic point groups,

only 21 are non-centrosymmetric structures, and apart from the 432 point group, the

other 20 groups are categorized as piezoelectric materials (Figure 1.1).

Page 24: Lead-free piezoelectric materials - Research Online

2

Figure 1.1 Symmetry hierarchy for piezoelectricity according to the crystallographic

point groups.

In the direct piezoelectric effect, the generated charge Q is proportional to the

applied force. The charge density D (dielectric displacement) is calculated using

Equation (1.1):

𝐷 = 𝑑𝑇 (1.1)

Where d is the piezoelectric coefficient in coulombs/Newton (C/N) and T in

Newtons/unit area is the applied stress. Changing from a compressive to a tensile

stress or vice versa reverses the sign of the direct piezoelectric coefficient. Figure

(1.2) shows the direct piezoelectric effect schematically in a piezoelectric material.

Page 25: Lead-free piezoelectric materials - Research Online

3

Figure 1.2 Schematic representation of the direct piezoelectric effect.

In the inverse piezoelectric effect, the induced strain S is proportional to the applied

electric field E

𝑆 = 𝑑𝐸 (1.2)

Where the inverse piezoelectric coefficient is expressed by d with units of meter/volt.

When the strain changes from a contraction to an expansion or vice versa, the sign

changes for the converse piezoelectric coefficient. The inverse piezoelectric

phenomenon is illustrated in Figure 1.3.

Figure 1.3 Schematic representation of the inverse piezoelectric effect.

Since the piezoelectric response is an anisotropic phenomenon, the piezoelectric

constant and other related parameters are represented by ij indices such as dij, in

which i denotes the direction of the polarization or electrical input/output and j

Page 26: Lead-free piezoelectric materials - Research Online

4

represents the direction of the strain or mechanical deformation developed or

applied. The subscripts are defined in a Cartesian coordinates system including X, Y,

and Z (Figure 1.4). The numbers 1-3 are assigned to the longitudinal parameters and

4-6 represent the shear mode.

Figure 1.4 Directions in a piezoelectric unit cell according to the Cartesian

coordinates system.

For example, the d31 mode identifies a mechanical deformation (strain) along the 1

(X) axis when an electric field is applied along the 3 (Z) axis, and d33 describes the

generated charge (electric field) along the 3 (Z) direction while a mechanical stress is

applied along the same direction as the electric field in the direct piezoelectric

phenomenon and vice versa in the inverse piezoelectric response.

1.1.1 Pyroelectricity

A subgroup of the piezoelectric materials consisting of 10 point groups possesses a

unique polar axis in a certain crystal direction due to an existing electric dipole. The

dipole moment can be changed by exposing these materials to uniform heat that leads

to a surface electric charge. Changing the magnitude of the dipole with temperature

Page 27: Lead-free piezoelectric materials - Research Online

5

is demonstrated in this subgroup of the piezoelectric materials as the pyroelectricity.

An electric dipole is constructed by separation of positive and negative charges in a

unit cell. Pyroelectric materials possess a spontaneous polarization, PS, which cannot

be reoriented by an external electric field. When these materials are heated,

increasing the temperature leads to a change in the spontaneous polarization and can

be detected according to Equation (1.3)

𝑑𝑃𝑠 = −𝑝 𝑑𝑇 (1.3)

Where 𝑑𝑃𝑠 is the polarization change created by the 𝑑𝑇 temperature variation. The

pyroelectric effect is defined by 𝑝 . The polarization is suppressed in the pyroelectric

materials when the temperature increases, and it is represented by the negative sign

in Equation 1.3.

Table 1.1 Centrosymmetric and non-centrosymmetric point groups in crystals [2].

Crystal

structure Point group

Centro-

symmetric

Non-centrosymmetric (Piezoelectric)

Non-polar Polar

Triclinic 1�, 1 1� ― 1

Monoclinic 2, m, 2/m 2/m ― 2,m

Orthorhombic 222, mm2, mmm mmm 222 mm2

Tetragonal 4, 4/m, 422, 4mm, 4�,

4/mmm, 4�2m or 4�m2 4/m, 4/mmm

422, 4�, 4�2m

or 4�m2

4,

4mm

Rhombohedral 3, 3�, 32, 3m, 3�m 3�m, 3� 32 3,3m

Hexagonal 6, 6/m, 622, 6mm, 6/mmm,

6�, 6�m2 or 6�2m 6/m, 6/mmm

622, 6�, 6�m2

or 6�2m

6,

6mm

Cubic 23, m3, 432, m3m, 4�3m m3, m3m 23, 432, 4�3m ―

Page 28: Lead-free piezoelectric materials - Research Online

6

In piezoelectric crystals apart from the pyroelectrics, dipoles are arranged in

compensating directions in such a way that the net crystal dipole moment is zero.

These crystals are polarized when the load (stress) is applied, and a net crystal dipole

is developed in a favoured direction. Crystallographic point groups in polar and non-

polar piezoelectric materials are represented in Table 1.1. Quartz is a well know non-

polar piezoelectric material. Figure 1.5 demonstrates the piezoelectric response

mechanism in quartz crystal.

Figure 1.5 (a) Hexagonal structure of β-quartz and its coordination vectors. (b)

Schematic illustration of the piezoelectric phenomenon in quartz as a non-polar

piezoelectric material.

The favoured direction for the polarization in quartz is along the a-axis in its unit

cell. Regarding the hexagonal structure in the β-quartz, there are three polarization

directions separated by 120° from each other (Figure 1.5(a)). It should be noted that

in a non-polar crystal, when a uniform hydrostatic pressure is applied to the crystal,

the net polarization is zero and only by applying the pressure in an individual

polarization direction, is a non-zero net polarization developed (Figure 1.5(b)).

(a) (b)

Page 29: Lead-free piezoelectric materials - Research Online

7

1.1.2 Ferroelectricity

Ferroelectric materials are a subclass of the pyroelectrics and consequently, the

piezoelectric materials. The spontaneous polarization in ferroelectrics is reversible

and switchable under an alternated electric field, whereas in the non-ferroelectric

pyroelectric materials, it remains unchanged when an external electric field is

applied. In the other words, there must be more than one equilibrium state for the

spontaneous polarization in ferroelectrics, so that the polarization vector can be

switched between those states. In the non-ferroelectric pyroelectric materials,

however, there is only one possible state. It should be noted, as is demonstrated in

Figure 1.1, that all ferroelectric materials are pyroelectric and consequently

piezoelectric, but all pyroelectric and piezoelectric materials are not ferroelectric.

1.1.3 Other important piezoelectric parameters

Piezoelectric voltage constant

The piezoelectric voltage constant, g, is defined as the electric field developed by a

stress in a piezoelectric material in units of mV N-1 that can be calculated using

Equations (1.4) and (1.5)

𝑔33 = 𝑑33𝜀0𝐾3𝑇

(1.4)

Where K is the relative dielectric constant and

𝑔31 = 𝑑31𝜀0𝐾3𝑇

(1.5)

Mechanical quality factor Qm

The mechanical quality factor, Qm, is the ratio of the stored energy to the wasted

energy in a cycle when a piezoelectric material is subjected to a periodic vibration,

and Qm-1 represents the mechanical loss[3]. It is also demonstrated that mechanical

Page 30: Lead-free piezoelectric materials - Research Online

8

strain in a piezoelectric resonant is amplified by a factor proportional to Qm at its

resonance frequency compared to the strains in off-resonance frequencies. In a

highly efficient actuator, a high Qm and low loss are essential.

The mechanical quality factor can be calculated in piezoelectric materials by using

Equation (1.6)

𝑄𝑚 = 12𝜋𝑓𝑟𝑍𝑚𝐶0

� 𝑓𝑎2

𝑓𝑎2 − 𝑓𝑟2� (1.6)

Where fr and fa are the resonance and anti-resonance frequencies. Zm is the minimum

impedance at resonance frequency and C0 represents the low frequency capacitance.

Coupling factor K

The electromechanical coupling factor, K, demonstrates the efficiency of a

piezoelectric material in the conversion of the mechanical energy into electricity and

vice versa. Three important coupling factors in piezoelectric materials are the planar

coupling coefficient Kp, the length extensional coupling coefficient K31, and the

thickness extensional coupling coefficient K33. These coefficients are calculated by

Equations (1.7)-(1.10) using the resonance and anti-resonance frequencies.

𝐾332 = 𝜋2

1+𝑓𝑎−𝑓𝑟𝑓𝑟

tan �𝜋(𝑓𝑎−𝑓𝑟)

2𝑓𝑟

1+(𝑓𝑎−𝑓𝑟)𝑓𝑟

� (1.7)

And if

𝜓 = 𝜋2�1 + 𝑓𝑎−𝑓𝑟

𝑓𝑟� tan �𝜋(𝑓𝑎−𝑓𝑟)

2𝑓𝑟� (1.8)

Then

𝐾312 = 𝜓𝜓+1

(1.9)

And

𝐾𝑝 = 𝑓𝑎2 − 𝑓𝑟2

𝑓𝑟2 (1.10)

Page 31: Lead-free piezoelectric materials - Research Online

9

1.2 Lead-based piezoelectric materials

1.2.1 Lead titanate (PbTiO3)

Lead titanate is one of the essential basic compounds used in commercially

employed piezoelectric materials, however, its pure form is not of interest to

industry. PbTiO3 is crystallized in the tetragonal structure with c/a ratio of 1.06 in the

P4mm space group at room temperature[4].

There have been few practical results on the piezoelectric and ferroelectric properties

of the PbTiO3 single crystals due to their high electrical conductivity, which could

originate from a high concentration of Pb vacancies, especially in high temperature

processes. They possess a piezoelectric constant d33 = 84-117 pC N-1 and dielectric

constant k33 = 80-126 [2, 5, 6].

In PbTiO3 ceramics, low resistivity together with the mechanical fracturing caused

by thermal expansion anisotropy and large spontaneous strain during the cubic to

tetragonal phase transition results in low dielectric and piezoelectric properties. In

modified PbTiO3 structures, however, introducing a variety of additives, mainly rare

earths ((Pb1-3/2xRex) TiO3) and alkaline elements ((Pb1-xCax)TiO3), has improved the

electrical resistivity and reduced the spontaneous strain by decreasing the Curie

temperature and tetragonality of pure PbTiO3 and made the doped compounds

appropriate dielectrics with useful electromechanical properties[7-10]. In the modified

lead titanate ceramics, the piezoelectric coefficient has been improved up to more

than 90 pC N-1 .

1.2.2 Lead zirconate titanate ( Pb(Zr,Ti)O3)

Lead zirconate titanate (PZT) is a solid solution of lead zirconate (PbZrO3) and lead

titanate (PbTiO3). PbZrO3 is an antiferroelectric. According the definition presented

Page 32: Lead-free piezoelectric materials - Research Online

10

by Kittel, in the antiferroelectric materials, spontaneous electric polarizations are

arranged in the antiparallel direction, so that the net polarization is zero and they are

not piezoelectric [11]. Interestingly, a pseudo binary system PbZrO3-x PbTiO3 (Pb(Zr1-

xTix)O3) or PZT exhibits very useful piezoelectric and ferroelectric properties. The

phase diagram of Pb(Zr1-xTix)O3 is shown in Figure 1.6. The rhombohedral region

contains two symmetries, including a low temperature rhombohedral form with R3c

symmetry and a high temperature one with R3m symmetry. As is mentioned in

Figure 1.6, the Ti-rich tetragonal region undergoes a direct phase transition from

tetragonal to cubic, however, the Zr-rich rhombohedral phase experiences a phase

change from the low temperature rhombohedral to the high temperature

rhombohedral structure before crystallizing in a cubic structure.

Figure 1.6 Phase diagram of lead zirconate titanate[12].

PZT with molar ratio x = 0.47-0.50 is crystallized in a unique region in its phase

diagram called the morphotropic phase boundary (MPB), so that two polar structures,

including the rhombohedral (R3m) and tetragonal (P4mm), can coexist.

Page 33: Lead-free piezoelectric materials - Research Online

11

Enormous attempts have been undertaken to understand the role of the MPB in the

superior piezoelectric performance of PZT, where d33=290 pC N-1, Qm=1000, Kt =

0.47, ε33=1300, and tan 𝛿 = 0.005 [12-16].

The most plausible explanation refers to the large amount of active polarization

orientation that exists in the MPB region and the flattening of the free energy profile

for switching spontaneous polarization vectors. Figure 1.7 shows 14 possible domain

orientations: 6 directions in tetragonal <001> (Figure 1.7(a)) and 8 directions in

rhombohedral <111> (Figure 1.7(b)) structures. This means there are 14 (6+8 = 14)

variants for switching of spontaneous polarization vectors in the vicinity of the MPB.

Variability of polarization switching and flattening of the free energy profile provide

efficient polarization reorientation during the poling procedure and remarkably

enhance the piezoelectric properties in the MPB region [17-20].

Figure 1.7 Schematic illustration of possible domain orientations in (a) tetragonal

structure: 6 directions and (b) rhombohedral structure: 8 directions.

PZT-based materials exhibit outstanding piezoelectric and ferroelectric properties

and have been used widely in different aspects of science and technology.

Page 34: Lead-free piezoelectric materials - Research Online

12

1.2.3 Other lead-based materials

The outstanding piezoelectric and ferroelectric properties of PZT have attracted

many attempts to investigate the other possible lead-based piezoelectric materials.

Many compositions and complex systems have been studied, and some of them have

become objects of interest because of their interesting performances. Compositions

such as Pb2Nb2O6 [21], Pb(Zn1/3Nb2/3)O3–PbTiO3(PZN–PT) [22, 23], Pb(Mg1/3,Nb1/3)O3–

PbTiO3 (PMN–PT) [24], Pb (Ni1/3Nb1/3)O3–Pb(Zr,Ti)O3 (PNN–PZT) [25], and

Pb(Sc1/2Nb1/2)O3-PbTiO3 [26] have been investigated, and the results were published.

1.2.4 Environmental issues

Lead (Pb) is one of the most toxic materials known, and continuous exposure to an

environment contaminated by this element has potential hazards. Being in contact

with Pb can cause serious damages to vital human organs such as the kidney, heart,

and brain. Since most commercial piezoelectric materials are based on PZT

containing about 60 wt% lead, intensive efforts have been undertaken to eliminate

the Pb and find good replacement candidates[27].

1.3 Lead-free piezoelectric materials

There has been a growing interest in developing alternative lead-free piezoelectric

materials that can eventually replace the current lead-based ones. Intensive research

efforts have been spent on related studies all around the world for over two decades.

In the following sections, current lead-free piezoelectric materials will be reviewed.

1.3.1 Bismuth based layered perovskite structures

Bismuth layered perovskite structure ferroelectrics (BLSF) with the chemical

formula Bi2Ax-1BxO3x+3 is another group of lead-free piezoelectric materials. This

Page 35: Lead-free piezoelectric materials - Research Online

13

structure comprises perovskite layers periodically separated by (Bi2O2)2+ layers

(Figure 1.8). As demonstrated in Figure 1.8 the layered structure in the BLSF leads

to a plate-like morphology in its microstructure [28, 29].

Figure 1.8 Bismuth layered structure in BLSF [27].

The high Curie temperatures (600°C ─ 900°C) in BLSF compositions, much higher

than those of the lead-based materials (200°C ─ 400°C), mak e the BLSF materials

good candidates as pyroelectric sensors and high temperature piezoelectric materials.

Due to the anisotropic nature of their structures, however, their electrical

conductivity is highly anisotropic and is also high because of the Bi volatility at high

temperatures during the sintering. Moreover, the switching of the spontaneous

polarization within the materials during poling is limited to within a two-dimensional

plane. Thus, poling the BLSF ceramics is not efficient enough and leads to a low

piezoelectric constant with a d33 of ~ 20 pC N-1 and a large coercive field [28].

Bi4Ti3O13 (A = Bi, B = Ti, and x = 3) is one of the most studied compositions in this

group. Further studies on this system have indicated that doping with Nb5+ and V5+

ions could increase the resistivity in this structure, thus improving the piezoelectric,

dielectric, and ferroelectric properties (Figure 1.9) [28, 30].

Page 36: Lead-free piezoelectric materials - Research Online

14

Figure 1.9 (a) Effect of Nb5+ doping on the resistivity and (b) effect of V5+ on the

dielectric permittivity of tBi4Ti3O13 [28, 30].

Many attempts have been undertaken to enhance the electrical and piezoelectric

properties of BLSF materials to date, and several complex compositions have been

investigated, such as (Bi1/2Na1/2) TiO3, (Na1/2 Bi1/2)1-xCaxBi4Ti4O15, and SrBi2Ta2O9

[31-36] . Table 1.2 shows a comparison of the piezoelectric and electrical properties in

different BLSF compositions.

Table 1.2 Comparison of piezoelectric properties in different BLSF compositions.

Composition d33 (pC N-1) K Tc (°C) Ref.

(Bi0.5Na0.5)TiO3 57-64 240-

467

310-

450 [32-34]

(Bi0.5Na0.5)TiO3-0.02NaNbO3 88 624 - [32]

(Na0.5Bi0.5)0.94Ba0.06TiO3 125 625 288 [35]

(Na0.5Bi0.5)0.94–6BaTiO3 + 0.5 mol%

CeO2 + 0.5 mol% La2O3 162 831 - [36]

1.3.2 Potassium sodium niobates

Alkali niobates with general chemical formula ANbO3 (A: alkali metal) are another

family of successful lead-free piezoelectric and ferroelectric materials. In 1950s and

1960s, several new compositions in this family were proposed and explored[37, 38].

Page 37: Lead-free piezoelectric materials - Research Online

15

Shirane et al reported the dielectric properties and phase transitions of the NaNbO3

and (Na,K)NbO3 (NKN) systems in the forms of single crystals and ceramics [37].

Although the NaNbO3 did not show any evidence of a ferroelectric response, it was

demonstrated that introducing the KNbO3 into the NaNbO3 structure created

ferroelectricity in the NaNbO3- x KNbO3 system (x < 0.97). Further investigations on

the new alkaline niobate systems demonstrated that for x = 0.50 (Figure 1.10)[39],

NKN exhibited the highest piezoelectric and ferroelectric properties compared to the

other compositions.

Figure 1.10 Equilibrium phase diagram of the KNbO3-NaNbO3 system[12], [39].

The volatility of the alkaline elements reduces the final density of the NKN ceramics

sintered in air and leads to a suppression of the dielectric constant, piezoelectric

coefficient, and ferroelectric response. In order to improve the sintered density and

consequently, the piezoelectric properties in the NKN ceramics, different techniques

and processing methods have been undertaken, such as employing hot-pressing,

Page 38: Lead-free piezoelectric materials - Research Online

16

sintering aids, and isostatic pressing. It has been reported that hot-pressed

Na0.5K0.5NbO3 possessed high remnant polarization (Pr = 33 μC/cm2), a large

piezoelectric coefficient (d33 = 160 pC N-1) and high planar coupling factor (Kp =

0.45) [38].

Enormous efforts have been undertaken to improve the physical and electrical

properties of the NKN-based ceramics by modifying the fabrication procedure,

employing different dopants (i.e. Li1+, Ta+5, and Sb3+)[40], and combining the NKN

with other good piezoelectric materials such as BaTiO3 [41] .

It is evident that creating engineered textured structures in ceramics has led to

fascinating piezoelectric and electrical properties; however, this process requires

high cost and special facilities. Saito et al. have successfully textured a complex

Na0.5K0.5NbO3-based ceramic, (K0.44Na0.52Li0.04)(Nb0.86Ta0.10Sb0.04)O3, and achieved

the best piezoelectric coefficients in the NKN system: d33 = 416 pC N-1 and d31 = -

152 pC N-1 [42].

1.3.3 Barium titanates

Barium titanate (BaTiO3) is one of the most widely explored lead-free ferroelectric

materials, as it possesses good piezoelectric, nonlinear optical properties and voltage-

tuneable low loss dielectric properties. It experiences several phase transitions from

low temperature to high temperature, including orthorhombic-rhombohedral at ~199

K, rhombohedral-tetragonal at ~ 285 K, tetragonal-cubic at ~ 393 K, and a drastic

crystal structure transition at ~ 1733 K from perovskite to hexagonal structure before

melting at 1891 K (Figure 1.11) [43].

Page 39: Lead-free piezoelectric materials - Research Online

17

(b)

Figure 1.11 (a) Barium titanate crystalline structures at different temperatures. (b)

Dielectric constant of BaTiO3 as a function of temperature [43].

Because of its fascinating dielectric, ferroelectric, and piezoelectric properties, the

perovskite BaTiO3 has been employed in various applications, such as piezoelectric

sensors, capacitors, memories, and optical devices [12] .

The piezoelectricity in perovskite structurea such as that of BaTiO3 is due to an off-

center atom located in an octahedral position in the centre of the unit cell. In the

BaTiO3 unit cell, the Ti displacement along the polarization direction changes the

charge balance and creates an internal electric field and charge separation (Figure

1.12). Spontaneous polarization developed by the Ti displacement results in strong

piezoelectric and ferroelectric properties in the BaTiO3 and other perovskite

structures such as PZT.

Figure 1.12 (a) Tetragonal perovskite structure of BaTiO3. (b) Piezoelectric effects

Page 40: Lead-free piezoelectric materials - Research Online

18

in a BaTiO3 unit cell due to Ti displacement under an external electric field E [44].

It is well known that doping is an effective method to improve the material

performance in electroceramics. BaTiO3 doped with Ca and Zr demonstrates high

dielectric permittivity, enhanced temperature stability, and high reliability compared

to pure BaTiO3. The effects of Ca and Zr on the piezoelectric and electrical

properties of BaTiO3 will be discussed in the following section.

Effect of Zr on BaTiO3

The Zr4+ ion is dissolved in the BaTiO3 structure via a substitutionally solid solution,

where it is placed in the Ti4+ site and forms barium zirconate titanate (Ba(Ti,Zr)O3).

The Zr4+ ion (ionic radius = 87 pm) is larger and chemically more stable than the Ti+4

ion (ionic radius = 68 pm), and thus replacement of Ti+4 by Zr+4 suppresses the

conduction developed by the electronic hopping between Ti+4 and Ti+3, and leads to

an enhanced dielectric constant and reduced leakage current in the BaTiO3 structure.

The piezoelectric and electrical properties in the Ba(Ti1-xZrx)O3 system have been

investigated for different Zr concentrations, and it has been demonstrated that for

compositions 0 ≤ x ≤ 0.1 the ceramics show normal ferroelectric behaviour,

however, for compositions 0.10 ≤ x ≤ 0.42 relaxor properties are indicated. It has

been demonstrated that using a Zr/Ti ratio of 20/80 in the Ba(Ti,Zr)O3 ceramics

creates good piezoelectric and electrical properties in this system.

Effect of Ca on BaTiO3

There have been several reports indicating that partial replacement of Ba2+ by Ca2+

ions enhances the dielectric constant, piezoelectric properties, electromechanical

response, and ferroelectric properties of the BaTiO3 composition [45-48]. In the barium

calcium titanate system ((Ba1-xCax) TiO3), the piezoelectric coefficient has been

Page 41: Lead-free piezoelectric materials - Research Online

19

(a)

improved from 180 up to 310 pC N-1 for 0.02 ≤ x ≤ 0.34 [48]. The Curie temperature

is increased to 136°C by adding the Ca2+ ions up to x = 0.08 and then reduced for

higher contents. [47].

Figure 1.13 (a) Equilibrium phase diagram of the BaTiO3-CaTiO3 system [45]. (b)

Piezoelectric constants of <100>c oriented (Ba,Ca)TiO3 single crystals and zone-melt

biphasic polycrystals [48].

Recently, the lead-free pseudobinary (1-x) Ba(Zr0.20Ti0.80)O3 - x (Ba0.70Ca0.30)TiO3

system[49] has been shown to possess a very high piezoelectric response (Figure

1.14). In this system, the optimal composition with x = 0.5 at room temperature at the

morphotropic phase boundary possesses the highest piezoelectric performance: d33 =

620 pC N-1, g33 = 15.38 × 10-3 Vm N-1, electromechanical coupling factor K33 = 65%,

dielectric permittivity as high as εг ~ 3060, and electromechanical (converse

piezoelectric) response of 1140 pm V-1 [49, 50].

(b)

aj393
Highlight
Page 42: Lead-free piezoelectric materials - Research Online

20

Figure 1.14 (a) Hysteresis loops of 40BCT, 40BCT,and 60BCT. (b1) Saturation

polarization, Pm, (b2) remnant polarization, Pr, (b3) coercive field, Ec, (b4)

permittivity, (b5) piezoelectric coefficient, d33, and (b6) converse piezoelectric

coefficient d = S/E. Values of various PZTs are also shown as a reference. (c)

Comparison of d33 among BZT-50BCT and other non-Pb piezoelectrics, and yjr PZT

family. (d) Comparison of the electrostrain versus electric field among BZT-50BCT

and various PZTs. (e1) Temperature dependence of the direct piezoelectric

coefficient in BZT-50BCT. (e2) Temperature dependence of the direct piezoelectric

coefficient in BZT-45BCT (f) Phase diagram of pseudo-binary ferroelectric system

Ba(Zr0.20Ti0.80) O3-(Ba0.70Ca0.30)TiO3, abbreviated as BZT-BCT [49].

It was demonstrated that there was a tricritical point (TCP) in the phase diagram at x

= 0.35 and T = 57 °C. Here, the cubic-paraelectric, ferroelectric rhombohedral, and

tetragonal phases meet each other. In the vicinity of the tricritical point, the

polarization anisotropy vanishes, so that the dielectric permittivity and piezoelectric

coefficient experience very strong enhancement (Figure 1.15). The piezoelectric

parameter is comparable to 500-600 pC N-1 in PZT and considerably higher than

(f)

Page 43: Lead-free piezoelectric materials - Research Online

21

(a)

those in Pb-free piezoelectric compositions such as alkaline niobate ceramics with

300 pC N-1 [42] and bismuth based layered ferroelectric compositions with about 100

pC N-1 [27]. The high piezoelectric performance was obtained in bulk BTZ-BCT

ceramic, which makes it a promising candidate in piezoelectric devices.

Figure 1.15 (a) Schematic illustration of the tilted MPB developed in the BZT-xBCT

system. Point b represents the TCP and c, d, and e identify three different molar

ratios x at different temperatures along the MPB. (b),(c),(d), and (e) represent the 1D

free energy barrier plot against the polarization rotation from rhombohedral (PR) to

tetragonal (PT) and vice versa. (f) Dielectric constant vs. temperature change in

Ba(Zr0.15Ti0.85)O3 (rhombohedral phase) and (Ba0.80Ca0.20)TiO3 (tetragonal phase) [51].

1.3.4 Other lead-free piezoelectric materials

There are other lead-free piezoelectric materials including wurtzite structures and

quartz (Figure 1.16), which are not ferroelectric. The piezoelectric behaviour in this

group has been discussed previously. Thus, they are employed either as single

crystals or as oriented polycrystalline samples and textured thin films.

(f)

(b) (c) (d) (e)

Page 44: Lead-free piezoelectric materials - Research Online

22

Figure 1.16 ZnO unit cell in the wurtzite structure [52].

Although most of them exhibit low piezoelectric performance, their stability and

compatibility in terms of integration into electronic systems have caused them to

emerge as useful piezoelectric materials in electronic devices. The most commonly

studied materials in this group are ZnO , AlN, and GaN in the forms of textured thin

films and single crystals [52-55]. Their piezoelectric coefficients are compared in Table

1.3.

Table 1.3 Piezoelectric coefficients of ZnO, AlN, and GaN in the forms of thin films

and single crystal.

Materials d33 (pm V-1)

Ref. Single crystal Thin film

ZnO 3.0 4.41 [54]

AlN 5.6 3.4 [55]

GaN 3.7 2.8 [53]

1.4 Low-dimensional piezoelectric properties measurements

Scanning probe microscopy (SPM) techniques, such as Kelvin probe force

microscopy (KPFM), scanning tunnelling microscopy (STM), scanning capacitance

microscopy (SCM), scanning spreading resistance microscopy (SSRM), and

piezoresponse force microscopy (PFM) have emerged as powerful tools to study

Page 45: Lead-free piezoelectric materials - Research Online

23

surface electrical properties, ferroelectric domains, and the piezoelectric response in

low-dimensional materials [56-62].

Figure 1.17 (a) Schematic illustration of the PFM experimental set up. (b) Surface

displacement of a piezoelectric sample due to the inverse piezoelectric effect

exhibited by the ferroelectric domains under applied electric field.

PFM has been established as a reliable approach to study the dynamic behaviour,

piezoelectric properties, switching mechanism, and configuration of the ferroelectric

domains in the piezoelectric materials. This technique is based on the contact-mode

in SPM, for which the instrument is equipped with a function generator, lock-in

amplifier, and a conductive cantilever [63-70]. Figure 1.17 schematically illustrates the

PFM setup.

(a)

(b)

Page 46: Lead-free piezoelectric materials - Research Online

24

The piezoelectric response of the surface is acquired by applying DC and AC

voltages to the tip scanning over the surface in the contact mode:

Vtip = Vdc + Vac cos(ωt) (1.11) [63]

Where the electric AC component causes deformation of the surface through the

converse piezoelectric effect, and leading to the tip deflection (Figure 1.18) as

follows:

d = d0 + A 𝑐𝑜𝑠(𝜔𝑡 + 𝜑) (1.12) [71]

Where A represents the amplitude of the piezoelectric response and φ defines the

phase shift generated by the ferroelectric domain located below the tip. The different

configurations of the ferroelectric domains include parallel to the sample surface (in-

plane or a-domain) and vertical with respect to the sample surface (out-of-plane or c-

domain) (Figure 1.18(a)). Lateral and vertical displacements are created by in-plane

and out-of-plane domains, respectively. Since the tip is in contact with the surface of

the sample, all types of domains exhibit corresponding movements in the tip,

consisting of vertical, longitudinal and lateral movements. Considering the vertical

and lateral movement of the tip (as longitudinal movement is objected in vertical

displacement), vertical PFM (VPFM) and lateral PFM (LPFM) have been developed

to study the in-plane and out-of-plane ferroelectric domains [72].

Page 47: Lead-free piezoelectric materials - Research Online

25

Figure 1.18(a) Schematic illustration of the a-domains and c-domains in a crystal.

(b) Possible movements of the cantilever due to a force developed by the interaction

of different domains with the applied AC signal, in which Fdefl results in deflection,

Fbuck leads to buckling, and Ftor creates torsion in the cantilever. (c) Side- and top-

view of the cantilever movement. (d) Possible movements of the laser spot on the

photo detector. Fdefl and Fbuck result in a vertical signal, while only Ftor results in a

lateral signal [70]

1.5 Research motivation

Low-dimensional ferroelectric materials, including nanofibers, nanotubes,

nanowires, nanobelts, and thin films, have emerged as a hot research topic due to

their novel properties and applications. The state-of-the-art piezoelectric bulk

materials, such as PZT, which have the best piezoelectric performance among all the

piezoelectric compounds, have been used practically as key components in electronic

sensors, actuators, transducers, electro-mechanical energy conversion devices, etc.

Integration of high performance piezoelectrics in piezoelectric devices and

micro/nano electromechanical systems (MEMS, NEMS) is a viable approach to

enhance their efficiency. Low-dimensional piezoelectric and ferroelectric materials

(a) (b) (c)

(d)

Page 48: Lead-free piezoelectric materials - Research Online

26

have been exploited widely in nanogenerators, sensors, transducers, MEMS devices,

and other applications such as microwave varactors and ferroelectric field effect

transistors.[73-78] For instance, high performance piezoelectric nanostructures can

enhance the output power of piezoelectric generators that convert kinetic energy of

vibrations, displacements, or applied force to electricity.[79-83]

There are three challenging issues, however, that need to be addressed for

nanostructured piezoelectric materials and their devices: 1) Although scaling down

the devices to the nanoscale provides significant benefits such as suppressing the

energy consumption, the ferroelectric and piezoelectric properties are often

suppressed in small dimensions due to intrinsic or extrinsic effects, depending on

such factors as lateral size, geometry, particle size, fraction of parallel (a-domains)

and perpendicular (c-domains) domains on the surface, in-plane stress, and domain

wall mobility. 2) Confinement of the nanosized piezoelectric components by the

substrate develops internal stress in its interface with the substrate (clamping effect),

which results in a suppression of its piezoelectric performance. 3) Although

piezoeloectric single crystals possess remarkably high piezoelectric coefficients,

growing single crystals on the nanoscale such as in nanowires and

nanorods/nanoribbons, while retaining their superior performance, is very difficult in

practice due to complex phase formation dynamics, in particular for ternary

piezoelectric systems.

Enormous efforts have been undertaken to find innovative new materials and

improve the piezoelectric response by varying their compositions and shapes in the

form of thin films, nanowires, and nanofibers. To the best of our knowledge,

however, the piezoelectric response reported for the polycrystalline nanostructured

materials, so far, has been less than 160 pm V-1 [58, 84-94] Grain boundaries create

Page 49: Lead-free piezoelectric materials - Research Online

27

pinning sites for ferroelectric domain wall motion. The high density of grain

boundaries restricts domain wall mobility in nanostructured materials. As a

consequence, the electromechanical response becomes smaller compared to bulk

ceramics containing coarse grains and single crystals, with a lower density of grain

boundaries.[95]

Recently, a lead-free Ba(Ti0.80Zr0.20)O3-x(Ba0.70Ca0.30)TiO3 ((1-x)BTZ-xBCT)

piezoelectric system with optimal composition x = 0.5 was reported to show superior

room temperature piezoelectricity, with the piezoelectric coefficient d33 = 620 pC N-

1, the piezoelectric voltage constant g33 = 15.38×10-3 Vm N-1, and the

electromechanical (converse piezoelectric) response as high as 1140 pm V-1. These

superior piezoelectric properties are comparable to or higher than those of state-of-

the-art PZT or other lead-free piezoelectric compounds, due to the low polarization

anisotropy and low energy barrier for lattice distortions in the MPB region.[49] The

large d33 obtained in bulk BTZ-BCT ceramics makes this compound a promising

candidate as a component for low-dimensional piezoelectric devices. It is expected

that nanowires/nanofibers or thin films of this compound should have better

piezoelectric properties than that of any existing piezoelectric materials.

In this research, two different aspects of the low dimensional piezoelectric materials

were investigated.

First, low-dimensional counterparts of BTZ-BCT in forms of thin films and

nanofibers were synthesized, and their structures and their piezoelectric and

ferroelectric properties were investigated and compared with the BTZ-BCT

ceramics. Different scanning probe microscopy techniques, including atomic force

microscopy (AFM), piezoresponse force microscopy (PFM), scanning capacitance

microscopy (SCM), and scanning spreading resistance microscopy (SSRM) were

Page 50: Lead-free piezoelectric materials - Research Online

28

employed to study the piezoelectric, ferroelectric, and electric properties of the thin

films and nanofibers. Then, biocompatible NKN nanofibers as a piezoelectric

scaffold was synthesized and characterized.

1.6 References

[1] B. Jaffe, Cook W.R. Jr., J. H.L., Piezoelectric Ceramics, Academic Press

Inc., 1971.

[2] A. Safari, E. K. Akdoğan, Piezoelectric and Acoustic Materials for

Transducer Applications, Springer London, Limited, 2008.

[3] D. Ensminger, F. B. Stulen, Ultrasonics: Data, Equations and Their

Practical Uses, Taylor & Francis, 2010.

[4] G. Shirane, J. D. Axe, J. Harada, J. P. Remeika, Phys. Rev. B 1970, 2, 155.

[5] E. Lines, A. M. Glass, Principles and Applications of Ferroelectrics and

Related Materials, OUP Oxford, 1977.

[6] K. Kushida, H. Takeuchi, Appl. Phys. Lett. 1987, 50, 1800.

[7] F. M. Pontes, D. S. L. Pontes, E. R. Leite, E. Longo, E. M. S. Santos, S.

Mergulhao, A. Chiquito, P. S. Pizani, F. Lanciotti, T. M. Boschi, J. A. Varela, J.

Appl. Phys. 2002, 91, 6650.

[8] R. Poyato, x, M. L. Calzada, L. Pardo, Appl. Phys. Lett. 2004, 84, 4161.

[9] H. Schmitt, J. Mendiola, F. Carmona, C. Alemany, B. Jimenez,

Microelectron. Eng. 1995, 29, 181.

[10] H. Takeuchi, S. Jyomura, C. Nakaya, "Highly Anisotropic Piezoelectric

Ceramics and Their Application in Ultrasonic Probes", presented at IEEE 1985

Ultrasonics Symposium, 16-18 Oct. 1985, 1985.

[11] C. Kittel, Physical Review 1951, 82, 729.

Page 51: Lead-free piezoelectric materials - Research Online

29

[12] B. Jaffe, Piezoelectric Ceramics, Elsevier Science, 1971.

[13] C. A. Randall, N. Kim, J.-P. Kucera, W. Cao, T. R. Shrout, J. Am. Ceram.

Soc. 1998, 81, 677.

[14] B. Noheda, D. E. Cox, G. Shirane, J. A. Gonzalo, L. E. Cross, S. E. Park,

Appl. Phys. Lett. 1999, 74, 2059.

[15] S. Mabud, J. Appl. Crystallogr. 1980, 13, 211.

[16] P. Ari-Gur, L. Benguigui, Solid State Commun. 1974, 15, 1077.

[17] R. Guo, L. E. Cross, S. E. Park, B. Noheda, D. E. Cox, G. Shirane, Phys. Rev.

Lett. 2000, 84, 5423.

[18] D. Damjanovic, Appl. Phys. Lett. 2010, 97, 062906.

[19] M. Budimir, D. Damjanovic, N. Setter, Phys. Rev. B 2006, 73, 174106.

[20] M. Ahart, M. Somayazulu, R. E. Cohen, P. Ganesh, P. Dera, H.-k. Mao, R. J.

Hemley, Y. Ren, P. Liermann, Z. Wu, Nature 2008, 451, 545.

[21] M. Pastor, P. K. Bajpai, R. N. P. Choudhary, Bull. Mater. Sci. 2005, 28, 199.

[22] J. Kuwata, K. Uchino, S. Nomura, Ferroelectrics 1981, 37, 579.

[23] A. Halliyal, A. Safari, Ferroelectrics 1994, 158, 295.

[24] E. K. Akdogan, A. Hall, W. K. Simon, A. Safari, J. Appl. Phys. 2007, 101,

024104.

[25] G. Robert, M. Demartin, D. Damjanovic, J. Am. Ceram. Soc. 1998, 81, 749.

[26] Y. Yamashita, K. Harada, T. Tao, N. Ichinose, Integrated Ferroelectrics

1996, 13, 9.

[27] J. Rödel, W. Jo, K. T. P. Seifert, E.-M. Anton, T. Granzow, D. Damjanovic,

J. Am. Ceram. Soc. 2009, 92, 1153.

[28] H. S. Shulman, M. Testorf, D. Damjanovic, N. Setter, J. Am. Ceram. Soc.

1996, 79, 3124.

Page 52: Lead-free piezoelectric materials - Research Online

30

[29] T. Takenaka, H. Nagata, J. Eur. Ceram. Soc. 2005, 25, 2693.

[30] Y. Noguchi, M. Miyayama, Appl. Phys. Lett. 2001, 78, 1903.

[31] T. Takenaka, K. Sakata, Ferroelectrics 1991, 118, 123.

[32] Y. Li, W. Chen, J. Zhou, Q. Xu, H. Sun, R. Xu, Materials Science and

Engineering: B 2004, 112, 5.

[33] C. Zhou, X. Liu, J Mater Sci 2008, 43, 1016.

[34] P. K. Panda, J Mater Sci 2009, 44, 5049.

[35] K. M. T. Takenaka, K. Sakata, , Jpn. J. Appl. Phys., Part 1 1991, 30, 4.

[36] X. X. Wang, H. L. W. Chan, C. L. Choy, Appl. Phys. A 2005, 80, 333.

[37] G. Shirane, R. Newnham, R. Pepinsky, Physical Review 1954, 96, 581.

[38] R. E. Jaeger, L. Egerton, J. Am. Ceram. Soc. 1962, 45, 209.

[39] E. Ringgaard, T. Wurlitzer, J. Eur. Ceram. Soc. 2005, 25, 2701.

[40] N. Marandian Hagh, B. Jadidian, A. Safari, J. Electroceram. 2007, 18, 339.

[41] Y. T. Lu, X. M. Chen, D. Z. Jin, X. Hu, Mater. Res. Bull. 2005, 40, 1847.

[42] Y. Saito, H. Takao, T. Tani, T. Nonoyama, K. Takatori, T. Homma, T.

Nagaya, M. Nakamura, Nature 2004, 432, 84.

[43] A. von Hippel, Reviews of Modern Physics 1950, 22, 221.

[44] T. Ryoko, F. Desheng, I. Mitsuru, D. Masahiro, K. Shin-ya, J. Phys.:

Condens. Matter 2009, 21, 215903.

[45] M.-S. Yoon, S.-C. Ur, Ceram. Int. 2008, 34, 1941.

[46] D. Fu, M. Itoh, S.-y. Koshihara, T. Kosugi, S. Tsuneyuki, Phys. Rev. Lett.

2008, 100, 227601.

[47] T. Mitsui, W. B. Westphal, Physical Review 1961, 124, 1354.

[48] D. Fu, M. Itoh, S.-y. Koshihara, Appl. Phys. Lett. 2008, 93, 012904.

[49] W. Liu, X. Ren, Phys. Rev. Lett. 2009, 103, 257602.

Page 53: Lead-free piezoelectric materials - Research Online

31

[50] D. Xue, Y. Zhou, H. Bao, C. Zhou, J. Gao, X. Ren, J. Appl. Phys. 2011, 109,

054110.

[51] B. Huixin, Z. Chao, X. Dezhen, G. Jinghui, R. Xiaobing, J. Phys. D: Appl.

Phys. 2010, 43, 465401.

[52] R. Prashanth, K. Sriram, R. Siddharth, N. W. Gregory, Smart Mater. Struct.

2012, 21, 094003.

[53] I. L. Guy, S. Muensit, E. M. Goldys, Appl. Phys. Lett. 1999, 75, 4133.

[54] D. A. Scrymgeour, T. L. Sounart, N. C. Simmons, J. W. P. Hsu, J. Appl.

Phys. 2007, 101, 014316.

[55] M.-A. Dubois, P. Muralt, Appl. Phys. Lett. 1999, 74, 3032.

[56] P. Eyben, F. Seidel, T. Hantschel, A. Schulze, A. Lorenz, A. U. De Castro, D.

Van Gestel, J. John, J. Horzel, W. Vandervorst, physica status solidi (a) 2011, 208,

596.

[57] X. Ou, P. D. Kanungo, R. Kögler, P. Werner, U. Gösele, W. Skorupa, X.

Wang, Nano Lett. 2009, 10, 171.

[58] A. Bernal, A. Tselev, S. Kalinin, N. Bassiri-Gharb, Adv. Mater. 2012, 24,

1159.

[59] Y. Kim, C. Bae, K. Ryu, H. Ko, Y. K. Kim, S. Hong, H. Shin, Appl. Phys.

Lett. 2009, 94, 032907.

[60] J. Y. Son, S. H. Bang, J. H. Cho, Appl. Phys. Lett. 2003, 82, 3505.

[61] L. Y. Kraya, R. Kraya, J. Appl. Phys. 2012, 111, 013708.

[62] N. Nobuyuki, Y. Takuma, S. Hiroyuki, S. Yoshihiko, M. Masayuki, W.

Shunji, Nanotechnology 1997, 8, A32.

[63] J. Stephen, P. B. Arthur, V. K. Sergei, Nanotechnology 2006, 17, 1615.

Page 54: Lead-free piezoelectric materials - Research Online

32

[64] S. V. Kalinin, A. Gruverman, Scanning Probe Microscopy Electrical and

Electromechanical Phenomena at the Nanoscale: Fundamentals and Applications,

Springer Science+Business Media, LLC, 2007.

[65] E. B. Araujo, E. C. Lima, I. K. Bdikin, A. L. Kholkin, J. Appl. Phys. 2013,

113, 187206.

[66] Y. Liu, K. H. Lam, K. K. Shung, J. Li, Q. Zhou, J. Appl. Phys. 2013, 113,

187205.

[67] V. Y. Shur, E. A. Mingaliev, V. A. Lebedev, D. K. Kuznetsov, D. V. Fursov,

J. Appl. Phys. 2013, 113, 187211.

[68] F. Borodavka, I. Gregora, A. Bartasyte, S. Margueron, V. Plausinaitiene, A.

Abrutis, J. Hlinka, J. Appl. Phys. 2013, 113, 187216.

[69] A. N. Morozovska, E. A. Eliseev, O. V. Varenyk, S. V. Kalinin, J. Appl.

Phys. 2013, 113, 187222.

[70] S. Elisabeth, J. Phys. D: Appl. Phys. 2011, 44, 464003.

[71] S. V. Kalinin, D. A. Bonnell, Phys. Rev. B 2002, 65, 125408.

[72] B. J. Rodriguez, A. Gruverman, A. I. Kingon, R. J. Nemanich, J. S. Cross, J.

Appl. Phys. 2004, 95, 1958.

[73] Z. L. Wang, J. Song, Science 2006, 312, 242.

[74] M. Bhaskaran, S. Sriram, S. Ruffell, A. Mitchell, Adv. Funct. Mater. 2011,

21, 2251.

[75] S. C. Masmanidis, R. B. Karabalin, I. De Vlaminck, G. Borghs, M. R.

Freeman, M. L. Roukes, Science 2007, 317, 780.

[76] P. Muralt, J. Micromech. Microeng. 2000, 10, 136.

[77] C.-R. Cho, J.-H. Koh, A. Grishin, S. Abadei, S. Gevorgian, Appl. Phys. Lett.

2000, 76, 1761.

Page 55: Lead-free piezoelectric materials - Research Online

33

[78] S.-M. Koo, S. Khartsev, C.-M. Zetterling, A. Grishin, M. Ostling, Appl. Phys.

Lett. 2003, 83, 3975.

[79] K.-I. Park, M. Lee, Y. Liu, S. Moon, G.-T. Hwang, G. Zhu, J. E. Kim, S. O.

Kim, D. K. Kim, Z. L. Wang, K. J. Lee, Adv. Mater. 2012, 24, 2999.

[80] X. Chen, S. Xu, N. Yao, Y. Shi, Nano Lett. 2010, 10, 2133.

[81] Y. Hu, Y. Zhang, C. Xu, G. Zhu, Z. L. Wang, Nano Lett. 2010, 10, 5025.

[82] H. D. Espinosa, R. A. Bernal, M. Minary-Jolandan, Adv. Mater. 2012, 24,

4656.

[83] B. J. Hansen, Y. Liu, R. Yang, Z. L. Wang, ACS Nano 2010, 4, 3647.

[84] M. Fan, W. Hui, Z. Li, Z. Shen, H. Li, A. Jiang, Y. Chen, R. Liu,

Microelectron. Eng. 2012, 98, 371.

[85] J. Wang, J. B. Neaton, H. Zheng, V. Nagarajan, S. B. Ogale, B. Liu, D.

Viehland, V. Vaithyanathan, D. G. Schlom, U. V. Waghmare, N. A. Spaldin, K. M.

Rabe, M. Wuttig, R. Ramesh, Science 2003, 299, 1719.

[86] J. J. Urban, W. S. Yun, Q. Gu, H. Park, J. Am. Chem. Soc. 2002, 124, 1186.

[87] J. J. Urban, J. E. Spanier, L. Ouyang, W. S. Yun, H. Park, Adv. Mater. 2003,

15, 423.

[88] L. Sagalowicz, F. Chu, P. D. Martin, D. Damjanovic, J. Appl. Phys. 2000, 88,

7258.

[89] P. M. Rørvik, T. Grande, M.-A. Einarsrud, Adv. Mater. 2011, 23, 4007.

[90] Y. Guo, K. Suzuki, K. Nishizawa, T. Miki, K. Kato, J. Cryst. Growth 2005,

284, 190.

[91] A. L. Kholkin, M. L. Calzada, P. Ramos, J. Mendiola, N. Setter, Appl. Phys.

Lett. 1996, 69, 3602.

Page 56: Lead-free piezoelectric materials - Research Online

34

[92] S. Fujino, M. Murakami, V. Anbusathaiah, S. H. Lim, V. Nagarajan, C. J.

Fennie, M. Wuttig, L. Salamanca-Riba, I. Takeuchi, Appl. Phys. Lett. 2008, 92,

202904.

[93] Y. Q. Chen, X. J. Zheng, X. Feng, Nanotechnology 2010, 21, 055708.

[94] S. H. Xie, J. Y. Li, Y. Qiao, Y. Y. Liu, L. N. Lan, Y. C. Zhou, S. T. Tan,

Appl. Phys. Lett. 2008, 92, 062901.

[95] A. Gruverman, A. Kholkin, Rep. Prog. Phys. 2006, 69, 2443.

Page 57: Lead-free piezoelectric materials - Research Online

35

Chapter 2. BTZ-BCT Ceramics

In this chapter, the impact of lattice parameters variations on the ferroelectric

properties in the barium titanate zirconate ─ barium calcium titanate ( (1-x)BTZ-

xBCT) system is investigated.

2.1 Introduction

Ferroelectricity in crystalline structures intrinsically is created by the electric dipole

developed by an off-centre atom located in the unit cell. Any change in the lattice

parameters can alter the electric dipole moment in the unit cell and consequently

changes the ferroelectric response. Several factors, such as doping, temperature

variation, crystal defects, internal/external strain, and electric field, can induce the

deformation in the unit cell which results in an enhancement or suppression of the

ferroelectric properties in the crystalline structures.

Several studies have been undertaken to explore the piezoelectric, pyroelectric, and

electromechanical properties of Ba (Ti0.8Zr0.2)TiO3- (Ba0.7Ca0.3)TiO3 (BTZ-BCT)

since it was introduced as a promising lead-free piezoelectric and ferroelectric

compound [1-6]. Since the (1-x)BTZ-xBCT system shows ferroelectricicity in addition

to piezoelectricity, a detailed study on the ferroelectricity and the compositional

dependency in this system is very much indicated. In this Chapter, the effects of Zr

and Ca on the lattice parameters and ferroelectric properties in the (1-x)BTZ-xBCT

system are investigated.

aj393
Highlight
aj393
Highlight
aj393
Highlight
aj393
Highlight
Page 58: Lead-free piezoelectric materials - Research Online

36

2.2 Experimental procedure

(1-x)Ba(Ti0.8Zr0.2)TiO3-x(Ba0.7Ca0.3)TiO3 ((1-x)BTZ-xBCT) ceramics with x = 0.0,

0.1,0.2,0.3,0.4,0.5 were prepared using the conventional solid-state reaction method.

BaCO3, CaCO3, TiO2, and ZrO2 were used as raw materials. Stoichiometric ratios of

the materials were mixed via planetary ball milling. Zirconia balls were used, and

distilled water was employed as the ball-milling medium to increase the efficiency.

The resultant slurries were dried in an oven at 100oC for 12 hrs. The mixed powders

were pressed into disks 12 mm in diameter and 3 mm in thickness under 150 MPa

using a uniaxial pressure and calcined at 1300°C for 2 h. Sintering of samples was

carried out at 1450°C for 2 h in the air. The polished surfaces of the samples were

coated with silver paste for electrical and piezoelectric measurements. The crystal

structure and phase content of the ceramics were studied by X-ray diffraction (XRD;

GBC MMA, CuKα radiation, 40 kV, 25 mA) and Raman spectroscopy (HORIBA

Jobin Yvon, HR800) using a He-Ne laser with 632.8 nm wavelength. A field

emission scanning electron microscope (FE-SEM, JEOL JSM 7500FA) was used to

study the morphology and particle size in the ceramics. The electric field dependence

of the polarization (P vs. E dependence) of the ceramics was measured using an HP

42980 LCR-meter. The piezoelectric coeffcient d33 of the ceramics were measured

using a d33 meter ( YE2730A, APC International. Ltd.) at room temperature.

2.3 Results and discussions

2.3.1 Crystalline structure and vibration modes

Figure 2.1 shows XRD patterns of (1-x)BTZ-xBCT ceramics with x = 0.1, 0.2, 0.3,

0.4, and 0.5 molar ratios sintered at 1450°C for 2 h in the air. The XRD patterns

indicate that all samples were fully crystallized in a perovskite structure. Introducing

aj393
Highlight
aj393
Highlight
Page 59: Lead-free piezoelectric materials - Research Online

37

the BCT into the BZT structure, which means reducing the Zr4+ concentration and

increasing the Ca2+ concentration in the primary BaTiO3 structure, leads to a shift in

the peak positions in the XRD patterns, as is shown in Figure 2.1 for a peak located

at 2θ ≈ 31.5º.

Figure 2.1 XRD patterns (left) of the (1-x)BTZ-xBCT ceramics sintered at 1450°C for 2 h in

air and enlarged peaks located at 2θ ≈ 31.5º fitted by the Lorentzian function for different x

values. Coexistence of two polar phases, both tetragonal (T) and rhombohedral (R), confirms

the crystallization of BTZ-BCT ceramics in the vicinity of the MPB region. The peak shifts

observed in different samples can be attributed to the internal strain developed by the

substitution of Ca2+ in the Ba+2 sites and Ti+4 in the Zr4+ sites in the BTZ ceramics.

Substitution of the divalent Ca+2 (1.06 A°) with smaller radius than Ba+2 (1.43 Å) on

the A-site and removing the equivalent Zr+4 (0.87 Å), which has a larger radius than

Ti+4 (0.64 Å), from the B-site of the ABO3 structure results in smaller lattice

constants as the concentration of the BCT increases. Figure 2.2 displays the Rietveld

aj393
Highlight
Page 60: Lead-free piezoelectric materials - Research Online

38

refinement results for the XRD patterns. The lattice constants and unit cell volumes

in the (1-x)BTZ-xBCT ceramics extracted from the refinement of the XRD patterns

are presented in Table 2.1. The results indicate that introducing the BCT into the

BTZ structure leads to a contraction in the unit cells in the system and reduces the

unit cell volume.

Figure 2.2 Refinement patterns of the (a) BTZ, (b) 0.9BTZ-0.1BCT, (c) 0.8BTZ-0.2BCT,

(d) 0.7BTZ-0.3BCT, and (e) 0.6BTZ-0.4BCT ceramics sintered at 1450°C for 2 h in air. The

symbols mark the experimental results for the XRD pattern that is fitted in red. The short

blue vertical lines mark the line positions of the standard, and the green spectrum at the

bottom is the difference spectrum between the fit and the experimental results.

BTZ (a) (b) 0.9BTZ-0.1BCT

(c) 0.8BTZ-0.2BCT (d) 0.7BTZ-0.3BCT

(e) 0.6BTZ-0.4BCT

aj393
Highlight
Page 61: Lead-free piezoelectric materials - Research Online

39

Crystal structures in the ceramics were determined by analysing of the intense peak

located at 2θ ≈ 31.5º in the XRD patterns of all the samples (Figure 2.1).

A Lorentzian peak fitting function was employed to study the crystalline phase

configuration and changes in the samples. It is demonstrated that by increasing the

BCT in the system, the primary structure experiences more distortion and lattice

changes, which are reflected in the XRD peak positions and full-width-half-

maximum values (FWHMs) for the peaks in the samples. Evolution of the tetragonal

structure in the BTZ-BCT ceramics is evident in Figure 2.1.

The coexistence of two ferroelectric phases, the tetragonal and rhombohedral,

enhances the polarizability of the system and elininates the energy barrier against

switching of the polarization direction [1]. Crystallization in the vicinity of the

morphotropic phase boundary (MPB) and approaching the triple critical point (TCP)

in the BTZ-BCT ceramic dramatically improve its piezoelectric and ferroelectric

properties, as reported earlier by Liu et.al. [1].

Field emission scanning electron microscope images of the BTZ-BCT ceramic are

shown in Figure 2.3 (a) and (b). The images were tcollected without a conducting

coating to preserve the inherent topography of the grains. Figure2.3 (a) shows

diverse particle sizes ranging from the submicron to about 3 µm. Different crystal

defects such as edge and screw dislocations form suitable sites for the crystal growth

in BTZ-BCT ceramic. Figure 2.3(b) displays the growth steps on grains.

aj393
Highlight
aj393
Highlight
Page 62: Lead-free piezoelectric materials - Research Online

40

Table 2.1 Lattice parameters and unit cell volumes of the (1-x)BTZ-xBCT ceramics

extracted from the Rietveld refinement results on the ceramics.

X= 0 X= 0.1 X= 0.2 X= 0.3 X= 0.4 X= 0.5*

Symmetry R3m R3m R3m R3m R3m R3m P4mm

Lattice

constants

a, b, c (Å) 4.043 4.034 4.029 4.016 4.009 4.000 3.998,

4.016

α (deg.) 89.72 89.87 89.93 89.97 89.90 89.97 90.00

Unit cell

volume (Å3)

66.08 65.64 65.41 64.77 64.44 64.11

Atomic

positions

(Ba, Ca)

x 0.0 0.0 0.0 0.0 0.0 -

y 0.0 0.0 0.0 0.0 0.0 -

z 0.0 0.0 0.0 0.0 0.0 -

(Ti, Zr)

x 0.4872 0.4850 0.4995 0.4893 0.4872 -

y 0.4872 0.4850 0.4995 0.4893 0.4872 -

z 0.4872 0.4850 0.4995 0.4893 0.4872 -

O

x 0.5109 0.5131 0.5152 0.5364 0.5109 -

y 0.5109 0.5131 0.5152 0.5364 0.5109 -

z 0.0193 -0.0251 -0.0312 0.0301 0.0193 -

Rwpa 15.47 18.63 19.19 16.83 19.45 -

GOFb 3.29 2.26 2.46 3.94 2.84 -

a) The R-weighted pattern (Rwp) is used to estimate the agreement between the observations and the

model during the course of the refinement (5% < Rwp < 10%: very good agreement; 10% < Rwp <

20%: typical).

b) Goodness of fitting (GOF).

*Lattice parameters and unit cell volume were calculated using corresponding equations for the

tetragonal and rhombohedral structures [7].

aj393
Highlight
Page 63: Lead-free piezoelectric materials - Research Online

41

Figure 2.3 (a) FE-SEM image of the particle size distribution and morphology of the grains

in the BTZ-BCT ceramic sintered at 1450°C for 2 h in the air. (b) Higher magnification

image of the area indicated by the square in (a) that shows the growth steps in individual

grains.

Different vibration modes in the (1-x)BTZ-xBCT system were investigated by

Raman spectroscopy at room temperature (Figure 2.4). Spectra of the ceramics are

compared with BaTiO3 as their primary structure.

Figure 2.4 Different vibration modes in the (1-x)BTZ-xBCT system were investigated by

Raman spectroscopy at room temperature.

aj393
Highlight
aj393
Highlight
Page 64: Lead-free piezoelectric materials - Research Online

42

All features characteristic of the perovskite structure of the bulk BaTiO3 can be

verified in the Raman spectra of all the ceramics [8, 9] .

According to the vibration modes in BaTiO3, the Raman peaks located at about 520

and 724 cm-1 are generated by the (Ba,Ca)—O bonds, and the peaks at about 300 cm-

1 are due to the (Ti,Zr)—O bond vibrations in the (1-x)BTZ-xBCT system [8-10].

According to the ionic radii of the elements that exist in the system, increasing the

amount of Ca2+, with an ionic radius smaller than its host ion, Ba2+, in unit cells leads

to shorter (Ba, Ca)—O bonds, and higher vibration frequencies and a blue shift in the

bands located at about 520 and 724 cm-1. By increasing the BCT concentration in the

system, the concentration of the Zr4+, which has a larger ionic radius compared to

Ti4+, decreases, and a red shift in the band generated by the (Ti,Zr)—O bonds occurs.

2.3.2 Ferroelectric and piezoelectric properties

In order to study the ferroelectric properties of the ceramics, the polarization versus

electric field (P-E loop) changes of the bulk ceramics were studied. Generally, the

switching of the spontaneous polarizations in ferroelectrics yields a hysteresis loop

when they are imposed by a strong enough alternating electric field from positive to

negative. Figure 2.5 illustrates a typical P-E loop in a ferroelectric material. When an

electric field is applied to a ferroelectric crystal, in the OA part, the spontaneously

polarized ferroelectric domains begin to rotate, but they do not enter a permanent

state due to the weak electric field and behave similarly to a typical dielectric

material.

aj393
Highlight
Page 65: Lead-free piezoelectric materials - Research Online

43

Figure 2.5 Typical polarization versus electric field hysteresis loop (P-E loop) in a

ferroelectric material[11].

As the electric field is increased, a dramatic enhancement occurs (AB) in the

allignment of the domains, which reaches a maximum at point C. In this state, almost

all of the domains are aligned in the direction of the electric field (but with the

opposite polarity) and net polarization in the system is at its saturation level. This

value of the polarization is called the saturation polarization (OE). Even when the

electric field is then reduced to zero, a part of the aligned domains keep their

arrangement and a remanent polarization (Pr) at zero electric field is achieved (OD).

Rotation of the domains continues by reversing the electric field, and a similar

phenomenon can be seen in the opposite arrangment. The required electric field for

switching the spontaneous polarization/ferroelectric domains is called the coercive

field (OF) which exists for both positive and negative electric fields. The G and H

points in the negative region correspond to the C and D points in the positive region.

In order to investigate the effects of Zr and Ca variations on the ferroelectric

properties of the (1-x)BTZ-xBCT, system the P-E loops of the ceramics were

collected at 100 Hz sweep of the electric field. Figure 2.6(a) demonstrates that the

maximum or saturation polarization (Ps) is increased by introducing the BCT into the

aj393
Highlight
Page 66: Lead-free piezoelectric materials - Research Online

44

BTZ structure, and rthe maximum Ps= 13.32 μC cm-2 at x = 0.5 in the (1-x)BTZ-

xBCT system.

The composition dependence of the polarization elements leads to an enhancement of

about two times in the BTZ-BCT ceramic compared to the BTZ (Ps = 6.96 μC cm-2)

under the same electric field.

The correlation between the spontaneous polarization Ps and the volume of the unit

cell is as follows:

𝑃𝑠 = 1𝑣

∑ 𝛿𝑖𝑖 𝑍𝑖 (2.1)

Where 𝑣 is the unit cell volume, 𝛿𝑖 is the displacement of an atom 𝑖 from a

centrosymmetric site in the unit cell, and 𝑍𝑖 is the effective nuclear charge of tatom 𝑖.

Reducing the bonding distances in the Ba—O and Zr—O bonds, as was discussed

earlier, in the (1-x)BTZ-xBCT system by Ca2+ and Ti4+ replacements results in a

shrinkage in the unit cell volume and provides more space (Ti4+ has a smaller ionic

radius compared to Zr4+) for the displacement of an off-centre atom inside the unit

cell, and therefore, according to Equation (2.1), the spontaneous polarization

increases with substitution of the Ba2+ and Zr4+ ions by Ca2+ and Ti4+.

aj393
Highlight
Page 67: Lead-free piezoelectric materials - Research Online

45

Figure 2.6 (a) Maximum polarization Ps is increased by introducing BCT into the BTZ

structure, and the maximum Ps= 13.32 μC cm-2 at x = 0.5 in the (1-x)BTZ-xBCT system. (b)

Correlation of the polarization saturation enhancement with the decrease in the unit cell

volume in the (1-x)BTZ-xBCT ceramics. (c) The BTZ-BCT ceramic achieved the maximum

spontaneous polarization Ps = 13.86 μC cm-2 under an electric field E = 40 kV cm-1 at room

temperature. (d) Piezoelectric response enhancement in the BTZ-BCT ceramic compared to

other compositions in this system.

Figure 2.6(b) displays the correlation between the polarization saturation

enhancement and the decrease in the unit cell volume in the (1-x)BTZ-xBCT

ceramics. It must be taken into the account that approaching the MPB in the BTZ-

BCT also eliminates the anisotropy of polarization, and this has an important impact

on the enhancement of the ferroelectric properties in this composition compared to

the other ceramics in this system [1].

(a) (b)

(c) (d)

aj393
Highlight
aj393
Highlight
Page 68: Lead-free piezoelectric materials - Research Online

46

The BTZ-BCT ceramics achieved the maximum spontaneous polarization Ps = 13.86

μC cm-2 under an electric field E = 40 kV cm-1 at room temperature (Figure 2.6(c)).

The (1-x)BTZ-xBCT ceramics used for the ferroelectric measurements were

polarized for the piezoelectric coefficient measurements. Poling was performed by

employing a 3 kV mm-1 DC electric field at room temperature to align the

spontaneous polarizations developed in the ceramics and create a non-zero-net

polarization.

Significant enhancement in the piezoelectric coefficient d33 = 356 pC N-1 was

observed in the BTZ-BCT ceramics. Referring to the other studies carried out on this

system, the same trend in the (1-x)BTZ-xBCT piezoelectric behaviour is observed in

our samples, however, the measured d33 values are less than the maximum

piezoelectric coefficients reported by W.Liu et.al.[1-3, 12, 13]. Figure 2.6(d) presents the

piezoelectric response enhancement in the BTZ-BCT ceramics compared to other

compositions in this system. Regarding the effects of the MPB on the piezoelectric

properties explained in Chapter 1, approaching the triple critical point (TCP),

crystallization in the vicinity of the MPB, and flattening of the energy barrier against

rotation of the polarization into different possible directions, results in high

piezoelectric performance in BTZ-BCT ceramic.

2.4 Conclusions

(1-x)BTZ-xBCT ceramics were synthesized using the conventional solid state

reaction at 1450°C in air. Crystal structures, lattice parameters, and unit cell volumes

of the ceramics were determined by employing the Rietveld method to refine the

corresponding XRD patterns. Structural analysis results, combined with the changes

in ferroelectric properties observed in the ceramics, reveal that reducing the unit cell

aj393
Highlight
aj393
Highlight
Page 69: Lead-free piezoelectric materials - Research Online

47

volume in combination with enhancement of displacement of toff-centre atoms in

this system leads to an enhancement in the ferroelectric performance in the (1-

x)BTZ-xBCT system as x is increased to 0.5.

2.5 References

[1] W. Liu, X. Ren, Phys. Rev. Lett. 2009, 103, 257602.

[2] B. Huixin, Z. Chao, X. Dezhen, G. Jinghui, R. Xiaobing, J. Phys. D: Appl.

Phys. 2010, 43, 465401.

[3] D. Xue, Y. Zhou, H. Bao, C. Zhou, J. Gao, X. Ren, J. Appl. Phys. 2011, 109,

054110.

[4] J. Gao, D. Xue, Y. Wang, D. Wang, L. Zhang, H. Wu, S. Guo, H. Bao, C.

Zhou, W. Liu, S. Hou, G. Xiao, X. Ren, Appl. Phys. Lett. 2011, 99, 092901.

[5] S. K. Ye, J. Y. H. Fuh, L. Lu, Appl. Phys. Lett. 2012, 100, 252906.

[6] Y. Shanshan, R. Wei, J. Hongfen, W. Xiaoqing, S. Peng, X. Dezhen, R.

Xiaobing, Y. Zuo-Guang, J. Phys. D: Appl. Phys. 2012, 45, 195301.

[7] M. De Graef, M. E. McHenry, Structure of Materials: An Introduction to

Crystallography, Diffraction and Symmetry, Cambridge University Press, 2007.

[8] U. D. Venkateswaran, V. M. Naik, R. Naik, Phys. Rev. B 1998, 58, 14256.

[9] Y. Shiratori, C. Pithan, J. Dornseiffer, R. Waser, J. Raman Spectrosc. 2007,

38, 1300.

[10] M. C. Chang, S.-C. Yu, J. Mater. Sci. Lett. 2000, 19, 1323.

[11] W. Zhang, H.-Y. Ye, R.-G. Xiong, Coord. Chem. Rev. 2009, 253, 2980.

[12] J. Wu, D. Xiao, W. Wu, Q. Chen, J. Zhu, Z. Yang, J. Wang, Scripta Mater.

2011, 65, 771.

Page 70: Lead-free piezoelectric materials - Research Online

48

[13] P. Wang, Y. Li, Y. Lu, J. Eur. Ceram. Soc. 2011, 31, 2005.

Page 71: Lead-free piezoelectric materials - Research Online

49

Chapter 3. BTZ-BCT Nanofibers

This chapter presents structural properties investigations, quantitative piezoelectric

and electrical characterizations, and resistance distribution of high performance Ba

(Ti0.8Zr0.2)TiO3- (Ba0.7Ca0.3)TiO3 (BTZ-BCT) piezoelectric nanofibers using different

microscopy techniques. Crystalline structure development and phase content of the

nanofibers have been studied, and effects of the crystallization in morphotropic phase

boundary on the piezoelectric performance of the nanofibers are discussed.

3.1 Introduction

Quasi-one-dimensional (1D) ferroelectric materials including nanofibers, nanotubes,

nanowires, and nanobelts have emerged as a hot research topic due to their novel

properties and applications. Owing to their low-dimension and high aspect ratio, they

exhibit distinct properties from their respective bulk materials. They could be

exploited at very high frequencies and tolerate significantly high strains. For

example, by utilizing piezoelectric nanofibers in nanogenerators, energy harvesting

and self-powered devices can exhibit more efficiency and output voltage in very

small sizes [1-9]. There have been several reports on fabrication and application of the

piezoelectric nanofibers, however, lack of information on their quantitative electrical

properties and structural states have hindered further attempts towards improving

physical and electrical properties characterizations. Understanding the structure and

electrical properties of the ferroelectric materials on the nanoscale is crucial for the

improvement of these devices and requires accurate quantitative characterizations.

Scanning probe microscopy (SPM) techniques such as Kelvin probe force

microscopy (KPFM), scanning tunnelling microscopy (STM), scanning spreading

aj393
Highlight
Page 72: Lead-free piezoelectric materials - Research Online

50

resistance microscopy (SSRM), and piezoresponse force microscopy (PFM) have

emerged as powerful tools to study surface electrical properties, ferroelectric

domains, and piezoelectric response in the low-dimensional materials [1, 10-14].

In the study of piezoelectricity, conventional piezoresponse force microscopy (PFM)

has been widely used to measure the piezoelectric response, and the domain and

domain wall dynamics. It gives accurate quantitative results on the nanoscale.

The electrical resistance is an essential property in the piezoelectric and ferroelectric

materials. High resistance reduces the dissipation of the generated charges on the

surfaces and decreases the leakage current in ferroelectric materials, thus improving

the piezoelectric response and ferroelectric performance[15]. The low leakage current

reduces the risk of early breakdown in the piezoelectric materials when an electric

field is applied to create a displacement.

Distribution of the resistance over the nanofibers must be taken into the account. A

non-uniform distribution of the resistance may lead to an inhomogeneous distribution

of the electric field when a voltage is applied to the nanofibers and creates uneven

stresses and strains in the nanofibers. These can suppress the piezoelectric

performance of the nanofibers and reduce their lifetime.

SSRM is a powerful technique to map the resistance distribution of a sample. A DC-

bias voltage applied between the tip and the sample results in a current which passes

through the surface. The topographic and current distribution images of the surface

collected in SSRM provide valuable information on the nanoscale.

3.2 Experimental procedure

The BTZ-BCT nanofibres were synthesized using sol-gel assisted electrospinnig

technique. A solution of (Ba0.85Ca0.15)(Ti0.90Zr0.10)O3 precursor (10 mL ) was

aj393
Highlight
Page 73: Lead-free piezoelectric materials - Research Online

51

prepared by dissolving barium acetate and calcium acetate in glacial acetic acid at 60

°C and cooling down the solution to room temperature, followed by mixing the Zr

and Ti solution prepared from dissolved titanium isopropoxide and zirconium

propoxide in ethanol. In order to prepare the BTZ-BCT solution for the

electrospinning, 350 mg of polyvinylpyrrolidone (PVP, MW= 1,300,000) was

dissolved in the BTZ-BCT precursor solution and stirred for 1 hour in a tightly

capped 20 ml glass bottle. The transparent and viscous solution was transferred into a

plastic syringe.

Electrospinning was carried out by applying 15 kV between the metallic needle of

the syringe pump and the aluminium foil collector. The solution was fed at a rate of 1

ml h-. A non-woven and bead-free nanofiber mat containing hydrolyzed BTZ-BCT

precursor and PVP were collected from the surface of the Al-foil collector placed at

9 cm below the needle. The mat was dried at 100 °C for 12 hours followed by

calcination at 500, 600, 700 and 800 °C for 1 hour in air. The heating/cooling rate

was 5 °C min-1.

Evolution of the crystal structure of the nanofibers calcined at different temperatures

was studied by X-ray diffraction (XRD, GBC MMA powder diffractometer, CuKα

radiation, 40 kV, 25 mA). Raman spectra were collected (HORIBA Jobin Yvon,

HR800 spectrometer) at room temperature in air by pumping samples with 632.8 nm

He-Ne laser light. A field emission scanning electron microscope (FE-SEM, JEOL

JSM 7500FA) was used to collect images of the nanofibers. Transmission electron

microscopy (TEM) was performed using a JEOL 2011 200 keV analytical

instrument. Samples were prepared by dispersion onto “Quantifoil” holey carbon

support film so that sample regions located over holes in the support film could be

examined.

aj393
Highlight
Page 74: Lead-free piezoelectric materials - Research Online

52

Piezoresponse force microscopy (PFM, Asylum Research, MFP-3D) for

piezoelectric properties measurements and a SPM operating in the SSRM and

scanning capacitance microscopy (SCM) modes (Digital Instrument, Dimension

3100) were used. A highly boron doped diamond coated Si tip for the SSRM and a

PtIr-coated Si tip for the PFM and SCM were employed. The bottom electrode was

an Ir layer on Si wafer, and the top electrode was the conducting tip.

3.3 Results and discussion

3.3.1 Microstructure and crystalline phase evolution

Field-emission scanning electron microscope images of the BTZ-BCT nanofibers

calcined at different temperatures, ranging from 500°C to 800 °C for 1 hour in air,

are displayed in Figures 3.1(a) to (d) .The evaporation and burning of the volatile

organic compounds during the calcination process lead to a significant shrinkage in

the nanofibers. The average diameter of nanofibers, which are tens of microns in

length, is about 200 nm for the sample annealed at 700 °C, but it is reduced to about

150 nm by annealing at 800 °C. Such sintering shrinkage occurs, as it is believed,

through the mass transport along grain boundaries to the neck between adjacent

grains [16, 17].

aj393
Highlight
Page 75: Lead-free piezoelectric materials - Research Online

53

Figure 3.1 FE-SEM images of the BTZ-BCT nanofibers (a) calcined at 500°C, (b) at 600°C

(c) at 700°C. And (d) at 800°C for 1 hour. Insets: highly magnified views of annealed BTZ-

BCT nanofibers. All images were collected under 0.5 kV acceleration voltage and 3.7 mm

working distance without conductive coating.

Microanalysis of the nanofibers annealed at 700°C in air that are displayed in Figure

3.2 confirms the presence of the essential constituent elements Ba, Ca, Ti, and Zr in

the nanofibers. The quantitative analysis results shown in the inset table in Figure 3.2

present the concentrations of each element. Considering the low stoichiometric

concentrations of the Ca and Zr in the BTZ-BCT formulation, the quantitative results

are reasonable.

aj393
Highlight
Page 76: Lead-free piezoelectric materials - Research Online

54

Figure 3.2 Energy dispersive X-ray spectroscopy (EDS) of the BTZ-BCT nanofibres

annealed at 700°C. The inset table shows the concentrations of the different elements.

Crystalline structure and phase evolution were examined by X-ray diffractometery

on the nanofibers annealed at different temperatures from 500°C to 800°C (Figure

3.3(a)). For 500 °C, no Bragg reflections corresponding to a crystalline structure

were observed. The broad peak at 2θ ~ 26 ° indicates the presence of amorphous

carbon [18]. BaCO3 nanocrystals formed at 600 °C ( 2θ = 26.7 °) [19] were no longer

stable at higher temperatures, and no residual BaCO3 phase was present due to its

thermal decomposition at temperatures above 600 °C. After annealing the nanofibers

at 700 °C and 800 °C, the onset of BTZ-BCT perovskite crystallization becomes

evident in the XRD patterns. Bragg reflections in Figure 3.3 are indexed according to

the PDF reference cards 75-2116 for tetragonal and 85-0368 for rhombohedral

structured BaTiO3 [30]. No detectable impurity phases were observed. In Figure 3.3

(b) and (c) enlarged scans of the main Bragg diffraction reflections for {110} at 2θ =

31.7 ° and {200} at 2θ = 45.5 ° for the BTZ-BCT nanofibers annealed at 700°C were

fitted using Lorentzian functions.

aj393
Highlight
aj393
Highlight
Page 77: Lead-free piezoelectric materials - Research Online

55

Figure 3.3 (a) XRD patterns of the BTZ-BCT nanofibers annealed at different temperatures

with corresponding Miller indices of BaTiO3. Enlarged scans of Bragg diffraction

reflections (b) for {110} at 2θ = 31.7 ° and (c) {200} at 2θ = 45.5 ° for the BTZ-BCT

nanofibers annealed at 700 °C. The peaks are fitted by Lorentzian functions. Coexistence of

rhombohedral and tetragonal phases proves that crystallization of the nanofibers has

occurred in the vicinity of the morphotropic phase boundary.

Coexistence of the tetragonal and rhombohedral crystalline structures indicates that

the nanofibers were successfully crystallized in the morphotropic phase boundary

(MPB) region. This occurs at remarkably lower temperature compared to the bulk

sample prepared by W. Liu et al. at ~1450 °C[20]. We attribute this to the much higher

surface-area-to-volume ratio of the nanosized fibers compared to micron-sized grains

in ceramics grown by solid-state reaction. The crystallite size of the nanofibers

calcined at 700°C and 800 °C was calculated using the Scherrer equation from the

aj393
Highlight
Page 78: Lead-free piezoelectric materials - Research Online

56

full-width-at-half-maximum (FWHM) values of the Bragg peaks at 2θ ≈ 31.70 °. The

tetragonal crystallite size increased from about 43 nm to 58 nm and the

rhombohedral crystallite size from 17 nm to 30 nm on average with 100°C increment

in calcination temperature from 700°C.

3.3.2 Vibration modes

The local atomic environment in the BTZ-BCT nanofibers was examined by Raman

spectroscopy at room temperature (Figure 3.4).

Figure 3.4 Raman scattering spectra of the BTZ-BCT nanofibers annealed at 500, 600, 700,

and 800 °C for 1 hour in air. Since the [A1(TO)] peak at 164 cm-1 exists only in the

rhombohedral phase of BaTiO3 nanocrystals, it indicates the coexistence of rhombohedral

(R) and tetragonal (T) phases in the MPB region at temperatures above 700°C.

All features characteristic of the perovskite structure of bulk BaTiO3 were verified in

aj393
Highlight
aj393
Highlight
Page 79: Lead-free piezoelectric materials - Research Online

57

the Raman spectra of bulk BTZ-BCT ceramics in different forms.[21, 22] Two A1

transverse optical modes [A1(TO)] and a longitudinal optical mode [A1(LO)],

together with a combined mode [B1,E(TO+LO)], are located in the range of 100 to

300 cm-1. Raman shifts at around 260 and 300 cm-1 are attributed to the Ti-O bond

vibrations. Ba-O bonds produce two mixed modes, [A1,E(TO)] and [A1,E(LO)], at

the high frequencies of about 520 and 725 cm-1. [21, 23] The A1 and E modes are

present in both Raman and infrared spectra, however, the B1 mode is only Raman

active. [21] The low intensity peak at 164 cm-1, assigned to the transverse optical

mode [A1(TO)], has been observed only in the rhombohedral phase of

nanocrystalline barium titanate, [24] while the other vibrational modes exist in both

tetragonal and rhombohedral phases. The coexistence of these two ferroelectric

phases in the Raman spectra confirms the crystallization of the BTZ-BCT nanofibers

and thin films in the MPB region.

Strain developed at the BTZ-BCT nanostructure-substrate interface together with the

substitution of Ca2+ and Zr4+, with ionic radii of 114 and 86 pm, respectively, for

Ba+2 and Ti+4 (with ionic radii of 146 pm and 74.5 pm, respectively) could be

responsible for the shift in the Raman spectra of BTZ-BCT bulks and nanofibers

compared to the polycrystalline BaTiO3. [22]

The signal appearing at 820 cm-1 could be generated by lattice defects caused by A or

B site vacancies in the ABO3 perovskite structure [25]. Three high frequency peaks in

the range of 1600 cm-1 to 1800 cm-1 are made faint by increasing the calcination

temperature up to 800 °C. The Raman revealed crystallization of the BTZ-BCT

nanofibers above 700 °C in coexisting tetragonal and rhombohedral symmetries is in

a good agreement with the XRD results and indicates the crystallization of the

nanofibers in the MPB region.

aj393
Highlight
aj393
Highlight
Page 80: Lead-free piezoelectric materials - Research Online

58

TEM studies were conducted for further investigation of the crystal structure and

phase content of the BTZ-BCT nanofibers. Regarding to crystallite sizes calculated

for tetragonal (43 nm) and rhombohedral phases (17 nm) at 700°C, large particles

correspond to the tetragonal phase, and small particles indicate rhombohedral phase

in Figure 3.5(a).

3.3.3 Crystalline phase observations by TEM

Figure 3.5 TEM results obtained from sample heat-treated at 700° C: (a) low magnification

image of a nanofibre containing larger tetragonal and smaller rhombohedral particles, with

inset selected area electron diffraction pattern; (b) indexing of the regions indicated in red

and green in (a) of {110} group reflections according to the indicated rhombohedral (R) and

tetragonal (T) phase reflections; (c) a second selected area diffraction pattern with

rhombohedral and tetragonal reflections as indicated; (d) high magnification image with

region containing fine twins indicated.

Page 81: Lead-free piezoelectric materials - Research Online

59

Analysis of the electron diffraction pattern acquired from the same fibre displayed in

Figure 3.5(a) shows that the nanofibers are composed of two crystalline structures,

tetragonal (T) and rhombohedral (R) (Figure 3.5(b) and(c)). Nanotwins about 10 nm

in length and 1-2 nm in width were observed in the high magnification image

displayed in Figure 3.5(d) as previously reported for BaTiO3 [26-28]. These twins could

be formed during the grain growth (growth twins) at high temperatures in the cubic

structure that is stable in ferroelectric phases, or they are generated by strain that

arises in the cubic-tetragonal or tetragonal-rhombohedral phase transitions

(deformation twins) below the Curie temperature. [29-31]. Twinning occurs to release

some of the elastic strains induced by the volume change that subsequently reduces

the total system energy. Lattice fringes near the twin boundaries show a symmetrical

mirror arrangement of the atoms in that area. Crystal defects such as twins, stacking

faults, and edge and screw dislocations act as pinning sites against the movement of

the ferroelectric domains and reduce their mobility. Suppression of the domain wall

mobility, either by the high density of grain boundaries due to the nanometer size

particles or crystal defects in the BTZ-BCT nanofibers reduces its piezoelectric

response [32, 33].

3.3.4 Piezoelectric and ferroelectric measurements by PFM

PFM was employed to quantify the piezoelectric coefficient in the BTZ-BCT fibers.

The deflection signal was calibrated by using the force-distance curve for the

cantilever (Figure 3.6) used for PFM measurements.

aj393
Highlight
Page 82: Lead-free piezoelectric materials - Research Online

60

Figure 3.6. Force-distance curve recorded from the surface of a silicon substrate using the

cantilever employed in our PFM measurements. This plot was used for the optical sensitivity

calibration of the cantilever.

Hysteretic bias voltage dependencies of the phase of the piezo-signal, the

displacement, and the piezoelectric coefficient in Figure 3.7(a)-(c) manifest the

switching of local ferroelectric domains in the BTZ-BCT thin film and nanofiber. A

DC-bias voltage, ranging from -5 V to +5 V, was employed. The PFM phase

hysteresis loops in Figure 3.7(a) show the 180° rotation of the electrical polarization

of domains in the BTZ-BCT nanofiber. Switching of polarization occurs due to an

inhomogeneous nucleation and anisotropic growth of domains [34] in external electric

field applied between the bottom electrode (Ir layer on Si wafer) and the top

electrode (PtIr-coated Si tip). The asymmetric shape of the hysteresis loops and their

shift toward the positive field (PtIr tip as anode) is due to the difference in work

functions, respectively, in the bottom and top electrodes: 4.23 eV for Ir and 5.6 eV

for PtIr. [35, 36]

aj393
Highlight
Page 83: Lead-free piezoelectric materials - Research Online

61

Figure 3.7 (a) Local PFM phase hysteresis loops and (b) piezoelectric butterfly loops of the

BTZ-BCT nanofiber obtained by switching the dc-bias voltage from -5 to +5 volts. 180°

domain switching in the phase hysteresis loop shows the switching of spontaneous electrical

dipoles built up in the BTZ-BCT fiber. (c) d33 hysteresis loop calculated by using the

converse piezoelectric equation, Δz = d33V, which indicates a significant enhancement of d33

= 180 pmV-1 in the BTZ-BCT nanofiber. (d) 3D topographic atomic force microscope

(AFM) image of the nanofiber used for the PFM measurements.

Figure 3.7(b) shows the displacement ∆z of the tip caused by the deformation of the

ferroelectric sample under applied electric field developed by the bias voltage.

Hysteretic curves exhibit an irreversible converse piezoelectric effect in BTZ-BCT

nanofiber. According to the definition of the converse piezoelectric effect, the

piezoelectric coefficient d33 can be calculated from the equation:

𝑑33 = ∆zV

(3.1) [37]

aj393
Highlight
aj393
Highlight
Page 84: Lead-free piezoelectric materials - Research Online

62

d33 values as high as 180 pmV-1 in fiber were obtained from the slope of the

linear part of the Δz vs. bias voltage curve at very low voltage.

The observed piezoelectric coefficient in our nanofibers exceeds that of all the

reported piezoelectric nanofibres, and comparable to the PZT ones. Table 3.1

compares the piezoelectric coefficient d33 of lead-containing and lead-free

piezoelectrics fabricated in the form of nanofibres.

Table 3.1 Comparison of the piezoelectric coefficient d33 of the BTZ-BCT nanofibres with

lead-based and lead-free piezoelectric nanofibres.

Materials Orientation d33(pm V-1) Ref.

BTZ-BCT Random 180 Current study

Pb(Zr0.3Ti0.7)O3 Random 83 [38]

0.65Pb(Mg1/3Nb2/3)O3–0.35PbTiO3 Random 50 [39]

CoFe2O4-Pb(Zr0.52Ti0.48)O3 Random 157 [40]

Bi3.4Ce0.6Ti3O12 Random 158 [41]

(Na0.82K0.18)0.5Bi0.5TiO3 Random 96 [42]

Bi3.15Nd0.85Ti3O12 Random 89 [43]

ZnO-V Random 121 [44]

(Na0.5Bi0.5)0.94TiO3-Ba0.06TiO3 Random 102 [42]

Their piezoelectric performances are comparable to d33 = 83-100 pm V-1 for PZT

nanofibers and thin films [38, 45] and considerably higher than those in Pb-free

piezoelectric materials such as (Na,K)NbO3 thin films with d33 = 40 pm V-1 [46] and

bismuth based layered ferroelectrics compositions in form of nanofibers and thin

films with d33 = 40-110 pm V-1 [43, 47]. Compared to PZT nanofibers, BaTiO3 thin

films, ZnO nanorods/pillars, and NaNbO3 nanowires exploited recently in high-

power piezoelectric generators [8, 48-51] (see Figure 3.8), our fibers seem to be capable

of significantly increased efficiency and output power in self-powered nanodevices

aj393
Highlight
aj393
Highlight
Page 85: Lead-free piezoelectric materials - Research Online

63

and energy harvesting systems. For example, according to Figure 3.8, d33 in BTZ-

BCT nanofibers exhibits more than 100% enhancement compared to the PZT

nanofibers. [38]

Figure 3.8 Comparison of piezoelectric coefficient d33 in nanostructured components used to

build piezoelectric generators with d33 in our BTZ-BCT films and fibers.

The reason for the high d33 values in our nanofibers is because the BTZ-BCT bulk

ceramics exhibit a very large piezoelectric effect compared to other piezoelectric

bulks. The large value (d33 = 1146 pm V-1) in the BTZ-BCT bulk ceramics is caused

by crystallization in the vicinity of the rhombohedral-cubic-tetragonal critical point

(critical triple point) which approaches the MPB region in the BTZ-BCT phase

diagram. The effects of the MPB and critical triple point on the piezoelectric

performance are discussed in details in Chapter 1. For this composition, the

polarization anisotropy vanishes, leading to a strong dependence of the electrical

polarization on elastic deformations, thus significantly increasing the

piezoelectricity. It is important to note that even a very small change in the

aj393
Highlight
aj393
Highlight
Page 86: Lead-free piezoelectric materials - Research Online

64

stoichiometry of the BTZ-BCT, as low as 5%, can result in drastic suppression of d33

in the final system of up to 40%.[20]

The equation (3.2) relates T (stress), E (electric field ), s13E (elastic compliance)

and d33 (inverse piezoelectric coefficient) to S (strain) in an unclamped piezoelectric

crystal.

𝑆3 = 𝑆13𝐸 (𝑇1 + 𝑇2) + 𝑆33𝐸 𝑇3 + 𝑑33𝐸3 (3.2) 40

It is well known that clamping to the substrate suppresses the strain and

consequently the measured converse piezoelectric coefficient in thin films and

nanofibers. The normal component of the electric field induces in-plane elastic stress,

σeffective. This affects normal deformations of the sample if a ferroelectric material is

tightly bonded with a rigid substrate. We assume the clamped nanofiber as the one-

dimensional object. Without applying stress on its free surfaces, T2=T3=0. The rigid

substrate can be considered as an isotropic object, therefore T1 =T in the nanofibers.

The T represents the effective stress.

Re-arranging the equation (3.2) with respect to the stress distributions in the

nanofibers, gives the effective piezoelectric coefficient 𝑑33,𝑓 for the one-dimensional

(equation (3.3)) objects.

𝑆3𝐸3

= 𝑑33,𝑓 = 𝑑33 + 𝑆13𝐸 σeffective𝐸3

(3.3)

Using d33 = 1140 pm V-1 [20] and s13E = 7.4 × 10-12 m2 N-1 [52], experimentally obtained

in bulk BTZ-BCT ceramics, we present the effective piezoelectric coefficient 𝑑33,𝑓

and the effective stress σeffective calculated from equation (3.2) in Table 3.2.

Page 87: Lead-free piezoelectric materials - Research Online

65

Table 3.2 Calculated effective stress induced by electric field in the nanofiber-substrate

interface.

Geometry

Maximum 𝑑33,𝑓

(pm V-1)

EcA

(kV cm-1 )

𝑑33,𝑓

(pm V-1)a

σeffective

( MPa)

Nanofibre 180 20 151 284 a d33 measured in coercive field

3.3.5 Ferroelectric nanodomain imaging by SCM

Since electrical domains in ferroelectric materials manifest themselves as a local

distribution of surface charges, domain switching can be detected by the ac-field

scanning capacitance microscopy (SCM) technique. SCM is more ideal for

visualizing the ferroelectric domains in nanofibers with curved surfaces than PFM.

The topography of a single BTZ-BCT nanofiber (atomic force microscope (AFM)

image) and its SCM image obtained simultaneously are displayed in Figure 3.9(a)

and (b).

The sign and magnitude of the C-V slope in the SCM image is recorded and

converted to the dC/dV signals when the AC bias is applied to a local capacitor

constructed by the nanofiber and conductive tip, and the substrate, which function as

top and bottom electrodes, respectively. In order to record the differential

capacitance (dC/dV) image, a PtIr-coated Si tip was used. dC/dV signals were

achieved at small 1 V amplitude ac-bias between the tip and conducting substrate at

frequency of 90 kHz. The lock-in 180° phase mode was selected to observe the

opposite polarity of domains in the form of a contrast in the dC/dV images of

nanofibers.

aj393
Highlight
Page 88: Lead-free piezoelectric materials - Research Online

66

Figure 3.9 (a) Topographical (AFM), (b) SCM images, obtained in contact mode, of a single

BTZ-BCT nanofiber annealed at 700°C on a Si/SiO2/Ti/Ir conductive substrate. The two

distinct types of regions in black and white represent opposite out-of-plane electric domains,

averaging 24 nm in size and (c) schematic illustration of the SCM measurement. The sign

and magnitude of the C-V slope in the SCM image is recorded while the AC bias is applied

to a local capacitor constructed by the nanofiber and conductive tip, and the substrate as top

and bottom electrodes, respectively. The ferroelectric domain polarities and their

configurations are developed in the dC/dV image in the SCM as a result of different trends in

the dC/dV with respect to the polarization states.

Figure 3.9(b) shows the sharp contrast in a SCM image recorded simultaneously with

the topographical image of the same nanofiber in Figure 3.9 (a). Regarding the

schematic illustration of the SCM, measuring capacitance versus voltage and dC/dV

versus voltage in ferroelectric materials (see Figure 3.9(c)), while sweeping the

voltage from positive to negative, creates two distinct regions in the C-V plot. The

two peaks that exist in the C-V curve in the ferroelectric materials represent the

aj393
Highlight
Page 89: Lead-free piezoelectric materials - Research Online

67

spontaneous polarization switching. The ferroelectric domain polarities and their

configurations are developed in the dC/dV images in the SCM as a result of different

trends in the dC/dV with respect to the polarization states. Since in SCM imaging, the

dC/dV signal recorded from the surface is converted to a voltage output, which is

different from dC/dV, thus, the SCM image is scaled in volts, and ferroelectric

domains with the opposite polarizations create the sharp contrast in this image. Two

distinct types of region, displayed as black and white areas, were obtained in zero

volt dc-bias, which we interpret as opposite out-of-plane electric domains. They have

an average size of about 24 nm.

3.3.6 Electric resistance measurements by SSRM

2D resistivity of the BTZ-BCT nanofiber/substrate was studied using SPM operating

in the SSRM mode. A highly boron doped Si tip as the conductive tip scans in

contact mode over the nanofiber and substrate surface. In order to collect the

resistivity map of the BTZ-BCT nanofiber, a -5 V DC bias voltage was applied

between the tip and the sample. The resulting current passed through the fibre and

conductive layer of the substrate.

The topographic image (Figure 3.10(a)) and current distribution image (Figure

3.10(b)) of the nanofiber were recorded in real-time. The contrast revealed in the

SSRM image was induced by the difference in conductivity/resistivity of the

conductive layer of the substrate and the BTZ-BCT nanofiber. The bright area of the

SSRM image represents the conducting surface of the substrate, and the dark area

shows the high resistance BTZ-BCT nanofiber, which allows the current toflow

through by applying a -5 V DC-bias voltage. The uniform SSRM image reveals an

even current distribution in the nanofiber.

aj393
Highlight
Page 90: Lead-free piezoelectric materials - Research Online

68

Figure 3.10 (a) Topographical (AFM), (b) SSRM images, obtained in contact mode, of a

single BTZ-BCT nanofiber annealed at 700°C on a Si/SiO2/Ti/Ir conductive substrate. The

two distinct types of regions in dark and bright colours represent the current distribution in

the conducting substrate and ferroelectric nanofiber. (c) Section analysis of the SSRM

image, and (d) reference transfer curve of the SSRM logarithmic current amplifier in

different biases.

This develops a highly homogeneous electric field distribution when a voltage is

employed on the fiber and impedes the non-uniform development of stresses/strains

in the nanofibers. These homogeneities in the current, electric field, and

consequently, the stress/strain distribution result in more accuracy in displacement,

higher output voltage, and longer lifetime of a device based on the BTZ-BCT

nanofibers. A section analysis of the SSRM image is shown in Figure 3.10(c).

According to the reference transfer curve of the SSRM logarithmic current amplifier

(Figure 3.10(d)) and the maximum output voltage extracted from the section

aj393
Highlight
Page 91: Lead-free piezoelectric materials - Research Online

69

analysis, the average DC resistance of the BTZ-BCT nanofiber is 1010 ohms. This

high resistance reduces the dissipation of the generated charges and leakage current,

thus improving the piezoelectric response and ferroelectric performance in the BTZ-

BCT nanofibers[15] .

Regarding the Ohm’s law

𝐼 = 𝑉𝑅 (3.3)

where I is current (A), V is potential difference (applied bias voltage (V)) and R is the

resistance (Ω), a 50 pA current is passed through the BTZ-BCT nanofiber by

applying difference potential difference of 5 V.

3.4 Conclusions

Crystalline phase evolution and structural analysis of the high performance

piezoelectric Ba (Ti0.80Zr0.20) O3- (Ba0.70Ca0.30)TiO3 (BTZ-BCT) nanofibers were

investigated. Coexistence of two ferroelectric phases, tetragonal and rhombohedral,

and crystallization of the fibers in the vicinity of the morphotropic phase boundary

(MPB) region has been demonstrated by employing different techniques, including

XRD, Raman spectroscopy, and TEM. We report very large piezoelectricity in

Ba(Ti0.80Zr0.20)O3- (Ba0.70Ca0.30)TiO3 (BTZ-BCT) lead-free nanofibers (d33 = 180 pm

V-1). These values are the highest among all the reported piezoelectric nanofibers,

more than two times higher than for PZT nanofibers. The SSRM results show a

uniform distribution of the resistance and very high resistance of 1010 ohms in the

nanofibers. Understanding the structural and electrical properties of the BTZ-BCT

nanofibers helps to improve the performance of different devices such as

nanogenerators, sensors, or voltage tuneable micro- and acoustic-wave devices.

aj393
Highlight
aj393
Highlight
Page 92: Lead-free piezoelectric materials - Research Online

70

3.5 References

[1] A. Bernal, A. Tselev, S. Kalinin, N. Bassiri-Gharb, Adv. Mater. 2012, 24,

1159.

[2] P. M. Rørvik, T. Grande, M.-A. Einarsrud, Adv. Mater. 2011, 23, 4007.

[3] B. J. Hansen, Y. Liu, R. Yang, Z. L. Wang, ACS Nano 2010, 4, 3647.

[4] Y. Bourquin, R. Wilson, Y. Zhang, J. Reboud, J. M. Cooper, Adv. Mater.

2011, 23, 1458.

[5] A. Jalalian, A. M. Grishin, Appl. Phys. Lett. 2012, 100, 012904.

[6] S. V. Kalinin, N. Setter, A. L. Kholkin, MRS Bull. 2009, 34, 634.

[7] Z. L. Wang, Adv. Funct. Mater. 2008, 18, 3553.

[8] X. Chen, S. Xu, N. Yao, Y. Shi, Nano Lett. 2010, 10, 2133.

[9] J. Chang, M. Dommer, C. Chang, L. Lin, Nano Energy 2012, 1, 356.

[10] P. Eyben, F. Seidel, T. Hantschel, A. Schulze, A. Lorenz, A. U. De Castro, D.

Van Gestel, J. John, J. Horzel, W. Vandervorst, physica status solidi (a) 2011, 208,

596.

[11] X. Ou, P. D. Kanungo, R. Kögler, P. Werner, U. Gösele, W. Skorupa, X.

Wang, Nano Lett. 2009, 10, 171.

[12] Y. Kim, C. Bae, K. Ryu, H. Ko, Y. K. Kim, S. Hong, H. Shin, Appl. Phys.

Lett. 2009, 94, 032907.

[13] J. Y. Son, S. H. Bang, J. H. Cho, Appl. Phys. Lett. 2003, 82, 3505.

[14] L. Y. Kraya, R. Kraya, J. Appl. Phys. 2012, 111, 013708.

[15] D. A. Scrymgeour, J. W. P. Hsu, Nano Lett. 2008, 8, 2204.

[16] K. Fujihara, A. Kumar, R. Jose, S. Ramakrishna, S. Uchida, Nanotechnology

2007, 18, 365709.

[17] A. Shui, L. Zeng, K. Uematsu, Scripta Mater. 2006, 55, 831.

Page 93: Lead-free piezoelectric materials - Research Online

71

[18] Y. Li, Y. Qian, H. Liao, Y. Ding, L. Yang, C. Xu, F. Li, G. Zhou, Science

1998, 281, 246.

[19] J. T. McCann, J. I. L. Chen, D. Li, Z.-G. Ye, Y. Xia, Chem. Phys. Lett. 2006,

424, 162.

[20] W. Liu, X. Ren, Phys. Rev. Lett. 2009, 103, 257602.

[21] U. D. Venkateswaran, V. M. Naik, R. Naik, Phys. Rev. B 1998, 58, 14256.

[22] Y. Shiratori, C. Pithan, J. Dornseiffer, R. Waser, J. Raman Spectrosc. 2007,

38, 1300.

[23] M. C. Chang, S.-C. Yu, J. Mater. Sci. Lett. 2000, 19, 1323.

[24] C. J. Xiao, C. Q. Jin, X. H. Wang, Mater. Chem. Phys. 2008, 111, 209.

[25] Y. Shiratori, C. Pithan, J. Dornseiffer, R. Waser, J. Raman Spectrosc. 2007,

38, 1288.

[26] E. Hamada, W.-S. Cho, K. Takayanagi, Philos. Mag. A 1998, 77, 1301.

[27] M. Fujimoto, J. Cryst. Growth 2002, 237–239, Part 1, 430.

[28] H. Oppolzer, H. Schmelz, J. Am. Ceram. Soc. 1983, 66, 444.

[29] W. Cao, L. E. Cross, Phys. Rev. B 1991, 44, 5.

[30] B.-K. Lee, S.-J. L. L. Kang, Acta Mater. 2001, 49, 1373.

[31] M. Daraktchiev, R. J. Harrison, E. H. Mountstevens, S. A. T. Redfern,

Materials Science and Engineering: A 2006, 442, 199.

[32] A. Gruverman, A. Kholkin, Rep. Prog. Phys. 2006, 69, 2443.

[33] L. Sagalowicz, F. Chu, P. D. Martin, D. Damjanovic, J. Appl. Phys. 2000, 88,

7258.

[34] C. T. Nelson, P. Gao, J. R. Jokisaari, C. Heikes, C. Adamo, A. Melville, S.-H.

Baek, C. M. Folkman, B. Winchester, Y. Gu, Y. Liu, K. Zhang, E. Wang, J. Li, L.-Q.

Chen, C.-B. Eom, D. G. Schlom, X. Pan, Science 2011, 334, 968.

Page 94: Lead-free piezoelectric materials - Research Online

72

[35] S. Brimley, M. S. Miller, M. J. Hagmann, J. Appl. Phys. 2011, 109, 094510.

[36] Z. Li, J. A. Bain, Z. Jian-Gang, L. Abelmann, T. Onoue, IEEE Transactions

on Magnetics 2004, 40, 2549.

[37] A. L. Kholkin, C. Wutchrich, D. V. Taylor, N. Setter, Rev. Sci. Instrum. 1996,

67, 1935.

[38] M. Fan, W. Hui, Z. Li, Z. Shen, H. Li, A. Jiang, Y. Chen, R. Liu,

Microelectron. Eng. 2012, 98, 371.

[39] S. Xu, G. Poirier, N. Yao, Nano Energy 2012, 1, 602.

[40] S. H. Xie, J. Y. Li, Y. Qiao, Y. Y. Liu, L. N. Lan, Y. C. Zhou, S. T. Tan,

Appl. Phys. Lett. 2008, 92, 062901.

[41] B. Jiang, M. Tang, J. Li, Y. Xiao, H. Tao, Y. Zhou, J. He, J. Electron. Mater.

2012, 41, 651.

[42] Y. Q. Chen, X. J. Zheng, X. Feng, S. H. Dai, D. Z. Zhang, Mater. Res. Bull.

2010, 45, 717.

[43] M. Liao, X. L. Zhong, J. B. Wang, S. H. Xie, Y. C. Zhou, Appl. Phys. Lett.

2010, 96, 012904.

[44] Y. Q. Chen, X. J. Zheng, X. Feng, Nanotechnology 2010, 21, 055708.

[45] D. V. Taylor, D. Damjanovic, Appl. Phys. Lett. 2000, 76, 1615.

[46] C. W. Ahn, S. Y. Lee, H. J. Lee, A. Ullah, J. S. Bae, E. D. Jeong, J. S. Choi,

B. H. Park, I. W. Kim, J. Phys. D: Appl. Phys. 2009, 42, 215304.

[47] S. Fujino, M. Murakami, V. Anbusathaiah, S. H. Lim, V. Nagarajan, C. J.

Fennie, M. Wuttig, L. Salamanca-Riba, I. Takeuchi, Appl. Phys. Lett. 2008, 92,

202904.

[48] K.-I. Park, M. Lee, Y. Liu, S. Moon, G.-T. Hwang, G. Zhu, J. E. Kim, S. O.

Kim, D. K. Kim, Z. L. Wang, K. J. Lee, Adv. Mater. 2012, 24, 2999.

Page 95: Lead-free piezoelectric materials - Research Online

73

[49] K.-I. Park, S. Xu, Y. Liu, G.-T. Hwang, S.-J. L. Kang, Z. L. Wang, K. J. Lee,

Nano Lett. 2010, 10, 4939.

[50] Y. Hu, Y. Zhang, C. Xu, G. Zhu, Z. L. Wang, Nano Lett. 2010, 10, 5025.

[51] J. H. Jung, M. Lee, J.-I. Hong, Y. Ding, C.-Y. Chen, L.-J. Chou, Z. L. Wang,

ACS Nano 2011, 5, 10041.

[52] D. Xue, Y. Zhou, H. Bao, C. Zhou, J. Gao, X. Ren, J. Appl. Phys. 2011, 109,

054110.

Page 96: Lead-free piezoelectric materials - Research Online

74

Chapter 4. BTZ-BCT Thin Films

In this Chapter, the synthesis and characterization of Ba(Ti0.80Zr0.20)O3-

(Ba0.70Ca0.30)TiO3 (BTZ-BCT) thin films will be presented. Crystalline phase content

and topography of the films will be examined. Ferroelectric domain configuration,

and piezoelectric and ferroelectric properties of the thin films will be investigated

using different techniques, including piezoelectric force microscopy (PFM) and

polarization versus electric field hysteresis loops.

4.1 Introduction

Integration of high performance piezoelectrics in piezoelectric devices and micro-

/nano-electromechanical systems (MEMS, NEMS) is a viable approach to enhance

their efficiency. Low-dimensional piezoelectric and ferroelectric materials have been

exploited widely in nanogenerators, sensors, transducers, MEMS devices, and other

applications, such as microwave varactors and ferroelectric field effect transistors.[1-6]

For instance, high performance piezoelectric nanostructures can enhance the output

power of piezoelectric generators that convert the kinetic energy of vibrations,

displacements, or applied force to electricity.[7-11]

In this study, observations on the high piezoelectric response achieved in BTZ-BCT

thin films are reported, along with visualization of ferroelectric nanodomains with

high spatial resolution using PFM. The influences of lateral size, geometry, and the

clamping effect on the piezoelectric performance were also investigated for thin

films and compared with BTZ-BCT nanofibers.

aj393
Highlight
aj393
Highlight
Page 97: Lead-free piezoelectric materials - Research Online

75

4.2 Experimental procedure

Ba(Ti0.80Zr0.20)O3- (Ba0.70Ca0.30)TiO3 thin film was fabricated using the spin-coating

technique. In order to prepare the BTZ-BCT precursor solution, barium acetate,

calcium acetate monohydrate, titanium butoxide, and zirconium (IV) propoxide were

used as starting materials. Barium acetate and calcium acetate monohydrate were

dissolved in glacial acetic acid at 80°C for 1 hour and cooled down to room

temperature. Titanium butoxide and zirconium propoxide were chelated by acetyl

acetone and added to the Ba and Ca solution. 2-methoxyethanol was used to adjust

the concentration of the BTZ-BCT precursor solution to 0.2 M. Spin-coating at 3000

rpm for 30 seconds was carried out to prepare the BTZ-BCT thin film on Si and

Si/SiO2/Ti/Ir substrates. The spin-coated thin films were dried and annealed at 100°C

and 600°C for 10 minutes, respectively, and the above procedure was repeated for 4

cycles until the films reached a thickness of about 200 nm, followed by calcination at

700°C for 1 hour in air.

Field-emission scanning electron microscopy (FE-SEM, JEOL JSM 7500FA) was

used to examine the surface morphology and microstructure of the cross-section of

the thin film with a 3.8 mm working distance under 1 kV acceleration voltage. A

Raman spectrometer (HORIBA JobinYvon, HR800 spectrometer) was employed to

study the state of crystallization and lattice vibration modes. Un-polarized spectra

were collected at room temperature using 632.8 nm light pumping from a He-Ne

laser and a charge coupled device (CCD) detector in this experiment.

Piezoresponse force microscopy (PFM, Asylum Research, MFP-3D) and scanning

capacitance microscopy (Digital Instrument, Dimension 3100) were used for our

experiments. Our PFM measurements were performed at a low frequency of 10 Hz

aj393
Highlight
aj393
Highlight
Page 98: Lead-free piezoelectric materials - Research Online

76

and in electric field, ranging from -300 kV/cm to +300 kV/cm. The bottom electrode

was an Ir layer on Si wafer and the top electrode was a PtIr-coated Si tip with

deflection sensitivity of 73.6 nm V-1 for our experiments, respectively.

4.3 Results and discussion

4.3.1 Morphology studies using FE-SEM and AFM

Figure 4.1 (a) FE-SEM image of BTZ-BCT thin film deposited on the Si/SiO2/Ti/Ir

substrate and annealed at 700°C for 1 hour in air. Inset: magnified view of the thin film

surface. (b) FE-SEM image of cross-section of BTZ-BCT thin film about 200 nm in

thickness with an average particle size of 33 nm. (c) 3D AFM image of topography of BTZ-

BCT thin film annealed at 700°C for 1 hour, showing 2 nm rms surface roughness.

(a) (b)

(c)

aj393
Highlight
aj393
Highlight
Page 99: Lead-free piezoelectric materials - Research Online

77

Field emission scanning electron microscope (FE-SEM) images from the surface and

cross-section of the BTZ-BCT thin film are displayed in Figures 4.1(a) and (b). The

image of the 200 nm thick BTZ-BCT film in Figure 4.1(b) reveals randomly oriented

particles. The images were collected at 3.8 mm working distance under 1 kV

acceleration voltage.

A typical three-dimensional (3D) atomic force microscope (AFM) image of a 400 ×

400 nm2 surface area of a BTZ-BCT thin film spin-coated on Si/SiO2/Ti/Ir substrate

and heat-treated at 700°C for 1 hour is shown in Figure 4.1(c).

The films have a smooth surface with a surface root mean square (rms) roughness

equal to 2.1 nm, with an average particle size of 33 nm.

4.3.2 Phase content investigation using XRD and Raman spectroscopy

The X-ray diffraction pattern of the thin film annealed at 700°C is presented in

Figure 4.2(a). The Bragg reflections demonstrate no preferred directional growth in

the film, which is similar to the case of the BTZ-BCT nanofibers that were annealed

under the same conditions using the same precursor solution.

Local atomic environments in the BTZ-BCT films were examined by Raman

spectroscopy at room temperature. A typical spectrum is displayed in Figure 4.2(b).

All features characteristic of the perovskite structure of bulk BaTiO3 were verified in

the Raman spectra of the thin film and are in good agreement with the results for the

other BTZ-BCT counterparts, which were analyzed in previous chapters in detail [12,

13].

aj393
Highlight
aj393
Highlight
Page 100: Lead-free piezoelectric materials - Research Online

78

Figure 4.2 (a) X-ray diffraction pattern of the BTZ-BCT thin film annealed at 700°C on the

Si substrate. (b) Raman spectrum of BTZ-BCT 200 nm thick BTZ-BCT film spin-coated on

Si/SiO2/Ti/Ir substrate annealed at 700°C for 1 hour in air. The spectrum of the substrate

alone is included for reference. All features are characteristic of the polar structure of

BaTiO3 and indicate successful evolution of the perovskite structure in all samples. The

existence of a peak at 164 cm-1 [A1(TO)] is only observed in the rhombohedral symmetry,

while the other peaks, which appear in both tetragonal and rhombohedral symmetries,

indicate the crystallization in morphotropic phase boundary (MPB) region.

The transverse optical mode [A1(TO)] located at 164 cm-1 is recorded in higher

intensity compared to the nanofibers. This peak has been observed only in the

rhombohedral phase of nanocrystalline barium titanate [14], while the other

vibrational modes exist in both tetragonal and rhombohedral phases. The coexistence

of these two ferroelectric phases in the Raman spectra confirms the crystallization of

the BTZ-BCT thin films in the MPB region.

4.3.3 Piezoelectric and ferroelectric properties investigation by PFM

Figure 4.3(a) and (b) presents 3D- PFM patterns of the surface of a 200 nm thick

BTZ-BCT film. Simultaneous recording of 3D profiles of the PFM phase and

amplitude signals enables a visualization of ferroelectric domains. These 3D images

(a) (b)

Page 101: Lead-free piezoelectric materials - Research Online

79

were acquired using a lock-in amplification technique with a PtIr-coated silicon

cantilever in contact mode, which was ac-biased with a voltage of 1.0 V.

Figure 4.3 3D PFM phase (a) and amplitude (b) images of the surface of 200 nm thick BTZ-

BCT film. Ferroelectric domain patterns in the phase image (a) display the opposite polarity

of ferroelectric domains 20 to 40 nm in size. These domains are revealed by 180° phase

difference contrast. The amplitude image (b) reproduces the domain shape. Sharp contrast

around domains visualizes domain boundaries. In (c), phase and amplitude profiles recorded

along the marked lines in (a) and (b) are overlaid on an AFM topographical image. All three

images were collected simultaneously under 1.0 V ac modulation voltage.

Ferroelectric domains have a lateral size of 20 to 40 nm. 180° PFM phase contrast in

Figure 4.3(a) distinctly shows that ferroelectric domains have antiparallel out-of-

plane polarization. The PFM amplitude pattern in Figure 4.3(b) reproduces the shape

(a)

(b)

(c)

Page 102: Lead-free piezoelectric materials - Research Online

80

of the domains. In Figure 4.3(c), phase and amplitude cross-sectional profiles

recorded along the marked line are overlaid on the topographical image taken from

Figure 4.4 (a) Local PFM phase hysteresis loops and (b) piezoelectric butterfly loops of the

BTZ-BCT thin film obtained by switching the bias voltage from -5 to +5 volts (±300 kV cm-

1 electric field). 180° domain switching in the phase hysteresis loop shows the switching

phenomenon of spontaneous electrical dipoles built up in the BTZ-BCT thin film. (c) d33

hysteresis loop calculated by using the converse piezoelectric equation, Δz = d33V, which

indicates the piezoelectric coefficient d33 = 141 pmV-1 in the BTZ-BCT thin film. (d)

Schematic illustration of lateral size and contact area in the thin film and nanofiber.

Figure 4.1(c). Sharp 180° PFM phase contrast occurs when the conducting cantilever

crosses narrow domain boundaries.

PFM was employed to quantify the piezoelectric coefficient in the BTZ-BCT films

under the same operational parameters used for the nanofibers. Figure 4.4(a) and (b)

aj393
Highlight
Page 103: Lead-free piezoelectric materials - Research Online

81

demonstrates the ferroelectric domain switching and displacement of the thin film

respectively developed via the inverse piezoelectric effect. It should be noted that the

PFM signals were calibrated using a force-distance curve before each measurement.

The calculated piezoelectric coefficient in the BTZ-BCT thin films (Figure 4.4(c))

shows smaller value compare to the nanofibers. We explain this by the much smaller

lateral size of the fiber-substrate interface and the fiber-to-substrate contact area, as

shown in Figure 4.4(d). Reduction of the residual ferroelectric-substrate interface

strain increases the piezoelectric response in the nanofibers.

Table 4.1 Comparison of the piezoelectric coefficient d33 in some Pb-based and Pb-

free piezoelectric thin films.

Materials Orientation d33

(pm V-1) Ref.

BTZ-BCT Random 141 Current

study

BaTiO3 (100) 30 [15]

BaTiO3 Single crystal 80-100 [16]

Pb0.76Ca0.24TiO3 Random 70 [17]

Bi0.83Sm0.17FeO3 (002) 110 [18]

BiFeO3 (001) 70 [19]

0.34BiScO3-0.66PbTiO3 (001) 130 [20]

0.948(K0.5Na0.5)NbO3–0.052LiSbO3 Random 50 [21]

(K0.44,Na0.52,Li0.04)(Nb0.84,Ta0.1,Sb0.06)O3 (001) 53 [22]

Pb(Zr0.6Ti0.4)O3

(100) 100 [23] (111) 63

Random 77

0.7Pb(Mg1/3Nb2/3)O3–0.3PbTiO3

(111)

79 [24] 0.9Pb(Mg1/3Nb2/3)O3–0.1PbTiO3 64

Pb(Zr0.6Ti0.4)O3 85

(K0.5Na0.5)NbO3 with 20 mol% extra K and Na. Random 40 [25]

aj393
Highlight
Page 104: Lead-free piezoelectric materials - Research Online

82

Similar to the BTZ-BCT nanofibers, the observed piezoelectric coefficients in our

thin film is exceeding those of lead zirconate titanate (PZT). Table 4.1 compares the

piezoelectric coefficient d33 for lead-containing and lead-free piezoelectric thin films

with textured and random crystal structures. According to Table 4.1, d33 in BTZ-BCT

thin films exhibits more than 100% enhancement compared to the PZT thin films

with a random orientation structure [25].

Since the same precursor solution as for the nanofibers was used for the thin films,

and the same crystal structure and phase content were developed as for the

nanofibers, crystallization in the vicinity of the MPB suppresses the free energy

barrier against the polarization rotation from rhombohedral to tetragonal and vice

versa. Reducing the energy barrier and eliminating the polarization anisotropy

enhanced the piezoelectric performance in the BTZ-BCT constrained structures,

including the thin films.

Table 4.2 Calculated effective stress induced by electric field in the thin film −

substrate interface.

Geometry Maximum𝑑33,𝑓

(pm V-1)

EcA

(kV cm-1 )

𝑑33,𝑓

(pm V-1)a

σeffective

( MPa)

Thin film 141 45 122 302.7 a d33 measured in coercive field

Suppression of the piezoelectric coefficient in the thin films compared to the BTZ-

BCT ceramics occurs because of the same reason as for the other constrained

structures such as nanofibers, as discussed in Chapter 3. We assume the clamped thin

films as the two-dimensional objects. Without applying stress on their free surfaces,

aj393
Highlight
Page 105: Lead-free piezoelectric materials - Research Online

83

T3=0 for the the thin films, therefore T1=T2=T. The T represents the effective stress.

The equation (4.1) gives the effective piezoelectric coefficient 𝑑33,𝑓 for the one-

dimensional objects.

𝑆3𝐸3

= 𝑑33,𝑓 = 𝑑33 + 2𝑆13𝐸 σeffective𝐸3

(4.1)

The calculated effective stress (σeffective) in the thin film ─ substrate interface is listed

in Table 4.2.

Figure 4.5 Polarization versus electric field (P-E) loop obtained at 1 kHz of the BTZ-BCT

thin film with about 200 nm thickness deposited on the Si/SiO2/Ti/Ir substrate.

The polarization versus electric field loop is displayed in Figure 4.5. The

ferroelectric hysteresis loop was recorded at 1 kHz. The ferroelectric domain

switching was saturated under an average electric field of Ec = 45.8 kV cm-1 and

achieved maximum polarization Ps = 6.9 µC cm-1, and remanent polarization Pr = 2.4

aj393
Highlight
Page 106: Lead-free piezoelectric materials - Research Online

84

µC cm-1 in the BTZ-BCT thin film with about 200nm thickness deposited on the

Si/SiO2/Ti/Ir substrate.

4.4 Conclusions

In conclusion, we report very large piezoelectricity in Ba(Ti0.80Zr0.20)O3-

(Ba0.70Ca0.30)TiO3 (BTZ-BCT) lead-free thin films with piezoelectric coefficient d33

= 141 pm V-1 . This value is the highest among all the reported piezoelectric thin

films, about two times higher than for PZT films with randomly oriented structure.

4.5 References

[1] Z. L. Wang, J. Song, Science 2006, 312, 242.

[2] M. Bhaskaran, S. Sriram, S. Ruffell, A. Mitchell, Adv. Funct. Mater. 2011,

21, 2251.

[3] S. C. Masmanidis, R. B. Karabalin, I. De Vlaminck, G. Borghs, M. R.

Freeman, M. L. Roukes, Science 2007, 317, 780.

[4] P. Muralt, J. Micromech. Microeng. 2000, 10, 136.

[5] C.-R. Cho, J.-H. Koh, A. Grishin, S. Abadei, S. Gevorgian, Appl. Phys. Lett.

2000, 76, 1761.

[6] S.-M. Koo, S. Khartsev, C.-M. Zetterling, A. Grishin, M. Ostling, Appl. Phys.

Lett. 2003, 83, 3975.

[7] K.-I. Park, M. Lee, Y. Liu, S. Moon, G.-T. Hwang, G. Zhu, J. E. Kim, S. O.

Kim, D. K. Kim, Z. L. Wang, K. J. Lee, Adv. Mater. 2012, 24, 2999.

[8] X. Chen, S. Xu, N. Yao, Y. Shi, Nano Lett. 2010, 10, 2133.

[9] Y. Hu, Y. Zhang, C. Xu, G. Zhu, Z. L. Wang, Nano Lett. 2010, 10, 5025.

aj393
Highlight
aj393
Highlight
aj393
Highlight
Page 107: Lead-free piezoelectric materials - Research Online

85

[10] H. D. Espinosa, R. A. Bernal, M. Minary-Jolandan, Adv. Mater. 2012, 24,

4656.

[11] B. J. Hansen, Y. Liu, R. Yang, Z. L. Wang, ACS Nano 2010, 4, 3647.

[12] U. D. Venkateswaran, V. M. Naik, R. Naik, Phys. Rev. B 1998, 58, 14256.

[13] Y. Shiratori, C. Pithan, J. Dornseiffer, R. Waser, J. Raman Spectrosc. 2007,

38, 1300.

[14] C. J. Xiao, C. Q. Jin, X. H. Wang, Mater. Chem. Phys. 2008, 111, 209.

[15] Y. Guo, K. Suzuki, K. Nishizawa, T. Miki, K. Kato, J. Cryst. Growth 2005,

284, 190.

[16] Y.-B. Park, J. L. Ruglovsky, H. A. Atwater, Appl. Phys. Lett. 2004, 85, 455.

[17] A. L. Kholkin, M. L. Calzada, P. Ramos, J. Mendiola, N. Setter, Appl. Phys.

Lett. 1996, 69, 3602.

[18] S. Fujino, M. Murakami, V. Anbusathaiah, S. H. Lim, V. Nagarajan, C. J.

Fennie, M. Wuttig, L. Salamanca-Riba, I. Takeuchi, Appl. Phys. Lett. 2008, 92,

202904.

[19] J. Wang, J. B. Neaton, H. Zheng, V. Nagarajan, S. B. Ogale, B. Liu, D.

Viehland, V. Vaithyanathan, D. G. Schlom, U. V. Waghmare, N. A. Spaldin, K. M.

Rabe, M. Wuttig, R. Ramesh, Science 2003, 299, 1719.

[20] H. Wen, X. Wang, C. Zhong, L. Shu, L. Li, Appl. Phys. Lett. 2007, 90,

202902.

[21] J. Ryu, J.-J. Choi, B.-D. Hahn, D.-S. Park, W.-H. Yoon, Appl. Phys. Lett.

2008, 92, 012905.

[22] M. Abazari, T. Choi, S. W. Cheong, A. Safari, J. Phys. D: Appl. Phys. 2010,

43, 025405.

[23] D. V. Taylor, D. Damjanovic, Appl. Phys. Lett. 2000, 76, 1615.

Page 108: Lead-free piezoelectric materials - Research Online

86

[24] R. Herdier, M. Detalle, D. Jenkins, C. Soyer, D. Remiens, Sensor. Actuat. A:

Phys. 2008, 148, 122.

[25] C. W. Ahn, S. Y. Lee, H. J. Lee, A. Ullah, J. S. Bae, E. D. Jeong, J. S. Choi,

B. H. Park, I. W. Kim, J. Phys. D: Appl. Phys. 2009, 42, 215304.

Page 109: Lead-free piezoelectric materials - Research Online

87

Chapter 5. Biocompatible Piezoelectric Nanofibers

In this chapter, piezoelectric properties and ferroelectric nano-domains developed in

the NKN nanofibers are investigated.

Biocompatible and piezoelectric (Na,K)NbO3 (NKN) nanofibers are synthesized by

sol-gel assisted electrospinning technique. X-ray diffraction pattern of the nanofibers

reveals a pure and single phase of polar structure after annealing at 700°C.

Transmission electron microscope images and electron diffraction pattern show the

growth of NKN single crystals in the form of nanofibers. The ferroelectric domain

switching and piezoelectric response of the nanofibers are investigated using

piezoelectric force microscopy (PFM) technique. A significantly higher piezoelectric

response is achieved in NKN nanofibers compare to its thin films. Owing

permanently charged regions in the NKN nanofibers known as ferroelectric domains

and generating the electrical signals via piezoelectric effect in them provide a new

opportunity for construction of a smart biocompatible scaffold which can be used for

repair, engineering and regeneration of damaged tissues.

5.1 Introduction

Electrically active materials have been approved to accelerate tissue growth and

improve the adaptation of the surrounding tissue of an implant or damaged tissues [1-

4]. Surface charges could be generated either in ionic biomaterials (e.g.

hydroxyapatite (HA)) due to the cation and anion separation in opposite orientations

under an electric field or as a result of piezoelectricity in piezoelectric materials such

as quartz that surface charges are created on the crystal surfaces under appropriate

Page 110: Lead-free piezoelectric materials - Research Online

88

mechanical stress. Possessing permanently charged regions in ferroelectric materials

(a subgroup of the piezoelectrics) known as ferroelectric domains and generating the

electrical signals via piezoelectric effect, have attracted intensive interest to exploit

them in biological applications [3, 5]. Electrical signals generated on the surface of

these materials stimulate nerves resulted in an enhanced blood flow in tissues[1]. In

terms of bone in growth, surface charges created on the implant surface, enhance the

protein adsorption onto the implant surface which leads to improve the

osteoconduction on the implant surfaces compare to neutral surfaces[6]. Beloti et al

demonstrated that the attachment of the human osteoblatic cells to the implants was

noticeably promoted in piezoelectric implants consists of P(VDF-TrFE)/BaTiO3

composite compare to the non-piezoelectric compounds [3]. Piezoelectric materials

can imitate the bone reaction [7, 8] when a force is applied and produce electrical

charges. There have been several reports that confirm piezoelectric phenomenon

improves the bone tissue growth and biological response [9-11].

Hydroxapatite Ca10(PO4)6(OH)2 has a composition similar to the natural bone (

Calcium phosphate) and excellent biocompatibility, however lack of permanent

charged regions and naturally neutral surfaces states further demand for new

compositions and structures[12]. There have been several efforts to merge

piezoelectricity with biocompatibility in composite and complex structures such as

HA- BaTiO3, polyvinylideneflouride (PVDF) - HA and PVDF-(Pb(Zr0.53Ti0.47)O3) [5,

13, 14]. In most cases addition of piezoelectric materials to the bioceramics only

increased the dielectric constant and did not show an improvement in the

piezoelectric performance due to the non-continues structure in piezoelectric segment

or limited their application due to the toxicity of additional part.

Page 111: Lead-free piezoelectric materials - Research Online

89

Nilsson et al Patented piezoelectric (Na, K) NbO3 ceramics (NKN) as a

biocompatible piezoelectric material for medical implants [15]. According to

toxicology examination, NKN mentioned excellent biocompatibility and acceptable

durability in form of ceramics and thin films. This invention inspired us to design

and fabricate a 3D structure constructed with NKN nanofibers which can be

employed as a scaffold for stimulation and encouraging the tissue growth.

In this paper, piezoelectric properties and ferroelectric nano-domains developed in

the NKN nanofibers are investigated.

5.2 Experimental procedure (Na, K)NbO3 precursor solution was prepared via sol-gel method. Potassium acetate

and sodium acetate were mixed in 2-methoxyethanol at room temperature and stirred

for 1 hour followed by adding a niobium ethoxide solution dissolved in acetyl

acetone. Polyvinylpyrrolidone (PVP, 0.035 g ml-1) as a binder was added to the NKN

sol to prepare the solution for electrospinning. The solution was transferred to a

plastic syringe with a metallic needle. Electrospinning was carried at 1.8 kVcm-1

electric field strength between the metallic needle and aluminium foil collector

located 8cm below the needle. A non-woven and bead-free nanofibers mat was

collected from the surface of the collector after drying at 100 °C in nitrogen

atmosphere for 12 hours followed by annealing at 600, 700 and 800°C for 1 hour in

air. Heating/cooling rate was 5 °C min-1.

Crystal structure evolution of the nanofibers calcined at different temperatures was

studied by X-ray diffraction (GBC MMA powder diffractometer, CuKα radiation, 40

kV, 25 mA). Raman spectra were collected (HORIBA jobin yvon, HR800

spectrometer) at room temperature in air by pumping samples with 632.8 nm He-Ne

Page 112: Lead-free piezoelectric materials - Research Online

90

laser light. Field emission scanning electron microscope (FE-SEM, JEOL JSM

7500FA) was used to collect images of nanofibers. Transmission electron

microscopy (TEM) was performed using a JEOL 2011 200 KeV analytical

instrument. Samples were prepared by dispersion onto “Quantifoil” holey carbon

support film with examination of sample regions located over holes in the support

film.

Piezoresponse force microscopy (PFM, Asylum Research, MFP-3D) for

piezoelectric properties measurements was used. A PtIr-coated Si tip for the PFM

was employed. Bottom electrode is Ir layer on Si wafer and the top electrode is the

conducting tip.

5.3 Results and discussions

5.3.1 Morphology and Crystalline phase evolution

Figure 5. 1 FE-SEM images of (Na,K)NbO3 nanofibers mat (a) annealed at 600 °C,

Page 113: Lead-free piezoelectric materials - Research Online

91

(b) 700 °C and (c) 800 °C in air. (d) Thermogravimetric analysis of NKN nanofibers

from room temperature up to 800 °C in air. 62 wt% of the NKN precursor evaporates

or burns out during the annealing process at different stages.

Field-emission scanning electron microscopy images of the NKN nanofibers

annealed at different temperatures ranging from 600°C to 800 °C for 1 hour in air are

displayed in Figures 5.1(a) to (c). It can be observed that crystallization of the

nanofibers results in a significant shrinkage in the nanofibers.

The average diameter of nanofibers is about 250 nm for the sample annealed at 700

°C and reduced to about 200 nm by annealing at 800 °C with tens of microns in

length. It is approved that the mass transport along grain boundaries to the neck

between adjacent grains develops corresponding shrinkage at the sintering procedure

[16, 17]. Thermogravimetric analysis (TGA) of the NKN nanofibers (Figure 5.1(d))

demonstrates a 62 wt% weight loss in the as-spun nanofibers during the heat

treatment from room temperature up to 800°C in air.

Figure 5.2 Energy dispersive x-ray spectroscopy (EDS) spectrum of the NKN

nanofibers annealed at 800°C. Inset: quantitative analysis of the EDS that approves

Page 114: Lead-free piezoelectric materials - Research Online

92

the existence of the constructive elements in the (Na, K) NbO3 nanofibers which

follow the stoichiometric ratios with reasonable accuracy.

The volatile organic compounds such as solvents in the NKN precursor nanofibers

evaporate and develop a 8 wt% weight loss when temperature is raised up to 150°C

and 15 wt% weight loss from 150 to 250°C can be due to the degradation of the PVP

in air. A drastic mass reduction at 300°C is attributed to the burning out of the binder

and other residual organic compounds which continues up to 600°C. In order to keep

the shape of nanofibers during the annealing, a heating regime of 5°C min-1 is under

taken.

Figure 5.3 XRD patterns of the (Na, K) NbO3 nanofibers annealed at 700 and 800°C

reveal the crystallization of the nanofibers in a monoclinic structure. All peaks are

indexed according to the PDF-card number 77-0038 corresponds to the (Na0.35K0.65)

NbO3.

Page 115: Lead-free piezoelectric materials - Research Online

93

Energy dispersive x-ray spectroscopy (EDS) spectrum of the NKN nanofibers

annealed at 800°C is displayed in Figure 5.2. Inset table demonstrates the

quantitative analysis of the EDS that approves the existence of the constructive

elements in the (Na, K) NbO3 nanofibers which follow the stoichiometric ratios with

reasonable accuracy.

Crystalline structure and phase evolution were examined by X-ray diffractometery

on the nanofibers annealed at different temperatures at 700°C and 800°C (Figures

5.3(a) and (b)). After annealing nanofibers at 700 °C and 800 °C, the existence of

NKN perovskite crystallization becomes evident in the XRD patterns. Bragg

reflections in Figure 3 are indexed according to the

PDF-card number 77-0038 for monoclinic (Na0.35K0.65) NbO3 and No detectable

impurity phases were observed.

Figure 5.4 Experimental Raman spectrum of the NKN-nanofibers annealed at 800

°C for 1 hour in air.

Page 116: Lead-free piezoelectric materials - Research Online

94

Unpolarized backscattered Raman spectra of NKN nanofibers were recorded at room

temperature (Figure 5.4). According to Kakimoto [18] , Raman scatterings in a wide

range from 200 to 900 cm-1 are attributed to internal vibrations of NbO6 octahedral in

NKN: 1A1g(ν1), 1Eg(ν2), 1F1u(ν3) stretching and 1F1u(ν4), F2g(ν5) bending modes.

Stretching vibration of the O―Nb―O bonds generate strong peaks at 562 and 616

cm-1. The bands located at 270 and438 cm-1 wave numbers are attributed to the

bending vibration of the O―Nb―O bonds and Raman signals at higher frequencies

including 804 and 860 cm-1 are assigned to the fundamental vibration modes ν1 and

ν2 generated by Nb=O short bonds in the NbO6 octahedral [19-21] .

5.3.2 Nanostructure studies by TEM

The TEM image displayed in Figure 5.5 (a) represents a NKN nanofiber constructed

by nano-single crystals grown in different lengths and similar lateral size of about

200nm. The selected area electron diffraction (SAED) pattern collected from the

same fiber (shown in the inset) reveals the single crystal structure for each segment.

Enlarged interface between two single crystals is mentioned in Figure 5.5 (b).

Page 117: Lead-free piezoelectric materials - Research Online

95

Figure 5.5 TEM results obtained from sample heat treated at 800° C; (a) low

magnification image of a nanofiber containing single crystals. (b) Selected area

electron diffraction pattern which reveals the single crystal. (c) High resolution TEM

image of the NKN nanofiber. (d) Schematic of a monoclinic structure.

5.3.4 Piezoelectric and ferroelectric properties

These 3D PFM images were acquired using a lock-in amplification technique with a

PtIr-coated silicon cantilever in contact mode, which was ac-biased with a voltage of

1.2 V. Figures 5.6(a) and (b) presents 3D PFM image of the surface of a NKN

nanofiber annealed on a conductive substrate Si/SiO2/Ti/Ir. The PFM phase and

amplitude signals recorded simultaneously with the topography image yield

ferroelectric domain configuration and piezoelectric response of the nanofiber.

The ferroelectric domains manifest themselves as permanent charged regions which

in Figure 5.6(b) they are configured in form of anti-parallel out-of plane polarizations

Page 118: Lead-free piezoelectric materials - Research Online

96

with 180° phase difference. The PFM amplitude pattern in figure 5.6(c) represents

the displacement of the NKN nanofiber located below the tip as a result of the

inverse piezoelectric phenomenon.

The PFM images confirm the existence of surface charges developed by the

ferroelectric domains ability of the surface charge creation via direct piezoelectric

effect which electric charge is generated by applying force or deformation on the

NKN nanofibers. It is approved that being electrically active stimulates the tissue

growth and improves the osteoconduction [1, 2, 5, 6, 22-25]. Several mechanisms have

been proposed to describe how surface charges affect the tissue growth such as

enhancement of the surface wettability [22], protein adsorption since it is

negatively charged [26], improving the ion exchanges [27] and stimulating the nerve

regeneration [1].

Nakamura et al proposed a mechanism that negative and positive surface charges

promote the colony formation of the essential elements for the tissue growth and

osteoconduction (Figure 5.6.(d)) [6]. Fibrin is an essential constructive of the blood

and evolves ―COOH groups are attracted by the positive surface charges and

negative charged regions attract Ca2+. This mechanism was approved in other reports

published later on [24, 28].

Figure 5.7 displays the ferroelectric domain switching and piezoelectric behavior in

the NKN nanofibers. A bias voltage, ranging from -5 V to +5 V, was employed. The

PFM phase hysteresis loops in Figure 7 (a) show the switching of the ferroelectric

domains below the tip and 180° rotation of the electrical polarization of domains in

the NKN nanofiber. The switching ability of the ferroelectric domains in the NKN

nanofiber lets us to create a favorite surface charge arrangement either positive or

Page 119: Lead-free piezoelectric materials - Research Online

97

negative and employ engineered surfaces with different electrical activates

compatible to their biologic environment.

Figure 5.7(b) displays the deformation of the fiber under electric field developed

between a conductive tip ( Si tip coated by the PtIr conducting coating) and substrate

( Silicon wafer coated by Ir) which occurs as a result of the inverse piezoelectric

effect and existence of the ferroelectric domains in the NKN nanofibers.

The piezoelectric coefficient was calculated using to the definition of the inverse

piezoelectric effect, Δz = d33V [29] and plotted in Figure 5.7(c) at the same range of

the applied voltage used for the displacement vs. voltage in Figure 5.7(b). The PFM

signal was calibrated by using the force-distance curve for the cantilever used for the

measurements.

The average of the piezoelectric coefficient d33 reached as high 58 pmV-1 in the NKN

nanofiber. This value was obtained from the slope of the linear part of the Δz vs.

voltage curve at very low voltage. The piezoelectric coefficient obtained in the NKN

nanofibers demonstrates stronger response compare to the NKN thin films (d33=40

pm V-1) reported by Ahn et al [30].

Enhancement of the piezoelectric response in the nanofibers can be attributed to the

effect of geometry and suppressing of the clamping effect induced by the contact

area between the sample and substrate. It is evident that the attachment of the human

osteoblatic cells to the implants was noticeably promoted in piezoelectric implants

consists of P(VDF-TrFE)/BaTiO3 composite compare to the non-piezoelectric

compounds [3]. Piezoelectric materials can imitate the bone reaction [7, 8] when a force

is applied and produce electrical charges on their surfaces. The electric signals

generated by the piezoelectric phenomenon can stimulate the surrounded

Page 120: Lead-free piezoelectric materials - Research Online

98

Figure 5.6 3D-PFM (a) topography, (b) phase and (c) amplitude images of the NKN

nanofiber. Ferroelectric domain pattern in the phase image (b) displays the opposite

polarity of ferroelectric domains. These domains are revealed by 180° phase

difference contrast. The amplitude image (c) represents the piezoelectric response of

the NKN nanofiber generated by the movement of the nanofiber under alternate

electric field developed between the tip and conducting substrate. The PFM-phase

and amplitude images are overlaid on the topography image and collected

simultaneously under 1.2 V ac modulation voltage. (d) Displays the interaction of

ferroelectric domains with essential elements for the tissue growth and

osteoconduction including fibrins and calcium ions respectively.

nerves and enhances the blood flow[1]. There have been several reports that confirm

piezoelectric phenomenon improves the bone tissue growth and biological response

[9-11]

Page 121: Lead-free piezoelectric materials - Research Online

99

Figure 5.7 (a) Local PFM phase hysteresis loop and (b) piezoelectric butterfly loop

of the NKN nanofiber obtained by switching the bias voltage from -5 to +5 volts at

25 Hz. 180° domain switching in the phase hysteresis loop shows the switching

phenomenon of spontaneous electrical dipoles built up in the NKN nanofiber. (c) d33

hysteresis loop calculated by using the converse piezoelectric equation, Δz = d33V,

indicates a significant enhancement of d33 = 58 pmV-1 in the NKN nanofiber. (d) 3D

topographical image of the NKN nanofiber with 200 nm width and 100 nm height,

used in PFM measurements.

5.4 Conclusions

Biocompatible piezoelectric scaffold constructed by the NKN nanofibers is

fabricated via sol-gel assisted electrospinning technique. The high piezoelectric

response d33=58 pm V-1 and ferroelectric domains developed in the NKN nanofiber

during its crystallization above 700°C are demonstrated using PFM technique.

Ferroelectric domains which manifest themselves as the permanently charged regions

Page 122: Lead-free piezoelectric materials - Research Online

100

and switchable provide an opportunity to create an engineered charged surface with

favorite polarity. Electrical signals generated by the piezoelectric phenomenon in the

NKN nanofibers emerge them as the biocompatible electrically active materials

which can improve the osteoconduction, tissue growth, healing and regeneration of

the damaged biological organs.

5.5 References

[1] E. G. Fine, R. F. Valentini, R. Bellamkonda, P. Aebischer, Biomaterials

1991, 12, 775.

[2] R. Costa, C. Ribeiro, A. C. Lopes, P. Martins, V. Sencadas, R. Soares, S.

Lanceros-Mendez, J Mater Sci: Mater Med 2013, 24, 395.

[3] M. M. Beloti, P. T. de Oliveira, R. Gimenes, M. A. Zaghete, M. J. Bertolini,

A. L. Rosa, J. Biomed. Mater. Res. A 2006, 79A, 282.

[4] T. Kobayashi, S. Nakamura, K. Yamashita, J. Biomed. Mater. Res. 2001, 57,

477.

[5] F. Baxter, I. Turner, C. Bowen, J. Gittings, J. Chaudhuri, J Mater Sci: Mater

Med 2009, 20, 1697.

[6] M. Nakamura, Y. Sekijima, S. Nakamura, T. Kobayashi, K. Niwa, K.

Yamashita, J. Biomed. Mater. Res. A 2006, 79A, 627.

[7] A. Marino, R. Becker, S. Soderholm, Calc. Tis Res. 1971, 8, 177.

[8] G. W. Hastings, F. A. Mahmud, J. Biomed. Eng. 1988, 10, 515.

[9] J. B. Park, B. J. Kelly, G. H. Kenner, A. F. von Recum, M. F. Grether, W. W.

Coffeen, J. Biomed. Mater. Res. 1981, 15, 103.

[10] C. C. Silva, A. F. L. Almeida, R. S. de Oliveira, A. G. Pinheiro, J. C. Góes,

A. S. B. Sombra, J Mater Sci 2003, 38, 3713.

Page 123: Lead-free piezoelectric materials - Research Online

101

[11] J. B. Park, A. F. von Recum, G. H. Kenner, B. J. Kelly, W. W. Coffeen, M. F.

Grether, J. Biomed. Mater. Res. 1980, 14, 269.

[12] J. R. Woodard, A. J. Hilldore, S. K. Lan, C. J. Park, A. W. Morgan, J. A. C.

Eurell, S. G. Clark, M. B. Wheeler, R. D. Jamison, A. J. Wagoner Johnson,

Biomaterials 2007, 28, 45.

[13] F. J. C. Braga, S. O. Rogero, A. A. Couto, R. F. C. Marques, A. A. Ribeiro, J.

S. d. C. Campos, Mater. Res. 2007, 10, 247.

[14] S. Firmino Mendes, C. M. Costa, V. Sencadas, J. Serrado Nunes, P. Costa, R.

Gregorio, Jr., S. Lanceros-Méndez, Appl. Phys. A 2009, 96, 899.

[15] Kenth Nilsson, Johan Lidman, Karin Ljungstrom, C. Kjellman, United States

of America Patent US6526984 B1, 2003.

[16] K. Fujihara, A. Kumar, R. Jose, S. Ramakrishna, S. Uchida, Nanotechnology

2007, 18, 365709.

[17] A. Shui, L. Zeng, K. Uematsu, Scripta Mater. 2006, 55, 831.

[18] Ken-ichi Kakimoto, Koichiro Akao, Yiping Guo, Hitoshi Ohsato, Jpn. J.

Appl. Phys. 2005, 44.

[19] H. J. Trodahl, N. Klein, D. Damjanovic, N. Setter, B. Ludbrook, D. Rytz, M.

Kuball, Appl. Phys. Lett. 2008, 93, 262901.

[20] F. Rubio-Marcos, J. J. Romero, D. A. Ochoa, J. E. García, R. Perez, J. F.

Fernandez, J. Am. Ceram. Soc. 2010, 93, 318.

[21] H. Ge, Y. Hou, X. Rao, M. Zhu, H. Wang, H. Yan, Appl. Phys. Lett. 2011,

99, 032905.

[22] S. Bodhak, S. Bose, A. Bandyopadhyay, Acta Biomater. 2010, 6, 641.

[23] F. R. Baxter, C. R. Bowen, I. G. Turner, A. C. E. Dent, Ann Biomed Eng

2010, 38, 2079.

Page 124: Lead-free piezoelectric materials - Research Online

102

[24] T. Kobayashi, S. Itoh, S. Nakamura, M. Nakamura, K. Shinomiya, K.

Yamashita, J. Biomed. Mater. Res. A 2007, 82A, 145.

[25] K. S. Hwang, J. E. Song, J. W. Jo, H. S. Yang, Y. J. Park, J. L. Ong, H. R.

Rawls, J Mater Sci: Mater Med 2002, 13, 133.

[26] B. D. Ratner, Surface Characterization of Biomaterials: Proceedings of the

Symposium on Surface Analysis of Biomaterials , Ann Arbor, Michigan, June 21-24,

1987 ; Ed. by Buddy D. Ratner, Elsevier Science Publishers, 1988.

[27] M. Ohgaki, T. Kizuki, M. Katsura, K. Yamashita, J. Biomed. Mater. Res.

2001, 57, 366.

[28] W. Wang, S. Itoh, Y. Tanaka, A. Nagai, K. Yamashita, Acta Biomater. 2009,

5, 3132.

[29] A. L. Kholkin, C. Wutchrich, D. V. Taylor, N. Setter, Rev. Sci. Instrum. 1996,

67, 1935.

[30] C. W. Ahn, S. Y. Lee, H. J. Lee, A. Ullah, J. S. Bae, E. D. Jeong, J. S. Choi,

B. H. Park, I. W. Kim, J. Phys. D: Appl. Phys. 2009, 42, 215304.

Page 125: Lead-free piezoelectric materials - Research Online

103

CONCLUSIONS

In this thesis, two lead-free piezoelectric and ferroelectric compounds, (1-x)Ba

(Ti0.80Zr0.20) O3-x (Ba0.70Ca0.30) TiO3 ((1-x)BTZ-xBCT) in various forms and (Na, K)

NbO3 (NKN) nanofibers were investigated.

In Chapter 2, (1-x)BTZ-xBCT ceramics were synthesized using the conventional

solid state reaction at 1450°C in air. Crystal structure and vibration modes were

studied using X-ray diffraction (XRD) and Raman spectroscopy, respectively. The

XRD patterns were refined by employing the Rietveld method. The refinement

results revealed that Ca+2 and Ba+2 substitutions in the BTZ reduced the unit cell

volume and led to a red shift in Raman peaks located at frequencies below 300 cm-1

and a blue shift in peaks appearing at higher frequencies. Ferroelectric properties

measurements indicated that increasing the Ca+2 and Ba+2 ions in the BZT enhanced

the maximum polarization (Ps) and remanent polarization (Pr) in the ceramics due to

the unit cell volume reduction.

In Chapter 3, BTZ-BCT nanofibers 150-250 nm in diameter and several microns

long were fabricated via the sol-gel assisted electrospinning technique at the low

temperature of 700°C. XRD patterns, Raman spectra, and transmission electron

microscope (TEM) examination results demonstrated the coexistence of two polar

structures, both tetragonal and rhombohedral, in the annealed nanofibers.

Observation of two ferroelectric phases revealed the crystallization of the fibers in

the vicinity of the morphotropic phase boundary (MPB) region. Different scanning

probe microscopy (SPM) techniques, including piezoresponse force microscopy

(PFM), scanning capacitance microscopy (SCM), and scanning spreading resistance

aj393
Highlight
aj393
Highlight
aj393
Highlight
Page 126: Lead-free piezoelectric materials - Research Online

104

microscopy (SSRM), were employed to study the piezoelectric, ferroelectric, and

electrical properties of the nanofibers.

The piezoelectric coefficient and ferroelectric domain switching behaviour of the

BTZ-BCT nanofibers were investigated by PFM. 180° domain switching in the phase

hysteresis loops demonstrated the switching of spontaneous electrical dipoles built

up in the BTZ-BCT nanofibers. The d33 hysteresis loop calculated by using the

converse piezoelectric equation, Δz = d33V, indicated that the piezoelectric

coefficient d33 = 180 pmV-1 in the BTZ-BCT nanofibers. The large piezoelectric

constant achieved in the BTZ-BCT nanofibers exceeded the piezoelectric

performance of Pb (Zr, Ti) O3 (PZT) nanofibers and those of other lead-free

piezoelectric nanofibers. SCM examination showed the configuration of the

antiparallel ferroelectric domains in the nanofibers. The SSRM results revealed a

uniform resistance distribution in the BTZ-BCT nanofibers with a high value of 1010

ohms. For this composition, the polarization anisotropy vanishes, leading to a strong

dependence of the electrical polarization on elastic deformation, thus significantly

increasing the piezoelectricity.

In Chapter 4, BTZ-BCT thin films with a thickness of about 200 nm were deposited

by the spin-coating method and annealed at 700°C on Si and Si/SiO2/Ti/Ir substrates.

The crystal structure and Raman spectra of the thin films contained all the features

characteristic of the MPB region, similar to the BTZ-BCT nanofibers. Surface

topography analysis revealed a very smooth surface with 2 nm rms surface

roughness. In the ferroelectric domain patterns recorded by the PFM, the opposite

polarity of ferroelectric domains 20 to 40 nm in size was revealed by 180° phase

difference contrast. The piezoelectric coefficient d33 = 141 pmV-1 in the BTZ-BCT

aj393
Highlight
aj393
Highlight
aj393
Highlight
Page 127: Lead-free piezoelectric materials - Research Online

105

thin film was comparable with those of PZT thin films and other lead-free

piezoelectric thin films.

The calculated piezoelectric coefficient in the BTZ-BCT thin films had a smaller

value compare to the nanofibers. This is attributed to suppression of the residual

ferroelectric-substrate interface strain due to the much smaller lateral size of the

fiber-substrate interface and the fiber-to-substrate contact area.

In Chapter 5, biocompatible and piezoelectric (Na, K) NbO3 (NKN) nanofibers

synthesized by the sol-gel assisted electrospinning technique were explored. The

XRD pattern and Raman spectrum of the nanofibers revealed a pure single phase

with polar structure after annealing at 700°C. Transmission electron microscope

images and electron diffraction patterns showed cube-on-cube growth of NKN single

crystals in the form of nanofibers. The ferroelectric domain switching and

piezoelectric response of the nanofibers were investigated using PFM. A higher

piezoelectric response was achieved in NKN nanofibers (d33 = 58 pm V-1) than in its

thin films (d33 = 40 pm V-1). Owing to the existence of permanently charged regions

in the NKN nanofibers known as ferroelectric domains, electrical signals can be

generated in them via the piezoelectric effect to provide a new opportunity for

construction of a smart biocompatible scaffold that can be used for repair,

engineering, and regeneration of damaged tissues.

aj393
Highlight
Page 128: Lead-free piezoelectric materials - Research Online

106

RECOMMENDATIONS

It has been demonstrated that low-dimensional Ba (Ti0.80Zr0.20) O3- (Ba0.70Ca0.30)

TiO3 (BZT-0.5BCT) structures can be employed in applications such as energy

harvesting systems, microelectromechanical systems (MEMS), and

nanoelectromechanical systems (NEMS) due to their high piezoelectric performance.

The BTZ-BCT nanofibers and thin films can be utilized as replacements for their

PZT counterparts in some applications and resolve the health risks due to the toxicity

of the Pb-containing piezoelectric materials.

It is suggested that, In order to increase the efficiency and output of a piezoelectric

generator, a fibrous geometry could be used rather than the usual film shape

(rectangular cross section) to reduce the effective stress developed in the

piezoelectric material-substrate interface. Fabrication of piezoelectric energy

harvesting systems and sensors employing lead-free piezoelectric BTZ-BCT

nanofibers can be considered as a future plan in continuation of the study presented

in this thesis.

Moreover, exploring the effect of the biocompatible piezoelectric NKN nanofibers

on tissue growth and healing rates of damaged tissues can be considered as the next

step. The growing of live cells on scaffolds composed of NKN nanofibers could be

of interest to scientists working on multidisciplinary research projects.

aj393
Highlight
aj393
Highlight
Page 129: Lead-free piezoelectric materials - Research Online

107

Publications

• A. Jalalian, A.M. Grishin, X. Wang, Z. Cheng and S. X. Dou, Large

piezoelectric coefficient and ferroelectric nanodomain switching in

Ba(Ti0.80Zr0.20)O3- (Ba0.70Ca0.30)TiO3 nanofibers and thin films, Appl. Phys.

Lett. 104, 103112 (2014).

• A. Jalalian, A. M. Grishin, X. Wang, and S. X. Dou, Fabrication of Ca, Zr

doped BaTiO3 ferroelectric nanofibers by electrospinning, Phys. Status Solidi

C 9, No. 7, 1574–1576 (2012).

• A Jalalian, A M Grishin, and S X Dou, Ferroelectric and ferromagnetic

nanofibers: synthesis, properties and applications, Journal of Physics: C 352

(2012) 012006.

• Morphotropic phase boundary observation and quantitative electrical

characterizations of high piezoelectric coefficient Ba (Ti0.80Zr0.20) O3-

(Ba0.70Ca0.30)TiO3 nanofibers (prepared for the submission) .

• Effect of lattice constants on ferroelectric properties of (1-x)Ba (Ti0.80Zr0.20)

O3-x(Ba0.70Ca0.30)TiO3 ceramics ( prepared for the submission).

Attended Conferences

• Poster presentation, Elsevier-Third International Conference on

Multifunctional, Hybrid and Nanomaterials , 3-7 March 2013,Sorrento, Italy .

• Poster presentation, Asia-Pacific Interdisciplinary Research Conference

Toyohashi University of Technology (Toyohashi Tech) on 17-18 November

2011, Japan.

• Oral presentation, 16th Semiconducting and Isolating Materials Conference

(SIMC-XVI), June 19-23, 2011, Stockholm, Sweden.

• Oral presentation, Complex Fluids Symposium, , Micronic MyData AB, 1

Dec. 2011, Taby, Sweden.