photoelastic and electro-optic effects: study of pmn-29%pt

125
Photoelastic and Electro-Optic Effects: Study of PMN-29%PT Single Crystals by Na Di Submitted in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy Supervised by Professor David J. Quesnel Department of Mechanical Engineering Arts, Sciences and Engineering Edmund A. Hajim School of Engineering and Applied Sciences University of Rochester Rochester, New York 2009

Upload: others

Post on 02-Oct-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

Photoelastic and Electro-Optic Effects: Study of

PMN-29%PT Single Crystals

by

Na Di

Submitted in Partial Fulfillment

of the

Requirements for the Degree

Doctor of Philosophy

Supervised by

Professor David J. Quesnel

Department of Mechanical Engineering Arts, Sciences and Engineering

Edmund A. Hajim School of Engineering and Applied Sciences

University of Rochester Rochester, New York

2009

Page 2: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

ii

Curriculum Vitae

The author was born in Shenyang, Liaoning province, China on November 28,

1977. She attended Liaoning Key High School and graduated in 1996. She

enrolled at Fudan University in 1996 and finished her B.S. degree program in

Theory and Applied Mechanics in 2000. Thereafter she continued her graduate

study at Fudan University and graduated with a Master’s degree in Engineering

Mechanics in 2003.

In fall 2003, she was accepted into the doctoral program at the University of

Rochester under the supervision of Professor David J. Quesnel. She received

her second Master’s degree in Mechanical Engineering from the University of

Rochester in 2005.

In May of 2005 she attended the U.S. Navy Workshop on Acoustic

Transduction Materials and Devices where she became familiar with the issues

constraining the behavior of next generation piezoelectric single crystals. Shortly

thereafter, she conceived of the idea of using photoelastic methods to

characterize the stress distributions in these materials from which this thesis

developed. While pursing her thesis research, she regularly participated in the

U.S. Navy Workshop on Acoustic Transduction Materials and Devices by making

the presentations that are listed below.

“Photoelastic study of PMN-29%PT single crystals”, U.S. Navy Workshop on

Acoustic Transduction Materials and Devices, May 2006.

“Photoelastic study of PMN-29%PT single crystals”, U.S. Navy Workshop on

Acoustic Transduction Materials and Devices, May 2007.

“Photoelastic study of PMN-PT single crystals under electric fields”, U.S. Navy

Workshop on Acoustic Transduction Materials and Devices, May 2008.

Page 3: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

iii

Acknowledgements

I would like to thank my advisor, Professor David J. Quesnel first. I am

thankful for his diligent guidance and constant encouragement. I have learned a

lot from him, from academic knowledge to language and life. Without his

financial and academic support throughout my graduate studies, I would never

have been able to finish my thesis.

Next, I would like to thank Mr. John C. “Jace” Harker and Mr. Stephen R.

Robinson. Jace is a fantastic lab mate, who always has a lot of brilliant ideas,

and will share them with me without reservation. We have held many meaningful

discussions over my research problems, and he helped a lot with my writing.

The strong technical skills of Stephen, who prepared samples and took the

photographs, are very much appreciated. He also helped me to improve my

English writing. Without Jace and Stephen’s help, I also would never have been

able to write out my thesis.

Many thanks to Professor Sheryl M. Gracewski, Professor Paul D.

Funkenbusch, Professor James C. M. Li, Professor John C. Lambropoulos,

Professor Stephen J. Burns, Professor Renato Perucchio and Professor Ahmet T.

Becene for their guidance and the knowledge I have learned from their classes.

I would like to thank Chris Pratt for helping me in conducting X-Ray

experiments. Thank you also to Jill Morris and Carla Gottschalk for all their help

along the way.

Final words of thanks go to my parents for their love and support.

Portions of this thesis are derived from publications that appear in the archival

literature. In particular, Chapter 2 draws from: Na Di and David J. Quesnel,

“Photoelastic effects in Pb(Mg1/3Nb2/3)O3-29%PbTiO3 single crystals investigated

by three-point bending technique”, J. Appl. Phys. 101, 043522 (2007); and

Chapter 3 is derived from: Na Di, John C. Harker, and David J. Quesnel,

Page 4: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

iv

“Photoelastic effects in Pb(Mg1/3Nb2/3)O3-29%PbTiO3 single crystals investigated

by Hertzian contact experiments”, J. Appl. Phys. 103, 053518 (2008); In this

work, John Harker’s contribution was through editing of the initial draft to a form

suitable for publication, with the technical discussion necessary to get the

meanings as intended.

Chapter 4 and Chapter 5 will be submitted for consideration as journal

articles. This is reflected in the format selected for these chapters. They will be

coauthored with my advisor, David J. Quesnel.

Page 5: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

v

Abstract

Relaxor ferroelectrics PMN-PT single crystals exhibit extra-high dielectric and

piezoelectric properties compared with conventional piezoelectric ceramics.

They are becoming widely used in high performance electromechanical devices.

However, PMN-PT single crystals are elastically softer than PMN-PT

polycrystalline ceramics. Mechanical loads and electric fields interact to produce

fractures at relatively low stresses, and cracks grow under both AC and DC

electric fields. To prevent the failure of the electromechanical devices, we need

to have a better understanding of the mechanisms of fracture in this material

when it is subjected to mechanical and electrical loadings.

Photoelasticity is an efficient and effective method to measure the internal

stress distributions of materials that result from both internal residual stress and

external loading. I report the exploration of the use of this classic technique to

study internal stresses inside PMN-PT single crystals through bending and

Hertzian contact experiments. Effects under electric field loading were also

investigated using birefringence techniques.

Page 6: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

vi

Table of Contents

Chapter 1 Introduction - piezoelectric single crystals & photoelasticity

1.1 PMN-29%PT single crystals……….……………………..………….1

1.2 Photoelasticity………………………………………………... ……...7

1.2.1 Discovery of the phenomenon of Photoelasticity…………7

1.2.2 Mathematical formulation of Photoelasticity…..................9

1.2.3 Plane polariscope and circular polariscope….… ………10

1.3 Preliminary three-point bending experiments……………... …….13

1.3.1 Experimental setup……………………..……..… ………..13

1.3.2 Fringe pattern……………………..……...….....................15

1.3.3 Deflection versus fringe order……………...…....……….17

1.3.4 Summary…………… …..………………………….………18

1.4 References..……………………………….… ………....................20

Chapter 2 Photoelastic study using three-point bending technique

2.1 Introduction……………………………………………..………….. .26

2.2 Experimental procedure..…..…………………………….. ……….29

2.3 Results and discussion……………………………… …….………32

2.3.1 Fringe pattern……………………………………………....32

2.3.2 Loading force versus deflection………………................34

2.3.3 Stress-optical coefficient…………………..…..………….36

2.3.4 Young’s modulus……………………….…………………..38

2.4 Summary………………………….….…………….............. ……...39

2.5 References…………………….……………………………............40

Page 7: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

vii

Chapter 3 Photoelastic study using Hertzian contact experiments

3.1 Introduction…………………………………………..…..…..…….. 42

3.2 Experimental procedure..……………………………...… …….....44

3.3 FEM modeling methods………………….……………….. ………46

3.4 Results and discussion……………………………….…… ………50

3.5 Conclusions……………………………...………….......................53

3.6 References…………………………….…………………………….54

Chapter 4 Photoelastic study using four-point bending technique

4.1 Introduction………………………………..………………..............57

4.2 Experimental procedure..…………………………..……………....60

4.3 Results and discussion…………………………………….............63

4.3.1 Fringe pattern………………………………………………63

4.3.2 Fiber stress versus fringe order………………............... 64

4.3.3 Fringe-stress coefficient…………………..…..…............ 66

4.3.4 Mechanical poling effect……………….………………… 69

4.4 Conclusions……………………………….… .……...................... 69

4.5 References………………………………….……… ….…. ……… 70

Chapter 5 Electrical field induced optical effects in PMN-29%PT single crystal

5.1 Introduction…………………………… ……………..…….............73

5.2 Experimental procedure..………………………………................77

5.3 Results and discussion………………………………...................79

5.3.1 “Hertzian Contact” electric field loading effects………...79

5.3.2 Electric poling effects………….………………................ 85

5.3.3 Mechanical poling versus electrical poling……………...89

5.4 Conclusions……………………………….………..……................91

Page 8: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

viii

5.5 References……………………………….…………… .…... ……...92

Chapter 6 Summary

6.1 Summary…………………….………………..………..…...............95

6.2 References……………...………………………..………………….99

Appendices

Appendix Ⅰ Basic theory of optical properties of crystals…………………101

Appendix Ⅱ Basics of photoelasticity……………...………………..………104

Appendix Ⅲ Calibration of in-situ loading frame…………………….……..108

References for Appendices………..……………………………..………… . ..111

Page 9: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

ix

List of Tables

Table 3.1 Elastic stiffness constants of PMN-30%PT single crystals (10Dijc 10

N/m2)………………………………………………………………………...47 Table 3.2 Input parameters used in ANSYS®. The elastic stiffness constants:

(10ijc 10 N/m2). Young's modulus of glass: E (1010 N/m2). Poisson's ratio

of glass: ν ………………………………………………………………….48

Page 10: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

x

List of Figures

Figure 1.1 (a) ABO3 Perovskite Structure. If one shifts the A-site ions to the center of the cell, the structure will appear differently with 12 oxygen atoms at the center of each cell edge, an arrangement often shown in geology texts, (b) Spontaneous polarization for the R phase in unpoled PMN-29%PT; Illustration redrawn from a similar figure in reference [5].………..............................................................................…...……….4

Figure 1.2 Phase diagram of PMN-PT single crystals; Illustration redrawn from

a similar figure in reference [33]….………………………………......……5 Figure 1.3 PMN-29%PT single crystal as received……………………………….8 Figure 1.4 (a) Dark field plane polariscope set-up. (b) Light vector

representation; Illustrations redrawn from a similar figure in reference [38]………………………………………………………………………….11

Figure 1.5 ZeissTM Microscope set-up……………………………………..….… .14 Figure 1.6 Preliminary three-point bending set-up………………………………14 Figure 1.7 Three-point bending image at 450 to both the polarizer and the

analyzer……………………………………………………………………... ..15 Figure 1.8 Principal Stress Vectors from ANSYS® simulation of three-point

bending. Only left half of sample is shown…….……………………………16

Figure 1.9 Three-point bending image at zero degree to both the polarizer and the analyzer………………………...……………………………………….16

Figure 1.10 (a) Three-point bending of unpoled PMN-29%PT; (b) Three-point

bending image of isotropic materials [39]………………………………….17

Figure 1.11 Deflection versus fringe order……..…………………………………..17 Figure 1.12 3D CAD model of loading frame. BimbaTM cylinder is mounted

through a hole in the aluminum frame.………………………………….…..19

Figure 2.1 (a) in situ loading frame working under microscope and (b) in situ

Page 11: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

xi

loading frame with three-point bending set-up as indicated by the arrow……………………………………………………………………. ...30

Figure 2.2 3-point bending schematic. P is the loading force;  c and t are the

compression and tension fiber stress. and are the reaction

loads…………………………………………………..……………………..30

1 2R R

Figure 2.3 Unpoled PMN-29%PT single crystal beam viewed in crossed polars:

(a) (100) face, as-received; (b) (100) face, after annealing; and (c) (010) face, after annealing. Polarizer and analyzer are horizontal and vertical, respectively. Strong colors in (a) indicate regions of net birefringent retardation…………………………………………………………………31

Figure 2.4 (a) Initial fringe pattern showing 0.5 fringe at point A on the free

surface opposite the loading point. (b) Second-order fringes at A, (c) Sixth-order fringes at A, and (d) first-order fringe remaining at A after the load is released…………………………………………………………33

Figure 2.5 Force versus deflection during increasing load for three experimental

runs…………………………………………………...................................34 Figure 2.6 Force versus deflection with polynomial fit curve…………………….35

Figure 2.7 Fringe order versus fiber stress. The stress-optical coefficient is

calculated from the slope of the proportional region…………………….37 Figure 3.1 (a) Top view of the loading frame. (b) Hertzian contact experimental

set-up as indicated by the arrow…………………………………………..45 Figure 3.2 Initial birefringence patterns of three samples in three different

orientations under circularly polarized illumination………………46

Figure 3.3 The 3 differently oriented samples relative to {001}-oriented pseudo-cubic axes. Arrows a and c represent compression along <100> direction; arrows b and d represent compression along <110> direction……………………………………………………………………...46

Figure 3.4 ANSYS® model for use in computation of fringe pattern images.

Boundary conditions are shown. Contact elements are used at the interface between the Hertzian cylinder indenter and the rectangular

Page 12: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

xii

piezocrystal (light gray). Cyan symbols represent displacement constraints. Red arrow indicates the force applied to all of the coupled nodes (green)………………………………………………………………49

Figure 3.5 (a) Hertzian indentation along <100> direction on sample 1. (b) Stress

intensity contour from ANSYS®…………………………………………..50 Figure 3.6 (a) Hertzian indentation along <110> direction on sample 2. (b) Stress

intensity contour from ANSYS®……………………………………………51

Figure 3.7 Hertzian indentation along <100> direction on sample 3 is shown in (a); Hertzian indentation along <110> direction is shown in (c). The initial birefringence is responsible for the asymmetric fringe in (a) and the layers along the surface in (c). Stress intensity contour from ANSYS® are shown in (b) and (d) correspondingly……………….………..…..51

Figure 3.8 Residual butterfly fringes are fully annealed out at 400 oC for one

hour…………………………………………………………………………..53

Figure 4.1 (a) Overview of the in situ loading frame and (b) Four-point bending set-up; A represents the tilting bar, and B represents the beam sample………………………………………………………………........….61

Figure 4.2 Beam 1 (a) and beam 2 (b) after one hour annealing at 400 oC…….61

Figure 4.3 Four-point bending layout. P is the loading force, cσ and tσ are

compression and tension stresses respectively. The diagram under the sample shows the absolute value of the bending moment……………..62

Figure 4.4 (a) 3.5 order of fringes and (b) 2 order of fringes left after the load is

released……………………………………………………………………...64 Figure 4.5 Maximum fiber stress versus fringe order……………………..………65 Figure 4.6 Maximum fiber stress versus fringe order with polynomial fit curve...65 Figure 4.7 From the light intensity plot, displacement between fringes can be

measured. Each valley of the intensity curves represents a fringe (darkest field), and each peak of the intensity curves represents the half order of fringe (brightest field)………………………………………..……66

Page 13: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

xiii

Figure 4.8 Stress versus fringe order for different load level. The number label represents the maximum fringe order obtained for each load level. The slope of each data line represents the fringe-stress coefficient….…….67

Figure 4.9 Fringe stress coefficient versus maximum fiber stress………………68 Figure 4.10 Fringe patterns of pure bending region at different load levels. (a)

Totally 11 order of fringes; (b) totally 16 order of fringes………………..69 Figure 5.1 Initial birefringence patterns of four differently oriented samples under

circularly polarized illumination after one hour annealing at 400 oC………………………………………………………………………..77

Figure 5.2 (a) Overview of the in situ electrical loading frame and (b) Top view of

“Hertzian contact” electrical loading set-up………………………..…..…78 Figure 5.3 “Hertzian contact” electrical loading (electrical point load) experiments

on {100}-oriented beam 1: (a) -2.3 KV/cm DC were applied from the top rod electrode for 2 minutes; (b) 2.3 KV/cm for 2 minutes; (c) 2.3 KV/cm applied to resulting fringes of (a) for another 2 minutes; (d) 2.3 KV/cm for additional 2 minutes after (c). The arrows in the pictures represent the electric field direction……………………………………………………… 81

Figure 5.4 Fringe pattern comparison: (a), (c) and (e) are fringes induced by DC

electric field loading for beam 1 with 2.3KV/cm, sample 3 with 1.8KV/cm and sample 4 with 1.8KV/cm. (b), (d) and (f) are fringes under Hertzian mechanical loading for comparably oriented samples, as shown in Chapter 3.……….…………………………………………………………...82

Figure 5.5 “Hertzian contact” electrical loading experiments on differently

oriented samples using square waveform voltage with 0.5 Hz and 500 Hz, respectively: (a), (b), and (c) (top row) resulted from square waveform voltage of 0.5 Hz. (d), (e), and (f) (bottom row) were from square waveform voltage of 500 Hz. (a) and (d) are from beam 1; (b) and (e) are sample 3; (c) and (f) are sample 4. The magnitude of electric field is 2.3 KV/cm for beam 1 and 1.8KV/cm for both sample 3 and 4...84

Figure 5.6 From top to bottom, beam 2 is electrical poled with increment DC

voltage. The experiment set-up is with two block electrodes; both the top and bottom surfaces are plated with gold………………………….…….86

Page 14: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

xiv

Figure 5.7 (a) 2.8 KV/cm square waveform cyclic electric field with 0.5 Hz was applied to beam 2. (b) Birefringence of beam 2 after annealing. The arrow points at crack generated during the experiment………………..88

Figure 5.8 (a) 2.5 KV/cm DC electric field was applied to sample 4 using

electrical “Hertzian contact” experimental set-up. (b) Hertzian mechanical loading on poled region……………………………………....89

Figure 5.9 Mechanical poling and electrical poling representation………...……90 Figure A1 Representation of optical index ellipsoid; Illustration redrawn from a

similar figure in reference [1].……………………………………...….....101 Figure A2 Circular polariscope set-up, reproduced from a similar figure in

reference [3]………………………………………………………………..104 Figure A3 Top view of calibration stage…………………………………………. 108 Figure A4 Side view of calibration stage………………………………………….108 Figure A5 Overview of loading frame……………………………………………..109 Figure A6 Loading force versus pressure…………………………………...…...110

Page 15: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

1

1 Introduction - piezoelectric single crystals and

photoelasticity

1.1 PMN-29%PT single crystals

In 1880, the famous brothers Pierre and Jacques Curie first discovered direct

piezoelectric effects in quartz crystals [1, 2]. They found that when a weight is

placed on the surface of a quartz plate, electric charges are generated on both

surfaces of the quartz plate. The charge was measured to be linearly

proportional to the weight placed. Following the discovery of the direct

piezoelectric effect, Lippmann in 1881 theoretically predicted the converse

piezoelectric effect, which says a voltage applied to a piezoelectric crystal

produces elastic strains in the crystal [2, 3]. Later, general theory of

piezoelectricity was thoroughly accounted by Voigt [2, 4]. For the next 60 years,

extensive characterization was performed on BaTiO3 ceramics. In the 1950’s,

Pb(Zn1/3Nb2/3)O3 (PZT) ceramics were found to exhibit an exceptionally strong

piezoelectric response. Since then, modified PZT ceramics and PZT-based solid

solution systems have become the dominant piezoelectric ceramics for various

applications [5].

This defining characteristic of the piezoelectric materials is due to the fact that

the centers of positive and negative charges do not coincide. Namely the crystal

structure does not have a center of symmetry. Such materials possess a

spontaneous polarization. When the spontaneous polarization can be reversed

by an applied electric field, the material is called a ferroelectric. Thus

ferroelectrics are a subset of piezoelectric materials.

In contrast to conventional piezoelectric ceramics, single crystal relaxor

ferroelectrics Pb(Mg1/3Nb2/3)O3-xPbTiO3 (PMN-xPT) and

Page 16: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

2

Pb(Zn1/3Nb2/3)O3-xPbTiO3 (PZN-xPT) exhibit extra-high dielectric and

piezoelectric properties and have become a new generation of piezoelectric

materials, attracting constant attention in recent years [6-10]. Both of them are

widely used in high performance applications such as medical imaging, active

noise suppression, and acoustic signature analysis.

Because PMN-PT has relatively high field-induced strain response and a

small hysteresis loop compared to PZN-PT, PMN-PT is more attractive than

PZN-PT [10]. Furthermore, relaxor-based ferroelectric single crystals PMN-PT,

with compositions near the morphotropic phase boundary (MPB) between the

ferroelectric rhombohedral and tetragonal phases, have ultimate

electromechanical coupling factors (k33 >90%), high piezoelectric coefficients

(d33>2000 pC/N) and high strain levels up to 1.7% [11-12]. They also have

potential to be used in electro-optical technology for their high electro-optical

coefficients [13-14]. Thus my study focuses on PMN-29%PT (close to MPB)

single crystals. The origins of PMN-PT single crystals’ excellent performance

have been attributed to the polarization rotation induced by the external electric

field [15].

However, these materials face crack problems which will reduce the

performance of the devices. Some researchers have studied the fracture

problems of piezoelectric materials, both theoretically [16-21] and experimentally

[22-26], however, most of them focused on using an AC electric field to drive the

growth of existing cracks. Because internal stress plays a significant role in

causing cracks and also in the propagation of cracks, further study of these

internal stresses induced either by mechanical loading or by electrical loading is

an important research topic. This will enable us to better understand and control

the internal stresses in relaxor ferroelectrics devices.

What are the possible sources of internal or residual stress in PMN-29%PT?

Page 17: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

3

Residual stresses are induced by inhomogeneous strain. Inhomogeneous strain

may be produced by thermal gradients during crystal growth [27-28], by phase

transitions during cooling [27, 29], and by mechanical operations, such as cutting,

grinding and polishing during each step of the machining processes [29]. Each

step, therefore, has the potential to produce more residual stresses in the single

crystals.

The topic is significant, but also very difficult due to the complicated internal

domain structures and the intrinsic coupling effects between mechanical and

electric fields. PMN-xPT single crystals have a simple perovskite ABO3 structure

above Curie temperature (for PMN-29%PT, the Curie temperature is about 135

°C), pictured in Figure 1.1(a), and it may readily have complex perovskite

structure A(B1/3B’2/3)O3, as well. X-ray diffraction (XRD) shows unpoled

PMN-xPT single crystals have a tetragonal-rhombohedral MPB (morphotropic

phase boundary). When x is under 30%, PMN-xPT is in rhombohedral (R) R3m

phase at room temperature; when x is above 33%, it begins to transform to

tetragonal (T) P4mm phase through monoclinic (M) or orthorhombic (O)

symmetries [30-33]. The spontaneous polarization direction of the R phase is

<111> and that of the T phase is <001>. The piezoelectric effect is observed to

peak at the morphotropic phase boundary. The enhancement in the

piezoelectric effect at the morphotropic phase boundary has been attributed to the

coexistence of the different phases, whose polarization vectors become more

readily aligned by an applied electric field when mixed in this manner than may

occur in either of the single phase regions.

PMN-29%PT is in the R phase at room temperature and there are eight

possible directions for the spontaneous polarization as shown in Figure 1.1(b).

After an electric field poling, PMN-29%PT will transform from R to M phase first

[32]. With increasing poling field, M to T phase transition may occur. Phase

Page 18: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

4

diagram is shown in Figure 1.2 [33]. The coercive electric field is about 5 KV/cm.

With the fluctuation of chemical composition, macro-domains formed with different

polarization directions. Some researchers even found that within

macro-domains (μm scale), there are micro-domains (0.1 μm scale), and within

micro-domains, there are nano-domains (nm scale). This is called the domain

hierarchy [34]. Accordingly, internal stress study has different levels. In this

thesis, I examine internal stress at the μm level through the use of optical

methods.

Figure 1.1 (a) ABO3 Perovskite Structure. If one shifts the A-site ions to the center of the cell, the structure will appear differently with 12 oxygen atoms at the center of each cell edge, an arrangement often shown in geology texts, (b) Spontaneous polarization for the R phase in

unpoled PMN-29%PT; Illustration redrawn from a similar figure in reference [5].

The constitutive equations of the piezoelectric materials are [35-36]:

321321ricitypiezoelectconverseelasticity

kkijklijklij EesC −=σ (1-1)

321321ricitypiezoelectconverseelasticity

kkijklijklkl EdSs += σ (1-2)

321321typermittiviricitypiezoelectdirect

kikklikli EseD ε+= (1-3)

Page 19: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

5

where ijσ , , and are stress, strain, electric field and electric

displacement tensors, respectively. , ,

kls kE iD

ijklC ijklS ikε , and are the elastic

constant tensor, elastic compliance tensor, the dielectric constants, the

piezoelectric stress coefficients and the piezoelectric strain coefficients,

respectively; these tensors are material specific. So, both the external

mechanical loading and electrical loading will induce internal stress/strain and

polarization, accompanied by domain switching. If the loading is large enough, it

can even induce phase transitions. This coupling between electrical and

mechanical field variables in the constitutive equations will bring serious

mathematical difficulty to the internal stress analysis problem.

kije kijd

Figure 1.2 Phase diagram of PMN-PT single crystals; Illustration redrawn from a similar figure

in reference [33].

Traditionally in materials research any of several types of strain gages can be

employed to help measure the internal strain and, further, to analyze the internal

stress. Since the available samples are too small to use strain gages, optical

methods were adopted, i.e. photoelasticity techniques. Compared with other

stress measurement techniques, photoelasticity can offer efficient quantitative

determination as well as qualitative observation of the stress distributions

Page 20: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

6

resulting from both internal stress and external loading [37-46]. Initial

explorations show that PMN-29%PT single crystals can be polished to be optically

transparent and the application of external loads produces an extremely large

number of well-defined fringes when observed with a polarizing microscope.

This work is novel because it explores the usage of optical techniques in

measuring stresses in next generation piezoelectric materials, which will be an

important quality assurance tool to produce robust and reliable devices in the

years ahead.

Optical methods can help to analyze electrical loading effect in piezoelectric

materials as well. An electric field applied to the piezoelectric single crystals will

cause at least three effects. First, the refractive index changes in proportion to

the electric field. This is known as linear electro-optic (EO) effect. Second, the

electric field induces internal stress/strain; this is known as the converse

piezoelectric effect. These internal stresses will induce photoelastic effects.

When the electric field is large enough, it can also pole the sample (align domains)

to induce phase transformations. Third, the refractive index changes in

proportion to the square of the electric field. This is known as the quadratic

electro-optic or Kerr effect. All of these three effects contribute to the observed

birefringence. Recently the optical properties of piezoelectric materials such as

the refractive indices have been reported [47-50]. Unpoled PMN-29%PT singe

crystals have many domains with different orientation, retaining an optically

isotropic pseudocubic state. Under this assumption, the refractive index of

unpoled PMN-30%PT single crystals is reported to be 2.501 [47]. Poled

tetragonal PMN-38%PT single crystals have an effective EO coefficient of

42.8 pm/V as reported in reference [48]. However, because applied field will

cause domain rotation and phase transformation in unpoled PMN-PT single

crystals, precise determination of the pure EO coefficients is neither possible, nor

Page 21: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

7

useful. Thus EO coefficient determination is not included in this thesis.

Besides the electric-optic effect, several researchers studied the cyclic electric

field induced effects in ferroelectric ceramics or piezoelectric single crystals

[51-56], which including domain switching, phase transformation, micro-cracking,

and fracture. The observation and study are normally carried out through TEM,

dielectric measurements, and optical microscopy, etc. Crack growth is directly

observed under the optical microscope and micro-crack growth under TEM.

Phase transformation is studied by measuring change of the dielectric properties.

These studies help reveal what is going on when the piezoelectric materials are

under electrical loading. However, if we can get to know the internal stress state

of the materials during the electrical loading, we can better understand the crack

initiation condition and crack growth. Fortunately, unpoled PMN-PT single

crystals can be polished to be transparent and show colorful birefringence. I

focused on AC/DC field-induced birefringence of samples originally without a

crack. Crack initiation caused by electric fields and phase transformations were

examined. Optical observation of phase transformations is a field where not

much research has been done and further study is necessary.

Commercial FEM software ANSYS® was applied to model the experiments

and offer theoretical/computational results, helping to interpret the experiment

results that I obtained.

1.2 Photoelasticity

1.2.1 Discovery of the phenomenon of photoelasticity

Photoelasticity is a well-known efficient method to measure the internal

stresses in a variety of transparent materials. This phenomenon was first

discovered by Sir David Brewster in the year 1815 [40]. He presented a paper

before the Royal Society of London where he reported the effect. In his

Page 22: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

8

experiments, he placed a piece of glass in between two crossed polarizers. He

found that when the glass was stretched transversely to the direction of

propagation of light, the field of view grew brighter, therefore showing that an

artificial birefringence is induced in the glass by the mechanical stress.

Furthermore, he found in the case of solids which are initially birefringent, the

initial birefringence is altered by the stress. Thereafter, photoelastic techniques

were developed to study crystals and other transparent solids. It has become an

important experimental method for the measurement of internal stress.

Figure1.3 PMN-29%PT single crystal as received.

Employing this method is very simple. Namely using a crossed polarizer

set-up, we can “see” stresses. Bright colors such as magenta and green, as well

as closely spaced fringes imply high level stresses. With this simple rule, we can

already tell that the samples as received have large birefringence. Figure 1.3

shows an example of fringe pattern for the as received sample. This fringe

pattern results from the net birefringence associated with combining the initial

domain distribution and residual stresses. Broad faces are polished transparent

and the four edges are in as received condition. The big lobes on four corners

with many little bumps along the edges are most likely due to residual stresses

from machining operations. Usually machining stresses are compressive near

the surface and tensile inside the sample.

Page 23: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

9

1.2.2 Mathematical formulation of Photoelasticity

The stress-optic law of photoelasticity states that if there is a difference in

principal stresses along two perpendicular directions in an otherwise optically

isotropic material, the refractive index in these directions is different and the

induced birefringence is proportional to the difference. For a two-dimensional

stress state, the law simplifies to [38-39]:

( 21 σσλ

−= Ctn ) (1-4)

Here 1σ and 2σ are the maximum and minimum principal stresses, n is

fringe order, t is the sample’s thickness, and λ is the wavelength of the incident

light. C is the stress-optic coefficient which is a constant. From Eq. (1-4), we

have:

Ctn

2221 λσσ

=− (1-5)

Hence

Ctn

2maxλτ = (1-6)

The stress-optic coefficient C is useful to quantitatively analyze the internal

stresses, such as maxτ as a function of position for any fringe pattern if we know

the fringe order n.

Eq. (1-5) can also be rearranged to:

tnf

Ctn

==−λσσ 21 (1-7)

Here, we introduce another concept: the fringe-stress optical coefficient , where f

Cf /λ= represents the principal stress difference necessary to produce a one

fringe order change in a crystal of unit thickness. The fringe-stress coefficient

depends on the stress-optical coefficient C of the material and the wavelength of

the incident light. Therefore, C is more general and convenient than allowing f

Page 24: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

10

the use of incident light with different wavelength. However, to allow easy stress

calculation and also easy comparison of optical properties with mechanical

properties, need to be evaluated. When is used for stress calculation, a

single standard wavelength should be used. While any color monochromatic light

is acceptable, ~535 nm green light was selected for the experiments reported in

this thesis. The engineering units of the fringe-stress coefficient are N/m.

f f

For anisotropic crystals, the mathematical formulation of photoelasticity is

more complex. The relevant equations are in Appendix I. As shown in Figure

1.1(b), unpoled PMN-29%PT single crystals have eight possible polarization

directions. When the number of domains is large enough, the global structure of

the unpoled sample can be treated as pseudocubic for unpoled PMN-29%PT

single crystals [51] and pseudotetragonal for <100> poled crystals [58]. These

assumptions enable photoelastic methods to be applicable to these materials in

theory, though the experimental results may turn out differently due to the

complex domain hierarchy structures that can develop as a result of

thermal/mechanical processing’s history of the samples.

1.2.3 Plane polariscope and circular polariscope

The general arrangement of light fields to perform photoelastic experiments

consists of two typical arrangements: The plane polariscope and the circular

polariscope. In photoelasticity, stress fields are displayed through the use of

light. The basic arrangement of a plane polarized microscope includes a

polarizer and an analyzer, mounted with a 900 rotation between them to minimize

the transmission of light through the pair. If an isotropic material is placed

between the plates, it will not affect the intensity of the transmitted light regardless

of its angle to the polarizers. This setup, with polarizers crossed is called a dark

field plane polariscope, as shown in Figure 1.4(a). The other arrangement,

Page 25: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

11

called a bright field plane polariscope, features the polarizer and analyzer parallel

to one another and was not used in this analysis.

(a) (b) Figure 1.4 (a) Dark field plane polariscope set-up. (b) Light vector representation; Illustrations

redrawn from a similar figure in reference [38].

In Figure 1.4(b), consider polarized light coming out the polarizer aligned with

E1 parallel to the x axis:

)cos(1 tkE ω= (1-8)

When entering the sample, the light vector splits to two vectors along the principal

stress axes. As the two components of light propagate through the sample, a

phase difference of δ is generated. Let be the slow axis and be the

fast axis, we have:

2E 3E

)2

cos(cos

)2

cos(sin

3

2

δωθ

δωθ

+=

−=

tkE

tkE (1-9)

After the light passes out through the analyzer, only the y axis component of the

light is visible, so:

Page 26: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

12

)sin(2

sin)2sin(

)2

cos()2

cos(cossin

sincos 324

tk

ttk

EEE

ωδθ

δωδωθθ

θθ

=

⎥⎦⎤

⎢⎣⎡ +−−=

−=

(1-10)

The intensity of the light we see thus dependent on the orientationθ , the time

and the phase difference t δ for each point in the image. Regions where

)2sin( θ or 2

sin δ or )sin( tω are zero are dark. The overall appearance is

similar to a contour map. These black bands in the stress patterns are known as

fringes. Namely when intensity of the light is zero, there is a fringe. Intensity is

proportional to the square of the amplitude and the time dependent term is usually

not considered:

)2

(sin)2(sin 22 δθap II = (1-11)

Here represents the amplitude of the incident light and other factors

affecting the transmission light intensity. From Eq. (1-11), we can see, there are

two set of fringes superimposed over each other, isochromatics and isoclinics.

Isochromatics are caused by the incident light phase difference

aI

δ of 2m (here  

m is an integer), or as is often said, a retardation caused by the principal stress

difference at the point. Isoclinic fringes are contours of constant inclination,

when the polarizer axis coincides with one of the principal stress directions at the

point of interest, 2/,0 πθ = .

Use of a circular polariscope eliminates isoclinics. Two quarter-wave plates

are added to the plane polariscope with their axes at 450 and 1350 to one of the

polarizers to achieve circular polarized microscopy. The details of circular

polariscope are described in Appendix II with the basic set up illustrated in Figure

A2. The result is that the intensity of light transmitted for circular dark-field only

depends on the retardationδ , thus only isochromatics will be seen. Circular

Page 27: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

13

polarized microscopy is used in most experiments conducted here. Plane

polarized microscopy can help to define the phase of PMN-PT single crystals by

measuring the extinction angle of the light through the crystal relative to the

direction of the polarizer.

1.3 Preliminary three-point bending experiments

1.3.1 Experimental setup

A ZeissTM Axioskop2MAT microscope was used for all the photoelasticity

experiments conducted, as shown in Figure 1.5. This microscope was modified

by addition of a rotation stage normally found on polarizing microscopes to

facilitate rotation of the sample. Light goes straight up from the bottom. A 2

megapixel camera is used to transfer the images to the connected computer, so

we can observe the images on a large monitor. When quarter wavelength

retardation plates are applied, the microscope is configured as a circular

polariscope. In preliminary three-point bending experiment, the microscope was

configured as a plane polariscope without quarter wavelength plates.

For preliminary three-point bending experiments, a parallel clamp with jaws

only 1mm tall was designed to apply the force to a sample while observing with a

polarizing microscope. This device is shown in Figure 1.6. Screws A and B are

adjusted individually to keep the loading faces parallel to each other as they are

brought together. Three semi-cylinder shaped glass rods were cut and polished,

each with 1mm height and 1 mm radius to exert loading and supporting forces.

Page 28: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

14

Figure 1.5 ZeissTM Microscope set-up.

Figure 1.6 Preliminary three-point bending set-up.

An unpoled [001]-oriented PMN-29%PT single crystal bar was used in these

bending experiments. This crystal was obtained from H.C. Materials Corporation,

Bolingbrook, Illinois. Unless otherwise noted, crystals used for this research

were grown by H. C. Materials Corporation. The surfaces for light transmission

Page 29: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

15

were polished down to 0.05 μm grit size. The dimension of the bar is 11X1.8X1

mm as measured directly from micrographs taken under a 2.5X objective lens.

The experiments are performed between crossed polarizers – under plane

polarized dark field. A green light filter with 535nm wavelength was applied to

show well-defined fringes.

1.3.2 Fringe pattern

Figure 1.7 is taken at an angle of 450 to both polarizer and analyzer. This

image clearly shows isoclinic fringes on the neutral axis, caused because

principal stresses on the neutral axis are perpendicular to each other and at an

angle of 450 to the parallel and perpendicular edge directions of the three-point

bending specimen. To verify this, the principal stress field in a beam under

three-point bending was calculated using ANSYS® software. Figure 1.8 shows

the resulting principal stress field displayed as vectors modeled in an isotropic

material. It is evident that the principal stress direction is at 450 to both the

polarizer and analyzer inside the solid line circles, which should be a bright region,

and is 00 to both the polarizer and analyzer inside the dash line circle, which

should be all dark under the microscope according to the photoelastic theory.

This matches the fringe pattern shown in Figure 1.7, implying that the basic

photoelastic technique works on PMN-29%PT single crystals. This supporting

result is also verified in subsequent four-point bending experiments.

Figure 1.7 Three-point bending image at 450 to both the polarizer and the analyzer.

Page 30: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

16

Figure 1.8 Principal stress vectors from ANSYS® simulation of three-point bending. Only left

half of sample is shown.

Fringe picture taken at zero degree to both polarizer and analyzer is shown in

Figure 1.9, and it shows isoclinic fringes as well. This may be easily verified by

rotating the analyzer and polarizer coordinate 450 clockwise in Figure 1.8.

Figure 1.9 Three-point bending image at zero degree to both the polarizer and the analyzer.

When the applied loading force is increased, fringes were observed

generated at the central portions of the top and bottom edges and move towards

the neutral axis. This process continues until we cause the bar to snap, usually

with as many as 25 or more fringes. The fringe pattern seen in PMN-PT single

crystals is similar to that seen in typical isotropic materials, as illustrated in Figure

1.10. The exception is the fringes resulting from Hertzian contact loading of

three glass rods, which show a two-lobed fringe pattern. This will be further

discussed in Chapter 3.

Page 31: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

17

Figure 1.10 (a) Three-point bending of unpoled PMN-29%PT; (b) Three-point bending image

of isotropic materials [39].

1.3.3 Deflection versus fringe order

Stress vs. Fringe Order

0

10

20

30

40

50

60

70

0 5 10 15 20 25 30

Fringe Order

Def

lect

ion

( μm

)

Figure 1.11 Deflection versus fringe order

In preliminary three-point bending experiments, we purposely put a thin glass

plate (a cover slip) on the glass rods, to obtain an edge to be used as a reference

to measure the deflection during the bending process, as shown in Figure 1.7.

Deflection versus fringe order was plotted in Figure 1.11. It is obvious that

deflection is linearly proportional to the fringe order. This implies for unpoled

PMN-29%PT single crystals, the mechanical properties characterized by the

Page 32: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

18

deflection are linearly proportional to the optical properties which control fringe

order. If the samples were isotropic and homogeneous in their optical properties

and mechanical properties, then the birefringence would be proportional to strain.

1.3.4 Summary

Through preliminary three-point bending experiments, isoclinic fringes were

observed; fringe patterns were also comparable to those of typical isotropic

materials. This means the photoelastic technique is useful and can be further

used to study internal stresses of unpoled PMN-29%PT single crystals.

However, loading force was unknown so that quantitative calculations could not

be performed; fringe patterns were observed only qualitatively while the loading

was increased. Because the elastic constants of unpoled PMN-29%PT single

crystals are also unknown, there was no way to analyze the internal stress. The

only quantitative result directly obtained from preliminary three-point bending is

the linear relationship between deflection and the fringe order. It was necessary

to design a new device which provides the same function while also allowing a

known force to be applied. To solve this problem, a BimbaTM 5/16’’ bore air

cylinder is used to design an in situ loading frame as shown in Figure 1.12.

Details of calibration of the loading system are provided in Appendix III. The

calibration result is that the loading force obtained from reading of the pressure

gauge is within 2.5% of the applied value. ±

In the following three chapters, birefringence response and internal stresses of

unpoled PMN-29%PT single crystals under mechanical loading are studied for the

first time using photoelastic techniques in a series of sequential experiments

comprising: three-point bending experiments, four-point bending experiments,

and Hertzian contact loading experiments. In the three-point bending

experiments, the numerical value of the stress-optical coefficient of PMN-PT was

Page 33: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

19

first estimated. The apparent Young’s modulus along <100> direction of unpoled

PMN-PT single crystals was calculated. In Hertzian contact loading experiments,

orientation dependences of fringe patterns were observed, showing the

anisotropic properties of unpoled PMN-PT single crystals. ANSYS® simulations

of piezoelectric single crystals were performed, verifying that the anisotropic

elastic properties indeed cause the orientation dependence of fringe patterns that

were observed. The results were published in two papers, references [58] and

[59] respectively. To further examine the variations of stress-optical coefficients

with incremental mechanical stresses, four-point bending experiments were

performed. A paper has recently been submitted to report the results. Finally,

electric field loading experiments were performed; the results of which are

reported in Chapter 5.

Figure 1.12 3D CAD model of loading frame. BimbaTM cylinder is mounted through a

hole in the aluminum frame.

Page 34: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

20

1.4 References 1. J. Curie, and P. Curie, “Development by pressure of polar electricity in

hemihedral crystals with inclined faces”, Bull. Soc. Min. de France, 3, 90-93, 1880.

2. S. Katzir, “Measuring constants of nature: confirmation and determination

in piezoelectricity”, Studies in History and philosophy of modern physics, 34, 579-606, 2003.

3. G. Lippmann, “Principle de la conservation de l'électricité”, Ann. Chim.

Phys. Ser. 5, 24, 145-178, 1881.

4. W. Voigt, Lehrbuch der Kristallphysik, B. G. Teubner, Leipzig and Berlin, 1928.

5. K. Uchino, and J. R. Giniewicz, Micromechatronics, Marcel Dekker, Inc.,

2003.

6. J. Kuwata, K. Uchino, S. Nomura, “Phase transition in the Pb(Zn1/3Nb2/3)O3-PbTiO3 system”, Ferroelectrics, 37, 579-582, 1981.

7. S.-E. Park and T. R. Shrout, “Ultrahigh strain and piezoelectric behavior in

relaxor based ferroelectric single crystals”, J. Appl. Phys. 82, No. 4, 1804, 1997.

8. J. Zhao, Q. M. Zhang, N. Kim and T. Shrout, “Electromechanical

properties of relaxor ferroelectric lead magnesium niobate-lead titanate ceramics”, Jpn. J. Appl. Phys, 34, 5658, 1995.

9. Z.-G. Ye, B. Noheda, M. Dong, D. Cox and G. Shirane, “Monoclinic phase

in the relaxor-Based piezoelectric/ferroelectric Pb(Mg1/3Nb2/3)O3-PbTiO3 system”, Phys. Rev. B, 64, 184114, 2001.

10. C. S. Tu, F.-T. Wang, R. R. Chien, B. Hugo Schmidt and G. F. Tuthill,

“Electric-field effects of dielectric and optical properties in Pb(Mg1/3Nb2/3)0.65Ti0.35O3 crystal”, J. Appl. Phys. 97, No. 6, 064112, 2005.

Page 35: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

21

11. D. Viehland, and J. Powers, "Electromechanical Coupling Coefficient of <001>-Oriented Pb(Mg1/3Nb2/3)O3-PbTiO3 Crystals: Stress and Temperature Independence", Appl. Phys. Lett. 78, 3112, 2001.

12. S.-E. Park and T. R. Shrout, "Characteristics of relaxor-based

piezoelectric single crystals for ultrasonic transducers", IEEE Trans. Ultrason. Ferroelectr. Freq. Control. 44, 1140, 1997.

13. X. Wan, H. Xu, T. He, D. Lin, and H. Luo, "Optical properties of tetragonal

Pb(Mg1/3Nb2/3) 0.62Ti0.38O3 single crystal”, J. Appl. Phys. 93, 4766, 2003.

14. Y. Barad, Y. Lu, Z.-Y. Cheng, S.-E. Park, and Q. M. Zhang, “Composition, temperature, and crystal orientation dependence of the linear electro-optic properties of Pb(Zn1/3Nb2/3)O3-PbTiO3 single crystals”, Appl. Phys. Lett. 77, 1247, 2000.

15. H. Fu, and R. Cohen, “Polarization rotation mechanism for ultrahigh

electromechanical response in single-crystal piezoelectrics”, Nature 403, 281, 2000.

16. H. Y. Wang and R. N. Singh, “Crack propagation in piezoelectric ceramics:

effects of applied electric fields”, J. Appl. Phys. 81, No. 11, 7471, 1997. 17. S. Kumar and R. N. Singh, “Crack propagation in piezoelectric materials

under combined mechanical and electrical loadings”, Acta Mater., 44, No. 1, 173, 1996.

18. Z. Suo, C. M. Kuo, D. M. Barnett, and J. R. Willis, “Fracture mechanics for

piezoelectric ceramics”, J. Mech. Phys Solid, 40, No. 4, 739, 1992.

19. T. Y. Zhang and J. E. Hack, “Mode-III cracks in piezoelectric materials”, J. Appl. Phys. 71, No. 12, 5865, 1992.

20. S. B. Park and C. T. Sun, “Effect of electric field on fracture of

piezoelectric ceramics”, International Journal of Fracture, 70, 203, 1995. 21. Y. E. Pak, “Crack extension force in a piezoelectric material”, J. Appl.

Mech. 57, 647, 1990.

Page 36: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

22

22. F. X. Li, S. Li and D. N. Fang, “Domain switching in ferroelectric single crystal/ceramics under electromechanical loading”, Materials Science & Engineering B, 120, 119-124, 2005.

23. Z. K. Xu, "In situ tem study of electric field-induced microcracking in

piezoelectric single crystals", Materials Science & Engineering B, 99, 106-111, 2003.

24. C. S. Lynch, W. Yang, L. Collier, Z. Suo and R. M. McMeeking, “Electrical

field induced cracking in ferroelectric ceramics”, Ferroelectrics, 166, 11-30, 1995.

25. H. Makino and N. Kamiya, “Effects of DC electric field on mechanical

properties of piezoelectric ceramics”, Jpn. J. Appl. Phys, 33, 5323, 1994.

26. H. C. Cao and A. G. Evans, “Electric-field-induced fatigue crack growth in piezoelectrics”, J. Am. Ceram. Soc, 77, No. 7, 1783, 1994.

27. X. M. Wan, J. Wang, H. L. W. Chan, C. L. Choy, H. S. Luo and Z. W. Yin,

“Growth and optical properties of 0.62 Pb(Mg1/3Nb2/3)O3-0.38PbTiO3 single crystals by a modified Bridgman technique”, Journal of Crystal Growth, 263, 251, 2004.

28. X. J. Zheng, J. Y. Li and Y. C. Zhou, “X-ray diffraction measurement of

residual stress in PZT thin films prepared by pulsed laser deposition”, Acta Materialia, 52, 3313-3322, 2004.

29. Z. L. Yan, X. Yao and L. Y. Zhang, “Analysis of internal-stress-induced

phase transition by thermal treatment”, Journal of Ceramics International, 30, 1423, 2004.

30. S. W. Choi, T. R. Shrout, S. J. Jang and A. S. Bhalla, “Morphotropic phase

boundary in Pb (Mg1/3Nb2/3) O3-PbTiO3 system”, Mater. Lett., 8, 253-255, 1989.

31. X. M. Wan, X. Y. Zhao, H. L. W. Chan, C. L. Choy and H. S. Luo, “Crystal

orientation dependence of the optical bandgap of (1-x) Pb(Mg1/3Nb2/3) O3-x PbTiO3 single crystals”, Materials chemistry and physics, 92, 123-127, 2005.

Page 37: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

23

32. C. S. Tu, R. R. Chien, F. T. Wang, V. H. Schmidt and P. Han, “Phase stability after an electric-field poling in Pb(Mg1/3Nb2/3)1-xTixO3 crystals”, Physical review B, 70, 220103(R), 2004.

33. Y. Guo, H. Luo, D. Ling, H. Xu, T. He, and Z. Yin, “The phase transition

sequence and the location of the morphotropic phase boundary region in (1-x)Pb(Mg1/3Nb2/3)O3-xPbTiO3 single Crystal”, J. Phys.: Condensed Matter. 15, 77-82, 2003.

34. F. Bai, J. Li, and D. Viehland, "Domain hierarchy in annealed

(001)-oriented Pb(Mg1/3Nb2/3)O3-x%PbTiO3 single crystals", Appl. Phys. Letters, 85, No. 12, 2313-2315, 2004.

35. C. M. Landis, “Non-linear constitutive modeling of ferroelectrics”, Current

opinion in Solid State and Materials Science, 8, 59-69, 2004. 36. H. Berger, U. Gabbert, H. Koppe, R. R.-Ramos, J. B.-Castillero, R.

G.-Diaz, J. A. Otero and G. A. Maugin, “Finite element and asymptotic homogenization methods applied to smart composite materials”, Computational Mechanics, 33, 61-67, 2003.

37. J. W. Dally and W. F. Riley, Experimental Stress Analysis, McGraw-Hill,

Inc., 1991. 38. K. Ramesh, Digital Photoelasticity Advanced Techniques and

Applications, Springer, 2000.

39. M. M. Frocht, Photoelasticity, VI, New York, John Wiley & Sons, Inc. 1941.

40. T. S. Narasimhamurty, Photoelastic and Electro-Optic Properties of

Crystals, Plenum Press, 1981. 41. J. Beynon, Introductory University Optics, Prentice Hall, 1996.

42. F. L. Pedrotti, S.J., L. S. Pedrotti, Introduction to Optics, Prentice Hall,

1993. 43. H. C. Liang, Y. X. Pan, S. Zhao, G. M. Qin and K. K. Chin,

“Two-dimensional state of stress in a silicon wafer”, J. Appl. Phys. 71, No. 6, 2863-2870, 1991.

Page 38: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

24

44. M. Lebeau, G. Majni, N. Paone and D. Rinaldi, “Photoelasticity for the investigation of internal stress in bgo scintillating crystals”, Nuclear Instruments & Methods in Physics Research A, 397, 317-322, 1997.

45. K. Higashida, M. Tanaka, E. Matsunaga and H. Hayashi, “Crack tip stress

fields revealed by infrared photoelasticity in silicon crystals”, Materials Science & Engineering A, 387, 377-380, 2004.

46. Y. Lu, Z. Y. Cheng, Y. Barad and Q. M. Zhang, “Photoelastic effects in

tetragonal Pb(Mg1/3Nb2/3)O3-x%PbTiO3 single crystals near the morphotropic phase boundary”, J. Appl. Phys. 89, No. 9, 5075-5078, 2001.

47. X. M. Wan, H. L. W. Chan and C. L. Choy, “Optical properties of

(1-x)Pb(Mg1/3Nb2/3)O3-xPbTiO3 single crystals studied by spectroscopic ellipsometry”, J. Appl. Phys. 96, No. 3, 1387-1391, 2004.

48. X. M. Wan, H. Q. Xu, T. H. He, D. Lin and H. S. Luo, “Optical properties of

tetragonal Pb(Mg1/3Nb2/3)0.62Ti0.38O3 single crystal”, J. Appl. Phys. 93, No. 8, 4766-4768, 2003.

49. Y. Barad, Y. Lu, Z. Y. Cheng, S. E. Park and Q. M. Zhang, “Composition,

temperature, and crystal orientation dependence of the linear electro-optic properties of Pb(Zn1/3Nb2/3)O3-PbTiO3 single crystals”, Applied Physics Letters, 77, No. 9, 1247-1249, 2000.

50. X. M. Wan, D.Y. Wang, X. Y. Zhao, H. S. Luo, H.L.W. Chan and C.L. Choy,

“Electro-optic characterization of tetragonal (1−x)Pb(Mg1/3Nb2/3)O3–xPbTiO3 single crystals by a modified sénarmont setup”, Solid State Communications, 134, No. 8 , 547-551, 2005.

51. C. S. Lynch, W. Yang, L. Collier, Z. Suo and R. M. McMeeking, “Electric

Field Induced cracking in ferroelectric ceramics”, Ferroelectrics, 166, 11-30, 1995.

52. C. S. Lynch, L. Chen, Z. Suo, R. M. McMeeking and W. Yang,

“Crack-growth in ferroelectric ceramics driven by cyclic polarization switching”, Journal of intelligent material systems and structures, 6, 191-198, 1995.

Page 39: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

25

53. Z. R. Li, Z. Xu, X. Yao and Z.-Y. Cheng, “Phase transition and phase stability in [110]-, [001]-, and [111]- oriented 0.68Pb(Mg1/3Nb2/3)-0.32Ti0.38O3 single crystal under electric field”, J. Appl. Phys. 104, 024112, 2008.

54. Z. K. Xu, “In situ TEM study of electric field-induced microcracking in

piezoelectric single crystals”, Materials Science and Engineering B, 99, 106-101, 2003.

55. E. T. Keve and K. L. Bye, “Phase identification and domain structure in

PLZT ceramics”, J. Appl. Phys. 46, 87, 1975. 56. F. X. Li, Shang Li, and D. N. Fang, “Domain switching in ferroelectric

single crystal/ceramics under electromechanical loading”, Materials Science and Engineering B, 120, 119-124, 2005.

57. R. Zhang, B. Jiang and W. W. Cao, “Elastic, piezoelectric, and dielectric

properties of multidomain 0.67Pb(Mg1/3Nb2/3)1-0.33TixO3 single crystals”, J. Appl. Phys. 90, 3471-3475, 2001.

58. N. Di, and D. J. Quesnel, “Photoelastic effects in

Pb(Mg1/3Nb2/3)O-29%PbTiO3 single crystals investigated by three-point bending technique”, J. Appl. Phys. 101, 043522, 2007.

59. N. Di, J. C. Harker and D. J. Quesnel, “Photoelastic effects in

Pb(Mg1/3Nb2/3)O-29%PbTiO3 single crystals investigated by Hertzian contact experiments”, J. Appl. Phys. 103, 053518, 2008.

Page 40: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

26

2 Photoelastic study using three-point bending technique

Abstract: Photoelastic effects in an unpoled PMN-29%PT single crystal

beam have been investigated using three-point bending experiments. A linear

relationship between the applied load and the measured displacement was

observed up to a proportional limit of ~30 MPa. Beyond this proportional limit,

yielding was observed. Samples were loaded as high as 77 MPa without fracture.

Young's modulus Y<001> ~1.9X1010 N/m2 was determined directly from the initially

linear region using beam theory. The photoelastic fringe order versus fiber

stress plot also displays an initially linear region up to a proportional limit of ~20

MPa, suggesting that optical measurements are a more sensitive measure of the

onset of microplasticity than mechanical measurements. Residual photoelastic

fringes associated with yielding were completely removable by annealing above

the Curie temperature, implying that plastic deformation occurs by reversible

processes such as domain switching and phase transformation. The

stress-optical coefficient for unpoled PMN-29%PT determined from the initially

linear region of the fringe order versus fiber stress curve is 104X10-12 Pa-1. This

value is large and comparable with the stress-optical coefficient of polycarbonate,

making unpoled PMN-29%PT single crystal a good candidate for optical stress

sensors and acousto-optic modulators.

2.1 Introduction

Relaxor ferroelectric single crystals exhibit ultrahigh dielectric and

piezoelectric properties compared with conventional piezoelectric ceramics.

Materials such as PMN-PT single crystals have become a next generation of

piezoelectrics that have attracted constant attention in recent years [1-5]. These

materials are finding wide-ranging applications in medical imaging, active noise

Page 41: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

27

suppression, and acoustic signature analysis. Residual stresses and internal

stresses in PMN-PT single crystals, however, can reduce the performance of

devices and lead to the initiation of cracks.

Residual stresses are induced by inhomogeneous strain. Inhomogeneous

strain may be produced by thermal gradients during crystal growth [6-7], by phase

transitions during cooling [6-8], and by mechanical cutting and finishing

operations during device fabrication [9]. When the size scale of the residual

stress distribution approaches the size scale of the microstructure, residual

stresses are often referred to as internal stresses or microstresses. Clearly, the

presence of a stress distribution within a component will influence its response to

applied loadings. To better understand and control stresses in relaxor

ferroelectrics devices, it is necessary to monitor internal stresses and residual

stresses.

Photoelasticity is an efficient and effective method to measure the residual

stresses and applied stresses in many transparent materials. It offers both

quantitative determination and qualitative observation of the stress distribution in

a sample [9-12]. Simply by examining a sample between crossed polarizers in

either the loaded or unloaded state, we can observe the influence of stress as a

result changes in optical birefringence. High-order pastel colors from the

Michel-Levy interference color chart, such as magenta and green, in combination

with closely spaced fringes, imply high stress levels and high stress gradients.

Quantitative evaluation of stress level requires that we measure the retardation

caused by stress and relate this to the stress-optical properties of the material

and the length of the optical path.

The stress-optic law of photoelasticity states that if there is a difference in

principal stresses along two perpendicular directions in an otherwise optically

isotropic material, the refractive index in these directions is different.

Page 42: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

28

Fundamentally, this effect arises from a change in the spacing between the atoms

due to strains induced by the principal stresses. The difference in refractive

index is an induced birefringence which is proportional to the difference in

principal stresses. The maximum and minimum refractive index directions are

aligned with the principal stresses.

For a two-dimensional plane stress state, the stress-optic law for an isotropic

material can be expressed as [12, 13]:

)( 21 σσλ

−= Ctn . (2-1)

Here 1σ and 2σ are the maximum and minimum principal stresses, is the

fringe order, and is the sample's thickness along the optical path.

n

t λ is the

wavelength of the incident light and is a constant known as the stress-optical

coefficient. From Eq. (2-1), we have

C

Ct

n22

21 λσσ=

− , (2-2)

where both sides are divided by 2 to produce the form of the maximum shear

stress,

Ct

n2max

λτ = (2-3)

Once the stress-optical coefficient C is known for a given material, it can be

used to quantitatively evaluate maxτ for a fringe pattern, provided we know the

fringe order. Photoelastic fringe patterns suitable for stress analysis are easily

recorded with the use of a circular polariscope and a monochromatic filter. The

fringe order may be found by locating a zero-order fringe and counting. Zero order

fringes in bending samples occur along the neutral axis. More complete details

of photoelastic methods for isotropic materials can be found in references [12,

13].

The purpose of this chapter is to explore photoelastic techniques for the

Page 43: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

29

investigation of stress distributions in unpoled PMN-29%PT single crystals. We

present the results of investigations performed using three-point bending

experiments. While the optical properties of unpoled PMN-29%PT are not

isotropic, fringe patterns are comparable with those typical of isotropic materials.

Experimental results show that there exists a linear relationship between loading

force and displacement and between fringe order and fiber stress within a

proportional limit. Beyond the proportional limit, yielding takes place. Yielding is

interpreted as stress-induced domain switching. Residual stress remaining after

unloading can be removed by annealing above the Curie temperature suggesting

that these switches are reversible. The linear relationships observed suggests

that photoelastic methods can be used more generally for these materials. The

use of optical techniques to measure stresses in next generation piezoelectrics

will be an important quality assurance tool to produce robust and reliable devices.

2.2 Experimental procedure

An in situ loading frame built to perform photoelastic measurements on

small-size beams is shown in Figure 2.1. Figure 2.1(a) illustrates the loading

frame below the objective of a ZeissTM microscope configured as a circular

polariscope while Figure 2.1(b) provides a top view of the three-point bending

set-up. Mechanical loading is applied using a BimbaTM pneumatic cylinder

shown in Figure 2.1(a), and three 1mm radius glass rods illustrated in Figure

2.1(b). The in situ load frame was calibrated so that the applied force could be

obtained directly from the reading of a pressure gauge within 2.5% accuracy.

The loading frame allows a 10X objective lens to be used to make deflection

measurements of the beam during bending.

±

A schematic of the three-point bending loading system is presented in Figure

2.2. Principal faces of the beam are parallel to the (100), (010), and (001)

Page 44: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

30

planes, respectively. The thickness and height are approximately equal at

t=1.06 mm and h=1.07 mm, enabling the beam to be bent in either direction by

changing the direction of the applied load P. An experiment in which P is aligned

with [100] is called [100] bending as shown in Figure 2.2. The small size of the

experimental set-up means that the exact placement of the loading rods will vary

from one run to the next. Specific values of the overall span length and the

numerical values of a and b were measured directly from micrographs taken

using a 5X objective lens.

Figure 2.1 (a) in situ loading frame working under microscope and (b) in situ loading

frame with three-point bending set-up as indicated by the arrow.

Figure 2.2 three-point bending schematic. P is the loading force, c and t are the

compression and tension fiber stress. and are the reaction loads. 1R 2R

Page 45: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

31

The unpoled PMN-29%PT single crystal beam (H.C. Materials Corporation)

was polished using an Allied High Tech Multi-PrepTM polishing system following

the rule of threes. Fixed abrasive diamond films in progressively finer sizes,

each removing a thickness of three times the diameter of the previous abrasive,

were followed by a final polish using 0.05 μm colloidal silica on each of the four

major faces. Sufficient material was removed between each abrasive step so

that no damaged material from the previous abrasive remained after each

polishing step. The final surfaces were of optical quality.

Figure 2.3 Unpoled PMN-29%PT single crystal beam viewed in crossed polars: (a) (100) face, as-received; (b) (100) face, after annealing; and (c) (010) face, after annealing.

Polarizer and analyzer are horizontal and vertical, respectively. Strong colors in (a) indicate regions of net birefringent retardation.

Residual stresses, net birefringence from initial domain distributions, or a

combination of these are apparent in the as-received sample as indicated by the

bright, low order color fringe patterns shown in Figure 2.3(a). For proper

photoelastic measurements, an initially stress-free sample with no net

birefringence is desired. It was found that annealing at 400 oC for 1 hour

substantially reduced the residual stresses and the apparent initial birefringence.

The observed fringe order looking through the (100) face was reduced to ~0.45

for the white regions [13] and 0 for black regions, as shown in Figure 2.3(b). The

fringe order of the (010) face shown in Figure 2.3(c) was reduced to ~0.28, the

fringe order associated with gray color as indicated in reference [13]. It is

important to note that the entire sample does not show uniform extinction as

would be the case for an isotropic material. Different faces show different

Page 46: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

32

extinction levels at nearly equal thicknesses implying the crystal is optically

anisotropic, perhaps a result of crystal growth [14].

Photoelastic experiments were performed using a ZeissTM optical microscope

configured as a dark field circular polariscope. This configuration eliminates

isoclinic fringes and thus produces only isochromatic fringes. Isochromatic

fringes depend only on the magnitude of the principal stress differences at each

point, greatly simplifying the analysis. A monochromatic green filter with

wavelength ~535 nm was used to record photographs for the evaluation of the

fringe order. Fringe order was counted from 5X objective lens images at the

point of maximum tensile stress on the free surface of the sample (point A) as

shown in Figure 2.2. Fractional fringes were estimated to the nearest 0.3 using

graphical intensity information from image analysis software. The displacement

of the sample relative to a fixed reference was measured directly from 10X

objective lens images at the same location A. Bending was performed on both

(100) and (010) faces for comparison, even though crystallographic symmetry

consideration suggests the results should be identical. Between each

experiment, the beam was annealed to remove all residual fringes, allowing the

same beam to be used again and again.

2.3 Results and discussion

2.3.1 Fringe patterns

Figure 2.4 shows photoelastic images obtained at different load levels for [100]

bending. Figure 2.4(a) shows the unloaded state, Figure 2.4(b) and Figure 2.4(c)

show examples of well-defined fringes obtained under increasing load, and Figure

2.4(d) shows residual fringes when the load is removed. These fringe patterns,

obtained using monochromatic green illumination under circular polariscope, are

comparable to those typical of isotropic materials [12]. During the experiments, it

Page 47: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

33

could be seen that fringes originated at the central portions of the top and bottom

edges and moved inward toward the neutral axis with increasing load. The

fringes formed in pairs, producing the relatively symmetric patterns shown in

these figures. Fringes of increasing fringe order distribute uniformly along the

height of the beam, corresponding to a linear variation of the principal stress along

the thickness, as shown in Figure 2.2. The upper half of the beam is in

compression, while the lower half is in tension. The zero-order fringe always lies

along the neutral axis which is stress free according to elementary beam theory.

We can see clearly from Figure 2.4(b) that the observed fringe pattern is

asymmetric at low loads: the zero- order fringe exists only on the right portion of

the neutral axis, corresponding to the region showing exactly zero fringe order in

the unloaded state. Thus the fringe patterns we observe from applied loading

are qualitatively consistent with fringe patterns that would have been obtained

from an isotropic sample. Note that in three-point bending experiments, the

principal stress xσ on the outer surface at point A is the maximum tension stress

maxσ , known as the fiber stress, and yσ equals zero as a result of the free

surface boundary condition. Thus, the fringe order is directly proportional to fiber

stress xσ according to Eq. (2-1).

Figure 2.4 (a) Initial fringe pattern showing 0.5 fringe at point A on the free surface opposite

the loading point. (b) Second-order fringes at A, (c) Sixth-order fringes at A, and (d) first-order fringe remaining at A after the load is released.

Page 48: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

34

2.3.2 Loading force versus deflection

Figure 2.5 illustrates the applied compressive load as a function of the

displacement measured at the center of the beam. Two data sets for [100]

bending and one data set for [010] bending are shown. The results are highly

repeatable, independent of the orientation of the bending, indicating that the

mechanical properties are the same for both orientations as we would expect

given the nearly identical dimensions of the samples. The force depends linearly

on the displacement over the initial portion up to a proportional limit which implies

that the loading induces elastic deformation in this regime. The proportional limit

is approximately 2 N, which corresponds to a fiber stress of 25 - 30 MPa, for the

sample geometries and span lengths used in these tests. Beyond the

proportional limit there is yielding after which the data appears to continue upward

with a reduced slope.

Figure 2.5 Force versus deflection during increasing load for three experimental runs.

Page 49: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

35

Figure 2.6 Force versus deflection with polynomial fit curve.

The slope in Figure 2.5 was obtained in terms of least-square fit based on

data below the proportional limit. The correlation coefficient is 0.988. As shown

in Figure 2.6, with a fourth-order polynomial fit for all the data, the resulting curve

provides a simple and reliable way to determine the proportional limit rather than a

simple visual inspection. The yield stress of 25 - 30 MPa is comparable to 20

MPa reported by Viehland for PMN-32%PT single crystals [15].

We interpret this yielding effect as the result of stress induced domain

switching that occurs throughout the sample, spreading from the high stress

surfaces as a result of the stress gradients. Plastic deformation represents a

reorientation of the polarization of the nanodomains distributed throughout this

otherwise unpoled sample. Essentially, the stress is changing the population of

the dipoles of the eight possible orientations of the unpoled sample [16], leading

to strain. At these modest stresses, the sample does not undergo large scale

mechanical poling or stress induced phase transformations that are possible in

this system. We can say this because the yielding was not accompanied by the

massive changes in optical properties expected from phase transitions. Rather,

Page 50: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

36

the development of the photoelastic fringe patterns proceeded smoothly through

the yielding region as described below. Thus, it seems appropriate to assign the

yielding phenomena shown in Figure 2.5 to distributed domain reorientation (i.e.

domain switching). By this process, the yielded portions of the sample have

been mechanically poled.

It is also possible that the apparent yielding we see is, in part, the result of

concentrated strains that occur at the loading points as a result of the Hertzian

contact stresses. These large contact stresses could be sufficient to trigger

stress induced phase transformations in the neighborhood at the loading points.

More work is needed to assess the relative importance of this contribution.

2.3.3 Stress-optical coefficient

From elementary beam theory, the fiber stress during three point bending is

expressed as:

2maxmax

max6LthPab

IyM

=−=σ (2-4)

Here is maximum bending moment at the location of point A and

is area moment of inertia of the beam cross section. P is the

loading force, while a , , , and represent dimensions as shown in Figure 2.2.

maxM

12/3thI =

b h t

L is the span length, namely ( +b ). is negative with a numerical value

equal to half the height h at location A. The stress

a maxy

maxσ calculated from Eq. (2-4)

represents the principal stress 1σ at point A since 2σ is zero there.

Page 51: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

37

Figure 2.7 Fringe order versus fiber stress. The stress-optical coefficient is calculated

from the slope of the proportional region.

Figure 2.7 displays the fringe order versus the fiber stress. The slope was

obtained in terms of a least-square fit based on data below the yield stress

obtained from Figure 2.6. The correlation coefficient of the linear fit is 0.9774.

Figure 2.7 shows the same trends between the fringe order and the fiber stress as

that between the loading force and the deflection, only the proportional limit

occurs earlier in the data set. The fringe order, characterizing the optical

properties, is a more sensitive indicator of the deviation from linearity than the

displacement. The proportional limit in fringe order versus stress is 20 MPa

compared to 25-30 MPa discussed earlier for load versus displacement.

Here again, the data is highly repeatable, particularly the data from the same

type of bending. Difference between the two orientations of the bending may be

attributed to the initially different birefringence at zero applied load. Namely, the

optical properties are different for [100] and [010] experiments because the initial

domain distributions are different for these two cases. At stresses below the

Page 52: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

38

proportional limit, fringe order is linear with the fiber stress showing a photoelastic

effect. The explanation for the ”optical yielding” should be the same as for the

mechanical yielding. Namely, it should be the result of distributed domain

switching and possibly locally phase transformation at the loading points.

From the slope of the linear region in Figure 2.7, 0.2077 MPa-1, we calculate

the stress-optical coefficient using Eq. (2-1):

MPa

12077.0=λtC (2-5)

Here is the thickness of the beam through which the light passes, 1.06 mm

to 1.07 mm, depending on the orientation.

t

λ is the wavelength of the green

filter, ~535 nm. The stress-optical coefficient C is calculated as 104X10-12 Pa-1.

Stated another way, approximately 2.4 MPa of shear stress (4.8 MPa of principal

stress difference) will induce one order of fringe in the nominally 1 mm thick

samples reported here. Fringes represent regions of constant shear stress for

each fringe order whose values can be determined using Eq. (2-3).

2.3.4 Young’s modulus

From beam bending theory, the load P and the displacement Aδ of the point

A are related by

)(6

222 baLabLEIP A

−−=

δ (2-6)

where E is Young's modulus along the [001] direction. From the slope of

Figure 2.5 below the proportional limit, the Young's modulus is 1.8-1.9X1010 N/m2.

Similar results were obtained by Viehland and Li [15], where the Young's modulus

for PMN-30%PT single crystal along <001> is reported to be 2X1010 N/m2, much

lower than the ~15X1010 N/m2 value reported for the <111> direction or the

~7.5X1010 N/m2 value reported for polycrystalline material of the same chemistry.

Page 53: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

39

After bending, it was usual to observe residual fringes remaining in the beam

after release of the loading force, as shown in Figure 2.4(d). The higher the

loading force applied, the higher the fringe order remaining after unloading. It is

hypothesized that domain switching and perhaps local phase transformations lock

the stresses inside the beam by producing inhomogeneous strains which are

larger in those regions further from the neutral axis.

Experimentally, it was found that annealing can remove the residual fringes.

Annealing at 400 oC for one hour was sufficient to remove all residual fringes and

restore the initial fringe pattern. This means that any stress induced domain

switching or possible phase transformations caused by the bending experiments

are reversible.

2.4 Summary

Three-point bending experiments were performed on an unpoled

PMN-29%PT single crystal. The crystal was restored to its initial condition

between bending experiments by annealing for one hour at 400 oC. The

relationship between the load and the displacement and between the fringe order

and the fiber stress is linear below a proportional limit. Beyond that proportional

limit, stress induced domain switching (mechanical poling) can explain the

apparent yielding. The stress-optical coefficient of the unpoled PMN-29%PT is

approximately 104X 10-12 Pa-1, higher than the values for materials used in

photoelastic stress analysis such as polycarbonate, 82X 10-12 Pa-1 [17]. Young's

modulus determined from the present experiment is 1.8 - 1.9X 1010 N/m2. Since

annealing removes all the residual fringes, the inhomogeneously distributed

domain switching responsible for the residual fringes must be reversible.

Page 54: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

40

2.5 References 1. S.-E. Park and T. R. Shrout, “Ultrahigh strain and piezoelectric behavior in

relaxor based ferroelectric single crystals”, J. Appl. Phys. 82, No. 4, 1804, 1997.

2. Y. Yamashita, “Large electromechanical coupling factors in perovskite

binary material system”, Jpn. J. Appl. Phys. 33, 5328, 1994. 3. Z.-G. Ye, B. Noheda, M. Dong, D. Cox and G. Shirane, “Monoclinic phase

in the relaxor-Based piezoelectric/ferroelectric Pb(Mg1/3Nb2/3)O3-PbTiO3 system”, Phys. Rev. B, 64, 184114, 2001.

4. C. S. Tu, F.-T. Wang, R. R. Chien, B. Hugo Schmidt and G. F. Tuthill,

“Electric-field effects of dielectric and optical properties in Pb(Mg1/3Nb2/3)0.65Ti0.35O3 crystal”, J. Appl. Phys. 97, No. 6, 064112, 2005.

5. X. Zhao, B. Fang, H. Cao, Y. Guo, and H Luo, “Dielectric and piezoelectric

performance of PMN-PT single crystals with compositions around the MPB: influence of composition, poling field and crystal orientation”, Materials Science and Engineering: B, 96, 254-262, 2002.

6. X. M. Wan, J. Wang, H. L. W. Chan, C. L. Choy, H. S. Luo and Z. W. Yin,

“Growth and optical properties of 0.62 Pb(Mg1/3Nb2/3)O3-0.38PbTiO3 single crystals by a modified Bridgman technique”, Journal of Crystal Growth, 263, 251, 2004.

7. X. J. Zheng, J. Y. Li and Y. C. Zhou, “X-ray diffraction measurement of

residual stress in PZT thin films prepared by pulsed laser deposition”, Acta Materialia, 52, 3313–3322, 2004.

8. Z. L. Yan, X. Yao and L. Y. Zhang, “Analysis of internal-stress-induced

phase transition by thermal treatment”, Journal of Ceramics International, 30, 1423, 2004.

9. H. C. Liang, Y. X. Pan, S. Zhao, G. M. Qin and K. K. Chin,

“Two-dimensional state of stress in a silicon wafer”, J. Appl. Phys. 71, No. 6, 2863-2870, 1992.

Page 55: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

41

10. M. Lebeau, G. Majni, N. Paone and D. Rinaldi, “Photoelasticity for the investigation of internal stress in BGO scintillating crystals”, Nuclear Instruments & Methods in Physics Research A, 397, 317-322, 1997.

11. K. Higashida, M. Tanaka, E. Matsunaga and H. Hayashi, “Crack tip stress

fields revealed by infrared photoelasticity in silicon crystals”, Materials Science & Engineering A, 387-389, 377-380, 2004.

12. M. M. Frocht, Photoelasticity, VI, New York, John Wiley & Sons, Inc.

1941.

13. K. Ramesh, Digital Photoelasticity Advanced Techniques and Applications, Springer, 2000.

14. A. Sehirlioglu, D. A. Payne, and P. Han, “Thermal expansion of phase

transformations in (1-x)Pb(Mg1/3Nb2/3)O3-xPbTiO3: evidence for preferred domain alignment in one of the (001) directions for melt-grown crystals”, Phys. Rev. B 72, 214110, 2005.

15. D. Viehland and J. F. Li, “Stress-induced phase transformations in

<001>-oriented Pb(Mg1/3Nb2/3)O3–PbTiO3 Crystals: bilinear coupling of ferroelastic strain and ferroelectric polarization”, Philosophical Magazine, 83, 53-59, 2003.

16. M. Abplanalp, D. Barosva, P. Bridenbaugh, J. Erhart, J. Fousek, P. Guter,

J. Nosek, and M. Sulc, “Ferroelectric domain structures in PZN-8/%PT single crystals studied by scanning force microscopy”, Solid State Commun. 119, 7, 2001.

17. G. D. Shyu, A. I. Isayev, and C. T. Li, “Photoviscoelastic behavior of

amorphous polymers during transition from the glassy to rubbery state”, J. Polym. Sci., Part B: Polym. Phys. 39, 2252, 2001.

Page 56: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

42

3 Photoelastic study using Hertzian contact experiments

Abstract: Photoelastic effects in small single-crystal PMN-29%PT samples

were investigated by Hertzian contact experiments. The experiments were

performed on samples having various crystallographic orientations. The

resulting photoelastic fringe patterns were observed to be strongly dependent on

the orientation of the samples, showing that pseudo-cubic unpoled PMN-29%PT

single crystals have highly anisotropic elastic properties. Annealing above the

Curie temperature was found to completely remove the fringe patterns created by

the Hertzian indentation experiments. In addition, three-dimensional simulations

of the Hertzian contact experiments were performed using ANSYS®. The

simulations used cubic-form elastic constants calculated from data on poled

PMN-30%PT single crystals. The ANSYS® modeling results were comparable to

the experimentally observed fringe patterns, suggesting that the elastic properties

of pseudo-cubic unpoled PMN-PT single crystals may resemble those of

pseudo-tetragonal poled PMN-PT single crystals. This resemblance is

considered significant because of the uniqueness of the growth direction during

seeded crystal growth. ANSYS® provided a reliable method for qualitative

simulation of photoelastic effects in unpoled PMN-PT single crystals under

mechanical loading.

3.1 Introduction

After being poled in the pseudo-cubic [001] direction, domain-engineered

PMN-PT single crystals exhibit much greater electromechanical properties than

conventional piezoelectric ceramics. Because of this, bulk single crystals of

relaxor ferroelectric PMN-PT show great promise as a replacement for ceramics

in many applications, such as sensors, actuators, and motors, and they are the

Page 57: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

43

focus of intense research activities [1-5]. However, PMN-PT single crystals are

mechanically softer than PMN-PT polycrystalline ceramics [6]. Therefore

PMN-PT single crystals more easily develop residual internal stresses induced by

preparation processes. In particular, when unpoled PMN-PT single crystals are

initially machined through cutting and polishing, the resulting damage can lead to

significant cracking problems later, reducing system performance and reliability.

To better understand and control the cracking problem, photoelastic

techniques have been applied to study the internal stresses of unpoled

PMN-29%PT single crystals. In a previous article (Chapter 2), we have reported

the photoelastic study of [001]-oriented unpoled PMN-29%PT single crystals

using three-point bending techniques [7]. The fringe pattern was observed to be

comparable to that of typical isotropic materials, so it was hypothesized that

unpoled PMN-PT single crystals might possess isotropic mechanical properties.

To verify this, in the current work photoelastic techniques were applied to

study the internal stress field of unpoled PMN-29%PT single crystals during

Hertzian contact experiments. We used several samples with different

crystallographic orientations to explore whether the elastic properties were

orientation dependent. The differently oriented samples showed very different

fringe patterns, especially when loaded from different directions. For example,

when loaded in the <100> direction, the samples displayed a two-lobed fringe

pattern, but when loaded in the <110> direction, the samples displayed a

one-lobed fringe pattern. This implies that the unpoled crystals possess highly

anisotropic elastic properties.

There are two possible explanations for the anisotropy of elastic properties.

First, Sehirlioglu, et al. found that, in unpoled multidomain PMN-PT single crystals,

the thermal expansion coefficient in one of the <100> directions is slightly larger

than in the other two <100> directions [8]. They attributed this effect to the

Page 58: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

44

seeded crystal growth method, in which the growth direction exhibits unique

properties. The uniqueness of the growth direction may also explain the highly

anisotropic elastic properties of unpoled PMN-PT single crystals observed here.

Second, Zhang, et al. reported that the domain wall motion may contribute to the

effective elastic constants, especially affecting the response of PMN-PT in <110>

directions [9]. This may explain why unpoled PMN-PT single crystals show

different fringe patterns when loaded from <100> and <110> directions.

In parallel with the experiments, ANSYS® simulations were created to allow a

qualitative interpretation of the experimental fringe patterns. Using cubic-form

elastic properties calculated from pseudo-tetragonal poled PMN-30%PT single

crystals, the ANSYS® modeling results were comparable to the fringe patterns of

unpoled PMN-29%PT single crystals. This implies that ANSYS® can provide a

reliable method for qualitative simulation of unpoled PMN-PT single crystals

under mechanical loading, and that the elastic properties of unpoled PMN-PT

single crystals are close to those of poled PMN-PT crystals.

3.2 Experimental procedure

To perform the photoelastic experiments, a mini-loading frame using a

pneumatic cylinder was designed to apply the load. The mini-loading frame with

the sample inside was placed in a ZeissTM optical microscope for in situ

observation of fringes during loading. The frame was thin enough to fit under the

10X objective lens with the sample in focus. The mini-loading system was

calibrated so that the applied force could be obtained directly from the reading of

a pressure gauge, with accuracy of ±2.5%. A slice from a glass rod with 1 mm

radius was used to exert the Hertzian cylinder load. Figure 3.1 shows the top

view of the loading frame with the Hertzian contact experiment set-up. The

maximum force applied in each experiment was approximately 10 N, while

Page 59: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

45

illustrations and simulations show typical loads around 4 N.

To explore the orientation effect, three samples were prepared as shown in

Figure 3.2: One with all six faces having {100} orientation; another one with all

four side faces having {110} orientation; the third one with two side faces {100}

and the other two side faces {110}. Figure 3.3 displays models for the three

samples in a cubic coordinate system, with the two loading directions <100> and

<110> labeled a, b, c and d corresponding to each experiment performed. The

same orientations were also used for ANSYS® modeling. The samples were

rectangular plates approximately 3 mm X 4 mm and 1 mm thick. The samples

were annealed at 400 oC for one hour before doing the experiments, in order to

remove any residual stresses and to reduce the initial birefringence.

Figure 3.1 (a) Top view of the loading frame. (b) Hertzian contact experimental

set-up as indicated by the arrow.

The optical microscope was configured as a circular polariscope, so only

isochromatic fringes were recorded. Fringe patterns observed in this way are

directly representative of the principal stress difference, according to

photoelasticity theory [10]. This will be discussed in detail below. Samples were

illuminated with green light of wavelength ~535 nm, obtained using a band-pass

filter.

Page 60: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

46

Figure 3.2 Initial birefringence patterns of three samples in three different

orientations under circularly polarized illumination.

Figure 3.3 The 3 differently oriented samples relative to {001}-oriented pseudo-cubic axes. Arrows a and c represent compression along <100> direction; arrows b and

d represent compression along <110> direction.

3.3 FEM modeling methods

The ANSYS® simulation was simplified using several assumptions, based on

the experimental parameters. Phase transformations at the contact point stress

singularity were neglected. The low loading force (around 4 N) and the resulting

Page 61: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

47

low stresses made phase transformations unlikely to occur in other regions [11].

The converse piezoelectric effect was also neglected, with the assumption that

the randomly oriented micro- and nanodomains in the unpoled sample would

produce zero net electric field on the macro-scale.

Therefore, with phase transformations and electric charge generation

neglected, the unpoled PMN-PT single crystals were treated as simple crystals.

Solid 5 (coupled-field eight-node brick) elements were used to model the PMN-PT

crystal, with the piezoelectric coefficient tensor set to zero. This was for later

convenience in developing the same finite element method (FEM) simulations for

poled PMN-PT single crystals. Solid 45 (eight-node brick) elements were used

to model the glass rod. Because of the above simplifying assumptions, only

elastic constants were required as inputs for the ANSYS® simulation. Although

the elastic, piezoelectric, dielectric, and even electro-optic properties of poled

PMN-PT single crystals have been studied and reported by several researchers

[12-15]; the elastic properties of unpoled PMN-PT single crystals are seldom, if

ever, reported. One possible reason for this is that samples must be poled for

the elastic properties to be measured using resonance methods. Consequently,

material properties reported by Cao et al. [16] for <001>-poled PMN-30%PT

single crystals (see Table 3.1) will be adapted in the present work to represent the

elastic properties of unpoled material.

Table 3.1: Elastic stiffness constants of PMN-30%PT single crystals

(10

Dijc

10 N/m2). Cao et al. [16]

Dc11 Dc12 Dc13 Dc33 Dc44 Dc66

11.8 10.4 9.5 17.4 7.8 6.6

Page 62: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

48

At room temperature, the PMN-29%PT multidomain single crystals are in the

rhombohedral phase. In each lightly skewed cubic unit cell, the dipole is along

one of eight <111> directions, thus there are eight possible domains with different

spontaneous polarization within the multidomain system. After being poled

along the <100> direction, four degenerate polarization orientations remain.

Therefore, from a macroscopic view, poled PMN-PT single crystals may be

treated as pseudo-tetragonal crystals with 4mm symmetry [12-14]. Cao et al.

treated <001>-poled PMN-30%PT single crystals as pseudo-tetragonal [16] and

their elastic constants (Table 3.1) take a tetragonal form. Similarly, in this

chapter we assume that unpoled PMN-29%PT single crystals can be treated as

having a pseudo-cubic structure. This pseudo-cubic assumption has been used

previously in the literature for measurement of electro-optical properties [15].

For cubic symmetry, there are three distinct elastic constants (Table 3.2),

while for tetragonal symmetry; there are six (Table 3.1) [17]. For this

ANSYS

cubijc

tetijc

® simulation, approximate cubic-form elastic constants were created by

averaging as follows: , , and

. The resulting elastic stiffness constants, and the elastic

data (

),,( 33111111tettettetcub cccavgc = ),,( 13131212

tettettetcub cccavgc =

),,( 66444444tettettetcub cccavgc =

E ,ν ) used for soda-lime-silica glass rods [18], are listed in Table 3.2.

Table 3.2: Input parameters used in ANSYS®. The elastic stiffness

constants: (10ijc 10 N/m2). Young's modulus of glass: E (1010 N/m2). Poisson's ratio of glass: ν .

11c 12c 44c E ν

13.7 9.8 7.4 7.4 0.21

In ANSYS®, three-dimensional (3D) models were built with the following

Page 63: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

49

symmetry-based constraints: Nodes on the bottom surface (X-Z plane) were

constrained in Y, nodes on the central surface (Y-Z plane) were constrained in X,

and nodes on the middle surface (parallel to X-Y plane) were constrained in Z.

The X and Y constraints are shown in Figure 3.4. The constraints on the middle

surface were omitted for clarity. To reduce the number of elements and

calculating time, the circular glass rod was modeled as a half-circular glass rod,

with the top surface nodes coupled to behave as one node. The dimensions of

the block sample were 3x4x1, and the half glass rod had radius of 1. These

proportions corresponded 1 to 1 with the experimental samples.

The model for compression along the [001] direction on a {001}-oriented

model is shown in Figure 3.4. The other three orientations shown in Figure 3.3

were built as ANSYS® models by rotating the first model's local coordinate system

relative to the global coordinate system. With this technique, the material

property input parameters did not need to be changed.

Figure 3.4 ANSYS® model for use in computation of fringe pattern images.

Boundary conditions are shown. Contact elements are used at the interface between the Hertzian cylinder indenter and the rectangular piezocrystal. Triangle

symbols represent displacement constraints. Top arrow indicates the force applied to all of the coupled nodes (nodes on top surface).

The ANSYS® simulation output was plotted as a stress intensity contour plot,

where the “stress intensity” is defined by ANSYS® as the largest difference

Page 64: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

50

between the three principal stresses. We chose the stress intensity contour on

the front/back surface which is free of surface traction and therefore

representative of a two-dimensional (2D) stress state. For a 2D stress state, the

stress-optic law is expressed as [10]:

)( 21 σσλ

−= ChN (3-1)

where is the fringe order, h is the thickness of the material, and N λ is the

wavelength of the incident light. C is known as the relative stress-optic

coefficient. 1σ and 2σ are the maximum and minimum principal stresses.

Therefore, the stress intensity contour on the top surface should directly

correspond to the experimental fringe pattern.

3.4 Results and discussion

The value of the relative stress-optic coefficient is unknown for unpoled

PMN-PT single crystals with different orientations, and the exact value of the

elastic stiffness constants for unpoled PMN-29%PT are (as mentioned above)

also unknown. Therefore, we must perform a qualitative analysis of the

simulated and experimental results, as data required for a more precise

quantitative analysis is unavailable.

Figure 3.5 (a) Hertzian indentation along <100> direction on sample 1, and (b)

Stress intensity contour from ANSYS®.

Page 65: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

51

Figure 3.5, Figure 3.6 and Figure 3.7 display experimental fringe patterns with

the comparable ANSYS® results for the three samples, respectively. Each fringe

picture was taken under a different loading force, but simulation results all used

the same simulated loading force 4 N.

Figure 3.6 (a) Hertzian indentation along <110> direction on sample 2, and (b)

Stress intensity contour from ANSYS®.

Figure 3.7 Hertzian indentation along <100> direction on sample 3 is shown in (a);

Hertzian indentation along <110> direction is shown in (c). The initial birefringence is responsible for the asymmetric fringe in (a) and the layers along the surface in (c).

Stress intensity contour from ANSYS® are shown in (b) and (d) correspondingly.

Page 66: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

52

The pictured fringe patterns are not like the typical water-drop-shaped fringes

of isotropic materials [19]. The fringes are all different for the three samples with

two loading directions: On sample 1 (loaded in the <100> direction), the fringes

have a two-lobe shape; on sample 2 (loaded in the <110> direction), the fringes

have a narrow single-lobe shape. On sample 3, when loaded in the <100>

direction, the fringes have a wider two-lobe shape compared to sample 1; when

loaded in the <110> direction, the fringes have a wider single-lobe shape

compared to sample 2. The different fringe patterns imply that unpoled

PMN-29%PT single crystals are highly anisotropic materials.

The anisotropy ratio can be calculated as [20]:

1211

442cc

cA−

= (3-2)

An anisotropy ratio of 1 indicates a perfectly isotropic material. Using the

estimated elastic constants from Table 3.2, the anisotropy ratio is 3.79, which

indicates fairly high anisotropy.

The ANSYS® simulation results were scaled to have dimensions comparable

with the fringe pattern pictures. It is easy to see that the contours from ANSYS®

simulation strongly resemble the experimental fringe patterns. This suggests

that the estimated cubic-form elastic constants are reasonable approximations to

the true values. As the constants were calculated from poled PMN-PT single

crystal properties, one possible conclusion could be that unpoled PMN-29%PT

single crystals have tetragonal-like elastic properties, which may be explained by

the seeded crystal growth method [8].

Finally, residual fringes remained in the crystal after the loading force was

released, as shown in Figure 3.8(a). It seems obvious that plastic deformation

occurred in these samples. The deformation mechanism can be explained as

mechanical-loading-induced domain switching. After one hour annealing at

Page 67: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

53

400 oC, the residual fringes were totally removed, as shown in Figure 3.8(b).

This means that this mechanically induced domain switching can be removed by

annealing.

Figure 3.8 Residual butterfly fringes are fully annealed out at 400 oC for one hour.

3.5 Conclusions

Photoelastic effects in small PMN-29%PT block samples were investigated by

Hertzian contact experiments. It was found that unpoled PMN-29%PT single

crystals display extremely different fringe patterns under Hertzian compression for

differently oriented samples, which implies that this crystal has extremely

anisotropic elastic properties. The anisotropy ratio may be as high as 3.90,

using cubic-form elastic constants calculated from data on poled PMN-30%PT

pseudo-tetragonal single crystals. Annealing above the Curie temperature can

completely remove the residual fringe patterns caused by

mechanical-loading-induced domain switching. Also, 3D simulations of the

Hertzian contact experiments were performed using ANSYS®. Stress intensity

contours from the ANSYS® models were strikingly similar to the experimentally

Page 68: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

54

observed fringe patterns of corresponding experimental samples, showing that

the elastic properties of unpoled PMN-PT single crystals are close to the

properties of tetragonal poled PMN-PT single crystals. This similarity may be

caused by the seeded crystal growth method and the possible tetragonal

symmetry a special direction would imply. Also, ANSYS® is a reliable method for

qualitative simulation of the photoelastic response of unpoled PMN-PT single

crystals under mechanical loading.

3.6 References 1. S.-E. Park and T. R. Shrout, “Ultrahigh strain and piezoelectric behavior in relaxor based ferroelectric single crystals”, J. Appl. Phys. 82, No. 4, 1804, 1997. 2. Y. Yamashita, “Large electromechanical coupling factors in perovskite binary material system”, Jpn. J. Appl. Phys. 33, 5328, 1994.

3. S.-E. Park, P. Lopath, K. Shung, and T. R. Shrout, “Ultrasonic transducers using piezoelectric single crystal perovskites”, Ferroelectrics 2, 543, 1996.

4. S.-E. Park and T. R. Shrout, “Characteristics or relaxor-based piezoelectric single crystals for ultrasonic transducers”, IEEE Trans. UFFC 44, 1140-1147, 1997.

5. T. R. Shrout, Z. P. Chang, N. Kim, and S. Markgraf, “Dielectric behavior of single-crystals near the (1-x)Pb(Mg1/3Nb2/3)O3-xPbTiO3 morphotropic phase-boundary”, Ferroelectr. Lett., 12, 63, 1990.

6. D. Viehland and J. F. Li, “Young’s modulus and hysteretic losses of 0.7 Pb(Mg1/3Nb2/3)O3-0.3PbTi O3”, J. Appl. Phys. 94, 7719, 2003.

7. N. Di, and D. J. Quesnel, “Photoelastic effects in Pb(Mg1/3Nb2/3)O-29%PbTiO3 single crystals investigated by three-point bending technique”, J. Appl. Phys. 101, 043522, 2007.

Page 69: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

55

8. A. Sehirlioglu, D. A. Payne, and P. Han, “Thermal expansion of phase transformations in (1-x)Pb(Mg1/3Nb2/3)O3-xPbTiO3: evidence for preferred domain alignment in one of the (001) directions for melt-grown crystals”, Phys. Rev. B 72, 214110, 2005.

9. R. Zhang, W. Jiang, B. Jiang, and W. Cao, “Elastic, dielectric and piezoelectric coefficients of domain engineered 0.7Pb(Mg1/3Nb2/3)O-0.3PbTiO3 single crystal”, Fundamental Physics of Ferroelectrics 2002, 188-197, 2002.

10. K. Ramesh, Digital Photoelasticity Advanced Techniques and Applications, Springer, 2000.

11. D. Viehland and J. F. Li, “Stress-induced phase transformations in <001>-oriented Pb(Mg1/3Nb2/3)O3–PbTiO3 Crystals: bilinear coupling of ferroelastic strain and ferroelectric polarization”, Philosophical Magazine, 83, 53-59, 2003.

12. R. Zhang, B. Jiang and W. W. Cao, “Elastic, piezoelectric, and dielectric properties of multidomain 0.67Pb(Mg1/3Nb2/3)1-0.33TixO3 single crystals”, J. Appl. Phys. 90, 3471-3475, 2001.

13. R. Zhang, B. Jiang, and W. Cao, “Orientation dependence of piezoelectric properties of single domain 0.67Pb(Mg1/3Nb2/3)O3-0.3PbTiO3 crystals”, Appl. Phys. Lett. 82, 3737, 2003.

14. W. Jiang, R. Zhang, B. Jiang, and W. Cao, “Characterization of piezoelectric materials with large piezoelectric and electromechanical coupling coefficients”, Ultrasonics 41, 55-63, 2003.

15. X. Wan, H. Xu, T. He, D. Lin, and H. Luo, “Optical properties of tetragonal Pb(Mg1/3Nb2/3)O30.62-PbTiO30.38 single crystal”, J. Appl. Phys. 93, 4766, 2003.

16. H. Cao, V. H. Schmidt, R. Zhang, W. Cao, and H. Luo, “Elastic, piezoelectric, and dielectric properties of 0.58Pb(Mg1/3Nb2/3)O3 -0.42PbTiO3 single crystal”, J. Appl. Phys. 96, 549, 2004.

17. J. F. Nye, Physical Properties of Crystals, Clarendon Press, Oxford, 1985

Page 70: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

56

18. D. G. Holloway, The Physical Properties of Crystals, Oxford University Press, New York, 1985.

19. K. L. Johnson, Contact Mechanics, Cambridge University Press, 1985.

20. J. P. Hirth and J. Lothe, Theory of Dislocations, McGraw Hill, 1968.

Page 71: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

57

4 Photoelastic study using four-point bending technique

Abstract: Photoelastic effects in unpoled PMN-29%PT single crystals have

been investigated using four-point bending experiments. Photoelastic fringes in

the constant moment region were observed to be uniformly parallel to the edge of

the beam, corresponding to a state of pure bending. The fiber stress versus

fringe order plot is consistent with results reported earlier using three-point

bending experiments and in addition, the fringe-stress coefficient has been

evaluated for several load levels. Fringe-stress coefficients varied from 3.5X103

N/m to 5.5X103 N/m. From an initial maximum of 5.5X103 N/m, the fringe-stress

coefficient decreases monotonically and nearly linearly with increasing stress until

~45 MPa. At this point, the fringe-stress coefficient begins to rise. The initial

decrease of the fringe-stress coefficient may be explained by mechanical poling,

i.e. plastic yielding via domain switching and/or phase transformation. The

subsequent increase of the fringe-stress coefficient at higher loading levels can

be explained as the saturation of these mechanical poling effects. The increased

fringe-stress coefficient also correlates with an increase in the observed optical

transmittance of the crystal. This chapter will discuss implications of these

observed mechanical poling effects.

4.1 Introduction

Domain-engineered (1-x)Pb(Mg1/3Nb2/3)O3-xPbTiO3 (PMN-PT) single crystals

possess extra-high piezoelectric coefficients, electromechanical coupling factor,

and field induced strain response compared with conventional piezoelectric

ceramics [1-6]. Thus relaxor ferroelectric PMN-PT single crystals are promising

to replace ceramics in high performance applications, such as higher sensitivity

ultrasonic transducers and large strain actuators. However, PMN-PT single

Page 72: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

58

crystals are very fragile compared with ceramic materials of the same chemistry,

and more easily develop cracks. This will reduce the performance and reliability

of devices. Because cracks are initiated more easily in samples containing

residual stresses that are induced during crystal growth [7-8] and preparation

processes, it is necessary to monitor and control the residual stresses throughout

the entire preparation processes. Careful attention to cutting and polishing

processes can limit residual stresses and subsurface damage, leading to

substantial improvements in the mechanical robustness of piezoelectric single

crystals.

As is well known, photoelasticity is an effective and easily implemented

technique to measure the internal stresses in transparent materials. It offers

both quantitative determination and qualitative observation of the stress

distribution in a sample [9-11]. However, implementation of quantitative

photoelastic stress analysis requires that the fringe-stress coefficient f be

determined by direct experiment. Once this key parameter is known for a given

material, it directly links the observed fringe patterns to numerical values of the

internal stress. The fringe-stress coefficient is contained within the stress-optic

law which, for a two-dimensional stress state, can be written as:

)( 21 σσλ

−= Ctn . (4-1)

Here 1σ and 2σ are the maximum and minimum principal stresses at the

point of observation. is the fringe order, is the sample thickness through

which the light travels, and

n t

λ is the wavelength of the incident light. C is the

stress-optical coefficient. Eq. (4-1) can be rearranged to

,21 ntfn

tC==−

λσσ (4-2)

where the fringe-stress coefficient Cf /λ= represents the principal stress

difference necessary to produce a one fringe order change in a crystal of unit

Page 73: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

59

thickness. The fringe-stress coefficient depends on the stress-optical coefficient

C of the material and the wavelength of the incident light. To allow easy

comparison and stress calculation via the fringe-stress coefficient, a single

standard wavelength should be used. While any color monochromatic light is

acceptable, ~535 nm green light was selected for current and previously reported

experiments. The engineering units of the fringe-stress coefficient are N/m.

In a previous paper (Chapter 2), we have reported the stress-optical

coefficient of [001]-oriented unpoled PMN-29%PT single crystals evaluated using

three-point bending technique [12]. However, in three-point bending

experiments, only the stress corresponding to the fringe at the bottom edge of the

beam can be calculated correctly using beam bending theory. Information from

other fringes is not useful because the stress state is a mixture of bending and

shear. In one experimental run with increasing load, for each load level we can

obtain only one datum. The stress-optical coefficient must then be estimated

from the linear region of the fringe order vs. fiber stress plot. As a result, from

one experimental run only one stress-optical coefficient can be calculated, and to

reduce the error, the same run needs to be repeated many times. Furthermore,

in three-point bending experiments, it is difficult to accurately measure how and

whether the stress-optical coefficient changes with increasing stress.

All of these problems can be solved in four-point bending experiments.

Four-point bending produces a pure bending region without any shear stress,

resulting in multiple parallel fringes that are aligned with the sample edges. For

each fringe, the corresponding tension (or compression) stress can be calculated

using beam bending theory. Since the sample contains multiple usable (pure

bending) fringes at each load, it is possible to plot multiple data points and

calculate the stress-optical coefficient once for each load level. Variations of the

stress-optical coefficient with increasing load can then be observed.

Page 74: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

60

The purpose of this chapter is to measure and report the stress optical

coefficient and the fringe stress coefficient of PMN-29%PT single crystals. The

four point bending method allows us to further determine and report the stress

dependence of these coefficients. A secondary purpose of this chapter is to

introduce the hypothesis of mechanical poling to characterize and explain the

complex sequence of crystallographic changes induced by mechanical loads

which comprises two states in sequence: a first state where domains freely

switch towards preferred directions to form larger domains, and a second state

where mechanical loads induce a rhombohedral to monoclinic phase

transformation. Mechanical poling can induce permanent, yet reversible,

changes in the domain structure as evidenced by residual photoelastic fringes

which can be recovered by annealing. The two states and their transitions

correspond to observed changes of the values of the fringe stress coefficient.

4.2 Experimental procedure

An in situ loading frame using a pneumatic cylinder as shown in Figure 4.1

was built to allow us to perform photoelastic measurements on small beam

samples under a ZeissTM microscope. The frame was calibrated so that the

applied force could be known directly from the reading of an air pressure gauge to

within 2.5%. The loading frame was designed with a low profile to allow a 10X

objective lens to be used to observe the experiments. Four glass rods with 1 mm

radius were used to exert the loading force. A tilting bar (labeled “A” in Figure

4.1 (b)) was used to provide symmetrical loading. A removable alignment fixture

was designed to position the glass rods before each experiment so that the upper

and lower glass rod spans were reliably set at 5 mm and 15 mm respectively.

±

Two (001)-oriented unpoled PMN-29%PT single crystal beam samples were

prepared as shown in Figure 4.2. Beam 1 with dimensions 1mm x 1mm x 17mm

Page 75: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

61

was previously subjected to three-point bending experiments, and was obtained

from H.C. Materials [12]; beam 2 with dimensions 1mm x 2.2mm x 18mm was

obtained from TRS Ceramics. During the experiments the samples were always

oriented so that light was transmitted through the 1 mm thick dimension. The

surfaces for light transmission were polished to optical quality using a graded

series of polishing papers down to 1um grit, then finished with

chemical-mechanical polishing using 0.05um colloidal silica. Before each

experiment, the beam samples were annealed at 400 oC for 1 hour to reduce any

residual fringes and minimize the initial fringe order. Figure 4.2 shows typical

fringe patterns after annealing; the fringe orders were close to 0th-order (black)

and typically less than 1st-order (rose) over the majority of the sample area.

Figure 4.1 (a) Overview of the in situ loading frame and (b) Four-point bending

set-up; A represents the tilting bar, and B represents the beam sample.

Figure 4.2 Beam 1 (a) and beam 2 (b) after 1 hour annealing at 400 oC.

Page 76: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

62

Figure 4.3 Four-point bending layout. P is the loading force, cσ and tσ are compression

and tension stresses respectively. The diagram under the sample shows the absolute value of the bending moment.

The schematic four-point bending loading system is represented in Figure

4.3. The symmetrical loading scheme stresses the beam with a constant

bending moment 6PLM = , producing a pure bending stress state in the loading

region. The resulted axial stress ( )yσ has a uniform stress gradient through

the beam height:

( ) 3

2thPLy

IMyy ==σ , (4-3)

where is the moment of inertia of the beam, and and are the

thickness and the height of the beam, respectively.

12/3thI = t h

L is 15 mm in all of the

experiments performed. is the perpendicular distance to the neutral axis. In

the pure bending region,

y

( )yσ is the principal stress 1σ at , and the principal

stress

y

02 =σ at , allowing us to simplify Eq. (4-2) to: y

( ) ntfy =σ . (4-4)

Therefore, fringe-stress coefficient can be obtained from the slope of

stress versus fringe order plot.

f

Page 77: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

63

To use Eq. (4-3) to calculate the stress for an observed fringe, the distance

between the fringe and the 0

y

th order fringe (neutral axis) needs to be measured.

During the experiments, pictures of fringe patterns were taken with the 10X

objective lens at several different load levels. The displacement of each

fringe could then be directly measured based on variations in light intensity, using

a method that is described in more detail below.

y

The experiments were performed using a dark field circular polariscope setup

to eliminate the isoclinic fringes and show only the isochromatic fringes, which

depend only on the internal principal stress difference. A narrow band-pass

green filter with central wavelength ~535 nm was used because this wavelength

produced a suitable number of fringes throughout the applied load range. Three

four-point bending experiments were performed on beam 1. Between the

experiments, beam 1 was heated to remove all of the residual fringes produced

by bending, allowing it to be reused; this “annealing” effect is discussed in more

detail below. Unfortunately, beam 2 was used only once, and was broken during

that experiment.

4.3 Results and discussion

4.3.1 Fringe Patterns

In the pure bending region, perfectly parallel fringes were observed,

comparable to those typical of isotropic materials [14]. As the applied load was

increased, fringes were generated from the top and bottom edges of the beam

and migrated towards the stress-free neutral axis where the 0th order fringe

always lies. Fringes formed in pairs (top and bottom), producing symmetric

patterns about the neutral axis. The orders of the fringes were thus typically

counted from the center as L,2,1 ±±=n . The fringes for each load level were

Page 78: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

64

found to be distributed uniformly along the height of the beam, corresponding to

the linear variation of the principal stress shown in Figure 4.3.

Figure 4.4 (a) and (b) show photoelastic images of beam 1 obtained while a

load was applied and after the load was released, respectively. After the load

was released, there were usually many residual fringes left. These residual

fringes could be fully removed by 1 hour of heating at 400 oC. The beam

sample could then be reused. This implies that any optical or mechanical effects

caused by the bending experiments were reversible.

Figure 4.4 (a) 3.5 order of fringes and (b) 2 order of fringes left after the load is released.

4.3.2 Fiber stress versus fringe order

Figure 4.5 shows a plot of maximum fiber stress versus maximum fringe order

for a range of different loads. Maximum fiber stress (in tension side) was

calculated using Eq. (4-3) with 2/ty = , the same method was used in our

three-point bending work [12]. Included in this figure are data from three runs for

beam 1 and one run for beam 2. In addition, one run of data for beam 1 under

three-point bending is included for comparison. The results are highly consistent

and independent of any changes in bending scheme or beam dimension. The

slope in Figure 4.5 was obtained in terms of least-square fit based on data below

the optical yield stress of 20 MPa obtained from Figure 4.6, that is comparable to

the value obtained from Figure 2.6 in Chapter 2. The correlation coefficient is

0.95.

Below 20-30 MPa, a linear relationship can be seen between the fiber stress

Page 79: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

65

and the fringe order. Above that limit some nonlinear behavior occurred. This

corresponds to results from our three-point bending work which demonstrated

mechanical yielding at higher load levels [12]. Using Eq. (4-4), from the slope of

the linear region of Figure 4.5 we calculated the fringe-stress coefficient to be

4.8X103 N/m, which equals to a stress-optical coefficient of 111.5X10-12 Pa-1.

Figure 4.5 Maximum fiber stress versus fringe order.

Figure 4.6 Maximum fiber stress versus fringe order with polynomial fit curve.

Page 80: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

66

4.3.3 Fringe-stress coefficient

The calculated fringe-stress coefficient of 4.8X103 N/m is only an approximate

result for low level stresses or loads. As discussed previously, four-point

bending provides a method to obtain the fringe-stress coefficient more accurately

at different levels of applied load. This method is illustrated in Figure 4.7. For

each load level, a microscope image is captured, and then a vertical section is

taken through the image. Using ImageProTM software, the green pixel intensity

along the section can be plotted as a function of distance. The vertical

displacement between each fringe and the 0

y

th order fringe can then be measured

in ImageProTM with an error of +/- 3 μm.

Given the displacement of each fringe from the center line, the corresponding

stress for each fringe order can then be calculated using Eq. (4-3). The result for

each microscope image is a line of stress vs. fringe order data points representing

the behavior of the sample at that load level.

Figure 4.7 From the light intensity plot, displacement between fringes can be measured. Each valley of the intensity curves represents a fringe (darkest field), and each peak of

the intensity curves represents the half order of fringe (brightest field).

Four of these data lines are plotted in Figure 4.8. The four load levels shown

had maximum fringe orders of 4, 7, 13, and 16, respectively. For each load the

data are essentially linear, and the slope of each line is the fringe-stress

coefficient at that load. It can be seen that with increasing load the fringe-stress

Page 81: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

67

coefficient (slope) decreases first, and then increases. This trend can be seen

more clearly in Figure 4.9, which shows the fitted slopes (fringe-stress coefficient)

for many load levels and includes both beam 1 and beam 2 data sets. The

fringe-stress coefficient decreases as the applied load increases, and reaches a

minimum at a max fiber stress of 45-50 MPa (with fringe order 14). Beyond this

point the fringe-stress coefficient begins to increase. The fringe-stress

coefficient varied in the range of (3.5~5.5) X 103 N/m.

In our previous three-point bending experiments [12] we observed mechanical

yielding as the beam was bent. This was evidenced by nonlinearity in the

force-displacement data. We observed similar nonlinearity in the stress vs.

fringe order data. Further, the max fringe order was linearly proportional to beam

displacement over the entire range of loads during the three-point bending

experiments as mentioned in the preliminary three-point bending experiment.

Therefore, we hypothesize that changes in the fringe-stress coefficient must

correlate with changes in the Young's modulus of the material.

Figure 4.8 Stress versus fringe order for different load level. The number label

represents the maximum fringe order obtained for each load level. The slope of each data line represents the fringe-stress coefficient.

Page 82: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

68

Figure 4.9 Fringe stress coefficient versus maximum fiber stress.

With this interpretation, the optical four-point bending data presented in Figure

4.8 clearly shows initial yielding of the bar, followed by stiffening at stresses above

~50 MPa. This behavior may be the result of internal stress-induced domain

switching/phase transformation. In unpoled PMN-29%PT single crystals with

rhombohedral structure, there are eight possible domain orientations along one of

eight <111> directions. Under a static stress state, domains are free to reorient.

Domain rotation may involve transformations from rhombohedral to monoclinic B

or tetragonal phase occurring locally according to the local stress level. This

domain switching partially absorbs the mechanical work of the externally applied

load, reducing the influence of the load and causing the material to behave as if

the elastic constant has decreased (softening). As the applied load increases,

eventually most domains have reoriented and the effect is saturated, causing the

material to resist further deformation and effectively stiffen. Similar effects

resulting from uniform compression stress have been reported by Viehland, et al.

[15] and Viehland and Li [16].

Page 83: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

69

4.3.4 Mechanical poling effect

During the experiments, when the applied load increased to the point where

the maximum fiber stress was ~45 MPa, the compression region began to

brighten. As shown in Figure 4.10(a), the brightness was relatively uniform

throughout the whole sample while the maximum fringe order was under 14.

However, once the fringe order went higher than 14 (Figure 4.10(b)), the

compression region became much brighter than other regions. This implies that

the transmission of that region was highly enhanced. As the applied load was

increased further, this brightening effect began to occur in the tension region as

well, and moved towards the neutral axis along with the fringes. Wan, et al. [7,

13] suggests that enhanced transmission is correlated with poling, so this effect

may indicate that the bright regions have become poled in some way. This

contributes to the hypothesis that PMN-PT undergoes mechanical poling - an

effect induced by mechanical loading.

Figure 4.10 Fringe patterns of pure bending region at different load levels.

(a) Totally 11 order of fringes; (b) totally 16 order of fringes.

4.4 Conclusions

Photoelastic effects in unpoled PMN-29%PT single crystals were investigated

using four-point bending experiments. Perfectly parallel fringes were distributed

uniformly through the height of the beam, implying a pure bending stress state

provided by four-point bending experiments. The fiber stress vs. fringe order plot

Page 84: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

70

is highly consistent with the result from previous three-point bending experiments.

Two methods were used to calculate the fringe-stress coefficient: first from the

slope of the linear region in Figure 4.5, was found to be 4.8X10f 3 N/m; Second,

the fringe-stress coefficient was measured from each birefringence image for

multiple loading levels, and varied in the range (3.5~5.5)X103 N/m depending on

the load. As shown in Figure 4.9, with increasing load, the fringe-stress

coefficient decreases to a minimum and then increases. This effect is

interpreted as evidence of internal domain switching and phase transformation, i.e.

a mechanical poling process. Increased optical transmission in high-stress

regions was also observed, a behavior which supports the hypothesis that a

poling process occurs.

Finally, we observed that after the samples were unloaded, many fringes and

the higher optical transmission effects remained. We found that heating above

the Curie temperature restored the samples to their initial optical state. This

implies that all of the mechanical poling effects that we observed are reversible.

4.5 References 1. S.-E. Park and T. R. Shrout, “Ultrahigh strain and piezoelectric behavior in

relaxor based ferroelectric single crystals”, J. Appl. Phys. 82, No. 4, 1804, 1997.

2. Y. Yamashita, “Large electromechanical coupling factors in perovskite

binary material system”, Jpn. J. Appl. Phys. 33, 5328, 1994.

3. Z G Ye, B Noheda, M Dong, D Cox and G Shirane, “Monoclinic phase in the relaxor-based piezo-/ferroelectric Pb(Mg1/3Nb2/3)O3-29%PbTiO3 system”, Physics Review B, 64, 184114, 2001.

4. X. Zhao, B. Fang, H. Cao, Y. Guo and H. Luo, “Dielectric and piezoelectric

performance of PMN-PT single crystals with compositions around the MPB: influence of composition, poling field and crystal orientation”, Materials Science and Engineering: B, 96, 254, 2002.

Page 85: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

71

5. R. Zhang, B. Jiang and W. Cao, “Elastic, piezoelectric, and dielectric properties of multidomain .67Pb(Mg1/3Nb2/3)O3 -0.33PbTiO3 single crystals”, J. Appl. Phys. 90, 3471, 2001.

6. M. Dong and Z. G. Ye, “High-temperature solution growth and

characterization of the piezo-/ferroelectric (1-x) Pb(Mg1/3Nb2/3)O3-x PbTiO3 [PMNT] single crystals”, Journal of Crystal Growth, 209, 81, 2000.

7. X. Wan, J. Wang, H. L. W. Chan, C. L. Choy, H. Luo and Z. Yin, “Growth

and optical properties of 0.62 Pb(Mg1/3Nb2/3)O3-0.32 PbTiO3 single crystals by a modified Bridgman technique”, Journal of Crystal Growth, 263, 251, 2004.

8. Z. Yan, X. Yao, and L. Zhang, “Analysis of internal-stress-induced phase

transition by thermal treatment”, Ceramics International, 30, 1423, 2004.

9. H. Liang, Y. Pan, S. Zhao, G. Qin, and K. K. Chin, “Two-dimensional state of stress in a silicon wafer”, J. Appl. Phys. 71, 2863, 1992.

10. M. Lebeau, G. Majnib, N. Paone, and D. Rinaldib, “Photoelasticity for the

investigation of internal stress in BGO scintillating crystals”, Nuclear Instruments and Methods in Physics Research A, 397, 317-322, 1997.

11. K. Higashida, M. Tanaka, E. Matsunaga, and H. Hayashi, “Crack tip stress

fields revealed by infrared photoelasticity in silicon crystals”, Materials Science and Engineering A, 387-389, 377, 2004.

12. N. Di, and D. J. Quesnel, “Photoelastic effects in

Pb(Mg1/3Nb2/3)O3-29%PbTiO3 single crystals investigated by three-point bending technique”, J. Appl. Phys. 101, 043522, 2007.

13. X. Wan, H. Luo, J. Wang, H. Chan, and C. Choy, “Investigation on optical

transmission spectra of (1-x) Pb(Mg1/3Nb2/3)O3-x PbTiO3 single crystals”, Solid State Communications, 129, 401, 2004.

14. M. M. Frocht, Photoelasticity, John Wiley and Sons, 1941.

15. D. Viehland, J. Powers, L. Ewart, and J. F. Li, “Ferroelastic switching and

elastic nonlinearity in <001>-oriented Pb(Mg1/3Nb2/3)O3-x%PbTiO3 and Pb(Zn1/3Nb2/3)O3-x%PbTiO3 crystals”, J. Appl. Phys. 88, 4907, 2000.

Page 86: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

72

16. D. Viehland and J. F. Li, “Stress-induced phase transformations in <001>-oriented Pb(Mg1/3Nb2/3)O3–PbTiO3 Crystals: bilinear coupling of ferroelastic strain and ferroelectric polarization”, Philosophical Magazine, 83, 53-59, 2003.

Page 87: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

73

5 Electric field induced optical effects in PMN-29%PT single

crystals

Abstract: Stress fields induced by electric fields in unpoled PMN-29%PT are

compared with those induced by mechanical loading using optical birefringence

visualization. Photoelastic experiments explore the differing effects of square

waveform cyclic electric field and DC field, as well as cyclic electric field frequency

effects. Experiments are performed on differently oriented samples to examine

the effects caused when electric fields are applied in different crystallographic

directions for comparison. The causes of crack initiation are examined.

Results show that it is easy for crystals to be locally poled around their edges by

cyclic electric field, and these edge regions are highly susceptible to crack

formation under a continuously applied cyclic electric field. Finally, the

mechanisms of mechanical poling and electrical poling are discussed.

5.1 Introduction

Piezoelectric single crystals possess stronger dielectric and piezoelectric

properties than those of conventional piezoelectric ceramics. They have

promising potential to replace ceramics and can be used in wide-ranging

applications, like transformers, actuators or sensors [1-6]. The working

condition of those devices is either under high magnitude cyclic or direct electric

field, under mechanical loading, or sometimes a mix of electrical loading and

mechanical loading. Unfortunately, piezoelectric single crystals are very

susceptible to fracture. This is why single crystals have not yet seen wide

implementations. They are elastically softer than ceramics and easily develop

cracks under even low levels of stress concentration. As is well known for

Page 88: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

74

piezoelectric materials, stress can be induced by both mechanical loading and

electrical loading, and there are numerous of processes that can cause stresses

in these single crystals. For example, during growth, an inhomogeneous

temperature field exists which may induce residual stress inside the crystals [7-8].

Then, during the machining process, mechanical loading causes increased

residual stresses inside the crystals, especially the surfaces, edges and corners.

These machining processes include cutting, grinding, and polishing. Finally,

crystals are normally used under either high frequency cyclic electric field or

under pulsed DC electric field loading. These working conditions can also

increase the internal stresses in the single crystals. Given the relatively high

residual stresses induced previously from crystal growth and machining process,

it is very easy for the crystals to fracture and finally fail to function. To strengthen

the crystals, we can develop new polishing techniques, and also apply annealing

to crystals to reduce the residual stresses. It is not clear if this will be sufficient

to allow crystals with small residual stresses to survive harsh working conditions

and obtain a long useful service life. Thus it is indispensable to investigate the

electric field effects. In this chapter, we will focus on studying electric field

induced internal stress in unpoled PMN-29%PT single crystals using optical

birefringence techniques.

Several researchers have already studied the cyclic electric field effects in

ferroelectric ceramics or piezoelectric single crystals such as Lynch et al. [9-10]

and other researchers [11-14]. Topics include domain switching, phase

transformation, microcracking, and fracture. The observation and study are

normally conducted through means such as TEM, dielectric measurements, optic

microscopy, and other similar techniques. Crack growth is directly observed

under an optical microscope, and microcrack growth is observed under TEM.

Phase transformation is studied by measuring changes in dielectric properties.

Page 89: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

75

These studies have helped gain increased understanding of what is going on

when the piezoelectric materials are under electrical loading. However, if we

can evaluate the internal stress state of the materials during the electrical loading,

we can better understand crack initiation and growth conditions. Fortunately,

unpoled PMN-PT single crystals can be polished to be transparent and show

colorful birefringence. Previously, using well-known photoelastic technique, we

already studied the internal stress fields of PMN-PT single crystals under

mechanical loading such as bending and Hertzian indentation [15-16]. Similarly,

we can use birefringence techniques to study the internal strain of PMN-PT single

crystals induced by electrical loading.

When electric field is applied to PMN-29%PT, several electro-optic effects are

induced. First, all transparent solids become birefringent when subjected to an

electric field, this phenomenon being known as the Kerr effect [17]. The

mechanism behind this phenomenon is the same as that of photoelasticity: the

electric field changes the index of refraction of the materials in the solid state. The

difference between the photoelastic phenomenon and the Kerr effect is that the

photoelastic birefringence is linearly proportional to the mechanical stress, while

the Kerr electro-optic birefringence is proportional to the square of the electric

field. That is why the Kerr effect is also called the quadratic electro-optic effect.

Second, there is a phenomenon known as the Pockels effect, which applies only

to crystals that are noncentrosymmetric such as piezoelectric crystals [17]. The

external electric field alters the dipole moment of the single crystals, causing the

Pockels effect. Similar to the photoelastic effect, birefringence of the Pockels

effect is linearly proportional to the electric field so that the Pockels effect is also

known as the linear electro-optic effect. Third, coming with the Pockels effect

and owing to the piezoelectric nature of PMN-29%PT single crystals, the external

electric field will induce strain and in turn cause a photoelastic effect. Thus, in

Page 90: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

76

general, the observed birefringence is induced by the combination of three effects,

one quadratic electro-optic effect and two linear electro-optic effects. Compared

to the linear electro-optic effects, the nonlinear Kerr effect is much weaker than

the Pockels effect in piezoelectric crystals, though it responds quickly to changes

in electric field, even for frequencies as high as 10 GHz. In our experiments, the

applied field is relatively small (under 3 KV/cm), so the response is primarily

induced by the linear electro-optic effects, which directly affects internal domain

switching and internal strain. This strain is constrained by the surrounding

material, resulting in a stress field. Therefore, the electric field induced

birefringence represents the internal stress field. This chapter is aimed at

gaining a fundamental understanding of the internal stress fields induced by

electric fields. Birefringence is used as a means to visualize the internal stress

state.

Unpoled PMN-PT single crystals were used in our experiments for two

reasons: first, unpoled PMN-PT single crystals are transparent while poled

PMN-PT single crystals are nearly opaque under microscopy; second, the

electrical poling process may induce residual stress, which is an important topic

studied here. As we know, even unpoled PMN-PT single crystals possess

anisotropic mechanical and optical properties [16]. To explore the orientation

dependence of electro-optic effects, several samples of different orientations

were used in this study. “Hertzian contact” electrical loading (the electrical

equivalent of mechanical point loading) and electrical poling experiments were

performed. Fringe patterns induced from “Hertzian contact” electrical loading

are highly orientation dependent, and are comparable with those induced by

Hertzian contact mechanical loading. In electrical poling experiments, cyclic

electric fields were found to induce cracks much more easily than DC electric

fields. The possible reason is that cyclic electric field can easily pole the crystals

Page 91: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

77

locally, thus degrading the crystal properties by introducing boundaries between

poled and unpoled regions where large strain gradients can cause cracking .

5.2 Experimental procedure

Figure 5.1 Initial birefringence patterns of four differently oriented samples under

circularly polarized illumination after 1 hour annealing at 400 oC. Single-crystal samples of unpoled PMN-29%PT in four different orientations

were prepared as shown in Figure 5.1. Samples were obtained from H.C.

Materials. Beam 1 with dimensions of 9mm x 1.7mm x1mm was “Hertzian

contact” electrically loaded. Beam 2 with dimensions of 8mm x 1.8mm X1mm

was used in electrical poling experiments. Both of these beams are

{100}-oriented. Figure 5.1(c) and (d) show sample 3 and sample 4 respectively,

which are used in “Hertzian contact” electrical loading experiments. Sample 3

has two side faces {100} and the other two side faces {110}, while sample 4 has

all four side faces having {110} orientation. When observed under the microscope,

light was always transmitted through the 1 mm thickness of all of the samples.

The surfaces for light transmission were polished to be optically transparent using

a sequence of graded abrasives with average grit size down to 0.05 μm. Before

Page 92: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

78

experiments, all samples were annealed at 400 oC for one hour to reduce residual

stresses.

As mentioned above, two different experiments have been performed to

explore the electric field loading effects. The first one is the “Hertzian contact”

electrical loading experiment. It was designed to imitate the mechanical

Hertzian contact experiments to explore whether an equivalent fringe pattern will

emerge to that observed in mechanical loading of unpoled PMN-29%PT single

crystals. An equivalent fringe pattern would allow a comparison of electrical and

mechanical loading in these crystals. The other experiment is an electrical

poling experiment.

Figure 5.2 (a) Overview of the in situ electrical loading frame and (b) Top view of

“Hertzian contact” electrical loading set-up. The “Hertzian contact” electrical loading experiment arrangement is shown in

Figure 5.2(b). The electrical poling experiments were performed with the same

experimental setup except the rod electrode was replaced by a block electrode.

The white loading frame was made using high density polyethylene. Different

size metal blocks and a 1mm radius metal rod were used as electrodes. In

“Hertzian contact” electrical loading experiments, the rod electrode was used to

apply drive voltage and block electrodes to provide ground. Furthermore, the

sample surface contacting the block electrode was coated with gold. In the

electrical poling experiment, both electrically contacted surfaces were coated with

Page 93: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

79

gold to provide a relatively uniform electric field through the samples. All

experiments are conducted at room temperature. The polyethylene load frame

allows a 10X lens to be used to observe small cracks as they initiate in the

crystals. A circular polariscope employing a green filter was used to eliminate

the isoclinic fringes and thus obtain well-defined isochromatic fringes in order to

characterize the electric field induced strains.

In the experiments, we applied DC and square waveform cyclic (AC) electric

field to the crystals with varying magnitude of the electric field and varying

frequency of the square waveform AC. Unless otherwise noted, voltage was

applied for a fixed 2 minute period for each experiment to allow simple

comparison of the resulting fringes. This is because the fringes will grow bigger

with time that the sample is exposed to the test voltage. We found the fringe

pattern is different for DC vs. square waveform AC electric field, and also, the

fringe pattern is dependent on the orientation of the sample. For a given test

time, the size of the fringe pattern is proportional to the magnitude of the electric

field; it is also related to the frequency of the AC field. Between the experiments,

samples are annealed to remove all the residual fringes so that they can be used

again.

5.3 Results and discussions

5.3.1 “Hertzian Contact” electric field loading effects

Hertzian contact loading experiments were conducted to explore

birefringence of unpoled PMN-29%PT single crystals under an electrical point

loading conditions. The experimental set-up, as shown in Figure 5.2(b), is

almost the same as the mechanical Hertzian contact loading set-up, except the

driving force: one is mechanical force, the other is electrical loading. Since

piezoelectric materials respond to both mechanical and electrical loading, to do

Page 94: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

80

the same type experiments, we can compare the different driving force effects on

strain generation by observing the birefringence with optical methods. These

types of experiments are also of practical importance because they characterize

single point electrical loading effects which may happen in several in-service

cases. During electrical poling, if the electrical-plating materials are non-uniform

or the surfaces are rough, the crystals will have an applied electrical Hertzian

contact loading. Also, during service, when cracks are generated there will be

electrical Hertzian contact loading near the cracks. This is the electrical

equivalent at a stress concentration representing an electric field concentration.

All pictures shown for electrical loading represent the residual fringes present

after the electric field is removed. Because there was almost no motion or

relaxation of the fringe pattern, these photos characterized the electrically loaded

state as well. In short, all the fringes stayed where they had grown to. This is a

substantial difference with the residual fringes observed under mechanical

loading, which relaxed significantly after the load was released.

In “Hertzian contact” electrical loading experiments, separate experiments

were performed where the DC electric field was applied in opposite directions to

examine if the internal domains have a preference for the DC field direction.

Figure 5.3 show the birefringence obtained from beam 1. There are two

experiments shown here: Figure 5.3(a), (c) and (d) are from one run of the

experiment, while Figure 5.3(b) is from another run of the experiment. In the first

experiment, first a -2.3 KV/cm electric field was applied for 2 minutes, then the

electric field was reversed to +2.3 KV/cm and applied for another 2 minutes

resulting in fringes as shown in Figure 5.3(c). Figure 5.3(d) shows fringe pattern

after additional 2 minutes of applying +2.3 KV/cm. For comparison, Figure 5.3(b)

was taken in another experiment after +2.3 KV/cm had been applied for 2

minutes.

Page 95: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

81

It is obvious that DC field induces two-lobed fringe patterns of very similar

shape regardless of the electric field direction. However, there are more fringes

from +2.3 KV/cm than those from -2.3 KV/cm field. This difference seems likely

to be induced by the initial internal domains switching preference.

Figure 5.3 “Hertzian contact” electrical loading (electrical point load) experiments on

{100}-oriented beam 1: (a) -2.3 KV/cm DC were applied from the top rod electrode for 2 minutes; (b) 2.3 KV/cm for 2 minutes; (c) 2.3 KV/cm applied to resulting fringes of (a) for

another 2 minutes; (d) 2.3 KV/cm for additional 2 minutes after (c). The arrows in the pictures represent the electric field direction.

Figure 5.3(c) shows that after the electric field was reversed, the fringe pattern

immediately develop two bumps around the region contacting the rod electrode

making the two-lobed shape more like butterfly wings. When voltage is reversed

(compare Figure 5.3(c) and 5.3(a)), parts of the fringes seem to pull back toward

the origin, which upper portions of the fringe moved outwards, resulting in the

butterfly wing shape. Figure 5.3(d) shows that 2 more minutes of reversed DC

voltage grew the fringes larger and maintained the butterfly wing shape. It was

Page 96: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

82

observed that the spot under the rod electrode became extremely bright. This

enhanced transmission implies the region has been electric poled with a very

large domain size [18-19]. It is thought that larger domains with same

polarization of rhombohedral phase or monoclinic phase are formed around the

region so that the light scattering losses caused by discontinuous refractive index

at the boundaries of different domains is highly reduced [20].

Figure 5.4 Fringe pattern comparison: (a), (c) and (e) are fringes induced by DC electric

field loading for beam 1 with 2.3KV/cm, sample 3 with 1.8KV/cm and sample 4 with 1.8KV/cm. (b), (d) and (f) are fringes under Hertzian mechanical loading for comparably oriented

samples, as shown in Chapter 3. Figure 5.4 shows the comparison of electric field induced fringe pattern with

those induced by Hertzian contact mechanical loading. For a sample of the

same orientation, it is evident the fringe patterns are similar although they are not

Page 97: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

83

exactly the same. The {100}-oriented samples generate “two-lobed” fringe

pattern under either mechanical loading or electrical loading when loaded from

[100] direction. When loaded from [110] direction, “single-lobed” fringe patterns

are formed as shown in Figure 5.4(c)-(f), independent of which direction we view

the pattern from, [110] or [001].

In Di et al. [16], the orientation and loading direction dependence of fringe

patterns are explained as results from the inherent elastic property of the material

itself. Fringe patterns resulting from mechanical loading were the same as those

fringe patterns which were obtained using ANSYS simulations; this verified that

the elastic properties of unpoled PMN-PT single crystals are highly anisotropic.

To conduct ANSYS simulations of these electrical loading experiments,

piezoelectric strain coefficients need to be considered along with the elastic

properties. However, the samples are unpoled PMN-PT crystals; any

piezoelectric properties would be very weak and not constant for different

samples. Furthermore, measuring the properties of samples uses the

piezoelectric response to induce the strains so that the conventional

measurement methods are not available for unpoled piezoelectric material [21].

Therefore, there is no piezoelectric strain coefficient data reported for unpoled

PMN-PT single crystals. Thus the comparable fringe patterns observed between

electrical and mechanical loading must indicate that the elastic properties affect

the anisotropic response primarily through one of the linear electric-optical effects:

the electric field induced strain conversely causes photoelastic effects.

On the other hand, the difference between the electrical and mechanical

fringe patterns, such as the two lobes growing along different angles, must be

caused by the Pockels effect directly inducing a switching of dipole moments by

the electric field. Furthermore, the DC field induced much larger fringe size

without generating cracks than mechanical loading did, and residual fringes

Page 98: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

84

induced by the DC field didn’t relax much, while those induced by mechanical

loading did. This is also consistent with the Pockels effect: once a dipole

moment is switched, it is hard for them to switch back.

Next the cyclic electric field frequency effect will be discussed. During the

experiments, we observed that as the electric field is cycled at low frequency; the

fringes grow and shrink, giving a ratchet-like incremental growth. When the

frequency is raised to over 120 Hz, the fringes appear to grow at a constant rate.

Figure 5.5 “Hertzian contact” electrical loading experiments on differently oriented

samples using square waveform voltage with 0.5 Hz and 500 Hz, respectively: (a), (b), and (c) (top row) resulted from square waveform voltage of 0.5 Hz. (d), (e), and (f) (bottom row) were from square waveform voltage of 500 Hz. (a) and (d) are from beam 1; (b) and (e) are sample

3; (c) and (f) are sample 4. The magnitude of electric field is 2.3 KV/cm for beam 1 and 1.8 KV/cm for both sample 3 and 4.

As shown in Figure 5.5, the top three pictures show fringes induced by a

square waveform voltage with a frequency of 0.5 Hz; the bottom three pictures

show fringes from squarewave voltage with a frequency of 500 Hz. No matter

what orientation the sample is, the low frequency squarewave voltage induced

large fringes while the high frequency voltage only induced small fringes.

Furthermore, if compared with Figure 5.4 (a), (c) and (e), under the same

Page 99: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

85

magnitude of electric field, obviously the DC field was found to have deeper

penetration than the cyclic electric field. There is one possible explanation for

this phenomenon: as observed in reversed DC experiments, the reversed voltage

didn’t grow the fringe pattern immediately; instead it worked on reversing the

polarization directions which originally aligned to an opposite direction.

Therefore the fringe shape was changed to develop “bumps” near the electrode

side, however the size of the fringe didn’t increase much as shown in Figure 5.3

(a) and (c). In other words, a reversal voltage needs time to grow fringes.

When the frequency of cycled electric field is higher than 0.5 Hz, which means

there is less than one second for positive or negative fields to act on the domains.

Apparently, there is not enough time for fringes to grow large. The small fringes

and bright regions under the electrode imply that electrical energy is mainly

stored close to the surface, and poles (orients) the domains locally, as shown in

Figure 5.5 (d), (e) and (f).

Furthermore, through the linear Pockels effect, a cyclic electric field tends to

switch the internal domain’s polarization as fast as its own frequency. This is

actually cyclic straining process similar to a fatigue process. It is reasonable to

assure that the elastic properties of the “bright” regions are degraded so that

continually applying a cyclic electric field to these regions may easily initiate

cracks. The hypothesis of the degradation of properties in the “bright” region will

be experimentally examined in the following section.

5.3.2 Electrical poling effects

Beam 2 and sample 4 were used in electrical poling experiments. In these

experiments, both top and bottom surfaces were coated with gold. The

experiments were performed with incremental DC voltage starting with ~+1 KV/cm

and increased to ~+6 KV/cm. For comparison, square wave form voltage was

Page 100: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

86

applied to the samples after incremental DC field was applied.

As shown in Figure 5.6, beam 2 was electrically poled by applying incremental

DC voltage at room temperature for one minute intervals. Two sequential stages

were observed in the poling process: first electric field driven fringes growing

inwards from the top and bottom surfaces. This phenomenon only occurred at

very low electric field at a range of (1.1-1.3) KV/cm. Second, when the electric

field was above 1.3 KV/cm, most of the internal bulk region was slightly poled,

appearing dark and cloudy, except the two edge regions near the coated surfaces

which are fully poled and begin to become very bright. The higher the electric

field was, the stronger the cloudiness was and the brighter the edges became.

Figure 5.6 From top to bottom, beam 2 is electrical poled with incremental DC voltage.

The experiment set-up is with two block electrodes; both the top and bottom surfaces are plated with gold.

These poling effects are due to the internal domain switching and phase

transformation. We know that in unpoled {100}-oriented PMN-PT single crystals,

there are eight possible polarization directions along <111> directions. When a

low level electric field is applied, four of the polarization directions that are

Page 101: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

87

contrary to the electric field direction will be removed by switching states so that

only the other four dipole directions remain. The four-dipole domain structure is

called an engineered domain structure. As the magnitude of electric field is

increased, those four polarizations will first switch from <111> direction to <110>

direction, then to <100> direction which is parallel to the electric field direction.

Correspondingly, the phase transforms from rhombohedral to monoclinic to

tetragonal (or orthorhombic) phase. This phase transition path of R →M → T is

reported by Bai and Li using XRD and dielectric properties methods respectively

[22, 23]. Furthermore, in reference [22], it is found that at room temperature, it is

easy for the phase transition from R → M, however, it is difficult for the M → T to

occur because there is a high energy-barrier between the M and T phases. The

dark cloud internal region may be thought of as a mixture of both R and M phases.

Internal boundaries contribute to the cloudiness. Small domains are aligning

themselves to form larger domains. This will induce lots of strain in between

those domains. Furthermore, refractive index discontinuities arise due to the

variation of the orientation of successive domains. This results in considerably

higher light scattering than in the thermally depoled optically-uniform state,

causing the whole area to be less transparent. These are similar phenomena to

those explained with the light scattering theory of piezoelectric ceramics by [13,

24]. For the fully-poled regions which have formed a large domain along the

edge, the transmission of light will be highly promoted, so they become extremely

bright [18, 19]. The cloudiness is only getting darker and darker under

incremental DC field without becoming bright, because the M → T phase

transition is difficult.

In another experiment, beam 2 was first poled with incremental DC field up to

3.4 KV/cm, and then with a low frequency squarewave cyclic electric field of 2.8

KV/cm. The resulting fringes are shown in Figure 5.7(a). It appeared that the low

Page 102: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

88

frequency cyclic electric field immediately poled beam 2: generating dark strips

parallel with the electric field in the middle region and extremely bright areas along

the edge. The bright region is much larger than that formed under DC field. In

addition, cracks were initiated parallel to the electric field from under the block

electrode where displacements are constrained, and crack growth along the

edges was observed. The edge region is extremely bright because domain

switching saturation (large number of small domains which have consolidated into

fewer larger domains) and phase transformation which occurred there.

Consequently, the bright region will lose much of its elastic flexibility and the

elastic constant is increased. Elastic constant of [001]-oriented PMN-32PT

single crystal is seen to increase with increasing electric field reported by Viehland

[25]. Furthermore, a continually applied cyclic electric field has a strain switching

fatigue effect, which will generate cracks much more easily than a DC field, as we

describe below.

Figure 5.7 (a) 2.8 KV/cm square waveform cyclic electric field with 0.5 Hz was applied to

beam 2. (b) Birefringence of beam 2 after annealing. The arrow points at crack generated during the experiment.

To verify that bright region has lost much of its elastic flexibility, Hertzian

mechanical loading experiments using the parallel clamp fixture were performed

on sample 4 which was first poled with DC field. The experiment results are

Page 103: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

89

shown in Figure 5.8. In Figure 5.8(a), with a DC electric field of 2.5 KV/cm

applied, semi-circular fringes developed in the bright region near both the upper

and lower electrodes, which are very different from those shown in Figure 5.4(e).

In Figure 5.8(b), a glass rod was used to exert mechanical loading directly on the

bright region. Cracks developed along the loading direction without generating

any fringes, implying that the mechanical properties of the bright region have

changed and can no longer produce the strain responsible for the mechanical

fringes. It is evident that the bright region has different optical and elastic/plastic

properties than the original material.

Figure 5.8 (a) 2.5 KV/cm DC electric field was applied to sample 4 using electrical

“Hertzian contact” experimental set-up. (b) Hertzian mechanical loading on poled region.

5.3.3 Mechanical poling versus electrical poling

As discussed above, cyclic electric field was shown to pole PMN-PT single

crystals more easily than DC field. The elastic properties of bright regions poled

by electric field are degraded leading to a loss of resistance to cracking which was

verified experimentally. These bright regions are very fragile to cracks if

subjected to electric field or mechanical loading. There is yet another poling

effect induced by external mechanical loading called mechanical poling.

Compared to electrical poling, mechanical poling is less likely to produce bright

Page 104: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

90

regions because of the different poling mechanism. However, large strain testing

in bending has demonstrated a bright region developing as a result of purely

mechanical loading. Let us examine this in detail.

Figure 5.9 Mechanical poling and electrical poling representation.

For <001>-oriented PMN-PT single crystals, mechanical poling is a

two-dimensional effect compared to electrical poling, which is a one-dimensional

effect, as shown in Figure 5.9. The original eight possible polarization directions

will switch to find themselves in a single plane under mechanical loading. Under

compression, the plane is perpendicular to the stress direction, while under

tension, the plane is parallel to the stress direction. It is obvious that electrical

poling can align domain polarization in one direction parallel to the electric field.

Therefore, there are at least four polarization directions left after mechanical

poling, and only one polarization direction left after thorough electrical poling.

We already know that the mechanical property and the optical property are highly

dependent on the internal domain structure. After poling, large numbers of small

domains are aligned and form fewer large domains. During this process, the

elastic modulus is stiffened. The fewer domains of different orientations which

are left, the more stiffened the elastic modulus becomes. This means that after

electrical poling, the elastic stiffness is higher than that after mechanical poling,

Page 105: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

91

and it is higher because the density of the domains still available to switch to

accommodate strain gradients is reduced. Thus electric field loading which

gives rise to a more completely poled structure may initiate cracks more easily

than mechanical loading. Essentially, electrical poling eliminates the domain

switching capability of the material that is responsible for its apparent plastic

deformation capability so that poled crystals are more prone to crack.

5.4 Conclusions

Birefringence induced by electric fields on unpoled PMN-29%PT single

crystals was investigated using optical techniques. There are a total of three

electro-optic effects: the Kerr quadratic effect, the Pockels linear effect and the

piezoelectric strain-induced photoelastic effect. For piezoelectric materials, the

observed birefringence is primarily induced by the later two linear electric-optical

effects. “Hertzian contact” electrical loading experiments and electrical poling

experiments were performed. In Hertzian electrical loading experiments, the DC

field induced fringe pattern is highly orientation-dependent, which is comparable

to the Hertzian mechanical loading fringe pattern. Cyclic electric fields with high

frequency induced fringes which are much smaller than those induced by a DC

field or a low frequency cyclic electric field. However, a cyclic electric field,

regardless of frequency, may easily pole PMN-PT single crystals, especially the

local region close to the conducting coated surfaces, thus making these areas

likely to generate cracks. These phenomena are observed in electrical poling

experiments, as well. The explanation is that the time for each sign of field to act

to orient domains becomes extremely short with increment of the frequency of

cyclic electric field, so there is not enough time for fringe to grow big. Also

domain switching follows the frequency of the electric field, which reduces the

ability of the material to use domain switching to accommodate strain gradients.

Page 106: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

92

This, in effect, embrittles the material. To verify this, “Hertzian contact”

mechanical loading experiments were performed on a {110}-{110} oriented

sample that was first electrically poled. Cracks were observed to form

immediately along the loading direction under the loading cylinder: This implies

that the bright region poled by the electric field has less elastic flexibility because

the internal domains are fully-poled and, in effect, form a larger single domain with

one polarization direction. Reduction of the number of multiple domains

degrades the ability of the PMN-PT single crystals to accommodate strain and

strain gradients, reducing the toughness of the material. Finally, the

mechanisms of mechanical poling and electrical poling were discussed.

5.5 References 1. S.-E. Park and T. R. Shrout, “Ultrahigh strain and piezoelectric behavior in

relaxor based ferroelectric single crystals”, J. Appl. Phys. 82, No. 4, 1804, 1997.

2. Y. Yamashita, “Large electromechanical coupling factors in perovskite

binary material system”, Jpn. J. Appl. Phys. 33, 5328, 1994.

3. Z.-G. Ye, B. Noheda, M. Dong, D. Cox and G. Shirane, “Monoclinic phase in the relaxor-Based piezoelectric/ferroelectric Pb(Mg1/3Nb2/3)O3-PbTiO3 system”, Phys. Rev. B, 64, 184114, 2001.

4. X. Zhao, B. Fang, H. Cao, Y. Guo, and H Luo, “Dielectric and piezoelectric

performance of PMN-PT single crystals with compositions around the MPB: influence of composition, poling field and crystal orientation”, Materials Science and Engineering: B, 96, 254-262, 2002.

5. R. Zhang, B. Jiang and W. W. Cao, “Elastic, piezoelectric, and dielectric

properties of multidomain 0.67Pb(Mg1/3Nb2/3)1-0.33TixO3 single crystals”, J. Appl. Phys. 90, 3471-3475, 2001.

6. M. Dong, and Z. G. Ye, “High-temperature solution growth and

characterization of the piezo-/ferroelectric (1-x)Pb(Mg1/3Nb2/3)O3-xPbTiO3 single crystals”, Journal of Crystal Growth, 209, 81-90, 2000.

Page 107: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

93

7. X. Wan, J. Wang, H. L. W. Chan, C. L. Choy, H. Luo and Z. Yin, “Growth and optical properties of 0.62 Pb(Mg1/3Nb2/3)O3-0.32 PbTiO3 single crystals by a modified Bridgman technique”, Journal of Crystal Growth, 263, 251, 2004.

8. Z. Yan, X. Yao, and L. Zhang, “Analysis of internal-stress-induced phase

transition by thermal treatment”, Ceramics International, 30, 1423, 2004.

9. C. S. Lynch, W. Yang, L. Collier, Z. Suo, and R. M. McMeeking, “Electric field induced cracking in ferroelectric ceramics”, Ferroelectrics, 166, 11 – 30, 1995.

10. C. S. Lynch, L. Chen, Z. Suo, R. M. McMeeking, and W. Yang,

“Crack-growth in ferroelectric ceramics driven by cyclic polarization switching”, Journal of intelligent material systems and structures, 6, 191-198, 1995.

11. Z. Li, Z. Xu, X. Yao, and Z.-Y. Cheng, “Phase transition and phase stability

in [110]-, [001]-, and [111]- oriented 0.68 Pb(Mg1/3Nb2/3)O3-0.32 PbTiO3 single crystal under electric field”, J. Appl. Phys. 104, 024112, 2008.

12. Z. Xu, “In situ TEM study of electric field-induced microcracking in

piezoelectric single crystals”, Materials Science and Engineering B, 99, 106-101, 2003.

13. E. T. Keve and K. L. Bye, “Phase identification and domain structure in

PLZT ceramics”, J. Appl. Phys. 46, 87, 1975.

14. F.-X. Li, S. Li, D.-N. Fang, “Domain switching in ferroelectric single crystal/ceramics under electromechanical loading”, Materials Science and Engineering B, 120, 119–124, 2005.

15. N. Di, and D. J. Quesnel, “Photoelastic effects in

Pb(Mg1/3Nb2/3)O-29%PbTiO3 single crystals investigated by three-point bending technique”, J. Appl. Phys. 101, 043522, 2007.

16. N. Di, J. C. Harker and D. J. Quesnel, “Photoelastic effects in

Pb(Mg1/3Nb2/3)O-29%PbTiO3 single crystals investigated by Hertzian contact experiments”, J. Appl. Phys. 103, 053518, 2008.

Page 108: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

94

17. T. S. Narasimhamurty, Photoelastic and Electro-Optic Properties of Crystals, Plenum Press, 1981.

18. X. Wan, J. Wang, H. L. W. Chan, C. L. Choy, H. Luo, and Z. Yin, “Growth

and optical properties of 0.68 Pb(Mg1/3Nb2/3)O3-0.32 PbTiO3 single crystals by a modified Bridgman technique”, Journal of Crystal Growth, 263, 251–255, 2004.

19. X. Wan, H. Luo, J. Wang, H. L. W. Chan, and C.L. Choy, “Investigation on

optical transmission spectra of (1-x) Pb(Mg1/3Nb2/3)O-xPbTiO3 single crystals”, Solid State Communications, 129, 401-405, 2004.

20. Z. Feng, X. Zhao, and H. Luo, “Effect of poling field and temperature on

dielectric and piezoelectric property of <001>-oriented 0.7Pb(Mg1/3Nb2/3)O3-0.3PbTiO3 crystals”, Materials Research Bulletin, 41, 1133–1137, 2006.

21. W. Jiang, R. Zhang, B. Jiang, and W. Cao, “Characterization of

piezoelectric materials with large piezoelectric and electromechanical coupling coefficients”, Ultrasonics, 41, 55-63, 2003.

22. F. M. Bai, N. G. Wang, J. F. Li, D. Viehland, G. Xu, and G. Shirane, J. Appl.

Phys. 96, 1620, 2004.

23. Z. Li, Z. Xu, X. Yao and Z.-Y. Cheng, “Phase transition and phase stability in [110]-, [001]-, and [111]-oriented 0.68Pb(Mg1/3Nb2/3)O3-0.32PbTiO3 single crystal under electric field”, J. Appl. Phys. 104, 024112, 2008.

24. E. T. Keve and A. D. Annis, “Studies of phases, phase transitions and

properties of some PLZT ceramics”, Ferroelectrics, 5, 77-89, 1973.

25. D. Viehland, J. Powers, L. Ewart, and J. F. Li, “Ferroelastic switching and elastic nonlinearity in <001>-oriented Pb(Mg1/3Nb2/3)O3-PbTiO3 and Pb(Zn1/3Nb2/3)O3-PbTiO3 crystals”, J. Appl. Phys. 88, 4907-4909, 2000.

Page 109: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

95

6 Summary

6.1 Summary

Relaxor-based ferroelectric single crystals of composition PMN-29%PT have

ultimate electromechanical coupling factors (k33 >90%), high piezoelectric

coefficients (d33>2000 pC/N) and high strain levels up to 1.7%, and therefore are

promising as replacements for conventional ceramics in wide range of

applications. However, because of the nature of single crystals, PMN-PT single

crystals are mechanically softer than PMN-PT polycrystalline ceramics [1].

Therefore PMN-PT single crystals more easily develop residual internal stresses

as a result of preparation processes, poling processes, and working loads, both

electrical and mechanical. The probability of crack initiation is strongly related to

the residual internal stresses. For this reason, it is important to investigate the

internal stress field during mechanical/electrical loading to better understand and

control the cracking problem. This thesis provides fundamental study of

birefringence induced by mechanical or electric field loading of PMN-29%PT

single crystals using optical methods. According to classical photoelasticity theory,

the birefringence is directly related to the internal stress field though its

proportionality to strain, therefore providing a visualization method to observe the

internal strain field. Strains induced from both electrical and mechanical loading

cause birefringent effects.

In Chapters 2, 3 and 4, classical photoelastic experiments including bending

and Hertzian contact loading experiments were conducted, and the results were

reported. In bending experiments, not only was the stress-optical coefficient C

estimated from three-point bending experiments, but also the load dependent

variation of the fringe-stress coefficient Cf /λ= was obtained through

four-point bending experiments. In all experiments, incident light wavelength λ

Page 110: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

96

is 535 . C has a value of 104X10nm -12 Pa-1, equivalent to an of value 5.14

X10

f

3 N/m, while varied in a range of 3.5X10f 3 N/m to 5.5X103 N/m. Notice

that the change of isn't monotonic: as load is increased, first decreases,

then at around the 14

f f

th fringe obtained (~45 MPa fiber stress), reached its

minimum value and began to increase with increasing load, as shown in Figure

4.8.

f

As reported in preliminary three-point bending experiments, the optical

property (characterized by fringe order) is linearly proportional to the mechanical

property (characterized by deflection) of unpoled PMN-29%PT single crystals. In

three-point bending experiments, Young’s modulus along the <001> direction was

calculated as 1.9X1010 N/m2, which is comparable to that obtained by Viehland

and Li [2]. However, this is only an average value. From observing the slope of

Figure 2.5, force versus deflection, the elastic property represented by the

nominal Young’s modulus is found first to have softened, then stiffened as load is

increased, which is consistent with the change of . Similar results were

reported by Viehland, et al. [3]. The explanations for these macroscopic

phenomena are microscopic strain-driven deformation processes, including

domain switching and phase transformation. This process consists of two

stages: first many small domains will freely switch in response to loadings to form

fewer large domains, and phase transformation from Rhombohedral to Monoclinic

phase may occur locally where internal strain is sufficiently high. During this

stage, the elastic modulus becomes softened. In the second stage, domain

switching and phase transformation saturate, thus the elastic modulus becomes

stiffened.

f

Because internal domain switching and phase transformation is induced by

external mechanical loading, the phenomenon is called mechanical poling.

Page 111: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

97

Regions being mechanically poled were observed to become very bright because

the transmission is greatly enhanced [4-5]. Notice that mechanical poling is a

2-dimensional effect compared to electrical poling, which is a 1- dimensional

effect, as shown in Figure 5.9. This means that after electrical poling, the

elastic stiffness is higher than that after mechanical poling. Thus electric field

loading may initiate or trigger cracks more easily than mechanical loading. This

is verified in Chapter 5.

In Chapter 5, electric field effects were investigated using birefringence

techniques. “Hertzian contact” electric field loading and electrical poling

experiments were performed with a DC field and a cyclic electric field, respectively.

Also for the cyclic electric field, frequency effects were explored.

From “Hertzian contact” experiments, the DC field was found to have a

deeper penetration than the cyclic electric field. In addition, the higher the

frequency of the cyclic field, the less penetration was observed in the fringe

pattern. This is due to extremely short time that each sign of the field is applied

to the material. Most of the electrical induced changes were close to the surface

and poled the local region to be very “bright”. The “bright” region has lost most of

its elastic flexibility and become susceptible to cracks. Thus cyclic field were

found more easily to initiate cracks in PMN-PT single crystals than DC field when

these samples were later subjected to mechanical loads.

From the electrical poling experiments, the DC field poling was observed to

first drive fringes to grow inside the sample at very low field magnitude of (1.1-1.3)

KV/cm, then, turning the entire sample turned dark and cloudy except regions

close to the surface as the field approached 5.4 KV/cm. This means that after

poling, a good practice would be to trim off a thin layer from the surface which has

been poled. When applying cyclic electric field to a sample previously poled with

DC field, the cyclic field immediately turned the sample to stripe patterns parallel

Page 112: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

98

to the electric field. These stripes are aligned domains. Furthermore cracks

were observed to develop from the edge towards the inside of the sample,

especially at highly strained regions, such as under the electrode, as shown in

Figure 5.7. From these results, we can draw the conclusion that a cyclic field

has a fatigue effect on PMN-PT single crystals, which may over-pole and crack

the material easily compared to DC field; thus a cyclic field shouldn’t be used for

electrical poling purposes.

Finally, orientation dependent optical and mechanical properties were

studied through Hertzian contact mechanical loading experiments and “Hertzian

contact” electrical loading experiments. Samples with three different orientations

were selected to perform the experiments: One is {100}-oriented on all six faces,

one with two side faces {100} and the other two side faces {110}, the last one with

all four side faces having {110} orientation. In Chapter 3, results of Hertzian

contact mechanical loading experiments were reported. Fringe pattern for all

three samples were totally different, and the loading direction was found to be an

important factor controlling the fringe pattern as well. Generally speaking, when

loaded from a <100> direction, the resulting fringe pattern is a two-lobed pattern;

when loaded from <110> direction, the resulting fringe pattern is a narrow

one-lobed pattern. Similar results were observed from “Hertzian contact”

electrical loading experiments. This shows that pseudo-cubic unpoled

PMN-29%PT single crystals have highly anisotropic elastic properties and that the

strain gradients giving rise to birefringence from mechanical and electrical field

loadings are similar.

Using cubic-form elastic constants calculated from data on poled

PMN-30%PT single crystals, ANSYS® simulation results were comparable to the

experimentally observed fringe patterns. This suggests that the elastic

properties of pseudo-cubic unpoled PMN-PT single crystals may resemble those

Page 113: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

99

of pseudo-tetragonal poled PMN-PT single crystals. There are two possible

explanations for the anisotropy of elastic properties. First, the seeded crystal

growth method with a unique growth direction may induce the highly anisotropic

elastic properties of unpoled PMN-PT single crystals, as reported by Sehirlioglu,

et al. [6]. Second, Zhang, et al. reported that domain wall motion may

contribute to the effective elastic constants, especially affecting the response of

PMN-PT in <110> directions [7]. This may explain why unpoled PMN-PT single

crystals show different fringe patterns when loaded from <100> and <110>

directions. That is, the fringe patterns depend not only on the elastic strain but

also on strains caused by domain switching.

As for the similar fringe pattern observed in the “Hertzian contact” electrical

loading experiments, because there is no piezoelectric strain coefficient reported

for unpoled PMN-PT single crystals, it is impossible to conduct ANSYS simulation.

Based on the similar fringe pattern, it is reasonable to conclude that the elastic

properties affect the anisotropic response through one of the linear electric-optical

effects: the electric field induced strain causes photoelastic effects which reflect

the anisotropy of the elastic properties that were explored by mechanical loading.

6.2 References 1. D. Viehland and J. F. Li, “Young’s modulus and hysteretic losses of 0.7

Pb(Mg1/3Nb2/3)O3-0.3PbTi O3”, J. Appl. Phys. 94, 7719, 2003. 2. D. Viehland and J. F. Li, “Stress-induced phase transformations in

<001>-oriented Pb(Mg1/3Nb2/3)O3–PbTiO3 Crystals: bilinear coupling of ferroelastic strain and ferroelectric polarization”, Philosophical Magazine, 83, 53-59, 2003.

3. D. Viehland, J. Powers, L. Ewart, and J. F. Li, “Ferroelastic switching and

elastic nonlinearity in <001>-oriented Pb(Mg1/3Nb2/3)O3-PbTiO3 and Pb(Zn1/3Nb2/3)O3-PbTiO3 crystals”, J. Appl. Phys. 88, No. 8, 4907-4909, 2000.

Page 114: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

100

4. X. Wan, J. Wang, H. L. W. Chan, C. L. Choy, H. Luo and Z. Yin, “Growth and optical properties of 0.62 Pb(Mg1/3Nb2/3)O3-0.32 PbTiO3 single crystals by a modified Bridgman technique”, Journal of Crystal Growth, 263, 251, 2004.

5. X. Wan, H. Luo, J. Wang, H. Chan, and C. Choy, “Investigation on optical

transmission spectra of (1-x) Pb(Mg1/3Nb2/3)O3-x PbTiO3 single crystals”, Solid State Communications, 129, 401, 2004.

6. A. Sehirlioglu, D. A. Payne, and P. Han, “Thermal expansion of phase

transformations in (1-x)Pb(Mg1/3Nb2/3)O3-xPbTiO3: evidence for preferred domain alignment in one of the (001) directions for melt-grown crystals”, Phys. Rev. B 72, 214110, 2005.

7. R. Zhang, W. Jiang, B. Jiang, and W. Cao, “Elastic, dielectric and

piezoelectric coefficients of domain engineered 0.7Pb(Mg1/3Nb2/3)O-0.3PbTiO3 single crystal”, Fundamental Physics of Ferroelectrics 2002, 188-197, 2002.

Page 115: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

101

Appendix Ι Basic theory of optical properties of crystals

This Appendix is to provide the background of basic theory of optical

properties of crystals, which is well documented in references [1] and [2].

Consider plane-polarized light passing through a crystal. There are in general

two waves, with different velocity, propagating through the crystal. The value of

for each wave is called the refractive index for that wave. Most

generally, if there are three principal axes

vc / n

zyx ,, of the dielectric constant tensor,

we can describe the refractive index by the refractive index ellipsoid, as shown

in Figure A1(a).

n

Figure A1: Representation of optical index ellipsoid; Illustration redrawn from a similar figure

in reference [1].

Now, for plane waves which propagate in the direction of any radius vector,

like the Z’ arrow in Figure A1(b), the two refractive indices are given by the

principal axes of the ellipse in which the ellipsoid is intersected by a plane

perpendicular to this radius vector, as the X’ and Y’ arrows shown in Figure A1(b).

The directions of these principal axes are the directions of the corresponding

electric vectors. The refractive index ellipsoid is defined by the equation:

Page 116: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

102

1=2

2

2

2

2

2

zyx nz

ny

nx

++ (A1-1)

xn , , are called the principal refractive indices, and yn zn ),,( zyxiKn ii == .

is the principal dielectric constant. Note the refractive index isn’t a tensor,

but the reciprocals of the square of can be treated as a tensor. If

iK n

n 2

1

ijn is

denoted by , the optical parameters, we have: ijB

(A1-2) 1=2332

222

11 zByBxB ++

which is called the optical index ellipsoid. In its most general form, it is given by:

(A1-3) 1=222 1231232

332

222

11 xyBzxByzBzByBxB +++++

Under an applied stress klσ , the Eq. (A1-3) changed to:

(A1-4) 1=222 1231232

332

222

11 xyBzxByzBzByBxB ′+′+′+′+′+′

So the changes are given by: ijBΔ

1,2,3),,,( ==−′=Δ lk jiBBB klijklijijij σπ (A1-5)

Here ijklπ is called the stress optical constant, which is a forth rank tensor with

units of m2/N. Since and ijB klσ are both symmetric tensors of second rank,

the ijklπ are not fully independent. ijlkijkljiklijkl ππππ == ; , therefore the number of

independent coefficients ijklπ is reduced from 81 to 36. Correspondingly,

because the strain is linear with the stress, photoelastic effects may also be

expressed in terms of the strains:

)3,2,1,,,( ==−′=Δ srjipBBB rsijrsijijij ε (A1-6)

Here is called the elasto-optical constant, and is dimensionless. and ijrsp ijrsp

Page 117: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

103

ijklπ are interrelated through and , the elastic constants and the

compliance constants. Using only one suffix let 11=1, 22=2, 33=3, 23=4, 31=5

and 12=6, Eq. (A1-5) and (A1-6) can be written as:

ijklC ijkls

)6...,,2,1,( ==−′=Δ nmBBB nmniim σπ (A1-7)

)6...,,2,1,( ==−′=Δ nmpBBB nmniim ε (A1-8)

For uniaxial (tetragonal, hexagonal and trigonal) crystals, the number of

refractive indices is two, which are those along the x and y axis; the refractive

indices are both the , while for the z axis, the refractive index is . and

are called ordinary and extraordinary refractive indices. Thus, the Eq. (A1-2)

can be written as:

on en on

en

12e

2

20

2

20

2

=++nz

ny

nx (A1-9)

If we let light travel along the z axis, i.e. the optical axis, the uniaxial crystals may

behave like an isotropic material, because the section of the ellipsoid

perpendicular to the light path is a circle.

Page 118: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

104

Appendix ΙΙ Basics of photoelasticity

This appendix is to provide the background of basic theory of photoelasticity,

which is also documented in references [3-4]. I will also summarize why a

circular polariscope can eliminate the isoclinic fringes. In photoelasticity, stress

fields are displayed through the use of light. Typically, plane-polarized light or

circular-polarized light are incident on the sample. The basic arrangement of

polarized microscope has already been discussed in Chapter 1. Here the

circular polariscope will be described, as shown in Figure A2(a). Two

quarter-wave plates, one with its axis at 450 and the other at 1350 were introduced

in the plane polarized microscopy configuration to achieve circular polarized

microscopy.

(a) (b)

Figure A2 Circular polariscope set-up, reproduced from a similar figure in reference [3].

To explain how a circular polariscope eliminates the isoclinic fringes, Jones

calculus needs to be introduced [3]. Generally speaking, an optical component

in a polariscope introduced both a rotation and retardation. In Jones calculus,

these can be represented as matrices.

First for a rotation, imagine an incident light ( , ), after passing through an

optical component, has been rotated by an angle

u v

θ . Then:

Page 119: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

105

θθ sincos vuu +=′ (A2-1)

θθ cossin vuv +−=′ (A2-2)

Namely, the light after rotation:

⎭⎬⎫

⎩⎨⎧

⎥⎦⎤

⎢⎣⎡−=

⎭⎬⎫

⎩⎨⎧

′′

vu

vu

θθθθ

cossinsincos (A2-3)

The following matrix is referred to as the rotation matrix:

(A2-4) ⎥⎦⎤

⎢⎣⎡− θθ

θθcossinsincos

Second, for representing retardation, let’s still assume an incident light (u , ).

To be general, assume:

v

)cos()cos(

22

11

αωαω

+=+=

tavtau

(A2-5)

If is the slow axis and is the fast axis, then the light coming out of the

medium with a retardation of

u v

δ can be represented as:

)

2cos(

)2

cos(

22

11

δαω

δαω

++=′

−+=′

tav

tau (A2-6)

Using complex number notations, the emerging light can be obtained as:

⎪⎭

⎪⎬⎫

⎪⎩

⎪⎨⎧

⎭⎬⎫

⎩⎨⎧

⎥⎥⎦

⎢⎢⎣

⎡=

⎭⎬⎫

⎩⎨⎧

′′ −

tii

i

i

i

eeaea

eeRv

u ωα

α

δ

δ

2

1

2

12

2

00

(A2-7)

Here R represents the real part, which we will deal with. Now, for a retarder,

with both rotation and retardation, we can represent it as follows:

⎥⎦⎤

⎢⎣⎡−⎥⎦

⎤⎢⎣

⎡ −

θθθθ

δ

δ

cossinsincos

00

2

2

i

i

ee (A2-8)

If we want to represent the light vector with respect to the original reference axes,

then:

Page 120: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

106

⎥⎥⎥

⎢⎢⎢

+−

−−=

⎥⎦⎤

⎢⎣⎡−⎥⎦

⎤⎢⎣

⎡⎥⎦⎤

⎢⎣⎡ − −

θδδθδ

θδθδδ

θθθθ

θθθθ

δ

δ

2cos2

sin2

cos2sin2

sin

2sin2

sin2cos2

sin2

cos

cossinsincos

00

cossinsincos

2

2

ii

ii

ee

i

i

(A2-9)

Namely any retarder can be represented by the matrix:

⎥⎥⎥

⎢⎢⎢

+−

−−

θδδθδ

θδθδδ

2cos2

sin2

cos2sin2

sin

2sin2

sin2cos2

sin2

cos

ii

ii (A2-10)

The retarder matrix can be used to represent a quarter-wave plate by substituting

and 0135=θ2πδ = , since quarter-wave plate provides a retardation of 2

π :

⎥⎦⎤

⎢⎣⎡

11

21

ii (A2-11)

In Figure A2(a), consider polarized light coming out the polarizer aligned parallel

to the x axis:

ti

y

x keEE ω

⎭⎬⎫

⎩⎨⎧=

⎭⎬⎫

⎩⎨⎧

01 (A2-12)

After entering the quarter-wave plate, we have:

ti

y

x keii

EE ω

⎭⎬⎫

⎩⎨⎧

⎥⎦⎤

⎢⎣⎡=

⎭⎬⎫

⎩⎨⎧

10

11

21 (A2-13)

After entering the sample, a phase difference of δ is generated for the light.

ti

y

x keii

ii

ii

EE ω

θδδθδ

θδθδδ

⎭⎬⎫

⎩⎨⎧

⎥⎦⎤

⎢⎣⎡

⎥⎥⎥

⎢⎢⎢

+−

−−=

⎭⎬⎫

⎩⎨⎧

01

11

2cos2

sin2

cos2sin2

sin

2sin2

sin2cos2

sin2

cos

21

(A2-14)

Just multiply the individual matrices of the various optical elements from the left

side of the original matrix in the order they are placed in the polariscope, we can

obtain the final output light. After another quarter-wave plate we have:

Page 121: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

107

ti

y

x keii

ii

ii

ii

EE ω

θδδθδ

θδθδδ

⎭⎬⎫

⎩⎨⎧

⎥⎦⎤

⎢⎣⎡

⎥⎥⎥

⎢⎢⎢

+−

−−

⎥⎦⎤

⎢⎣⎡−

−=⎭⎬⎫

⎩⎨⎧

01

11

2cos2

sin2

cos2sin2

sin

2sin2

sin2cos2

sin2

cos

11

21

(A2-15)

Upon simplification one gets:

ti

iy

x kee

EE ω

θδ

δ

⎪⎭

⎪⎬

⎪⎩

⎪⎨

=⎭⎬⎫

⎩⎨⎧

− 2

2sin

2cos

(A2-16)

In the dark field, when the analyzer is crossed with polarizer, the intensity of light

transmitted is obtained as the product of , which is simplified as: yy EE

2

sin 2 δad II = (A2-17)

Here represents the amplitude of the incident light. While in a bright field

with the analyzer parallel to the polarizer, we have the intensity of light , one

gets

aI

bI

2cos2 δ

ab II = (A2-18)

It is to be noted that no matter for dark field or bright field, the intensity equations

are independent of θ and hence the extinction condition is only a function of δ

and thus only isochromatics will be seen. The separating of isoclinics and

isochromatics is a significant achievement and greatly simplifies the analysis of

photoelastic fringe patterns.

Page 122: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

108

Appendix Ⅲ Calibration of in-situ loading frame

In this Appendix, I report the calibration of the cylinder loading system. To

calibrate the cylinder loading force from reading the pressure gauge, a calibration

system was designed and fabricated with all components as shown in Fig A3-5.

Figure A3 Top view of calibration stage

Figure A4 Side view of calibration stage

Page 123: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

109

Figure A5 Overview of loading frame

The loading system with the cylinder is first clamped on the stage; the loading

bar with a steel ball pinned on the head is put inside the front “hole” of the loading

system. As the pressure is increased, the cylinder pushes the brass to touch

the steel ball, thus the force is transferred to the load cell. Data for loading force

versus incremental pressure are obtained and plotted in Figure A5. Data clearly

shows a “straight line” as the pressure increases. All the data with pressure

increasing are displayed by black dots each with ± 2.5% error line. The fit line

falls within the error line, implying the force calculated

from function is within an error range of

5874.23254.0 −= xy

± 2.5%.

Page 124: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

110

Figure A6 Loading force versus pressure

Notice that the loading system can exert force over a range of 0-25 Newton.

In all of our experiments, the force exerted normally fell within a range of 0-15

Newton indicating the pressure we applied is within a range of 0-50 Psi. In

addition, to eliminate any hysteresis effects, we performed experiments with only

increasing pressure. Pressure was never decreased until the experiment was

completed.

Page 125: Photoelastic and Electro-Optic Effects: Study of PMN-29%PT

111

References for Appendices

1. T. S. Narasimhamurty, Photoelastic and Electro-Optic Properties of Crystals, Plenum Press, 1981.

2. J. F. Nye, Physical Properties of Crystals, Clarendon Press, Oxford,

1985.

3. K. Ramesh, Digital Photoelasticity Advanced Techniques and Applications, Springer, 2000.

4. M. M. Frocht, Photoelasticity, VI, New York, John Wiley & Sons, Inc.

1941.