hydrogen induced damage of lead zirconate titanate (pzt)

161
HydrogenInduced Damage of LeadZirconateTitanate (PZT) by ALI SHAFIEI MOHAMMADABADI A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY in THE FACULTY OF GRADUATE STUDIES (Materials Engineering) THE UNIVERSITY OF BRITISH COLUMBIA (Vancouver) April 2013 © Ali Shafiei Mohammadabadi, 2013

Upload: others

Post on 14-May-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

Hydrogen‐Induced Damage of Lead‐Zirconate‐Titanate (PZT)

by

ALI SHAFIEI MOHAMMADABADI

A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF

THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

in

THE FACULTY OF GRADUATE STUDIES

(Materials Engineering)

THE UNIVERSITY OF BRITISH COLUMBIA

(Vancouver)

April 2013

© Ali Shafiei Mohammadabadi, 2013

Page 2: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

ii

Abstract

Lead-Zirconate-Titanate Pb(Zr,Ti)O3 (PZT) based actuators are evaluated by

automotive industry for advanced fuel-injection systems, including hydrogen injection.

However, hydrogen can have deleterious effect on the PZT's functionality and properties.

The general objective of this work is to study the interactions between PZT and hydrogen.

The results of long-term (200-1200 hours) high-pressure (10 MPa) hydrogen exposure on the

PZT microstructure show that hydrogen has only superficial effects on the microstructure of

bare PZT. However, when an electrode is attached to PZT, the hydrogen damage increased; a

porous layer developed immediately adjacent to the electrodes on the PZT surface due to

hydrogen spillover. The kinetics of the PZT structural modifications due to hydrogen was

investigated by online monitoring of the electrical properties of PZT above the Curie

temperature, up to 650C. The results show that the structural changes can be described by

the classical nucleation and growth theory. The growth of the new structure appears to be

limited by the diffusion of protons into PZT, with a calculated activation energy of 0.440.09

eV, at 450-650C. Two relaxation peaks were observed in the dissipation factor curves of the

hydrogen-treated PZT. While the kinetics of one of the relaxation peaks obeys the classical

Arrhenius law with the activation energy of 0.66 eV, the other peak shows an unusual

relaxation kinetic. The mechanisms for the formation of these relaxation peaks are

determined. Low temperature (20C) diffusion of hydrogen into the PZT was also studied,

using the water electrolysis technique. Based on the microstructural observations, the

Page 3: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

iii

diffusion coefficient of hydrogen in PZT was calculated as 9×10-11 cm2/sec. The Maxwell-

Wagner polarization mechanism is determined to be responsible for the changes in the

hydrogen-affected PZT capacitance. In the last part of the project, alumina coatings were

applied to PZT plates using the sol-gel technique, to explore the possibilities of decreasing H2

damage to PZT. The functionality of the coating against hydrogen damage was evaluated by

the water electrolysis technique. Significant decrease of hydrogen damage was observed even

for highly porous coatings. The mechanisms by which the alumina coating decreases the

hydrogen damage were tentatively proposed.

Page 4: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

iv

Preface

This research work was conducted as part of a NSERC Strategic Project Grant

awarded to The University of British Columbia, Simon Fraser University and Westport

Innovations Inc. of Vancouver. The journal papers listed below have been prepared from the

work presented in the dissertation. I am the primary contributor to all of them, and the co-

authors contributions are as follows: my supervisor, Dr. Tom Troczynski, extensively

commented on the experimental and analysis methods, and the results interpretation in all

three papers. C. Oprea edited all three papers, and provided the SEM pictures in all three

papers. T. Nickchi contributed in the second paper by commenting on the experimental

setup. Dr. A. Alfantazi contributed in the second and third papers by providing and

commenting on the experimental setup.

1- A. Shafiei, C. Oprea, T. Troczynski, Investigation of the effects of high-pressure hydrogen

on Pb(Zr,Ti)O3 (PZT) ceramics, Journal of the American Ceramic Society, 2012, 95(2), 782–

787

2- A. Shafiei, T. Nickchi, C. Oprea, A. Alfantazi, T. Troczynski, Investigation of hydrogen

effects on the properties of Pb(Zr,Ti)O3 in tetragonal phase using water electrolysis

technique, Applied Physics Letters, 2011, 99 (21), 212903-212906

3- A. Shafiei, C. Oprea, A. Alfantazi, and T. Troczynski, In situ monitoring of the effects of

hydrogen on Pb(Zr,Ti)O3 structure, Journal of Applied Physics, 2011, 109 (11), 114108-

114116

Chapter 5-1 is based on paper “1”. Chapter 5-2 is based on paper “3”. Chapter 5-3 is

based on paper “2”. Please check the first pages of these chapters to see footnotes with similar

information.

Page 5: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

v

Table of Contents

Table of Contents

Abstract .......................................................................................................................................... ii

Preface ........................................................................................................................................... iv

Table of Contents ........................................................................................................................... v

List of Tables ................................................................................................................................ vii

List of Figures .............................................................................................................................. viii

Nomenclature .............................................................................................................................. xiii

Acknowledgements .................................................................................................................... xvi

Dedication .................................................................................................................................. xvii

1 Introduction ................................................................................................................................ 1

2 Literature Review ....................................................................................................................... 4

2.1 Lead Zirconate Titanate (Pb(Zr,Ti)O3) ............................................................................... 4

2.2 Hydrogen damage of PZT .................................................................................................... 9

2.3 Mechanisms of hydrogen damage of PZT ........................................................................ 12

2.3.1 Hydrogen incorporation into PZT ............................................................................. 12

2.3.2 Stable forms of hydrogen in PZT................................................................................ 22

2.3.3 Stable sites of H+ in PZT .............................................................................................. 24

2.3.4 Effect of hydrogen on PZT ......................................................................................... 26

2.4 Dielectric spectroscopy of PZT ......................................................................................... 31

2.5 High pressure hydrogen compatibility of PZT ................................................................. 33

2.6 Methods of decreasing the hydrogen damage to PZT ...................................................... 35

3 Scope and Objectives ................................................................................................................ 37

4 Materials and Methods ............................................................................................................. 40

4.1 Samples ............................................................................................................................... 40

4.2 Gas hydrogen treatment .................................................................................................... 42

4.3 Water electrolysis treatment of PZT ................................................................................ 47

4.4 Alumina sol-gel coating ..................................................................................................... 51

4.5 Characterization techniques .............................................................................................. 55

4.5.1 X-ray diffraction analysis (XRD) ................................................................................ 55

4.5.2 Scanning electron microscopy (SEM) ........................................................................ 55

4.5.3 Electrical properties measurements ............................................................................ 55

5 Results and Discussion .............................................................................................................. 57

Page 6: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

vi

5.1 High-Pressure conditions (T=100C, p=10 MPa, t=200-1200 hours)* ............................. 57

5.1.1 H2 effects on PZT microstructure ............................................................................... 57

5.1.2 H2 effects on the electrical properties of PZT ............................................................ 68

5.2 High-Temperature conditions (T=450-600C, p=0.013 MPa) * ....................................... 72

5.2.1 H2 effects on PZT microstructure ............................................................................... 72

5.2.2 H2 effects on PZT electrical properties ....................................................................... 75

5.3 Water-electrolysis treatment of PZT* ............................................................................... 96

5.3.1 Microstructure ............................................................................................................. 96

5.3.2 Electrical properties of PZT exposed to water electrolysis ..................................... 102

5.4 Ceramic Coatings for PZT Damage Protection .............................................................. 112

5.4.1 Alumina coatings microstructure ............................................................................. 112

5.4.2 Hydrogen resistivity of alumina-coated PZT .......................................................... 120

6 Conclusions ............................................................................................................................. 129

7 Future Work ............................................................................................................................ 133

References .................................................................................................................................. 136

Page 7: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

vii

List of Tables

Table 1- The heat and entropy of dissolution of hydrogen in different metals [22] ............... 17

Table 2- The EDX analysis for the bright particles in Figure 44b ............................................ 73

Table 3- Different values of exponents for the equation (28) [67] ............................................ 80

Table 4- The fitting values obtained for the equation (30) ....................................................... 82

Table 5- The fitting values obtained for the HN equation ........................................................ 91

Table 6- Microstructural charachteristics of alumina coatings a a function of TC [91]....... 117

Table 7- The EDX analysis for the -alumina coating (high concentration of Au is due to the

gold coating on the sample for SEM analysis) .......................................................................... 120

Page 8: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

viii

List of Figures

Figure 1- A typical polarization versus electric field (P-E) hysteresis for ferroelectrical

materials ......................................................................................................................................... 2

Figure 2- A schematic of an electronic fuel injector based on PZT actuators [3] ...................... 3

Figure 3- Binary phase diagram of PbZrO3-PbTiO3 system (refer to text for explanation of the

symbols) [5] .................................................................................................................................... 5

Figure 4- Unit cell for PZT (a) above and (b) below Curie temperature .................................... 6

Figure 5- TEM image showing domains in PZT [8] and a schematic of the 180 (1-1) and 90

(2-2) domain walls in PZT ............................................................................................................. 8

Figure 6- A ferroelectric ceramic with differently oriented domains inside each grain (a)

before and (b) after polarization, with remnant strain ................................................................ 9

Figure 7- Initiation and growth of hydrogen fissures during charging at 50 mA/cm2 for 2h (b)

and 4h (c); the black bar is 50 µm [12] ....................................................................................... 12

Figure 8- Different paths for the incorporation of hydrogen into PZT: (1) through the surface

of bare PZT (only at high temperatures) and (2) through the surface of electrodes (at low

temperatures). .............................................................................................................................. 14

Figure 9- Switching charge as a function of H2 annealing temperature with and without

upper Pt electrode [17] ................................................................................................................ 14

Figure 10- Remnant polarization as a function of the H2 annealing temperature for capacitors

with Pt, Pd, Au or Ag electrodes (applied voltage of 5V) [19] .................................................. 15

Figure 11- Solubility (cc of hydrogen per 100 g of metal) of hydrogen inside metals at 1

atmosphere pressure of hydrogen [24] ....................................................................................... 18

Figure 12- Schematic cross-section of the electrode/PZT assembly in H2 gas (it is assumed

that hydrogen diffuses only in the z direction).......................................................................... 19

Figure 13- Hydrogen diffusion coefficient in different metals [25] .......................................... 21

Figure 14- The stable positions of protons according to the reference [32] in a) perovskites

with large lattice constants, and b) perovskites with short lattice constants ........................... 25

Figure 15- PZT crystal structure showing the possible location of protons in the lattice of

PZT [31]; b) stable lattice site of protons for tetragonal PbTiO3 and c) cubic phase of PbTiO3

supposed by Park and Chadi [30] ................................................................................................ 26

Figure 16- Proposed PZT damage mechanisms can be categorized by the place where damage

occurs ............................................................................................................................................ 27

Figure 17- The microstructure of bare PZT plates .................................................................... 40

Figure 18- The microstructure of the PZT samples ................................................................... 41

Figure 19- Micrographs of the surface of the electrodes: silver (a) and silver-palladium (b) . 42

Figure 20- The time-temperature schedule of the ‘High-Pressure’ hydrogen treatment used

in this work .................................................................................................................................. 43

Figure 21- Schematic of an actuator made from PZT plates stacked together, b) the

equivalent electrical circuit of an actuator ................................................................................. 43

Page 9: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

ix

Figure 22- The schematic of the setup used in this work for online monitoring the electrical

properties ...................................................................................................................................... 44

Figure 23- The ‘High-Temperature’ hydrogen treatment used in this work ........................... 45

Figure 24- A typical Nyquist plot for the PZT at 500°C in air (the frequency is swiped

between 106 and 10-3 Hz) ............................................................................................................. 47

Figure 25- Schematic of the setup for the water electrolysis experiment ................................ 49

Figure 26- Schematic of the steps used for the preparation of the alumina coating ................ 52

Figure 27- Micrographs of PZT surface: a) the as-received sample; b) after 1200 h hydrogen

treatment ...................................................................................................................................... 58

Figure 28- Surfaces of PZT plates at higher magnifications: (a) before and (b) after 1200 h

hydrogen treatment ..................................................................................................................... 58

Figure 29- XRD results of bare PZT for as-received and after 1200 h hydrogen treatment ... 59

Figure 30- Cross-section of the as-received sample (a) in comparison to the cross section of

the sample after 1200 h hydrogen treatment (b) ....................................................................... 60

Figure 31- Cross section of the hydrogen treated sample for 1200 hours, close to the surface

....................................................................................................................................................... 60

Figure 32- Low magnification (a) and high magnification (b) images of damaged layer on the

PZT surface next to the Ag electrode after 400 hours hydrogen-treatment ............................ 61

Figure 33- The interface of the Ag electrode with PZT after grinding and polishing, for the

as-received sample (a) and for the sample hydrogen-treated for 400 hours (b); no detachment

of the electrode from the PZT and no damaged layer are visible ............................................. 63

Figure 34- The spillover mechanism of hydrogen atoms from the surface of the Ag electrode

to the surface of the PZT ............................................................................................................. 63

Figure 35- The detachment of the Ag electrode from the PZT for the sample treated for 600 h

....................................................................................................................................................... 64

Figure 36- Detachment of the Ag electrode from PZT; some cracks are present on the surface

of the electrode for the sample heat-treated for 1200 hours ..................................................... 64

Figure 37- Micrographs of the sample with Ag/Pd electrodes after hydrogen-treated for 200h:

the 1x10 mm side face (a) and its cross-section (b) .................................................................... 65

Figure 38- Surface of the side face of the sample with Ag/Pd electrode: (a) as-received, (b)

hydrogen-treated for 400 h; noticeable corroded area next to the electrode (c) ..................... 67

Figure 39- Micrograph of the cross-section of the sample shown in Figure 38 ....................... 68

Figure 40- Capacitance of PZT sample with Ag/Pd electrode in high-pressure hydrogen

atmosphere (a); at point ‘1’ the heater is on, and at point ‘3’ the heater is off. (b): capacitance

variation with temperature in hydrogen atmosphere ............................................................... 69

Figure 41- Capacitance of PZT sample with Ag electrode in high-pressure argon atmosphere

....................................................................................................................................................... 70

Figure 42- Capacitance of PZT sample with Ag electrode in high-pressure hydrogen

atmosphere ................................................................................................................................... 71

Figure 43- Image from the side surface of the PZT plate with Ag electrode for (a) as-received

and (b) after hydrogen treatment (for 2 h / 400C / p= 0.013 MPa) ......................................... 73

Page 10: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

x

Figure 44- Metallic lead in hydrogen treated PZT samples (for 2 h / 600C / p= 0.013 MPa) 73

Figure 45- The XRD pattern for as-received and hydrogen treated PZT with Ag electrodes 74

Figure 46- The XRD pattern for hydrogen treated PZT with Ag electrodes at 500C and

600C ............................................................................................................................................ 75

Figure 47- (a) The general trend of PZT capacitance variation with time in hydrogen

atmosphere at 500°C; (b) measurements for the real part of the impedance (ZRe) and the

calculated values of R according to equation (8) (the data is obtained at the constant

frequency of 1 kHz) ..................................................................................................................... 76

Figure 48- The ZRe -ZIm plot for PZT plate heat treated at 550C and the resistance

determined using the ZRe -ZIm plots for PZT; the noise in the ZRe-ZIm plots corresponds to the

times when the heater was on. ................................................................................................... 77

Figure 49- The variation of PZT capacitance at 530°C, 550°C and 600°C ................................ 78

Figure 50- The general trend for the isothermal α - time plots having different time steps,

time equal to zero shows the start of the reaction [68] ............................................................. 79

Figure 51- The results of fitting the capacitance data to equation (30) for the temperatures of

550C (a) and 600C (b) ............................................................................................................... 81

Figure 52- (a) The results of fitting the capacitance data to equation (30); (b) the activation

energy of hydrogen diffusion, obtained from the fit ................................................................. 82

Figure 53- The Grotthuss mechanism for diffusion of protons in PZT, including the

reorientation and hopping of protons between oxygen onions ................................................ 83

Figure 54- Hypothetical schematic of the different modes which can be assumed for the

dissolution of hydrogen in PZT; (a) where the diffusion of protons into PZT occur uniformly

from the surface; in this case the diffusion equation with proper initial and boundary

equation could be used for determining the total amount of protons in PZT; (b) where the

diffusion of protons can occur from limited places in the PZT; in this case the nucleation and

growth models can be used to describe the total amount of protons in PZT ........................... 84

Figure 55- Changes of capacitance C and dissipation factor DF of hydrogen-treated sample

(for 24 hrs / 550C / p= 0.013 MPa) versus temperature (The thick grey line shows the

changes of capacitance for as-received sample) ......................................................................... 86

Figure 56- Schematics of the dipolar polarization mechanisms, wherein direction of the

dipoles changes with changing the direction of applied voltage .............................................. 87

Figure 57- Schematics of the Maxwell-Wagner polarization mechanism, wherein differences

in the electrical properties of different regions cause charge accumulation at the interfaces

between the different regions, leading to the increase of capacitance ..................................... 88

Figure 58- Variations of ε’ and ε’’ for hydrogen-treated samples with the frequency of applied

voltage in the temperature range of 200-325C, with 25C increments .................................. 89

Figure 59- The results of fitting the ε´ and ε’’ data to the Debye equation for at T= 325C, for

PZT hydrogen-treated samples (for 24 hrs / 550C / p= 0.013 MPa) ........................................ 90

Figure 60- The results of fitting the ε´ and ε’’ data to the Havriliak–Negami equation for at

T= 325C, for PZT hydrogen-treated samples (for 24 hrs / 550C / p= 0.013 MPa) ................ 91

Page 11: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

xi

Figure 61- The activation energy for the ion jumping, obtained from the fits, for PZT

hydrogen-treated samples (for 24 hrs / 550C / p= 0.013 MPa) ................................................ 92

Figure 62- Variation of DF with frequency in the temperature range 22-42C, for PZT

hydrogen-treated samples (for 24 hrs / 550C / p= 0.013 MPa) ................................................ 94

Figure 63- Micrographs of the cross-section through PZT plate after water electrolysis: a)

low-magnification image (after 48 hours water electrolysis); b) microstructure of the

corroded layer (close to the electrode); c) microstructure in a region far from the corroded

layer .............................................................................................................................................. 97

Figure 64- The microstructure of PZT after water electrolysis just beneath the electrode

(after removing the electrode) .................................................................................................... 99

Figure 65- XRD pattern of the as-received PZT sample versus the water electrolyzed PZT

sample, using the following parameters: I=100 mA/cm2, t=48 hours ..................................... 100

Figure 66- The thickness of the corroded layer versus the square root of time of water

electrolysis .................................................................................................................................. 102

Figure 67- The changes of capacitance (C) and dissipation factor (DF) versus the duration of

water electrolysis at the frequency of 1 kHz ............................................................................ 103

Figure 68- Variations of electrical properties after 6 hours water electrolysis and subsequent

aging in air: a) capacitance (C); b) dissipation factor (DF) (: as-received,: after water

electrolysis, : after aging) ........................................................................................................ 104

Figure 69- Variations of electrical properties after 10 hours water electrolysis and subsequent

aging in air: a) capacitance (C); b) dissipation factor (DF) (: as-received,: after water

electrolysis, : after aging) ........................................................................................................ 104

Figure 70- Variations of capacitance (C) and dissipation factor (DF) after water electrolysis

for 48 hrs, and subsequent aging at room temperature in air (■: as-received, : after water

electrolysis, ▲:after 10 hours aging, : after 24 hours aging) ................................................ 105

Figure 71- The results of fitting the ε´´ (ε´´=DFε´) data to Debye and Havriliak–Negami

equation. An iterative MATLAB code was developed and used for the fitting procedure. .. 109

Figure 72- The changes of the capacitance (C) and dissipation factor (DF) for (a) a leaky

capacitor with electronic conduction, (b) for a capacitor with hopping charge carriers

adapted from [38] ....................................................................................................................... 109

Figure 73- Low magnification image of the coating on the surface of PZT (a) after dip coating

with pure boehmite sol (b) before and (c) after heat treatment of the coating in the furnace,

in some places on the surface of the coating, detachment of the coating was observed ....... 113

Figure 74- Low magnification image of the coating on the surface of PZT after dip coating

before the heat treatment of the coating in the furnace (comparing with Figure 73a, a smooth

uniform coating has formed on the surface of PZT with the addition of PVA to sol) .......... 114

Figure 75- Low magnification (a) and high magnification (b) images of the coating on the

surface of PZT after dip coating and after heat treatment of the coating in the furnace

(comparing with Figure 73b and c, a smooth uniform coating has formed on the surface of

PZT with the addition of PVA to sol) ...................................................................................... 115

Page 12: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

xii

Figure 76- Low magnification (a) and high magnification (b) images of the cross section of

the alumina coating. As it can be seen from (b), the coating had enough fluidity to fill out the

pores on the surface of PZT ...................................................................................................... 116

Figure 77- High resolution image of the cross section through the alumina coating processed

at 450C in air for 5 hours ......................................................................................................... 118

Figure 78- Transformation sequence of the different aluminum hydroxides with temperature

(adapted from [93]). ................................................................................................................... 119

Figure 79- XRD results for the as-received boehmite powder and after heat treatment at

450C for 5 hr ............................................................................................................................. 119

Figure 80- The cross section of the sample with Au-Pd electrodes and after 24 hours water

electrolysis. The thickness of the corroded layer is about 100 microns ................................. 121

Figure 81- The cross section of the sample with Au-Pd electrodes and alumina coating and

after 48 hours water electrolysis ............................................................................................... 122

Figure 82- Schematic for the reaction of hydrogen atoms with -alumina particles 1)

transformation of hydrogen atoms to hydrogen molecules which leave the system away from

the coating (i.e. as hydrogen bubbles during water electrolysis), 2) diffusion of hydrogen

atoms through the electrode and attachment to -alumina particles, followed by surface and

bulk diffusion through -alumina towards PZT ...................................................................... 123

Figure 83- An image of the cross section of PZT with alumina coating on top ..................... 124

Figure 84- The cross section of the sample with Au-Pd electrodes and thin alumina coating

and after 144 hours water electrolysis ...................................................................................... 125

Figure 85- Schematic image for the combination of hydrogen atoms at the interface of

metallic electrode with - alumina ........................................................................................... 125

Figure 86-Equivalent electrical circuit for PZT and PZT with coatings ................................ 128

Page 13: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

xiii

Nomenclature

Latin Symbols A Constant

a Area of electrodes

Hydrogen concentration

C Capacitance

C0 Concentration of hydrogen

D Diffusion coefficient

d Distance between electrodes

DF Dissipation factor

DPb Lead diffusion coefficient in PbTiO3

DM Diffusion coefficient of hydrogen in metallic electrode

DPZT Diffusion coefficient of hydrogen in PZT

E Electric Field

Ec Coercive Field

f Frequency

fMax Frequency at which the maximum occurs

fugacity of H2

g Gravity

G Conductance

h Coating thickness

hH Partial enthalpy

ΔH Activation Energy

ΔHM Heat of dissolution of hydrogen inside metal

i Imaginary number

ic Charging current density

JH Flux of hydrogen atoms

k Coefficient in equation (10)

KIC Fracture toughness of the un-affected PZT

KIH Fracture toughness of hydrogen-treated PZT

m Constant

n Constant

nis Number of interstitial sites per metal atom

P Polarization

p Pressure

P0 Constant

p0 Pressure at standard conditions

PH2 Partial pressure of hydrogen

Pr Remnant Polarization

Ps Spontaneous Polarization

Q Activation energy

R Ohmic resistance

SH Non-configurational part of entropy

ΔSM Dissolution of hydrogen in different metals

Page 14: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

xiv

t Time

T Temperature

Tc Curie temperature

Y Admittance

x Thickness of corroded layer

Z Impedance

ZIm Imaginary part of impedance

ZRe Real part of impedance

Greek Symbols θ Constant

α Fraction of the volume converted to the product of reaction

LV Sol-vapor surface energy

β Constant

ε Dielectric constant

ε0 Vacuum permittivity

τ Relaxation time

0 Constant

ε´ Real part of dielectric constant

ε´´ Imaginary part of dielectric constant

εs Dielectric constant when 0

ε Dielectric constant when

Viscosity of the sol

Density of the sol

Withdraw speed

( ) Chemical potential of gaseous hydrogen per molecule

( ) Chemical potential of hydrogen atom dissolved in the metallic electrode per atom

( ) Chemical potential at a given standard state

( ) Chemical potential at a given standard state

Angular frequency

Abbreviation

AO Orthorhombic phase

EDX Energy-dispersive X-ray spectroscopy

FCC Face-centered cubic lattice

FM Monoclinic phase

FR Rhombohedral phase

FT Tetragonal phase

KTN Potassium tantalate niobate

MPB Morphotropic phase boundary

MW Maxwell-Wagner polarization mechanism

PVA Polyvinyl alcohol

PZT Lead zirconate titanate

Page 15: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

xv

Abbreviation (continue)

Pc Cubic perovskite structure

SEM Scanning electron microscope

TEM Transmission electron microscope

TC Curie temperature

XRD X-ray diffraction

Page 16: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

xvi

Acknowledgements

I would like to take this opportunity to express my utmost gratitude toward my

supervisor, Dr. Tom Troczynski for his continuous trust, patience, and guidance throughout

the course of this work. I also acknowledge Prof. Akram Alfantazi, Prof. Guangrui Xia, and

Prof. Steve Cockcroft for their valuable comments.

My sincere thanks go to Carmen Oprea, who has been a valuable and amazing

colleague and friend, and has been continuously supporting and helping me throughout the

years I have been at UBC. She taught me to write “in comparison to” instead of “in compare

to”. She taught me to write “reach” instead of “reach to”. Thank you so much Carmen!

I would like to thank all staff members in the Department of Materials Engineering at

The University of British Columbia for their assistance with my research work. My special

thanks to all colleagues and officemates for providing a friendly environment that I was

always pleased to work in. Natural Sciences and Engineering Research Council of Canada

(NSERC) and Westport are greatly acknowledged for financial support.

Special thanks are owed to my parents, who have supported me throughout my years of

education.

Page 17: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

xvii

Dedication

To my patient parents

Page 18: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

1

1 Introduction

Lead Zirconate Titanate (Pb(Zr,Ti)O3) or PZT is the general name for the perovskite

solid solutions between PbZrO3 and PbTiO3. PZT is well known because of its unique

electrical properties such as high dielectric, piezoelectric and electro-optic coefficients. The

main reason for having such interesting electrical properties is the unique crystal structure

and the arrangement of ions inside unit-cell of PZT. In the unit-cell of PZT, the titanium or

zirconium ions reside off-center in the octahedral interstitial positions surrounded by six

oxygen ions. Therefore, the center of negative charge of oxygen ions will not coincide with

the center of positive charge of Ti or Zr ions and this arrangement results in permanent

dipoles inside the unit-cell of PZT. These built-in permanent dipoles in the unit-cell of the

crystal structure of PZT are the origin of the superior electrical properties of PZT.

Application of an electric field of sufficient magnitude will cause these built-in

dipoles inside the PZT to switch to a different, stable direction in accordance to the direction

of the applied electric field. Moreover, by removal of the electric field, the dipoles will not

return to their original direction. This brings up one of the very interesting properties of

PZT, which is the switchable Polarization (P)-Electric Field (E) hysteresis, as schematically

shown in Figure 1. This property of PZT is generally known as ferroelectricity, and it has

also been seen in other oxides like BaTiO3. This property of PZT has enabled the extensive

use of these materials in applications such as ferroelectric random access memories (FeRAM).

It is known that hydrogen treated PZT may not show this hysteresis anymore. Although the

Page 19: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

2

main reason for this phenomena is not known very well, this has been attributed to the

formation of [OH]– dipoles, which inhibits the switching of the spontaneous dipoles in PZT.

This effect of hydrogen on the polarization hysteresis of PZT has been the main reason for

investigating the effect of hydrogen on PZT since 1995 [1-2].

Figure 1- A typical polarization versus electric field (P-E) hysteresis for ferroelectrical materials

In addition to the ferroelectrical properties, a “poled” PZT ceramic (i.e. PZT with

oriented dipoles) can also show superior piezoelectric properties, which have caused the

extensive use of PZT in other applications, such as actuators and sensors. Recently, attention

has been paid to the effect of hydrogen on the piezoelectric properties of PZT as well, as it

has been suggested that hydrogen might also have deleterious effects on these properties [3].

For example, Figure 2 schematically shows a modern electronic fuel injector that uses PZT

E(V/m)

P (C/cm2)

Pr

Ps

Ec

Ps :spontaneous polarizationPr :remnant polarizationEc :coercive field

Ec

Page 20: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

3

actuators for valve opening, instead of the conventional solenoid technology. These fuel

injectors have been introduced by the leading engine manufacturers in recent years. One of

the issues of using such fuel injectors in a hydrogen atmosphere is the possible deleterious

effects which hydrogen may have on the functionality of the PZT actuators [3].

Figure 2- A schematic of an electronic fuel injector based on PZT actuators [3]

The above points provide rationale for the investigation of the interactions of

hydrogen with PZT, which is important topic both from practical and scientific points of

view. The objective of the present work is to address some of the issues regarding the

interaction of hydrogen and PZT. More specifically, we focused on the kinetics of

degradation of PZT properties by hydrogen. Attempts have also been made to propose

techniques to inhibit or decrease the hydrogen damage to PZT. In this regard, the sol-gel

technique was used to develop hydrogen barrier coatings on the surface of PZT.

Page 21: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

4

2 Literature Review

2.1 Lead Zirconate Titanate (Pb(Zr,Ti)O3)

PZT is the general name for the perovskite solid solutions between PbZrO3 and

PbTiO3, as shown in the binary phase diagram in Figure 3. Starting at PbTiO3 part of the

diagram with a ferroelectric tetragonal (FT) phase, by increasing the amount of PbZrO3 in the

solution, the composition Pb(Zr0.53Ti0.47)O3 is reached, where the ferroelectric tetragonal

phase starts to transforms into another ferroelectric phase, however, with a different crystal

structure (rhombohedral phase (FR)) [4]. The composition where the tetragonal phase (FT)

transforms to the rhombohedral phase (FR) is considered to be the morphotropic phase

boundary (MPB) in the PZT binary phase diagram (Figure 3). Instead of a sharp boundary,

the MPB is often observed in real systems as a region of phase coexistence whose width

depends on the compositional homogeneity and on the sample processing conditions [5].

Therefore, the location of this boundary has not been determined exactly, as its positions

changed from report to report. However, the recent work by Noheda et al. [5] has changed

our understanding of the MPB in PZT. Their studies show that instead of a boundary, a low-

symmetry monoclinic phase (FM) exists between the tetragonal and rhombohedral phases and

the superior electrical properties of PZT around MPB are actually due to the very high

polarization in this phase (Figure 3). In other words, the monoclinic phase acts as a bridge for

the phase transformation from tetragonal to rombohedral and vice versa.

Page 22: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

5

The rhombohedral part of the diagram itself consists of two other phases: a high-

temperature phase (FR(HT)) and a low-temperature ferroelectric rhombohedral (FR(LT)) phase.

The PZT structure close to the PbZrO3 part of the diagram up to about 10 %mole of PbTiO3

(Figure 3) has an antiferroelectric, orthorhombic (AO) structure. This phase of PZT has a very

complex structure, in the sense that the displacement of cations along the [110] direction is

coupled with octahedral tilts [4]. According to Figure 3, we can see that depending on its

composition, PZT can have different crystal structures, and consequently PZT can show

different piezoelectric, pyroelectric and electro-optic coefficients. The great technological

and commercial importance of PZT is actually due to such variations in the electrical

properties, which can be obtained by changing the PZT composition. Especially when the

PZT is doped with other secondary ions, it can show even more interesting electrical

properties [4]. The most widely used PZT ceramics today have the compositions near the

MPB composition. For example, Pb(Zr0.53Ti0.47)O3 is the composition for the PZT ceramics

used in this work.

Figure 3- Binary phase diagram of PbZrO3-PbTiO3 system (refer to text for explanation of the symbols) [5]

Tem

per

atu

re (

°C)

100

200

300

400

500

4020 60 80 1000

PbZrO3 PbTiO3

FT

PC

FR(HT)

FR(LT)

AO

MPB

Mole % PbTiO3

FM

Page 23: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

6

The cubic perovskite structure of PZT (Pc) above Curie temperature is shown in

Figure 4a, wherein each lead ion is surrounded by 12 oxygen ions. The oxygen ions plus the

lead ions form a face-centered cubic (FCC) lattice. The titanium or zirconium ions reside in

the octahedral interstitial positions surrounded by six oxygen ions. Temperature has a strong

effect on the cubic structure shown in Figure 4a. When the temperature decreases to about

375C, the structure changes to the tetragonal shown in Figure 4b. The octahedral site is now

distorted, with the Ti or Zr in off-center positions, resulting in a permanent dipole. This

"built-in" permanent dipole in the unit cell of the crystal structure of PZT is the origin of the

superior dielectrical and piezoelectrical properties of PZT. This is sometimes called a

spontaneous polarization. The temperature of transformation from the cubic to the

tetragonal phase is called the Curie temperature (Tc).

Figure 4- Unit cell for PZT (a) above and (b) below Curie temperature

Pb

O

Ti or Zr

Above TC

(a)Below TC

Dipole direction

(b)

Page 24: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

7

During the transformation from the cubic to the tetragonal phase, permanent dipoles

form inside the PZT crystal; however, the direction of these dipoles is not the same

throughout the whole crystal or inside each PZT grain. The regions inside PZT crystals

where the dipoles are aligned in the same direction are called ferroelectric domains. As soon

as spontaneous dipoles start to form during the phase transformation from the cubic to the

tetragonal, surface charges start to appear on the surface of PZT. Such surface charges

produce very high electric fields in the order of MV m−1 [7]. Therefore, the electrostatic

energy of the system increases due to the existence of such electric fields. The electrostatic

energy associated with these electric fields can be minimized if the PZT crystal can split into

separate ferroelectric domains, and this is one of the reasons why ferroelectric domains form

inside the PZT crystals [7].

Another reason for the formation of ferroelectric domains inside the PZT crystals is to

minimize the elastic energy associated with the mechanical constraints to which the PZT

crystal is subjected as it is cooled, through the paraelectric–ferroelectric phase transition [7].

To better understand this, assume that a part of the PZT crystal is subjected to compression

forces while cooling down from the cubic phase to the tetragonal phase. At the phase

transformation temperature, spontaneous dipoles start to form inside PZT; however, they

will form in the directions perpendicular to the compression forces, in order to minimize the

elastic energy of the system. Therefore, the mechanical forces, and the elastic energy of the

system are minimized by the fragmentation of the grains into separate domains [7]. As a

result, a complex domain structure develops in each grain, according to the allowed

Page 25: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

8

directions of the polarization in each domain. In the tetragonal phase, the so-called separated

180 and 90 domain walls form (Figure 5) [7].

Figure 5- TEM image showing domains in PZT [8] and a schematic of the 180 (1-1) and 90 (2-2) domain walls

in PZT

In each ferroelectric domain inside a grain of a ferroelectric ceramic, a uniform

orientation of the dipoles exists. It should be noted that since the grains and the domains

contained in them are randomly oriented, the properties of the ferroelectric ceramics are

isotropic both after the synthesis and after the cooling below the Curie temperature [7]. By

applying an electric field with sufficient magnitude, the spontaneous polarization inside the

ferroelectric ceramic can switch to a different, stable direction. By stable direction we mean

that the spontaneous polarizations will not return to its original direction and magnitude

when the electric field is removed, as shown in Figure 6. This brings up the most important

characteristic property of PZT, which is the switchable Polarization (P)-Electric Field (E)

hysteresis, as schematically shown in Figure 1.

1

1

2

2

Page 26: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

9

Figure 6- A ferroelectric ceramic with differently oriented domains inside each grain (a) before and (b) after

polarization, with remnant strain

2.2 Hydrogen damage of PZT

Four different kinds of degradation of the electrical properties after H2 treatment have

been reported for PZT: 1) loss of polarization hysteresis [9], 2) increase in leakage current [9],

3) drop in resistivity [10], and 4) decrease in dielectric constant [9]. To confirm that this

degradation is not just due to the reducing nature of the atmosphere (usually a mixture of

nitrogen and hydrogen), the ferroelectric capacitor characteristics were also measured after

the heat treatment in nitrogen and they were unchanged. Therefore, the degradation is due

to the hydrogen in the atmosphere entering PZT and altering atomic structure of PZT.

The decrease in resistivity after H2 treatment has also been reported: the resistivity of

as-grown PZT drops from 5×1011 Ωcm to 2×107 Ωcm after annealing in forming gas (a

mixture of up to 5.7% hydrogen and nitrogen) at 400°C for 30 min [10]. Although strong

changes have been observed in the polarization hysteresis characteristics and leakage current

(a) (b)

remnant strain

Page 27: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

10

of the hydrogen-treated Pt/PZT/Pt ferroelectric capacitors, their relative dielectric constant

measured for a low-voltage signal was not greatly affected; it was about half of the as-

received sample, i.e. 800. This may indicate that the PZT film was not damaged completely

through its thickness [9]. The degradation mechanisms will be discussed in the next section,

and it will be shown that the deterioration occurs mostly at the interface between PZT and

electrode.

There are just a few papers regarding the degradation of the mechanical properties of

the PZT after hydrogen treatment [11-13]. The general conclusions drawn from these papers

are the following:

-The cohesive strength of PZT decreases due to the presence of H or H+ in the lattice.

-Recombination of H or H+ at grain boundaries and in micro-voids may form molecular

hydrogen; when the internal pressure of H2 exceeds the cohesive strength of the grain

boundaries, fissures or micro-cracks appear.

Delayed hydrogen-induced failure has been reported for PZT ceramics during charging by

hydrogen under a constant load [11]. Therefore, it can be concluded that hydrogen atoms

incorporated into the structure of PZT can decrease the cohesive strength of PZT, although

they also have a considerable effect on ferroelectric properties. Wang et al. have studied

hydrogen induced delayed fracture of PZT [11]. Their results show that the strength of PZT

decreases with increasing the hydrogen concentration inside the specimen. Moreover, they

have also found that the KIH/KIC decreases with the hydrogen concentration in the specimen,

C0, in the form of KIH/KIC = 0.400-0.155ln(C0) where C0 is the concentration of hydrogen

Page 28: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

11

which can diffuse out from the samples after charging, KIH is the fracture toughness of

hydrogen-treated sample and KIC is the fracture toughness of the un-affected sample. They

have reported that the fracture mostly occurred in inter-granular mode, which shows that

cracks were mainly initiated at the grain boundaries.

Peng et al. have investigated the initiation and propagation of hydrogen fissures in a

PZT ferroelectric ceramic during charging by hydrogen without loading [12]. Their results

show that when hydrogen concentration in PZT exceeds a certain value (about 260 ppm),

hydrogen fissures or micro-cracks form within PZT. Figure 7, reproduced from [12], shows

the initiation of such micro-cracks at grain boundaries. Usually a typical sintered PZT

ceramic is not 100% dense, and there are many voids and porosities at the grain boundaries.

The recombination of hydrogen atoms at such porosity or voids results in increasing pressure

of the molecular hydrogen (H2) inside such holes, and when the hydrogen pressure inside

these voids or porosities becomes equal to the strength of PZT at the grain boundary, which

has been decreased by the presence of atomic hydrogen, hydrogen fissures or microcracks

form [12].

Page 29: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

12

Figure 7- Initiation and growth of hydrogen fissures during charging at 50 mA/cm2 for 2h (b) and 4h (c); the

black bar is 50 µm [12]

2.3 Mechanisms of hydrogen damage of PZT

2.3.1 Hydrogen incorporation into PZT

Hydrogen damage has been linked to structural modifications of PZT, which lead to

the changes in the properties of PZT [1-2]. This could occur due to chemical reactions

between hydrogen and constituents of PZT, or simply due to the presence of hydrogen (in

different forms of ions, atoms, or molecules) within PZT. Before discussing these

mechanisms, we first need to address two important issues: 1) the paths for the incorporation

of hydrogen in PZT and 2) the stable forms of hydrogen inside the lattice of PZT (H or H+).

Page 30: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

13

There are many uncertainties regarding the second issue, and in the following sections we

will review the results reported in other studies.

2.3.1.1 Mechanism

In general two different scenarios can be considered. The first mechanism is that

hydrogen molecules dissociate at the PZT surface, and as a result H atoms diffuse into the

structure (Figure 8 ).

Generally, such mechanism can be active in bare (no electrodes) PZT crystals, as

observed in a few experiments [9, 14-15]. However, most oxides have limited hydrogen

diffusivity [16], and the results show that hydrogen can be incorporated into the PZT

structure only at temperatures higher than 400°C [17].

Another possible mechanism, especially at temperatures as low as 200°C, is the

incorporation of H into the structure from the metallic electrode. Hydrogen molecule

dissociates at the surface of the electrode, and thus produced H atoms diffuse to the

electrode/PZT interface, then continue into the PZT crystal and modify the PZT structure

(Figure 8). Indeed, this is the most probable mechanism, and it has been reported in many

studies [9, 14, 18-21] as the degradation of PZT properties can be correlated with the

properties of electrodes. For example, the PZT-electrode assembly with In2O3 electrode

showed the least amount of degradation when subjected to hydrogen atmosphere, whereas Pt

is the worst electrode [21], due to the highly catalytic nature of Pt helping in dissociating the

hydrogen molecules into atoms.

Page 31: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

14

H2HH

HH

H2

H

H

Electrode

PZT

2

1

Figure 8- Different paths for the incorporation of hydrogen into PZT: (1) through the surface of bare PZT (only

at high temperatures) and (2) through the surface of electrodes (at low temperatures).

Figure 9 [17] shows that degradation of PZT with Pt electrodes is more severe in

comparison to the bare PZT, which clearly confirms the catalytic behavior of Pt. Figure 10

[19] shows the effects of different electrodes on the remnant polarization (Pr). It is clear that

capacitors with Au or Ag electrodes are much more stable during H2 annealing than those

using Pt or Pd. The difference in the level of H2 damage corresponds to the difference in the

catalytic activity in the hydrogenation reaction and the adsorptive properties of hydrogen by

metallic electrodes [19].

Figure 9- Switching charge as a function of H2 annealing temperature with and without upper Pt electrode [17]

Page 32: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

15

Figure 10- Remnant polarization as a function of the H2 annealing temperature for capacitors with Pt, Pd, Au or

Ag electrodes (applied voltage of 5V) [19]

2.3.1.1 Interactions between hydrogen and metallic electrodes

The incorporation of hydrogen in PZT through the metallic electrode includes the

steps of (i) hydrogen absorption into the metallic electrode and dissociation of hydrogen

molecules on the surface of the metallic electrode, and (ii) diffusion of hydrogen atoms

through the metallic electrode and into PZT (Figure 8). The absorption of hydrogen

molecules (H2) into metallic electrode via the gas phase can be described by the following

chemical reaction [22]:

( ) (1)

where [H] refers to hydrogen atoms dissolved in the metallic electrode. At equilibrium, the

chemical potential of hydrogen in the gas phase is equal to the chemical potential of

hydrogen dissolved in the metallic electrode, therefore:

( ) ( ) (2)

Page 33: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

16

where ( ) is the chemical potential of gaseous hydrogen per molecule and ( ) is the

chemical potential of hydrogen atom dissolved in the metallic electrode per atom.

The chemical potential of gaseous hydrogen can be written as follows:

( )

( ) (

) (3)

where ( ) refers to the chemical potential at a given standard state, is the Boltzmann

constant, is the fugacity of H2 and is the pressure at standard conditions (1 bar);

is

defined as the activity of hydrogen ( ). At pressure of hydrogen below 10 MPa [22-23]

hydrogen can be considered as an ideal gas; therefore, will be equal to

, the partial

pressure of hydrogen. Therefore, the chemical potential of gaseous hydrogen can be written

as follows:

( )

( ) (

) (4)

On the other hand, the chemical potential of hydrogen in the metallic electrode can be

written as [22]:

( ) ( ) ( ( )) (5)

where ( ) is the chemical potential at a given standard state, and ( ) is the activity of

hydrogen in the metal. The chemical potential of atomic hydrogen in metals can also be

written as [23]:

( ) ( (

)) (6)

where is the partial enthalpy, is the non-configurational part of entropy, is the

number of interstitial sites per metal atom, and is the number of hydrogen atoms per

Page 34: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

17

metal atom (i.e. hydrogen solubility, expressed as atomic fraction). Taking into account

equations (2), (4) and (6), the following relations can be obtained for the hydrogen solubility

into the metallic electrode:

(

)

( ) (7)

or

(

)

(8)

where is the heat of dissolution of hydrogen inside the metal, and is the entropy of

mixing. The values of and for the dissolution of hydrogen in different metals can be

obtained experimentally, and the literature data are reported in Table 1. The value of

depends on the electronic structure of the metal in which hydrogen is being dissolved, and

the is predicted by theoretical studies to be around -7.8 [22].

Table 1- The heat and entropy of dissolution of hydrogen in different metals [22]

Metal ([eV per atom]) / T(C)

Fe (bcc) 0.25 -6 <900

Al 0.70 −6 500

Ni 0.17 −6 350–1400

Pd 0.1 −7 −78–75

Pt 0.48 −7 −

Cu 0.44 −6 <1080

Ag 0.71 −5 550–961

Au 0.37 −9 700–900

U(α) 0.1 −6 <668

Page 35: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

18

For small hydrogen concentrations ( ), equation (7) can be written as:

(

) (

) (9)

or

(10)

This relation is known as Sieverts’ law [22-23], which states that the solubility of

hydrogen in metals is proportional to the square root of the partial pressure of the hydrogen

in equilibrium with the metal. is known as the solubility coefficient. Sieverts’ law is also

applicable for other diatomic gases (eg. N2, O2) [22]. Figure 11 shows the experimental data

on hydrogen solubility in various metals as a function of temperature [24].

Figure 11- Solubility (cc of hydrogen per 100 g of metal) of hydrogen inside metals at 1

atmosphere pressure of hydrogen [24]

Let us consider the role of the metallic electrode on the level of hydrogen absorption

by PZT. The case under consideration is the metallic electrode attached to PZT, and both are

4 8 12 16 20 24

10000/T K-1

-8

-4

0

4

8

Ln

cH

T (C)

Cu

Fe

Ag

Pt

Ni

Pd

Al

2004006001000

Page 36: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

19

in the hydrogen atmosphere (Figure 12). In this case, the hydrogen adsorption in PZT

includes the steps of (i) the hydrogen absorption into the metallic electrode and its

dissociation, and (ii) diffusion of hydrogen atoms through the metallic electrode and into

PZT (it is assumed that hydrogen diffusion is just in the z direction). The chemical potential

of hydrogen in the metallic electrode ( ) and PZT ( ( )) can be described by [25]:

( ) (

) (11)

where is the hydrogen concentration at the partial enthalpy of , is the gas constant

and stands for (metal) or .

Figure 12- Schematic cross-section of the electrode/PZT assembly in H2 gas (it is assumed that hydrogen diffuses

only in the z direction)

During hydrogen diffusion, the following boundary equilibrium conditions can be

considered [25]:

at interface A :

( ) ( ) (12)

at interface B : ( ) ( ) (13)

It should be noted that both ( ) and ( ) are functions of time and space; that is

because during the hydrogen diffusion, when the system is not in equilibrium, is changing

H2 (g)

electrode

PZT

A

Bz

Page 37: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

20

with time and space, and therefore both ( ) and ( ) are changing with time and

space [25]. Comparing equations (11) and (13) the following relation can be written:

(

) (

) (14)

By adding the term

( ) to both sides of equation (14), the following relation is obtained

[25]:

(

) (

) (15)

where and are the heat of hydrogen dissolution in metallic electrode and PZT,

respectively. Equation (15) shows that the , the hydrogen concentration inside PZT,

depends on , and . According to this equation, we expect that with changing

the metallic electrode, different amounts of hydrogen would dissolve in PZT. The different

level of hydrogen dissolution in PZT due to the different electrodes is therefore the reason

for the dependency of the damage affected by hydrogen on the metallic electrode (Figure

10). Another important boundary condition which must be satisfied in order to conserve the

H atoms is the equality of the flux of hydrogen atoms which leave the metallic electrode to

the flux of hydrogen atoms which enter the PZT at the interface B ( ) [25]. In other words,

at the interface B, the following equation must be satisfied:

( )

( )

(16)

where is the diffusion coefficient of hydrogen in the metallic electrode, and is the

diffusion coefficient of hydrogen in PZT. According to equation (16), and are other

parameters which control the hydrogen entry to PZT. Therefore, based on equations (15)

Page 38: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

21

and (16), one can conclude that the hydrogen incorporation, hence the damage to PZT,

indeed depends on , and , i.e. on the type of the electrode used.

According to equation (16), the rate of hydrogen atoms which leave the metallic

electrode increases with . Moreover, the time needed for the hydrogen atoms to reach the

interface between the electrode and PZT from the surface of the electrode depends also on

. However, often the electrodes used in making the PZT capacitors, e.g. for FeRAMs, are

very thin (about 10 nm). Therefore, we expect that these diffusion times (<1 sec) are much

smaller than the typical duration (30 min) of hydrogen treatments (the depth of hydrogen

diffusion ( ) from the surface of the electrode can be estimated by √ . If = 10

nm, and if we assume to be equal to 10-5 cm2/sec at 150C Figure 13, then the time for

hydrogen atoms to reach the interface between the electrode and PZT will be on the order of

microseconds). Figure 13 shows the hydrogen diffusion coefficient for different metals as a

function of temperature [25].

Figure 13- Hydrogen diffusion coefficient in different metals [25]

4 8 12 16 20 24 28

10000/T K-1

0

0.2

0.4

0.6

0.8

D (

cm2 /

sec)

10

-4

T (C)

Cu

Au

Ag

Pt

Ni

Pd

2004006001000

Page 39: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

22

In summary, the role of metallic electrodes on the damaging effect of hydrogen on

PZT can be evaluated by considering the system parameters such as hydrogen diffusivity ,

heat of hydrogen dissolution in the metal , and hydrogen concentration in the metal

. If we want to predict the amount of hydrogen damage with different electrodes, not

only the above parameters, but their interplay should be considered as well; moreover, the

respective PZT parameters ( , , ) should be considered. For example, by

considering Figure 11 and Figure 13, one can see that the hydrogen diffusion and hydrogen

solubility are higher in Pd than in Pt. Therefore, one may expect that Pd may cause more

degradation to PZT in comparison to Pt. However, greater damage has been observed in the

samples with Pt electrode (Figure 10), possibly due to the higher heat of hydrogen solubility

in Pt than in Pd.

2.3.2 Stable forms of hydrogen in PZT

In general there are three possible forms of hydrogen in oxides; it can exist as a

hydrogen atom (H), as a hydrogen ion (H+ or H-) or as a hydrogen molecule (H2). As an atom

H, it will just fill the interstitial sites of the lattice, with no interaction with other elements

of the structure, especially with oxygen anions. These hydrogen atoms can diffuse freely

through the structure without changing the electrical properties of the oxide [27]. On the

other hand, hydrogen atom can ionize to H+ and release one electron in the lattice. The

produced proton (H+) cannot exist as such, because it is very unstable in this form [28], so it

Page 40: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

23

will react with the oxygen anions of the lattice and form an O-H bond. This bond is

directional, and depending on the interatomic distances, the coordination number of the

proton would be one or two [28]. When there are large interatomic distances, the

coordination number of the proton is one. On the other hand, a proton can be shared

between two oxygen anions when the distance between the oxygen anions is short enough

[28] (the interatomic distances in the crystalline lattice strictly depend on the radii of cations

and anions which formed the structure). Finally, hydrogen can also exist as a molecule

within the structure of oxides. In this form, hydrogen might only exist at grain boundaries,

voids, or pores, where there is enough space for a hydrogen molecule. It has been suggested

that hydrogen atoms may form hydrogen molecules at grain boundaries, and this may cause

the formation of cracks in these regions [12].

The important question is “what is the stable form of hydrogen in the crystalline

lattice of PZT?” Xiong and Robertson [29] have investigated stable forms of hydrogen in the

structure of PbTiO3 and PbZrO3 using the first-principle of quantum mechanics or ab-initio

calculations. Their results show that a hydrogen atom will form a donor state in these

structures, i.e. the stable form of hydrogen was the proton (H+) [29]. This result can also be

applicable for PZT, as the formation of [OH]– bonds in PbTiO3 has also been confirmed by

Park and Chadi [30]. Using first-principles calculations, they have shown that hydrogen

impurities will act as shallow donors in the structure of PbTiO3. Raman spectroscopy results

of hydrogen treated PZT samples also confirmed the existence of O-H bonds in the structure

of PZT [31]. Therefore, it is reasonable to presume that the stable form of hydrogen in the

Page 41: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

24

crystalline PZT is H+. However, based on the above results, one cannot conclude that

hydrogen cannot exist in other forms (H or H2), but it is certain that some of the hydrogen

atoms will be ionized inside the PZT lattice.

2.3.3 Stable sites of H+ in PZT

Determining the stable sites of protons in PZT matters, as protons form directional

bonding with oxygen ions, and this could affect the built-in dipoles in PZT. Generally the

stable sites of H+ in the crystalline lattice of perovskite oxides depend on the binding

interactions between the H+ and oxygen anions [32]. In other words, since H+ will form an

OH- bond in the crystalline lattice, it is reasonable to expect that the position of H+ will

depend on the oxygen sites in the structure. Furthermore, one can say that “protons are

localized within the valence electron density of the oxygen” [32]. The positions of OH- bond

also will be determined by its interaction with cations existing in the structure [32]. Kreuer

has suggested two stable sites for protons [32], depending on the lattice constants of the

perovskite oxides. For the perovskites with large lattice constants, the stable sites of protons

will be the edges of octahedrals (Figure 14a). On the other hand, for the perovskites with

short lattice constants, the protons also could be shared between two oxygen ions of two

adjacent octahedrals (Figure 14b) [32].

Page 42: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

25

(a)

(b)

Figure 14- The stable positions of protons according to the reference [32] in a) perovskites with large lattice

constants, and b) perovskites with short lattice constants

Aggarwal et al. have also investigated the possible sites of protons in the structure of

PZT [31]. They considered four possible sites for protons, and based on the Raman spectra

obtained for hydrogen treated PZT samples, concluded that the more probable site for H+ is

between the apical oxygen ions and Ti (Figure 15a) [31]. They concluded that just one of the

apical oxygen ions will react with a proton [26]. Park and Chadi [30] have investigated the

stable sites of protons in the crystalline lattice of PbTiO3 using first-principles calculations.

For the tetragonal phase of PbTiO3 they have suggested different stable positions for protons,

as depicted in Figure 15b. They have also predicted that depending on the position of

protons, the [OH]– will either destroy, or enhance the polarization of the spontaneous

dipoles in PZT [30]. However, they have predicted that in the presence of the [OH]– dipoles,

the spontaneous dipoles cannot be switched by applying the electric field. On the other

Page 43: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

26

hand, for a cubic PbTiO3 they have found the most stable site to be the site shown in Figure

15c.

(a)

(b)

(c)

Figure 15- PZT crystal structure showing the possible location of protons in the lattice of PZT [31]; b) stable

lattice site of protons for tetragonal PbTiO3 and c) cubic phase of PbTiO3 supposed by Park and Chadi [30]

2.3.4 Effect of hydrogen on PZT

We can now focus on the possible mechanisms responsible for the degradation of PZT

by hydrogen. As seen in Figure 16, the structural degradation can be categorized by the place

Page 44: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

27

where it occurs: a) at the grain boundaries and at the interfaces between PZT and electrodes,

and b) in the crystalline lattice of PZT.

O and Pb vacancies(mostly at grain boundaries)

degraded layer due to [OH]–

(inside the crystalline lattice)

Figure 16- Proposed PZT damage mechanisms can be categorized by the place where damage occurs

Damage at the grain boundaries and interfaces of PZT and electrodes

When H atoms diffuse from the electrode and reach the interface between the

electrode and PZT, some may diffuse fast through the grain boundaries. Such hydrogen

atoms may affect the structure and properties of the PZT in different ways. The first

mechanism is that hydrogen atoms may undergo the following reaction [33]:

2H + O2- H2O + (17)

Where is the oxygen vacancy with two electrons. Oxygen vacancies produced by the

reaction (17) can further ionize and produce two free electrons; this can decrease resistivity

of PZT. However, because H2O cannot exist inside the PZT ceramic body, reaction (17) can

only occur on the surface of the PZT ceramic grains [11]. Therefore, for such a reaction to

continue at the interfaces, oxygen atoms must diffuse from the bulk of the grains to the

interfaces, and water molecules should also diffuse out from the sample along the grain

Page 45: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

28

boundaries. Generally, such diffusion, especially the oxygen diffusion in PZT, is slow. As

such, the above damage by H can only occur at the grain boundaries of PZT and the damage

is limited to the grain boundaries [11, 34]. Therefore, the damage that occurs by reaction (17)

is only at the surface of the ceramic particles, which may include electrode/PZT and grain

boundaries (Figure 16).

It should be noted that the oxygen vacancies might also be produced by the following

reaction, without a direct reaction with hydrogen:

+ ½ O2 (g) (18)

The above reaction occurs in reducing atmospheres. Experiments show that treatment of

PZT under a hydrogen deficient atmosphere (such as pure N2) has no noticeable effects on

the properties of PZT, while similar treatment in hydrogen has considerable effects on the

properties of PZT [16, 34]. It appears that the rate of the reaction (17) is too low to have

noticeable effects on the properties of PZT. Shimakawa and Kubo have proposed that PZT

color change (after exposure to H2) from white to black is due to the formation of such

oxygen vacancies [35]. It was also supposed that oxygen defects would produce donor levels

within the PZT band gap, which would account for the change in color and would increase

the leakage current in capacitors [35].

Another mechanism, by which H atoms could change the PZT structure, is the formation

of Pb vacancies. Indeed, the presence of metallic Pb at the electrode/PZT interface and at

grain boundaries after the hydrogen treatment was previously reported [34-36]. The

electrical properties of PZT may change due to the formation of Pb vacancies in PZT,

Page 46: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

29

because they can produce free holes in the valence band of PZT. The formation of lead

vacancies is limited by the diffusion of lead, which is very slow in PZT at temperatures less

than 600C (DPb in PbTiO3 =7.210-4exp(-181800/RT) (cm2/sec), equivalent to 9.9910-15 cm2/sec

at 600C and to 2.710-11 cm2/sec at 1000C [37]) and is therefore likely limited only to the

interfaces [35]. As the amount of metallic Pb is low in the H2-treated PZT (less than 0.3%

[35]), Pb vacancies cannot be considered the responsible mechanism for the alteration of

ferroelectric properties and color change from white to black [35]. On the other hand,

Ikarashi [36] has concluded that the reduction of PbO in PZT to metallic lead is the main

reason for the degradation of the ferroelectric properties of PZT [36]. The author has

suggested that the degradation of the properties of PZT by hydrogen annealing could be

avoided by using electrode materials that prevent the Pb diffusion from PZT [36]. It should

be noted that the reduction of PbO in PZT to metallic lead does not occur only due to

reducing atmosphere. H can diffuse into the PZT structure and change its atomic bonding

[34], leading to reduction of PbO in PZT to metallic lead. No reduction to metallic lead was

observed in the bare PZT plates up to 600C, while for PZT samples with Pt electrode the

reduction to lead was observed at temperatures as low as 320C [34]. The changes in the

atomic bonding of PZT can be due to the existence of H+ within the lattice and its bonding

with O2-. This is discussed in the next section.

Damage of the crystalline lattice of PZT

Another mechanism proposed for the changes of the electrical properties of PZT is

the ionization of hydrogen atoms inside the lattice of PZT [31]:

Page 47: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

30

H H+ + e- (19)

Upon ionization, hydrogen releases an electron, which decreases the resistivity of the films,

and the produced hydrogen ion will interact with an oxygen ion to form a polar hydroxyl

bond [OH]- [31]:

H+ + O2- [OH]- (20)

Aggarwal et al. presumed that the above mechanism is the main reason for the changes in

the ferroelectric properties of PZT [31]. They have concluded this from the fact that oxygen

diffusion coefficient in PZT at 200°C is not high enough to cause significant damage. Another

reason to support this idea is that the annealing of PZT films in oxygen deficient atmospheres

does not affect their ferroelectric properties [31]. It was concluded that the reactions (3) and

(4) during PZT annealing with forming gas are the primary avenues for the degradation of

ferroelectric properties [31]. The [OH]− ion acts as a fixed dipole, which does not allow the

switching of the ferroelectric domains [17]; this idea has also been supported by theoretical

studies [29-30]. As said before, it is supposed that the bonding of hydrogen with oxygen

anions may change the atomic bonding of oxygen with the other elements of PZT (Pb, Zr,

Ti) and as a result, it might cause some changes in the atomic bonding of PZT elements [35].

Among the reactions mentioned above, probably the reactions (19) and (20), occurring in the

crystalline lattice of PZT, are most likely responsible for the changes in the electrical

properties of PZT. That is because they affect the crystalline lattice of PZT, while the other

reactions just modify the interfaces.

Page 48: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

31

2.4 Dielectric spectroscopy of PZT

Dielectric spectroscopy has been used for studying the properties of a wide range of

materials, such as glasses, polymers, and ceramics [38]. This technique measures the

polarization response of a dielectric medium to an applied electric field; when an electric

field is applied to a dielectric medium, charges, (including ions and electrons), molecules, and

dipoles are displaced according to the applied electric field, which causes polarization in the

dielectric medium. By measuring the polarization of the dielectric medium, different

parameters of the dielectric medium, such as impedance (Z), admittance (Y), and dielectric

constant (ε) can be obtained. By analyzing these parameters, useful information can be

obtained regarding the charges, dipoles, and molecules inside the dielectric medium. The

parameters commonly used for analyzing insulators include dissipation factor DF and

dielectric constant ε, which are measured in a wide range of frequencies (from mHz to MHz)

and temperature, and are then used to analyze the data by fitting the results to one of the

available mathematical models [38].

The classical model of dielectric relaxation of a dielectric medium containing dipoles

is the Debye model. According to this model, the polarization ( ) of a dielectric medium

changes in accordance to following equation:

( ) (

) (21)

Page 49: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

32

where is a constant, t is time, and τ shows the relaxation time of the dipoles. Based on the

equation (21) for polarization, the complex dielectric constant ε (where ε= ε´+iε´´) can be

expressed as

( )

(22)

where ε is the dielectric constant when , εs is the dielectric constant when 0, is

the angular frequency, and i is the imaginary number [38]. A more flexible model,

commonly used for modeling the dielectric constant data, is the Havriliak–Negami equation

[38-39]. According to this model, the complex dielectric constants ε of a dielectric can be

evaluated by:

( )

( ( ) ) (23)

where θ and β are constants between 0 and 1 [39]. By fitting the dielectric data to one of the

above equations, one can obtain different information about the dipoles inside the dielectric

medium, such as the number of dipoles and the activation energy for moving (or hopping) of

ions. For example, Kamishima et al. [40] investigated the dielectric properties of proton

conductor Yb-doped SrZrO3 after hydration in water, and using this technique they were

able to anticipate the position of the [OH] bonds inside the SrZrO3. In this work, we have

attempted (for the first time according to our knowledge) to use this technique to assess the

effect of hydrogen on the properties of PZT.

Page 50: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

33

2.5 High pressure hydrogen compatibility of PZT

The high pressure hydrogen compatibility of PZT is an issue which has recently been

raised due to the possible application of piezoelectric actuators in the hydrogen fuel injectors.

Modern electronic fuel injectors that use lead-zirconate-titanate (PZT) based actuators for

valve opening, instead of the conventional solenoid technology, have been introduced by the

leading engine manufacturers in recent years (Figure 2). Since the valve is actuated quicker

(i.e. about 5 times faster [41]) by the piezoelectric actuators than the conventional solenoid

technology, very precise injection intervals become possible between the pre- and main

injection. Consequently, fuel consumption and emissions are noticeably reduced (by up to 15

percent) [42]. Piezoelectric actuators also facilitate an increase in the injection pressure, up

to 250 MPa; the higher the pressure and the more accurate the dosing and timing of the

injection, the more efficient (and therefore less polluting) the combustion event becomes

[42]. It should be noted that whereas the valve needle stroke was fixed in the previous

electromagnetic injection systems, in injectors where the piezoelectric actuator acts directly

on the needle, the needle stroke can be varied by changing the magnitude of the applied

voltage, thus enabling better control over the valve opening.

It was reported previously that PZT thin films lose their ferroelectrical properties

after hydrogen treatment [2]. However, the hydrogen environment used in the published

studies [2, 9-10] is not comparable to the hydrogen environment conditions in an engine,

where the pressure of the gas is up to 30 MPa, and the maximum operating temperature of

Page 51: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

34

the PZT ceramic is about 100C. Alvine et al., have investigated the effect of high pressure

hydrogen on the properties of PZT films [43]. They studied the structural and compositional

changes of PZT thin films (50 nm) after 24 hours of high pressure hydrogen treatment

(p=13.8 MPa, T=100C). The most important structural changes which they observed was the

hydrogen induced blistering on the surface of bare PZT films and PZT film with Pd

electrodes [38]. They have also observed “significant mixing of the Pd layer into the PZT film

along with migration of Pb into the Pd layer” [43]. The hydrogen absorption for bare PZT

films was about 10 at%. The Pd layer on the surface of PZT films had a considerable effect on

the amount of hydrogen absorption: due to the presence of the Pd layer the hydrogen

concentration in the PZT ceramic is increased to nearly 20 at% [43]. Other results from

their experiments are as followings:

(a) Piezo actuators can degrade in high-pressure hydrogen environments due to the

hydrogen uptake in the PZT plates.

(b) The amount of hydrogen absorption is a function of the metallic electrode and it increases

for different electrodes in the following order: Pd> Al> W> Ti> Cu.

(c) Lead migrates into all the above mentioned electrodes, with the possible exception of Ti.

One of the main issues not addressed in the work of Alvine et al. is the possible effects

of the high pressure hydrogen environment on the electrical properties of PZT. Therefore,

in the present work, in addition to investigation the effect of high-pressure hydrogen

environment on the microstructure of PZT, we also investigated the effect of hydrogen on

the electrical properties of PZT.

Page 52: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

35

2.6 Methods of decreasing the hydrogen damage to PZT

One such method is based on the experimental observations that the degradation of

properties of PZT is related to the hydrogen reactivity and absorption of the metallic

electrodes. For example, PZT-electrode assembly with Au electrodes showed the least

amount of degradation when subjected to an hydrogen atmosphere, whereas Pt was the

worst electrode [20], due to the catalytic nature of Pt in dissociating the hydrogen molecules

into atoms. Therefore it has been suggested to use electrodes with less catalytic activity such

as Au, instead of electrodes such as Pt [9]. Moreover, it was suggested that oxygen plasma

treatment of the Pt electrode can be used to reduce its catalytic activity [44], as this

procedure modifies the surface of the electrode. Conductive metal oxide electrodes like IrO2,

LaNiO3 have been tried, and it has been found that annealing in a hydrogen containing

atmosphere will not degrade the properties of PZT with such electrodes [20]. Abdolghafar

et. al. proposed IrO2 as the top electrode to prevent the hydrogen damage to the PZT during

hydrogen treatment [20]. Their results show that PZT capacitors with IrO2 electrodes have

poor hydrogen resistivity because of the reduction of the IrO2 to metallic Ir during hydrogen

treatment; however, after oxygen pre-annealing at 600C, the PZT capacitors with IrO2

electrode showed excellent hydrogen damage resistivity. The hydrogen damage resistivity

after oxygen annealing is attributed to the enhancement of the IrO2 structure by oxidation.

Another technique designed to improve the hydrogen resistivity of the PZT

capacitors is using a hydrogen diffusion-barrier layer on top of the capacitor. This technique

Page 53: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

36

prohibits hydrogen diffusion by encapsulating the whole capacitor (electrode/PZT/electrode)

in hydrogen barrier layer(s), sometimes called a hard mask. It has even been shown that even

sidewall diffusion barriers can dramatically enhance the hydrogen resistivity [40]. Both

conductive (e.g. TiAlN) and non-conductive layers (e.g. Al2O3, SiO2) can be used as a hard

masks. Saito et al. have investigated different oxides (Al2O3, HfO2, Bi3Ti4O12, ZrO2 and SiO2)

as encapsulation layers [46]. They suggested Al2O3 and SiO2 as promising hydrogen diffusion

barrier layers.

While the above techniques try to prohibit hydrogen damage during the hydrogen

treatment process, another idea is to recover the properties of PZT capacitors after hydrogen

annealing [21]. This could be done by high temperature (600-700C) annealing of the PZT

capacitor in atmospheres like N2 or air after the hydrogen gas treatment. It has been reported

that the Pt/PZT/Pt assembly will recover its properties after treatment in O2 gas [17 and 21].

The recovery of the properties has been suggested to be due to the diffusion of hydrogen

atoms out of the sample during the post-annealing treatment. These results confirm the

recently published theoretical and experimental results by Bjorheim et al., who investigated

the hydration thermodynamics of PbZrO3 and concluded that protons which are absorbed

inside PZT from the hydrogen gas can be removed from PZT by heat treatment in air at

temperatures higher than 700C, or at lower temperatures in dry oxidizing or inert

atmospheres [47].

Page 54: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

37

3 Scope and Objectives

Scope

The investigation of the interactions of hydrogen with PZT is an important topic both

from practical and scientific points of view and still many issues need to be addressed in this

regard. While a lot of attention has been paid to the effects of hydrogen on the ferroelectrical

properties of PZT, there are very few publications on the effects of hydrogen on the

piezoelectrical properties of PZT. This is an important issue for using the piezo-actuators in

the advanced internal clean-combustion engines, and indeed there is very limited

understanding of the performance and durability of such piezo-actuators in hydrogen

environments. The broad scope of this project is to examine the key research issues related to

the performance of the PZT-based piezo actuators exposed to hydrogen, and to develop

methods for preventing or limiting the damaging effects due to hydrogen, with the aim of

extending the lifetime of the actuators. Specifically, the scope of this work involves the

following activities:

(a) Bare PZT plates and similar plates including silver electrodes (Ag/PZT/Ag) are

exposed to hydrogen atmosphere at 10 MPa pressure for 200-1200 hours. The changes in the

microstructure of PZT plates are investigated using scanning electron microscopy combined

with energy-dispersive X-ray spectroscopy (SEM/EDX), and X-ray diffraction (XRD). The

effect of hydrogen on the electrical properties is investigated by measuring the changes in

the capacitance of Ag/PZT/Ag capacitors online during the hydrogen treatment.

Page 55: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

38

(b) The kinetics of PZT structural modification due to hydrogen exposure is

investigated using online monitoring of the electrical properties of PZT. Specifically, the

changes in the capacitance of the Ag/PZT/Ag capacitors are measured online during the

hydrogen treatment at temperatures higher than the Curie temperature. Furthermore, we

measure the dielectric constant and dissipation factor of PZT after the hydrogen treatment in

a wide range of frequency (from 12 Hz to 200 kHz) and temperatures (25 to 400C); these

data are then analyzed by fitting the results to one of the existing mathematical models.

(c) The effect of hydrogen on the microstructure and electrical properties of PZT in

the tetragonal phase (i.e. at room temperature) is investigated using the water electrolysis

technique, wherein PZT plates are exposed to atomic hydrogen generated in water

electrolysis for 6 and 48 hours. SEM is used for evaluating the changes in the microstructure.

The capacitance of Ag/PZT/Ag capacitors is measured on-line during the water electrolysis

and after finishing the hydrogen exposure (i.e. during PZT aging in air), to evaluate the

effects of hydrogen on the electrical properties of PZT.

(d) Alumina coatings are applied to PZT plates using sol-gel technique, to explore the

possibilities of decreasing H2 damage to PZT. The functionality of the coating against

hydrogen damage is evaluated by water electrolysis technique.

It is anticipated that the findings of this project will contribute to the fundamental

understanding of PZT-hydrogen interaction, and will also provide Canadian clean-engine

technology developers with strategies for improving hydrogen technology, thus becoming

more competitive in the global market.

Page 56: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

39

Objectives

The broad objective of the present work is to evaluate the interactions between the

PZT-based piezoelectric materials used in piezo actuators, and hydrogen. Based on the

improved understanding of such interactions and the resulting damage to PZT, we aim to

propose methods to decrease the negative effects of H2 on PZT. Within this broad goal of the

project, we have the following specific objectives:

1. To evaluate and quantify the effects of long-term high-pressure gaseous hydrogen

exposure on electrical properties and microstructure of PZT

2. To determine the parameters describing the kinetics of the interactions between

hydrogen and PZT, in particular the incorporation of hydrogen into PZT and the resulting

changes in the PZT properties

3. To evaluate and quantify the effects of hydrogen on the electrical properties and

microstructure of PZT below the Curie temperature, i.e. in the temperature range of the

actuator's use in the engine

4. To propose and develop methods to decrease and prevent the hydrogen damage to

PZT, such as through the deposition of protective coatings.

Page 57: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

40

4 Materials and Methods

4.1 Samples

“PZT” plates, 10×10×1 mm, (PIC 255, PI Ceramic GmbH, Lindenstrasse, 07589

Lederhose - Germany) of the composition Pb(Zr0.53Ti0.47)O3 with 1% Nb2O5 dopant were used

in this work. The effect of hydrogen on PZT microstructure was investigated using poled

PZT plates with electrodes, as well as bare, not poled plates. Two types of electrodes were

used: silver (which contained small additions of Bi) and silver-palladium alloy (76 wt% Ag +

24 wt% Pd, as given by SEM/EDX); they were screen-printed on the large faces of the plates,

in 20 µm layers. Silver and their alloys are usually used in making the actuators, and

therefore, we considered this type of electrodes. The microstructure of the bare PZT samples

is shown in Figure 17.

Figure 17- The microstructure of bare PZT plates

bare PZT

Page 58: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

41

The interface between the electrode and PZT is shown in Figure 18. As it can be seen

in this figure, there is a good adhesion between the electrodes and the PZT for inside section

(Figure 18a) and poorer on the edge section (Figure 18b). Later it will be shown how

hydrogen affects this interface.

Ag or Ag/Pd electrode

PZT

Figure 18- The microstructure of the PZT samples

Figure 19 shows micrographs of as-received electrodes' surfaces; the Ag-Pd electrodes

were much more porous than the Ag electrodes.

Ag or Ag/Pd electrode

PZT

Page 59: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

42

Figure 19- Micrographs of the surface of the electrodes: silver (a) and silver-palladium (b)

4.2 Gas hydrogen treatment

Two hydrogen gas treatment procedures were used in this work. The one which we

call “High-Pressure” hydrogen treatment is schematically shown in Figure 20. For this

condition of hydrogen treatment different exposure times (200-1200 hours) were used, while

the pressure and temperature were kept constant at 10 MPa (pure hydrogen), and 100C,

respectively. Before point (a), the chamber was purged several times with argon to remove

air. The temperature and pressure of the gas were chosen according to the practical condition

of the actuators in the engine [43]. Therefore, the results of this experiment can be used to

evaluate the possible effects of high-pressure hydrogen environment on the microstructure

of PZT plates used in fuel injectors. After finishing the experiments the samples were taken

Ag or Ag/Pd electrode

PZT

Page 60: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

43

out from the chamber and microstructural analysis was done using XRD and SEM

techniques.

Tem

per

atu

re

a bc

d

e

Time

Vacuum On

Cooling in HydrogenHydrogen

On

Heater on

Figure 20- The time-temperature schedule of the ‘High-Pressure’ hydrogen treatment used in this work

Fuel injection actuators are made of many (>50) single PZT plates stacked together, and

the electrical properties of each single layer determine the performance of the whole

actuator (Figure 21).

+-

electrodes

PZT plate

+-

R

C

(a) (b)

Figure 21- Schematic of an actuator made from PZT plates stacked together, b) the equivalent electrical circuit

of an actuator

Page 61: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

44

Therefore, measuring the capacitance of a single PZT plate is a suitable way to

investigate the effect of hydrogen on the performance of the actuator assembly in an

hydrogen atmosphere. To this end, two thin copper wires were attached to the electrodes,

and a GW Instek LCR meter (LCR-821) was used to measure online the capacitance of PZT

plates during hydrogen exposure. The capacitance measurements were performed with an

internal voltage of 0.125 V at a constant frequency of 1000 Hz. Figure 22 schematically

shows the experimental setup for measuring the electrical properties.

P= up to 10 MPaT= up to 700 °C

H2

LCR meteror

Potentiostat

Silver Electrodes

Copper wires connected to LCR meter

High-Pressure Vessel

PZT

PZT microstructure

(10×10×1 mm)

Connected to the vacuum pump and

hydrogen

Figure 22- The schematic of the setup used in this work for online monitoring the electrical properties

The other hydrogen treatment procedure, which we would call “High-Temperature”

hydrogen treatment, is schematically shown in Figure 23. From point “a” to “b” the sample

was heated in air up to the desired temperature in 450C-650C range. The chamber was then

vacuumed (700 mm Hg (0.09 MPa)) at point “b”, and at point “c” pressurized with 0.13 MPa

Page 62: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

45

of 90% Ar/10% H2 gas mix. A drop of about 5C was observed, but the temperature

recovered after 15 minutes.

Tem

per

atu

re

a

b c d

e

Time

Heating in Air

Vacuum On

Hydrogen On

Cooling in Hydrogen

Figure 23- The ‘High-Temperature’ hydrogen treatment used in this work

To measure the resistance and capacitance of PZT plates, two thin copper wires were

attached to the silver electrodes using high temperature silver paste. The same LCR meter

(LCR-821) was again used to measure the capacitance and the ‘real’ part of the impedance

(ZRe) of PZT plates (Figure 22). The frequency was 1000 Hz for all measurements. The

electrical properties were monitored online during the hydrogen treatment: capacitance and

ZRe values were collected every 0.896 second.

Knowing the relation between the ZRe and the capacitance, the total resistance of the

PZT plates can be estimated. Figure 24 shows a typical ZRe-ZIm curve measured for this type

of PZT plates in air. Curves with the same trend were obtained for other temperatures, and

also in the hydrogen atmosphere.

Page 63: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

46

According to Figure 24, the PZT plate can be considered as a parallel R-C circuit.

Therefore, the relation between the ZRe, R and C is as follows [49]:

(24)

Using equation (24) the ohmic resistance (R) of PZT can be calculated, but this method is not

without errors, especially at low temperatures. The reason is that with decreasing the

temperature, the Nyquist diagrams at high frequencies show a depressed semicircle; in other

words, equation (24) is no further valid [49]. Therefore, we used the R values obtained by the

equation (24) to just qualify the trend for the changes of R during the hydrogen treatment.

We cannot say that the R values obtained from the equation (24) are the exact values for R

during the hydrogen treatment; however, they can show the general trend for the changes of

the resistance. In order to measure the exact values for R, we used another technique,

detailed in the next paragraph. .

In order to determine the electrical resistance of PZT, the variation of resistance

from the ZRe-ZIm curves was also calculated. The ZRe-ZIm curves were recorded every 5

minutes during the hydrogen treatment; all ZRe-ZIm curves were semi-circles where the

diameter of the whole circle can be considered as the ohmic resistance of PZT. For this set of

experiments we used smaller PZT plates (4×4×1 mm), but with the same composition and

electrodes; the estimation of R for both sets of sample showed the same trend.

It should be noted again that we used the R values obtained either from equation (28)

or from the ZRe-ZIm curves just to identify the general trend for the changes of R during the

Page 64: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

47

hydrogen treatment. The electrical property used to monitor the structural changes was the

capacitance C, which we were able to read directly from the LCR meter with reasonable

accuracy. The accuracy of the LCR meter used in this work was 0.05% for measuring the

capacitance.

Figure 24- A typical Nyquist plot for the PZT at 500°C in air (the frequency is swiped between 106 and 10-3 Hz)

4.3 Water electrolysis treatment of PZT

The water electrolysis technique was used to charge the PZT samples with hydrogen,

following a previously reported methodology [33]. In comparison to the treatment in

hydrogen gas, which needs temperatures higher than room temperature, this technique has

the advantage that it can be done at room temperature; this inhibits the de-polarization of

PZT samples due to elevated temperatures. The basic idea for this technique is that one of

the silver electrodes attached to PZT is used as the cathode during water electrolysis. In this

0 100 200 300 400

ZRek

0

100

200

300

400

Im (k

)

R

C

Page 65: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

48

way, the hydrogen atoms which became released on the surface of the silver electrode can

further diffuse into the electrode and into the PZT attached to the electrodes. The water

electrolysis was performed in a 0.1 M NaOH solution, with a current density of 100 mA/cm2;

the current density was kept constant by varying the voltage, to ensure the same rate of

hydrogen release. The leakage of the solution to the interface between the electrode and PZT

and the release of hydrogen at the interface caused the detachment of the electrode from

PZT after very short experimental time (around 5 minutes). In order to restrict the leakage of

the solution to the interface between the Ag electrodes and PZT, the edges of the specimens

were encased in epoxy. After the samples were encased in epoxy, the electrodes did not

detach from the sample during the water electrolysis.

At different times after the beginning of the water electrolysis (ie. t= 10, 60, 120, 300,

600, 1200 minutes) the samples were taken out, dried, and the capacitance was measured. To

measure the capacitance of PZT plates, the same LCR meter (GW Instek-LCR-821) was used.

An internal bias of 0.125 V was used for measuring the capacitance, and the frequency was

kept constant at 1000 Hz during all measurements. Figure 25 schematically shows the setup

which is used for the water electrolysis experiment. After water electrolysis, the samples

were prepared for the microstructural analysis. In this order, they were grinded, polished

and then the cross section was studied by SEM.

Page 66: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

49

Figure 25- Schematic of the setup for the water electrolysis experiment

In order to estimate the equivalent pressure of the hydrogen above the electrode

during the water electrolysis, and to compare the results with the gas hydrogen treatment,

we must first quantify the hydrogen absorption into the Ag layer during the water

electrolysis. The electrolysis is actually a common and efficient technique for charging

metals with hydrogen, wherein very high equivalent pressures of hydrogen can be produced

above the metallic electrode [50]. It has been reported that very high fugacity of hydrogen

on the order of 106 atmospheres (corresponding to pressures of about 104 atmospheres) can be

obtained with cathodic charging [50]. Because of this high efficiency in introducing the

hydrogen atoms inside metals, this technique has being used for producing metal hydrides,

and for studying of hydrogen embrittlement in metals [51]. Wu measured the hydrogen

solubility in iron by hydrogen cathodic charging [52]. At a potential of 0.25 V and in 1M

PZTAg

H+

H+

H

H

H

H

H

H

H+

0.1 M NaOH

anode (+) cathode (-)

Page 67: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

50

H2SO4 solution at room temperature, he has found that the solubility of hydrogen in iron can

be described by the following relation:

( ) (25)

where is the charging current density. Therefore, for a typical value for (0.1A/cm2) at

room temperature, 0.53 ppm of hydrogen would dissolve in iron. On the other hand, the

solubility of gaseous hydrogen of pressure PH2 in Fe can be obtained from the following

relation: [53]

( ) (

)

(26)

Therefore, at 1 atmosphere partial pressure of hydrogen and at room temperature, the

solubility of hydrogen in Fe would be about 0.00076 ppm, much lower than the predicted

solubility of hydrogen in Fe during cathodic charging. Therefore, we can conclude that the

fugacity of hydrogen gas above the Fe electrode during water electrolysis is much higher

than the fugacity of gaseous hydrogen. This simple comparison confirms that high solubility

of hydrogen in metals can be obtained with cathodic charging. Danielson found that the

solubility of hydrogen in AA5083 aluminum alloy in cathodic charging (3.410-7 g-atoms

H/cm3) is higher than the solubility of hydrogen in gas atmosphere (110-11 g-atoms H/cm3),

by about 4 orders of magnitude [54]. He concluded that the cathodic charging of hydrogen

has a major effect on increasing the hydrogen solubility in Al.

There is no published data on the absorption of hydrogen atoms inside silver during

the cathodic charging. Therefore, it is impossible to estimate the equivalent pressure of the

Page 68: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

51

hydrogen above the electrode during the water electrolysis. However, it will be shown later

(Section 5-3) that noticeable damage occurred to PZT after water electrolysis treatment. We

therefore propose that this high degree of degradation is due to the higher hydrogen

solubility in silver during the water electrolysis, in comparison to the gas atmosphere.

4.4 Alumina sol-gel coating

Sol preparation

Boehmite (ALOOH) sol was used as precursor sol for the alumina (Al2O3) coatings.

Boehmite sol was prepared by dissolving 10 gr boehmite powder (Dispersal Sol P2, Condea

Chemie GmbH, Germany) into 100 ml distilled water. The solution was mixed for 20

minutes at room temperature using magnetic stirrer. To increase the viscosity of the sol, PVA

solution (10 wt% PVA) was added to the Boehmite sol (for the total amount of PVA equal

4.5 wt% of sol) and then solution was again stirred for 10 minutes to homogenize the sol

before using it for dip coating. Figure 26 schematically shows the procedure used for the

preparation of the sol.

Page 69: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

52

Figure 26- Schematic of the steps used for the preparation of the alumina coating

The coating process

The dimension of PZT plates used for coating was 10×10×1 mm. Before coating, the

surface of the samples was ground, and then polished with 1 micron diamond paste. The

samples were washed with acetone before coating. To apply the boehmite sol to the surface

of PZT plates dip coating technique is used. “Dip coating" is a common coating technique

100 ml Deionized water

Stir at room temperature

for 20 minutes

10 gr AlOOH

100 ml PVA (10 wt%)

Stir at room temperature

for 20 minutes

Stable sol

Final sol

Dip coating with withdraw speed

of 3 cm/min

Firing at 450°C for 5 hours

Page 70: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

53

used in applications such as optical coatings and membranes. During the dip coating process,

the substrate is submerged into the sol and carefully withdrawn out of the sol [55]. We

selected this technique for the coating as it is relatively simple, and the samples were

relatively small. The dip coating was done using the SCS PL3201 Dip Coater (manufactured

by SCS, 7645 Woodland Drive, Indianapolis, Indiana, 46278, USA) with a withdraw speed of

3 cm/minutes. The thickness of the coating can be evaluated by the Landau-Levich equation

[56]:

( )

( ) (27)

where is the coating thickness, is the viscosity of the sol, is the withdraw speed, is

the sol-vapor surface energy, is the density of the sol, and is the gravity. Considering the

above equation and using the and values for water, and considering the viscosity of the

sol equal to 20 mPa.s [57], the thickness of the single-layer of as-deposited coating (i.e. before

heat treatment) was calculated to be about 7 µm.

Firing treatment

Heat treatment is always needed for the densification and crystallization of the sol

layer which is applied on the surface on the PZT plates. In this way, a relatively dense film

can be obtained on the surface, depending on the temperature and time of the heat

treatment. The heat treatment used in this work was at 450C in air for 5 hours [58]. When

doing high temperature processing of the coating layer, we have to consider its interaction

Page 71: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

54

with the PZT (i.e. change of properties because of the changing chemistry). The temperature

that Boehmite starts to transforms to alumina is reported to be about 450C [58]. Therefore,

we selected this temperature for our heat treatment procedure. Furthermore, the lower the

heat treatment temperature, the lower the risk for cracking is during the cooling due to

thermal expansion. Based on the literature data [59-60] it is expected that after this treatment

the coating still contained about 40 vol% of porosity. Therefore, the presence of the coating

will not prevent access of molecular hydrogen to PZT surface, but it will prevent direct

contact between the metallic electrode and PZT. Thus it is expected that there will be no

access of atomic hydrogen to the coated PZT surface. The coating and firing steps were

repeated 3 times in order to achieve > 5 m thick coatings. As shown later in Section 5-4-1,

the actual thickness of the coatings after heat treatment, as determined by SEM, varied

between 5 and 10 µm.

Assessment of coatings effects

To assess the effects of the coatings on hydrogen penetration into PZT, water

electrolysis technique was used. In this regard, thin Au-Pd electrodes were sputtered on the

faces of the PZT plates coated with alumina. Au-Pd electrodes were also deposited on the

faces of a PZT plate without alumina coating and that sample was used as a reference sample.

The water electrolysis was performed in a 0.1 M NaOH solution, with a constant current

density of 100 mA/cm2 and with the voltage between 7-10 V. After water electrolysis, the

cross section of the sample was studied with SEM and the amount of hydrogen damage was

studied.

Page 72: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

55

4.5 Characterization techniques

4.5.1 X-ray diffraction analysis (XRD)

X-ray diffraction analysis (XRD) was used in this work for the phase analysis of PZT

plates after hydrogen treatment, using Rigaku Multifles diffractometer operated at 40kV, and

30 mA (X-ray:CuKα). To perform the XRD analysis, the metallic electrodes were detached

from the PZT plates, and the remaining bare PZT plate was used for the analysis. The

dimensions of the PZT plates were about 10101 mm. XRD spectrums were collected over

diffraction angles between 20 and 80 at a speed of 2 /min.

4.5.2 Scanning electron microscopy (SEM)

The microstructural investigation was performed with a Hitachi S 3000-N Scanning

Electron Microscope with an Electron Dispersive X-Ray Spectroscopy detector and Quartz

X-One X-ray post-processing software. The SEM/EDS investigations were performed under

low vacuum (20 kPa) - variable pressure mode, using the back scattered electron method,

which offers very good element contrast and allows the study of non-conducting specimens

without applying any conductive coating to avoid charging and contamination and surface

alterations of samples through additional processing and handling.

4.5.3 Electrical properties measurements

To measure the capacitance of the PZT plates, a GW Instek LCR meter (LCR-821) was

used. In online measurements, capacitance measurements were performed with an internal

Page 73: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

56

voltage of 0.125 V at a constant frequency of 1000 Hz. The data were collected every 0.896

second. Two modes were used for measuring: C-DF mode for the capacitance (C) and the

dissipation factor (DF) of samples, and Z-theta mode for the real part of the impedance (ZRe)

of samples. The dielectric constant () of the samples calculated through the equation of the

capacitance for parallel-plate capacitors: = (dC)/( 0a) where d is the distance between the

electrodes (1 mm in this work), C is the capacitance of parallel plate capacitors measured by

LCR meter, 0 is the permittivity of vacuum, and a is the area of the electrodes (1 cm2 in this

work). To measure the capacitance and the dissipation factor of PZT capacitors after water

electrolysis, the same equipment and internal voltage was used; however the frequency was

not constant, and it was changed between 12 and 200 kHz.

To measure the resistance of PZT capacitors, the ZRe-ZIm curves were recorded every 5

minutes during the hydrogen treatment; all ZRe-ZIm curves were semi-circles where the

diameter of the whole circle can be considered as the ohmic resistance of PZT. To obtain the

ZRe-ZIm curves (or Nyquist plot), a potentiostat VersaSTAT 4 was used, and the frequency of

the applied voltage (1 V) was swiped between 106 and 10-3 Hz.

Page 74: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

57

5 Results and Discussion

5.1 High-Pressure conditions (T=100C, p=10 MPa, t=200-1200 hours)

5.1.1 H2 effects on PZT microstructure

To address the high pressure hydrogen compatibility of PZT plates, a high-pressure

hydrogen treatment was considered as is described in Section 4-2. The results of this study

can be used to evaluate the possible effects of high-pressure hydrogen environment on the

microstructure of PZT plates used in fuel injectors. The results of this experiment are

brought in Sections 5-1-1 (microstructure) and 5-1-2 (electrical properties).

Bare PZT Plates

Figure 27 shows micrographs of the top surface (10x10 mm) of bare PZT plates before

(as-received) and after the exposure to the hydrogen atmosphere for 1,200 hrs at 10 MPa and

at 100C; no sample preparation was done before taking the pictures. No noticeable

structural changes were observed on the bare PZT plates in Figure 27; however, the higher

magnification images of the surfaces show that the grain boundaries of the samples treated

for 600 and 1200 hours form clear discontinuities, approximately 100 nm wide. As shown in

Figure 28b, the PZT grains in the sample treated for 1,200 hours are separated, suggesting

that the grain boundaries are affected by the hydrogen. The XRD results of the as-received

and hydrogen-treated bare PZT up to 1200 hours at 100°C did not indicate any new

diffraction peaks, Figure 29. However, it was observed that the XRD pattern of the hydrogen

treated samples has shifted to lower two-theta angles by about 0.4 degree.

Page 75: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

58

bare PZT

Figure 27- Micrographs of PZT surface: a) the as-received sample; b) after 1200 h hydrogen treatment

bare PZT

Figure 28- Surfaces of PZT plates at higher magnifications: (a) before and (b) after 1200 h hydrogen treatment

One of the reasons for such systematic shift could be change of internal stresses

during the treatment [61], particularly in the regions close to the surface of the PZT plates.

Page 76: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

59

As it can be seen in Figure 27, the grain boundaries on the surface of PZT plates are corroded

after the hydrogen treatment, which may affect the internal stress. Furthermore, changes of

the lattice parameters due to the dissolution of hydrogen atoms in PZT could also cause shifts

in the diffraction pattern [9, 18, 35]. While we were not able to measure the hydrogen

content of our samples after hydrogen treatment, it has been reported that high-pressure

hydrogen dissolves in PZT to a level of 4-10 at% [43].

Figure 29- XRD results of bare PZT for as-received and after 1200 h hydrogen treatment

The microstructure of the hydrogen-treated samples was also investigated in the 1x10

mm cross-sections of the plates. Figure 30 shows the cross-sections of the as-received sample

and the sample which was hydrogen treated for 1200 hours.

10 20 30 40 50 60 70 80 90

2

as-received

hydrogen treated

(00

1)

(10

0)

(11

0)

(11

1)

(00

2)

(20

0)

(10

2)

(21

0)

(11

2)

(21

1)

(02

2)

(22

0)

Page 77: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

60

bare PZT

Figure 30- Cross-section of the as-received sample (a) in comparison to the cross section of the sample after

1200 h hydrogen treatment (b)

According to this figure, one can conclude that no changes have occurred in the

regions far from the surface (below 10 µm depth). In other words, the hydrogen damage

appears to be limited to the regions close to the surface, while the center of the sample was

not affected (as observed under SEM). This is better shown in Figure 31.

Figure 31- Cross section of the hydrogen treated sample for 1200 hours, close to the surface

Page 78: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

61

The general conclusion from the SEM microstructural observations of the bare PZT

plates is that the high pressure hydrogen atmosphere would affect the microstructure of the

PZT only close to the surface (to 10 µm depth), and no differences were observed in the

microstructure of the samples farther from the surface.

PZT Plates with Ag Electrodes

An image of the unprocessed side face (10x1mm) of a PZT plate with silver electrode

after 400 hours of hydrogen treatment is shown in Figure 32, clearly showing a degraded

layer formed just adjacent to the electrode. Similar microstructures were also observed for

the 200 and 1200 hours hydrogen-treated samples; the degree of damage was proportional to

the duration of exposure. The higher magnification image of the degraded layer, Figure 32b,

shows a porous structure, where voids are left behind by individual PZT grains detached

from the surface of the sample.

Ag electrode

PZT

Figure 32- Low magnification (a) and high magnification (b) images of damaged layer on the PZT surface next

to the Ag electrode after 400 hours hydrogen-treatment

Page 79: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

62

Such a damaged layer was not observed in the cross-section of any of the treated

samples with Ag electrodes, as shown for example in Figure 33. The hydrogen treatment

conditions for the samples shown in Figure 32 and Figure 33 are the same; however, Figure

32 shows an image from the surface of the sample while Figure 33 shows an image from the

cross section of the sample. As this porous layer formed only on the surface of the PZT plates

and only next to the electrodes, we conclude that this damaged layer is probably due to the

spillover of hydrogen atoms from the surface of silver electrodes to the surface of PZT. By

hydrogen spillover we mean the formation of hydrogen atoms on the surface of the metallic

electrode attached to PZT and diffusion of those hydrogen atoms to the surface of the PZT.

Therefore hydrogen spillover is just limited to the surface and regions of PZT close to the

electrode. This is in accordance to our microstructural observations. This is schematically

shown in Figure 34. The spillover of hydrogen from the surface of metals to the surface of

oxides has been reported previously in systems such as Pd/SiO2 [62], Pt/Al2O3 [63], Pt/WO3

[64].

Another feature observed on these samples was the detachment of the electrodes

from the PZT surface after prolonged hydrogen treatment of 600 hours and 1200 hours, as

shown in Figure 35.

Page 80: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

63

Ag electrode

PZT

Figure 33- The interface of the Ag electrode with PZT after grinding and polishing, for the as-received sample

(a) and for the sample hydrogen-treated for 400 hours (b); no detachment of the electrode from the PZT and no

damaged layer are visible

Figure 34- The spillover mechanism of hydrogen atoms from the surface of the Ag electrode to the surface of

the PZT

H2HHH2 H

H

H

H

H

Electrode

PZTHydrogen spillover

Page 81: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

64

Ag electrode

PZT

Figure 35- The detachment of the Ag electrode from the PZT for the sample treated for 600 h

This could be due to the accumulation of hydrogen molecules at the interface of silver

electrode and PZT. For the sample treated for 1200 hours, the electrode was detached from

the PZT on almost half of the interface. The electrode itself was damaged at the edges, and

cracks in the silver electrode were also observed for the sample heat-treated for 1200 hours,

Figure 36.

Ag electrode

PZT

Figure 36- Detachment of the Ag electrode from PZT; some cracks are present on the surface of the electrode

for the sample heat-treated for 1200 hours

Page 82: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

65

PZT Plates with Ag/Pd Electrodes

Degradation of the properties of PZT after hydrogen treatment has been reported to

correlate with the type of electrode in contact with PZT [9, 21]. It was proposed [21] that H2

dissociates at the surface of the electrode, and H atoms diffuse to the electrode/PZT interface;

hydrogen atoms that reached the electrode/PZT interface could further diffuse into the PZT,

or re-combine to form molecular H2, leading to local blisters [21]. To investigate the effects

of electrode on the degradation of PZT, plates with Ag/Pd electrodes were also evaluated in

this work. A micrograph of the unprocessed side face of a PZT plate with Ag/Pd electrodes

after hydrogen treatment for 200 hours is shown in Figure 37a; a damaged layer similar to

that seen in Figure 37b formed in the vicinity of the electrode.

Ag/Pd electrode

PZT

Ag/Pd electrode

PZT

Figure 37- Micrographs of the sample with Ag/Pd electrodes after hydrogen-treated for 200h: the 1x10 mm side

face (a) and its cross-section (b)

Page 83: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

66

However, the layer was also limited only to the surface of the sample; no damaged

layer was visible in the cross-section (Figure 37b), but the electrode itself was degraded –

either detached or weakened to the extent that it was destroyed during the sample

preparation.

The surface of the side of the sample (1x10mm) hydrogen-treated for 400 hours is

shown in Figure 38b. Evidence of extensive corrosion is seen, especially in the immediate

vicinity of the electrodes (Figure 38b, c). Such a structure is an indication of damage to the

grain boundaries, probably caused by the diffusion of hydrogen, possibly involving

recombination of atomic to molecular H2 at the grain boundaries. The micrograph of the

cross-section of the sample hydrogen-treated for 400 hours is shown in Figure 39 wherein in

some parts of the cross section a noticeable portion (about 100 µm depth below the electrode)

was corroded after hydrogen treatment. We have not investigated the effect of hydrogen on

the PZT plates with Ag/Pd electrodes for 600 and 1200 hours. However, we expect that

increasing the time of hydrogen exposure will cause more severe damages to these samples as

well.

Page 84: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

67

Ag/Pd electrode

PZT

(b)

(b)

Ag/Pd electrode

PZT

(b)

(b)

Ag/Pd electrode

PZT

(b)

(b)

Figure 38- Surface of the side face of the sample with Ag/Pd electrode: (a) as-received, (b) hydrogen-treated for

400 h; noticeable corroded area next to the electrode (c)

When compared to the not-damaged sample with Ag electrodes (Figure 33b), the

depth of the hydrogen damage in the cross section of the sample with Ag/Pd electrodes

Page 85: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

68

indicates that the samples with Ag electrodes are more resistant to hydrogen damage. As

palladium is well-known for its ability to absorb atomic hydrogen [65], from a practical point

of view this means that it is better to use Ag electrodes instead of Ag/Pd electrodes. If the

replacement of the electrodes is not possible, the amount of palladium in Ag/Pd electrodes

should be decreased as much as possible or other electrodes with less hydrogen reactivity

should be used in the manufacturing of the actuators.

Ag/Pd electrode

PZT

Figure 39- Micrograph of the cross-section of the sample shown in Figure 38

5.1.2 H2 effects on the electrical properties of PZT

A typical curve showing changes of capacitance of PZT plates with Ag/Pd electrodes

in high pressure (10 MPa) hydrogen atmosphere is shown in Figure 40a. With increasing the

temperature from 20C (point 1) to 100C (point 2), the sample capacitance also increases

Page 86: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

69

from 1.67 nF (point 1) to 1.98 nF (point 2). This is most likely due to the fact that the

switching of the dipoles becomes easier at higher temperature [66]. Figure 40b shows the

variations in the capacitance and temperature; the increase of the capacitance follows the

increase of the temperature. Figure 41 shows the variation of capacitance of PZT plates with

Ag electrodes in argon atmosphere at the same pressure of 10 MPa. The changes of PZT

capacitance are similar to those in H2 atmosphere (the same trend for the capacitance

variation was also observed for the heat-treatment in air).

1 10 100

Time (hour)

1.6

1.7

1.8

1.9

2

2.1

C(n

F)

1

3

4

(a)2

0 4 8 12

Time (hour)

0

20

40

60

80

100

120

Tem

per

ture

(°C

)

1.92

1.94

1.96

1.98

2

2.02

C(n

F)

(b)

Figure 40- Capacitance of PZT sample with Ag/Pd electrode in high-pressure hydrogen atmosphere (a); at point

‘1’ the heater is on, and at point ‘3’ the heater is off. (b): capacitance variation with temperature in hydrogen

atmosphere

It seems that hydrogen has no noticeable effects on the capacitance of PZT plates

under the conditions of this work (T=100C, p=10 MPa, time=200 hours). The reason why

the capacitance drops after reaching the maximum value (point 2 in Figure 40a) could be the

Page 87: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

70

increase in the temperature of the samples. The dielectric constant of ferroelectric ceramics

shows an aging effect after any abrupt thermal changes, or application of strong mechanical

stress [66], due to the rearrangement of the ferroelectric domains. The experiment was done

just once for the samples with Ag/Pd electrodes. The error for measured capacitance values is

within 0.05% according to the specification of the LCR meter used. Similar results were also

seen for the sample with Ag electrodes (Figure 42), in repeated experiments. The general

conclusion from the above results is that the dielectric constant of PZT plates will not change

in high-pressure hydrogen condition (p=10 MPa, T=100C) after 200 hours. However, in the

next section we will see that considerable changes occur in PZT microstructure and electrical

properties in high temperature conditions.

1 10 100

Time (hour)

1.6

1.7

1.8

1.9

2

2.1

C(n

F)

Figure 41- Capacitance of PZT sample with Ag electrode in high-pressure argon atmosphere

Page 88: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

71

1 10 100

Time (hour)

1.6

1.7

1.8

1.9

2

2.1

C(n

F)

Figure 42- Capacitance of PZT sample with Ag electrode in high-pressure hydrogen atmosphere

Page 89: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

72

5.2 High-Temperature conditions (T=450-600C, p=0.013 MPa)

5.2.1 H2 effects on PZT microstructure

To address the kinetics of interactions of hydrogen with PZT plates, a high-temperature

hydrogen treatment was considered as it is described in Section 4-2. While for the samples

hydrogen-treated in the low-temperature / high-pressure conditions we did not observe

noticeable changes in the microstructure of PZT (except for the sample with Ag/Pd

electrodes hydrogen treated for 400 h, Figure 39), we observed considerable changes in the

microstructure of the samples hydrogen-treated at high temperatures. Figure 43 compares

the microstructure of the side surface of the PZT plate with silver electrode in the as-

received condition and after heat treatment at 400 C for 2 hours. According to this figure,

noticeable damage has occurred on the surface of the sample hydrogen-treated at high

temperatures. Another aspect which was observed in some of the samples was the lead

reduction on the surface of some of the samples, in the regions adjacent to the electrodes.

This is shown in Figure 44 for the sample hydrogen-treated at 600C for 2 hours, wherein

lead particles on the surface of PZT can be observed as the bright-contrasted sub-micron

particles. It was considered that these particles were metallic lead based on the much higher

ratio Pb/O (14.0), compared to the surrounding PZT particles (2.8) (Table 2). However, no

lead reduction was observed in the cross- section of the samples.

Page 90: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

73

Table 2- The EDX analysis for the bright particles in Figure 44b

Element O Ti Zr Pb

Concentration (wt%)

Bright particles 22 6 10 62

Concentration (wt%)

PZT grains 6 5 5 84

Ag electrode

PZT

Figure 43- Image from the side surface of the PZT plate with Ag electrode for (a) as-received and (b) after

hydrogen treatment (for 2 h / 400C / p= 0.013 MPa)

Ag electrode

PZT

Figure 44- Metallic lead in hydrogen treated PZT samples (for 2 h / 600C / p= 0.013 MPa)

Page 91: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

74

XRD analysis was also done on the samples to study the possible phase changes in the

PZT after hydrogen treatment. To do this analysis, the silver electrodes were removed from

the plates after the hydrogen treatment. The results are shown in Figure 45. Comparing the

XRD results, one can say that no new peaks have formed and PZT still has maintained its

tetragonal structure after hydrogen treatment.

Figure 45- The XRD pattern for as-received and hydrogen treated PZT with Ag electrodes

(for 24 h / 550°C /p=0.013 MPa)

According to Figure 45, a small shift can be observed in the XRD pattern of PZT after

hydrogen treatment. Moreover, it can be seen that the relative intensity of some peaks

changed after hydrogen treatment. The reason for these changes could be due to the

hydrogen dissolution and the formation of lattice defects such as lead and oxygen vacancies

[35]. Nevertheless, the important point is that no new peaks have formed in the XRD pattern

of PZT samples after hydrogen treatment, indicating that no new phases, detectable by XRD,

have formed inside PZT. We also investigated the XRD patterns for other samples treated at

10 20 30 40 50 60

2

as-received

after treatment

(00

1)

(10

0)

(11

1)

(00

2)

(20

0)

(10

2)

(21

0)

(11

2)

(21

1)

(11

0)

Page 92: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

75

500°C and 600°C, and again we did not observed any new peaks after hydrogen treatment

(Figure 46). Indeed, other studies about the structural changes in PZT after hydrogen

treatment also have not reported the formation of any new XRD peaks [9, 18, 35]. However,

changes in the lattice parameters of PZT after hydrogen treatment due to hydrogen

dissolution and formation of lattice defects such as lead and oxygen vacancies have been

reported [9, 18, 35]. These changes also could be the reasons for the observed small shifts in

the XRD pattern after hydrogen treatment (Figure 45).

Figure 46- The XRD pattern for hydrogen treated PZT with Ag electrodes at 500C and 600C

5.2.2 H2 effects on PZT electrical properties

Figure 47a shows a typical variation of the capacitance with the temperature and

duration of H2 exposure. The hydrogen treatment begins at t=0 in Figure 47 (point c in

Figure 23) and, as seen in the inset in Figure 47a (magnifying the effects in the first 40 min of

the exposure), the sample capacitance starts changing immediately after the exposure to H2

10 20 30 40 50 60

2

600C

500C

(00

1)

(10

0)

(11

1)

(00

2)

(20

0)

(10

2)

(21

0)

(11

2)

(21

1)

(11

0)

Page 93: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

76

began. The reference test of 1 hour heat treatment in Ar atmosphere did not result in

noticeable changes in the PZT resistance or capacitance.

0 400 800 1200Time (min)

0

10

20

30

40

C (

nF

)

(a)

0 20 401.6

2

2.4

2.8

0 400 800 1200

Time (min)

20

24

28

32

36

40

ZR

e (k

)0

100

200

300

R (

k

)

(b)

Figure 47- (a) The general trend of PZT capacitance variation with time in hydrogen atmosphere at 500°C; (b)

measurements for the real part of the impedance (ZRe) and the calculated values of R according to equation (8)

(the data is obtained at the constant frequency of 1 kHz)

Figure 47b shows that the variation of the real part of the impedance ZRe (as measured

by the LCR meter) corresponds to the variation of capacitance observed in Figure 47a; the

secondary Y axis shows the resistance calculated from equation (8). While resistance

variation with time shows the same trend as the variation of capacitance, the relative amount

of the change of resistance is different from the relative change of capacitance. For example,

while after 20 mins the capacitance increases from 2 nF to 2.75 nF, there is a decrease in

resistivity to 60% of the initial value (i.e. from 250 k to 100 k ). To examine the trend of

resistance decrease, ZRe vs. ZIm curves were re-plotted and a typical result is shown in Figure

48a for the sample hydrogen treated at 550C. The intersection of the semi-circles in Figure

Page 94: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

77

48a with the ZRe axis shows the ohmic resistance of the PZT plates [49]. In this way, we

measured the ohmic resistance of the PZT plates for different durations of hydrogen

exposure and the results are shown in Figure 48b. According to this figure, the resistance

decreases at almost the same rate as in Figure 47b, showing the general trend for the

resistance variation. The curves in Figure 48 also show the relatively fast decrease in

resistance in the first 20 min of H2 exposure. For the sample hydrogen-treated at 550C the

resistance decreases from 500 k to 200 k, and for the sample hydrogen-treated at 600C

the resistance decreases from 220 k to 100 k after 20 min of hydrogen exposure. The

decrease in resistivity after hydrogen treatment is also reported elsewhere [33].

0 200 400 600

ZRe(k)

0

200

400

600

ZIm

(k

)

0 min

5 mins

30 mins

Figure 48- The ZRe -ZIm plot for PZT plate heat treated at 550C and the resistance determined using the ZRe -ZIm

plots for PZT; the noise in the ZRe-ZIm plots corresponds to the times when the heater was on.

Figure 49 illustrates capacitance change for 530, 550 and 600C, showing a similar

trend, which suggests similar structural changes in PZT, although the time until the

capacitance reaches the maximum value decreases with increasing the temperature (600 min

0 50 100 150 200 250

Time (min)

0

100

200

300

400

500

R (

k

)

550 oC

600 oC

Page 95: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

78

at 530C, 250 min at 550C, and 40 min at 600C). Therefore, we may conclude that the

structural changes affected by hydrogen in PZT are thermally activated processes, as it is

proposed in [14].

0 200 400 600 800 1000

Time (min)

0

20

40

60

C (

nF

)

530oC

0 100 200 300

Time (min)

0

20

40

60

80

C (

nF

)

550°C

0 20 40 60Time (min)

0

200

400

600

800

1000

C (

nF

)

600°C

Figure 49- The variation of PZT capacitance at 530°C, 550°C and 600°C

Based on the above data, it is hypothesized that the variations observed in the

capacitance and resistance of PZT are due to the structural modifications caused by the

presence of hydrogen in PZT. These structural changes seem to conform to “Isothermal

Page 96: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

79

Solid-State Reactions” where the process occurs by the nucleation and the growth of the

product nuclei [68]. A typical plot α-time (t) for the solid state chemical reactions is shown

in Figure 50 [68], where α is the fraction of the volume converted to the product of reaction.

Before reaching point A, the short progress of the chemical reaction happens in the less

stable sites of the media in which reaction is occurring; (A-B) step shows the incubation time

needed for the development of growth nuclei; (B-C) is the much longer acceleratory period

related to the development of the stable nuclei formed in the previous step; in this period

new stable nuclei may also form; (C-D) is the step where the further expansion of the nuclei

is not possible due to the impingement and consumption of reactant and this leads to the

deceleratory or decay period [68]. Kinetics of such reactions can be written in the form of

f(α)=kt where k depends exponentially on temperature, k=A exp (-Q/RT), where Q is the

activation energy for the process [68-69], and A is a constant.

Time

0

0.2

0.4

0.6

0.8

1

1.2

A B

CD

start of reaction

completion of reaction

Figure 50- The general trend for the isothermal α - time plots having different time steps, time equal to zero

shows the start of the reaction [68]

Page 97: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

80

“The rate-determining step can be either (i) diffusion, i.e. the transportation of

participants to, or from, a zone of preferred reaction, or (ii) a chemical reaction, i.e. one or

more bond redistribution steps, generally occurring at a reaction interface” [68]. The model

which is frequently used to describe the sigmoid isothermal α – time plots is the Avrami-

Erofeev (A-E) relation, also known as Johnson-Mehl-Avrami-Kolmogorov, or JMAK,

equation [68-69]:

[–ln (1– α)](1/n)=kt (28)

or

α = 1 – exp (–ktn) (28a)

where n is a constant. Depending on the nucleation and growth conditions, n can have

different values, as summarized in

Table 3.

Table 3- Different values of exponents for the equation (28) [68]

Model Phase Boundary control (n) Diffusion control (m)

Three-dimensional growth (Spherical particles of reactant)

Nucleation rate

1. Constant

2. Zero (instantaneous)

3. Deceleratory

4

3

3-4

2.5

1.5

1.5-2.5

Two-dimensional growth (Laminar particles of reactant)

Nucleation rate

1. Constant

2. Zero (instantaneous)

3. Deceleratory

3

2

2-3

2

1

1-2

One-dimensional growth (Lath-shaped particles of reactant)

Nucleation rate

1. Constant

2. Zero (instantaneous)

3. Deceleratory

2

1

1-2

1.5

0.5

0.5-1.5

Page 98: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

81

The above equations are applicable to our condition if we define α as follows:

α= [C(t) – C()]/[Cmax – C()] (29)

where C(t) shows the capacitance at time t, Cmax shows the maximum capacitance, and is

the incubation time (up to point B in Figure 50). Considering the equations (28) and (29) we

have to fit the data to the following equation:

α= [C(t) – C()]/[Cmax – C()]=1 – exp (–k(t-)n) (30)

Therefore, to fit our data to the equation (30) we need to find the values for , n, and k. The

best fits were obtained with the values reported in Table 4. The results of the fitting for the

two temperatures of the 550C and 600C are shown in Figure 51 shows the results of the

fitting for all temperatures investigated in this work.

0 100 200 300

t- (min)

0

0.2

0.4

0.6

0.8

1

1.2

(a)

0 20 40 60

t-(min)

0

0.2

0.4

0.6

0.8

1

1.2

(b)

Figure 51- The results of fitting the capacitance data to equation (30) for the temperatures of 550C (a) and

600C (b)

Page 99: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

82

ln(t-)

-15

-10

-5

0

5

ln(-

ln(1

-))

500

550

600

(a)

530

0.001 0.0011 0.0012 0.0013 0.0014

1/T (K-1)

-20

-15

-10

-5

ln(k

)

Q=42,433 ± 8087 J/mol(b)

Figure 52- (a) The results of fitting the capacitance data to equation (30); (b) the activation energy of

hydrogen diffusion, obtained from the fit

Table 4- The fitting values obtained for the equation (30) Temperature n (sec) ln(k) R-squared

500 C 1.690.2 9000500 -17.480.72 0.9813

530 C 1.360.2 3700250 -12.910.5 0.9861

550 C 1.810.2 2100250 -15.760.46 0.9863

600 C 1.480.2 45050 -9.990.24 0.9850

Based on the obtained values for k, the activation energy for the limiting process was

obtained to be 0.440.09 eV, Figure 52b. This is close to the reported activation energy for

diffusion of H+ in zirconate and titanate perovskite oxides, i.e. 0.44-0.50 eV for BaZrO3-(2-

10)%Y [70], 0.83 0.62 eV for BaZrO3 [71], 0.50 0.22 eV for SrTiO3 [71], 0.420.30 eV for

CaTiO3 [71]. Therefore we can hypothesis that the rate-limiting phenomenon for the

structural changes observed in PZT is the diffusion of protons. As activation energy for the

oxygen diffusion is about 1 eV [71], it appears that protons have the main contribution to the

structural degradation of PZT.

Comparing the average value of n (1.56) determined in this work with the data in

Page 100: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

83

Table 3, one may hypothesize about the nucleation and growth mechanism by which

the structural changes occur in PZT (for example, three-dimensional, two-dimensional or

one-dimensional growth). During the growth, the coalescence of the developed nuclei and

the ingestion of undeveloped nucleation sites may occur, as the coalescence and ingestion are

characteristics of the sigmoid isothermal α – time plots [68]. It should be noted that the value

of n itself is not enough to confirm the specific nucleation and growth mechanism; some

independent confirmation, such as structural observations, are also needed. However, the

value of n suggests the rate-limiting phenomenon is hydrogen diffusion, and not its chemical

reaction with oxygen. The mechanism responsible for diffusion (or conduction) of protons in

oxides, especially in perovskite oxides, is believed to be the Grotthuss mechanism [72],

schematically shown in Figure 53.

Figure 53- The Grotthuss mechanism for diffusion of protons in PZT, including the reorientation and hopping

of protons between oxygen onions

According to this model, the diffusion of protons consists of (i) transfer from one

stable position to another stable position in the structure (hopping) and then (ii)

reorientation of the proton for transferring to another site. Therefore, it is probably the

proton reorientation and hopping between the different oxygen atoms which leads to the

Page 101: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

84

expansion of the new structure formed in PZT. These protons penetrate into PZT and form a

solid solution within PZT. Figure 54 schematically illustrates this process.

Another explanation for the variation of the electrical properties would be that a new

“phase”, with a new crystal structure is forming in PZT, where protons become a part of its

chemical composition. Formation of new phases has been reported after H2 exposure for

other oxides, such as Sr6Ta2O11, and Ba2In2O5 [73-76]. The formation of the new hydrate

composition was reported to occur by a disorder-order phase transition leading to saturated

solid solution of the oxide and protons [73]. We however did not see any differences in the

XRD spectrum of the as-received and hydrogen treated samples (Figure 45), either because

no phase transition has occurred or due to the relatively small volume of such phase, e.g. <0.5

vol% [75].

Figure 54- Hypothetical schematic of the different modes which can be assumed for the dissolution of hydrogen

in PZT; (a) where the diffusion of protons into PZT occur uniformly from the surface; in this case the diffusion

equation with proper initial and boundary equation could be used for determining the total amount of protons

in PZT; (b) where the diffusion of protons can occur from limited places in the PZT; in this case the nucleation

and growth models can be used to describe the total amount of protons in PZT

Page 102: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

85

To further study the effect of hydrogen on the electrical properties of PZT, we used

the dielectric spectroscopy technique. The technique measures the dissipation factor DF and

dielectric constant ε of the material in a wide range of frequency (from mHz to MHz) and

temperature, which are then used to analyze the data by fitting the results to one of the

physical or mathematical models [38] (the relevant relationships between dielectric

parameters of materials are listed in Chapter 2-3, equations (21-23)). We investigated the

dielectric constant of PZT after hydrogen treatment in the frequency range of 12-200 kHz

and in the temperature range of 25-400°C. The data obtained from these experiments were

analyzed to determine the effect of hydrogen on the dielectric properties of PZT. We have

also used this data to correlate the changes in dielectric properties with the effects of

hydrogen on the microstructure of the PZT plates.

Figure 55 shows the variations in capacitance (C) and dissipation factor (DF) after

hydrogen gas treatment versus temperature at the frequency of 1 kHz. A capacitance peak is

observed at 375C, which could be attributed to the phase transition of PZT from cubic to

tetragonal phase at the Curie temperature. Because the high dielectric constant of PZT at the

Curie temperature is due to the dipoles, it appears that these dipoles are still present in PZT

after the hydrogen treatment. However, the capacitance at the peak (20 nF) was only about

half of the capacitance before the hydrogen treatment (40 nF), therefore the dipoles inside

the PZT are probably affected by hydrogen. The decrease in dielectric constant after

hydrogen treatment has been reported before [31]. What should be pointed out is that the

existence of dipoles even after hydrogen treatment does not necessarily mean that their

Page 103: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

86

directions could be switched with changing the direction of electric field as it proposed by

Aggarwal et al. [31]. They suggested that the [OH]− group acts as a fixed dipole, which does

not allow switching of the ferroelectric dipoles and domains inside the PZT [31], and

therefore PZT may not show polarization hysteresis after hydrogen treatment even if it has a

tetragonal structure [31]. Accordingly, the lower value of the capacitance may be attributed

to the interaction of [OH]- dipoles with the dipoles inside the PZT.

Figure 55- Changes of capacitance C and dissipation factor DF of hydrogen-treated sample (for 24 hrs / 550C /

p= 0.013 MPa) versus temperature (The thick grey line shows the changes of capacitance for as-received

sample)

Figure 55 also shows two relaxation peaks (R1 and R2) in the dissipation factor, while

no such peaks were observed for the as-received sample. Generally, there are two conditions

which can lead to such relaxation peaks in the dissipation factor. First, when there are

dipoles in the dielectric medium and therefore, the dipolar polarization mechanism is active

[38] (Figure 56). For this mechanism, the relaxation peak occurs when the natural vibrational

100 200 300 400 500

Temp (°C)

0

10

20

30

40

50

C (

nF

)

0

1

2

3

4

5

DF

R1

R2

Page 104: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

87

frequency of such dipoles coincides with the frequency of the applied voltage [38].

Therefore, one might conclude that the relaxation peaks in Figure 55 may be due to the

dipoles formed in PZT after the hydrogen treatment.

a

+

-

+

- +

-

+

-

b

+

-

+

-

+

-

+

-

+

-

+

-

c +

-

+

-

+

-

+

-

+

-

+ + + + + + + + + +

_ _ _ _ _ _ _ _

+ + + + + + + + + +

_ _ _ _ _ _ _ _

Figure 56- Schematics of the dipolar polarization mechanisms, wherein direction of the dipoles changes with

changing the direction of applied voltage

Another reason for the occurrence of the relaxation peaks could be the Maxwell-

Wagner (MW) polarization mechanism, which is active in inhomogeneous systems, where

the dielectric material consists of regions with different electrical properties (Figure 57). This

"extrinsic" polarization mechanism can be explained by considering the heterogeneity of the

system without any microscopic polarization process inside the sample [38, 77]. When a

voltage is applied across the dielectric medium, due to the differences in electrical

conduction in different regions of a heterogeneous material, charge accommodation occurs at

Page 105: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

88

the interfaces of the different regions, leading to the increase in the capacitance of the

sample. The frequency response of such a system is similar to the frequency response of a

Debye relaxor [77], leading to a Debye type relaxation peak. The relaxation time in such

heterogenic systems depends on dielectric constant and conductivity of the different regions

in that medium. To further understand the physics behind such relaxation peaks, we have

changed the frequency of the voltage applied to the samples under the hydrogen atmosphere

at different temperatures, ranging from 200C to 325C for the first relaxation peak and from

22C to 42C for the second relaxation peak.

a

b

c

+ + + + + + + + + +

_ _ _ _ _ _ _ _

+ + + + + + + + + +

_ _ _ _ _ _ _ _

_ _ _ _ ___

_

+ + + + ++ + +

____

+++++

_+

++

___

Figure 57- Schematics of the Maxwell-Wagner polarization mechanism, wherein differences in the electrical

properties of different regions cause charge accumulation at the interfaces between the different regions,

leading to the increase of capacitance

Page 106: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

89

Relaxation Peak #1 (R1)

Figure 58 shows the changes of ε´ and ε´´ (where ε´´=DF ε´) versus frequency in the

temperature range of 200-325C, with 25C increments.

1 100 10000 1000000

f(Hz)

0

2000

4000

6000

8000

'

(a)

325C

200C

1 100 10000 1000000

f(Hz)

0

1000

2000

3000

''

(b)

325C

200C

Figure 58- Variations of ε’ and ε’’ for hydrogen-treated samples with the frequency of applied voltage in the

temperature range of 200-325C, with 25C increments

While decreasing the temperature, the frequency at which the relaxation peak occurs

also decreases. This may be because the polarization process which has led to such relaxation

becomes slower at lower temperatures, as the relaxation time is inversely proportional to the

frequency at which the maximum occurs (fMax=1/τ, where τ is the relaxation time). This is

typical condition for the dipolar polarization, where dipoles change their positions with

changing the direction of the applied electric field [38]. At higher temperatures, such dipoles

can change their direction faster and as such, the relaxation peak moves to higher

frequencies [38]. Therefore, one may conclude that the first peak observed in Figure 55 is

due to the dipoles formed in PZT after the hydrogen treatment. To further investigate the

Page 107: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

90

nature of such dipoles (such as the activation energy for the relaxation process), the results

should be fitted to one of the physical models of this polarization mechanism.

The model most frequently used for the description of such relaxation peaks is the

Debye model (Chapter 2-3, pages 32-33, equations (21-22)). We tried to fit our data to the

Debye equation, but the fit was poor (Figure 59). This is likely because the Debye equation

assumes that the dipoles do not interact with each other, but this is not usually valid for

dipoles inside a dielectric medium.

1 100 10000 1000000

f(Hz)

0

1000

2000

3000

'' Debye model

Figure 59- The results of fitting the ε´ and ε’’ data to the Debye equation for at T= 325C, for PZT hydrogen-

treated samples (for 24 hrs / 550C / p= 0.013 MPa)

A more flexible model, commonly used for modeling dielectric constant data, is the

Havriliak–Negami equation [38-39] (Chapter 2-3, page 33, equation (23)). The fit based on

the Havriliak–Negami equation is much better than the fitting results based on the Debye

equation, Figure 60 (Table 5 compiles the best fit parameters). An iterative MATLAB code

was developed and used for fitting procedure.

Page 108: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

91

1 100 10000 1000000

f(Hz)

0

2000

4000

6000

8000

'(a)

HN model

1 100 10000 1000000

f(Hz)

0

1000

2000

3000

''

(b)

HN model

Figure 60- The results of fitting the ε´ and ε’’ data to the Havriliak–Negami equation for at T= 325C, for PZT

hydrogen-treated samples (for 24 hrs / 550C / p= 0.013 MPa)

Table 5- The fitting values obtained for the HN equation

Temperature Δε=εs-ε (sec) θ β

325 C 7300 0.0019 0.8 1

300 C 4900 0.0029 0.8 1

275 C 4100 0.004 0.7 1

250C 3200 0.0078 0.8 1

225C 2700 0.025 0.8 0.8

200C 2000 0.05 1 0.6

The activation energy for the relaxation process behind the relaxation peak #1 can be

evaluated from the values obtained for . As is the average residence time of an ion at any

given site, it changes with temperature according to =0×exp(-ΔH/RT), where ΔH is the

activation energy for ions jumping from one position to another [38]. This equation shows

that the kinetics of the relaxation of the system follows the Arrhenius law, i.e. the relaxation

time of the system decreases with increasing the temperature (as vibration frequency of ions

increases, the probability of ions jumping from one position to another position increases,

hence the average residence time or decreases). The relaxation time fit to the above

equation (Figure 61) yields the activation energy of about 0.66 eV.

Page 109: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

92

0.0016 0.0018 0.002 0.00221/T (K-1)

-10

-5

0

ln(

)

Q=63,760 J/mol

Figure 61- The activation energy for the ion jumping, obtained from the fits, for PZT hydrogen-treated samples

(for 24 hrs / 550C / p= 0.013 MPa)

One of the microstructural changes reported for PZT after hydrogen treatment is the

lead reduction and owing to that, the presence of lead vacancies in PZT [35]. Indeed we also

observed lead reduction in our samples after hydrogen treatment (Figure 44). Therefore, the

relaxation peak #1 may be tentatively assumed to be due to the hopping of lead cations in the

PZT lattice. However, the reorientation of such dipoles should be relatively slow below

300C, where the diffusion of lead cations is very slow [37]. The activation energy for the

diffusion of lead ions in PbTiO3 has been reported to be about 1.89 eV [37], which is

significantly higher than the energy obtained here for the reorientation of the dipoles.

Therefore it can be concluded that the first relaxation peak is not due to the hopping of lead

cations. Another structural change proposed for PZT after the hydrogen treatment is the

existence of oxygen vacancies in the lattice of PZT [29, 35], so the relaxation peak could be

due to the hopping of oxygen anions in the PZT lattice. However, we believe that the

temperature is not high enough for the oxygen ions to have enough mobility in the PZT

Page 110: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

93

lattice. Therefore, the dipole reorientation or the relaxation peak cannot be due to the

hopping of oxygen anions. Kamishima et al. [40] investigated the dielectric properties of the

Yb-doped SrZrO3 after hydration in water atmosphere. While they did not observe any

relaxation peaks for the pure SrZrO3, they did observe a relaxation peak for the sample with

1 wt% Yb. For samples with more than 1 wt% Yb, they observed yet another relaxation

peak. The activation energy obtained for the first relaxation peak was about 0.58 eV, and

they attributed this activation energy to the Yb-OH dipoles. They also attributed the second

relaxation peak, observed in samples with more than 1%Yb, to Yb-OH dipoles in the Yb-

clusters. Therefore, we might tentatively conclude that the relaxation peak which we also

observed is due to the dipoles formed by the protons with the dopant (Nb) in the PZT. We

realize that more experimentation and analysis is needed to fully confirm this hypothesis.

Relaxation Peak #2

Figure 62 shows the variation of dissipation factor (DF=ε´´/ε´) with frequency in the

temperature range 22-47C. When temperature increases, the frequency at which the

maximum occurs moves to lower values. If we assume that this relaxation peak is due to the

reorientation of dipoles inside the PZT under the applied electric field, then the movement

of the peak to the higher frequency values with decreasing temperature means that the

hopping of ions becomes slower with increasing temperature. However, this cannot be true,

as the vibration of the ions increases at higher temperatures, hence the possibility for their

hopping increases. Therefore, we can conclude that this peak is not due to the dipoles inside

Page 111: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

94

the PZT. This unusual direct dependency of the relaxation time with the temperature was

also observed in other systems. For example, it has been reported for

Na58(AlO2)58(SiO2)136mH2O zeolite (NaY) of the faujasite type, and it was assigned to the

relaxation of the water molecules confined inside the molecular cages of NaY [78-79]. A

similar unusual relaxation process has also been observed for potassium tantalate niobate

(KTN) crystal doped with copper, where relaxation occurred below the ferroelectric phase

transition [80]. This relaxation process has been attributed to “the reorientation of virtual

dipoles provided by the Cu ions hopping between different states of local equilibrium” [81].

Figure 62- Variation of DF with frequency in the temperature range 22-42C, for PZT hydrogen-treated

samples (for 24 hrs / 550C / p= 0.013 MPa)

Different explanations could be considered for this non-monotonic relaxation

kinetics, e.g. [81] suggests that “this situation usually occurs for ‘small’ systems where

relaxing particles become able to participate in the relaxation due to the formation of some

1 100 10000 1000000

f(Hz)

0

0.1

0.2

0.3

0.4

0.5

DF

42C

32C

27C

22C

Page 112: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

95

‘defects’ in ordered structure”. Therefore, we hypothesise that such a relaxation peak could

indicate the formation of structural defects which hydrogen produces in PZT. Indeed,

different structural defects have been proposed for PZT after hydrogen treatment, such as

oxygen and lead vacancies, and the formation of [OH]- dipoles [31]. Therefore, the

interaction between such defects and [OH]- dipoles in PZT might be the reason for the

formation of this relaxation peak. Another explanation for the second relaxation peak and its

unusual kinetics behavior could be due to the Maxwell-Wagner polarization mechanism, as

first proposed in [81-82]. As mentioned before, this polarization mechanism is active in non-

homogenous systems. In the previous section, we showed that a new structure forms inside

the PZT during the treatment in hydrogen atmosphere. Therefore, it can be assumed that the

second relaxation peak is due to the formation of a new phase in PZT, with different

electrical properties. In other words, this relaxation peak confirms our idea that a new

structure with new electrical properties forms inside the PZT during the hydrogen

treatment. It is very difficult to pinpoint the exact reason for the formation of the second

relaxation peak in the dissipation factor curve of PZT after hydrogen treatment; further

study is needed to understand the physics behind the second relaxation peak and the details

of the structural changes induced by hydrogen in PZT.

Page 113: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

96

5.3 Water-electolysis treatment of PZT

5.3.1 Microstructure

The water electrolysis technique was used to charge the PZT samples with hydrogen,

following previously reported methodology [33]. In comparison to the treatment in

hydrogen gas (which needs elevated temperatures), water electrolysis technique can be

completed even at room temperature; this inhibits the de-polarization of PZT samples due to

the effects of temperature.

Figure 63 shows SEM images of the cross-sections through PZT plates in the region

adjacent to the electrode which functioned as the cathode. As seen in Figure 63a, the

structure of PZT immediately below the electrode is different from the structure in the

center, far from the electrode. Moreover, comparing Figure 63b and c, showing the same

magnification images, one can see that the grain boundaries of PZT after water electrolysis

are extensively corroded in comparison to the microstructure of the as-received samples.

Page 114: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

97

Figure 63- Micrographs of the cross-section through PZT plate after water electrolysis: a) low-magnification

image (after 48 hours water electrolysis); b) microstructure of the corroded layer (close to the electrode); c)

microstructure in a region far from the corroded layer

According to the Figure 63, a corroded layer was formed just beneath the electrode

after the water electrolysis. To see the possible changes in the crystal structure of PZT in the

Page 115: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

98

corroded layer, the silver electrode was detached from the surface of the PZT and the XRD

test was done on the top surface of as-received sample and on the corroded layer following

the treatment. Figure 64 shows the microstructure of the PZT surface right below the

electrode (after the electrode removal). As it can be seen, the grain boundaries are

extensively corroded, which is in accordance to Figure 63b. According to Figure 65, no new

peaks have formed in the XRD pattern of samples after the water electrolysis treatment;

therefore, we can conclude that no new phase has formed inside PZT after the water

electrolysis, and PZT still has its tetragonal structure after this treatment. However, a

systematic shift to lower two-theta values in the XRD pattern of samples after water

electrolysis is observed. Moreover, the tetragonal splitting of (100), (200) peaks became more

significant after the treatment. Huang et al. have investigated the changes in PZT structure

after water electrolysis using the XRD technique, and they observed a very small increase in

the lattice parameter of PZT after this treatment [83]. Therefore, the changes observed in

the XRD pattern of PZT after water electrolysis in this study can also be due to the

dissolution of hydrogen inside the PZT and the changes of the lattice parameters of PZT. It

should be noted that Huang et al. did not observe corrosion in grain boundaries of PZT after

water electrolysis, as we did in this work. Therefore, the corrosion of the grain boundaries

and changes in the microstructure of PZT could cause the difference in the XRD pattern

after water electrolysis.

Other parameter which should be considered here is the roughness of the surface of

samples. After the water electrolysis, the interface between the electrode and PZT was

Page 116: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

99

extensively degraded, and the electrode was easily detached from the PZT surface. After

removing the electrode, we did not further polished the surface because the corroded layer

was very thin, so the surface of the sample which was used for XRD was very rough, as it can

be seen in Figure 63. Therefore, the roughness of the surface could be another reason for the

changes in the XRD pattern after the water electrolysis [61]. The conclusion from the above

discussion is that the hydrogen has diffused trough the crystal lattice and/or grain boundaries

of PZT during water electrolysis, without the formation of any new phase detectable by

XRD, and PZT still maintained its tetragonal structure after the water electrolysis treatment;

however, the hydrogen presence inside the crystalline lattice of PZT could have caused some

modifications in the lattice parameters of PZT.

Figure 64- The microstructure of PZT after water electrolysis just beneath the electrode (after removing the

electrode)

Page 117: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

100

Figure 65- XRD pattern of the as-received PZT sample versus the water electrolyzed PZT sample, using the

following parameters: I=100 mA/cm2, t=48 hours

The damage to the grain boundaries in the corroded layer can be due to the diffusion

of hydrogen atoms from the silver electrode to the grains, or preferably grain boundaries

(with more open structure) of PZT, followed by the formation of hydrogen molecules at the

grain boundaries [12]. The formation of hydrogen molecules would be accompanied by local

increase of pressure, thus stresses along the grain boundaries, which leads to cracks. If it is

assumed that the damage in grain boundaries is due to the diffusion of atomic hydrogen

predominantly along the grains boundaries, the thickness of the corroded layer can be used

for estimating the diffusion coefficient of hydrogen atoms along the grain boundaries of PZT.

The thickness of the corroded layer ( ) is proportional to the diffusion coefficient of atomic

hydrogen (D) by the relation of √ where t is the duration of water electrolysis [83].

Because the thickness of the corroded layer was not the same along the cross section of the

sample, the area of the corrosion layer was first measured and then the average thickness of

the corroded layer was calculated by dividing the area by the width of the cross section (10

10 20 30 40 50 60 70 80 90

2

as-received

after treatment

(00

1)

(10

0)

(11

1)

(00

2)

(200

)

(10

2)

(21

0)

(11

2)

(21

1)

(02

2)

(22

0)

(11

0)

Page 118: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

101

mm). The results, Figure 66, indicate that the value for the diffusion coefficient of hydrogen

in PZT at room temperature is about 9×10-11 (cm2/sec). It should again be emphasized that we

believe that this value is related to the diffusion of atomic hydrogen along the grain

boundaries of PZT, and not necessarily inside the crystalline lattice of PZT. This is lower

than the values obtained by other researchers using the same technique: 4.9×10-8 cm2/sec

[83]. This discrepancy in data might be related to the slightly different composition of the

samples, or it might be related to the recombination of hydrogen atoms into hydrogen

molecules along the grain boundaries. Moreover, the value of D= 4.9×10-8 cm2/sec is obtained

based on the advancement of a layer of a different color (yellowish to grey) in the PZT

charged in NaOH solutions at room temperatures with a current of 50 mA/cm. In our work,

we did not observe any change in color in our samples and we measured the above value

based on the thickness of the corroded layer. Therefore, this discrepancy in data might be

related to the different interactions between hydrogen and PZT, as proposed by Alvine et al

[84]. They recently investigated the hydrogen diffusion inside PZT using the proton nuclear

magnetic resonance (1HNMR) and quasi-elastic neutron scattering (QENS) techniques after

charging the samples with high-pressure gaseous hydrogen (T=100C, p=32 MPa for 1HNMR

analysis, and T=100C, p= 17 MPa for QENS analysis). They obtained different values for the

hydrogen diffusion coefficient at room temperature using these techniques; using 1HNMR

they obtained D = 6×10-14 cm2/sec, and using the QENS technique they obtained D = 3×10-6

cm2/sec [84]. Because the diffusion results were several orders of magnitude different, they

concluded that there were different diffusive processes for hydrogen inside the PZT [84].

Page 119: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

102

Therefore, the discrepancies between different hydrogen diffusion coefficient values could

be due to the different interactions occurring between the hydrogen and PZT.

0 200 400 600

time1/2 (sec1/2)

0

0.1

0.2

0.3

x (m

m)

25°C

Figure 66- The thickness of the corroded layer versus the square root of time of water electrolysis

5.3.2 Electrical properties of PZT exposed to water electrolysis

It is reasonable to assume that the changes which occur in the microstructure of PZT

will affect the dielectric properties of PZT as well. Figure 67 shows the variation of the

capacitance and dissipation factor of PZT capacitors during water electrolyzes at the

frequency of 1 kHz, showing that both parameters increase with time of water electrolysis.

The increase in capacitance and dissipation factor after water electrolysis has also been

reported in other studies [85-89].

Page 120: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

103

Figure 67- The changes of capacitance (C) and dissipation factor (DF) versus the duration of water electrolysis at

the frequency of 1 kHz

To further study the effect of water electrolysis on the electrical properties of PZT,

the capacitance and dissipation factor of the treated samples were measured versus the

frequency of applied voltage. Figure 68a shows the variation of capacitance (C) and

dissipation factor (DF) versus frequency for the PZT samples before and immediately after

water electrolysis, for 6 hours. The capacitance right after water electrolysis was higher than

the initial value. The same trend is also observed for the dissipation factor above 200Hz.

After aging for 24 hours in air however, the capacitance decreased significantly below the

initial values (e.g. at 103 Hz, the capacitance values were 1.7 nF, 1.8 nF and 0.2 nF for the as-

received, hydrogen-treated for 6 hrs, and 24 hrs aged samples respectively), as also shown in

other researchers’ studies [33]. The dissipation factor was however higher for the aged

sample versus the hydrogen-treated sample, for frequencies below 20 kHz. While for the as-

0 20 40 60

time(hours)

1.4

1.5

1.6

1.7

1.8

C(n

F)

0.012

0.016

0.02

0.024

DF

Page 121: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

104

received sample the dissipation factor was almost constant, after hydrogen treatment and

aging, it increases with decreasing frequency, as seen in Figure 68b.

1 100 10000 1000000

f(Hz)

0

0.5

1

1.5

2

2.5

C(n

F)

(a)

Figure 68- Variations of electrical properties after 6 hours water electrolysis and subsequent aging in air: a)

capacitance (C); b) dissipation factor (DF) (: as-received,: after water electrolysis, : after aging)

Another interesting finding is the relaxation peak observed in the sample after water

electrolysis for 10 hours, and after 24 hours aging in air (Figure 69).

1 100 10000 1000000

f(Hz)

0

0.5

1

1.5

2

2.5

C(n

F)

(a)

1 100 10000 1000000

f(Hz)

0

0.1

0.2

0.3

DF

(b)

Figure 69- Variations of electrical properties after 10 hours water electrolysis and subsequent aging in air: a)

capacitance (C); b) dissipation factor (DF) (: as-received,: after water electrolysis, : after aging)

1 100 10000 1000000

f(Hz)

0

0.1

0.2

0.3

DF

(b)

Page 122: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

105

The formation of the relaxation peak is better seen in Figure 70, which shows the

changes of dissipation factor and capacitance for the sample which was water-electrolyzed

for 48 hours. According to Figure 70, no relaxation peak in DF was observed for the samples

tested immediately after water electrolysis; however after aging for 10 hrs, a relaxation peak

starts to form (at about 105 Hz), and can be clearly observed at about 103 Hz after ageing the

samples for 24 hrs. The relaxation peaks in the dissipation factor have also been reported for

other oxides [85-88], but not for PZT.

Figure 70- Variations of capacitance (C) and dissipation factor (DF) after water electrolysis for 48 hrs, and

subsequent aging at room temperature in air (■: as-received, : after water electrolysis, ▲:after 10 hours aging,

: after 24 hours aging)

Considering the previously published results, a few issues need to be addressed: the

first issue is why does the dielectric constant increase during water electrolysis, and then

decreases after aging in air. It should be noted that the increase in the dielectric constant

during water electrolysis and further changes in capacitance during aging were also reported

1 100 10000 1000000

f(Hz)

0

0.5

1

1.5

2

2.5

C(n

F)

(a)

1 100 10000 1000000

f(Hz)

0

0.1

0.2

0.3

0.4

0.5D

(b)F

Page 123: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

106

for oxides such as TiO2, SrTiO3, BaTiO3, CaCu3Ti4O12, BiFeO3, and WO3 [85, 88]. Chen et al.

proposed that the increase in dielectric constant could result from the dipoles formed by the

protons inside the oxide [86], related to the complexes formed by protons with structural

defects such as oxygen vacancies. It was assumed in [86] that during the aging in air, such

protons leave the oxide and this leads to the recovery of the electrical properties, or to

decrease of the dielectric constant, as it was observed for BaTiO3. Another mechanism which

could be considered for the increase of the dielectric constant after water electrolysis is in

accordance to that proposed in the work of Park and Chadi [30]; they investigated stable sites

of protons in PbTiO3 using first-principles calculations. For the tetragonal phase of PbTiO3,

their results show that “the direction of the [OH]– dipole is favorably aligned with the host

polarization” [30]. Thus [OH]– should enhance polarization of the spontaneous dipoles in

PZT, and therefore this could be the reason for the increase in the dielectric constant of PZT

after hydrogen treatment. As the bond between proton and oxygen is strong, protons

attached to oxygen atoms cannot easily diffuse out from the PZT bulk. Therefore, it might be

concluded that the [OH]– dipoles could not be the reason for the increase in the dielectric

constant right after water electrolysis.

The mechanism which we propose here for the increase of the dielectric constant, not

mentioned in the previous studies [86-89], is the Maxwell-Wagner (MW) polarization, active

in inhomogeneous systems where the dielectric material consists of regions with different

electrical properties [77]. As mentioned before (Chapter 5-2-2), this "extrinsic" polarization

mechanism can be explained by considering the heterogeneity of the system without any

Page 124: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

107

microscopic polarization (dipolar polarization) process inside the sample [38, 77]. Due to the

differences in electrical conduction in different regions of such heterogeneous material,

charge accommodation occurs at the interfaces, leading to increase of capacitance of the

sample. According to Figure 63a, although a corroded layer formed beneath the electrode,

most of the sample was unaffected. Therefore, the assumption of an inhomogeneous

dielectric medium after water electrolysis is reasonable for our samples. As such, we believe

that the increase in capacitance right after the water electrolysis is due to the formation of

the corroded layer, and to the difference in the electrical properties of this layer and the

unaffected layer of PZT. Furthermore, during the aging, hydrogen atoms diffuse out from the

corroded layer, and this leads to further changes in capacitance.

The formation of the relaxation peaks could also be explained by the MW

polarization mechanism. The microstructure of PZT after water electrolysis can be

considered as a dielectric made of two different layers perpendicular to the electric field. It

can be shown that the frequency response of such layered system is similar to the frequency

response of a Debye relaxor [77], leading to a Debye type relaxation peak. In our opinion,

this is the main reason for the observation of the relaxation peak in the dissipation factor.

The relaxation time () can be evaluated as = (C1+C2)/(G1+G2) where C1, C2 are the

capacitance, and G1, G2 are the conductance of corroded and un-affected layers, respectively

[77]. Accordingly, the frequency at which the relaxation occurs (f=1/) depends on the

electrical properties of the different regions, and variations in these electrical properties will

change the relaxation time and the frequency at which relaxation occurs [38, 77]. We

Page 125: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

108

propose that the changes in the capacitance and the movement of the relaxation peak to

lower frequencies with sample aging could be explained by the MW polarization mechanism.

When hydrogen atoms diffuse out of the PZT bulk during aging, the electrical properties of

the corroded layer change, and this changes the frequency at which the relaxation occurs.

The relaxation peak was however not observed for the samples exposed for <10 hours to

water electrolysis, likely because of the insufficient thickness of the corroded layer.

Chen et al [86] have proposed that the relaxation peak is due to dipoles related to the

complexes which protons form with structural defects in the oxide. If this is the case, the

intensity of the relaxation peaks should decrease with aging, when the protons diffuse out of

the sample and the number of dipoles inside the oxide decreases. However, the intensity of

the relaxation peaks increases as more hydrogen diffuses out from the sample while aging

continues (Figure 70), suggesting that the relaxation peak is not due to the dipoles, but rather

is due to the MW polarization.

Figure 71 shows the poor fit of the Figure 70 data to the Debye equation, possibly

because the thickness of the corroded layer was not uniform, leading to a distribution of

relaxation times (as it depends on the thickness for the corroded layer). Figure 71 shows that

the Havriliak–Negami equation (Chapter 2-3, equation 23) fits well our experimental results,

for ε = 450, εs = 2500, = 0.012 sec-1, θ = 0.9, and β = 0.4.

Page 126: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

109

f (Hz)

0

200

400

600

800

1000

''

10310 105

Debye

Havriliak-Negami

Figure 71- The results of fitting the ε´´ (ε´´=DFε´) data to Debye and Havriliak–Negami equation. An

iterative MATLAB code was developed and used for the fitting procedure.

Typical trends for the changes of the dielectric constant and dissipation factor of a

leaky capacitor (due to the electronic conduction) are shown in Figure 72a. Figure 72b also

shows the typical trend for the changes of the dielectric constant of a capacitor with mobile

charges which can move by hopping.

Figure 72- The changes of the capacitance (C) and dissipation factor (DF) for (a) a leaky capacitor with

electronic conduction, (b) for a capacitor with hopping charge carriers adapted from [38]

Page 127: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

110

The trends shown in Figure 72 are similar to the trends observed for our samples,

aged in room conditions after water electrolysis (i.e. compare Figure 68 and Figure 72). This

trend is not obvious in Figure 70 because of the existence of the relaxation peak, and because

of the limits on the frequency that can be used. Therefore, it can be concluded that the

ohmic resistance of the PZT plates decreases after being charged with hydrogen. The

increase in the electrical conduction of PZT after water electrolysis is also reported in other

studies [33, 83].

Different mechanisms could be considered for the decrease in the resistivity of the

samples. The first mechanism relates it to the formation of oxygen vacancies (2H + O2- →

H2O + ); ionization of the vacancies (

) will contribute up to two electrons available for

conduction [89]. Another possible mechanism includes the ionization of hydrogen atoms

inside the lattice of PZT (H → H+ + e-) [31], with the electrons produced available for

conduction through hopping [31]. While these mechanisms can explain the increase in the

dissipation factor, they cannot explain why the dielectric constant decreases after water

electrolysis. The reason could be damage to the grain boundaries of PZT, or reaction of

protons with PZT and [OH]– dipoles formation. These dipoles could hinder the switching of

the dipoles inside PZT, and therefore affect the movement of the domain walls, and

consequently decrease the dielectric constant of PZT [82]. The disappearance of switchable

polarization hysteresis of PZT after hydrogen treatment has been also attributed to the

formation of [OH]– dipoles, which inhibits the switching of the spontaneous dipoles in PZT

[88]. Protons bonded with oxygen ions cannot easily leave the PZT bulk during aging, i.e.

Page 128: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

111

high temperature treatment (>700C) is needed [47]. Thus the decrease in capacitance

demonstrated in our experiments is likely due to the persistence of [OH] – dipoles within the

PZT structure, obstructing the movement of the PZT domains, and the formation of oxygen

vacancies. The decrease of dielectric constant of BaTiO3 at high frequency after water

electrolysis was previously linked to interstitial hydrogen on the domain walls of BaTiO3

[88].

Page 129: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

112

5.4 Ceramic Coatings for PZT Damage Protection

5.4.1 Alumina coatings microstructure

Attempts have been made in the past few years to solve the issue of hydrogen damage

in PZT during the forming gas annealing [55]. In this section, we investigated the possibility

of coating the PZT with alumina using the sol gel technique. Alumina has been proposed

before as a hydrogen barrier layer for PZT, and it has been shown that the alumina layer can

successfully act as a hydrogen barrier layer [55]. However, the method which we propose

here is the simple sol-gel method (as described in Chapter 4-4), which is different from the

method used in the former works for alumina deposition. In this section we show that while

the developed coating is porous, it can still significantly decrease the amount of hydrogen

damage to PZT.

Figure 73a shows the surface of the coating obtained by dip coating with pure

Boehmite sol, before drying; some excessive sol accumulation occurred on the surface of

PZT, probably because of the surface porosity on the PZT. According to Figure 73b, c, cracks

and coating detachment were observed in the places where sol accumulation occurred after

firing. Generally, the quality of the coating obtained with pure Boehmite sol was not

sufficient to demonstrate its effect on H2 damage protection of PZT.

Page 130: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

113

Figure 73- Low magnification image of the coating on the surface of PZT (a) after dip coating with pure

boehmite sol (b) before and (c) after heat treatment of the coating in the furnace, in some places on the surface

of the coating, detachment of the coating was observed

Page 131: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

114

One of the ways to improve the quality of the coating is the addition of PVA to the

sol, which increases viscosity of the sol [90]. The microstructures of the coating processed

with 10wt% PVA are shown in Figure 74. Figure 74 shows the surface of the coating directly

after dip coating (before drying it in the furnace); it can be observed that now a smooth

uniform coating has formed on the surface of PZT. Comparing the Figure 73 with Figure 74,

one can see that the quality of the coating was noticeably enhanced. The small pits which

can be seen on the coating are due to the pores on the surface of the PZT. As it will be shown

later, the sol had enough fluidity to fill out the pores on the surface of the PZT, but not

completely, so small dimples formed the surface of the coating.

Figure 74- Low magnification image of the coating on the surface of PZT after dip coating before the heat

treatment of the coating in the furnace (comparing with Figure 73a, a smooth uniform coating has formed on

the surface of PZT with the addition of PVA to sol)

Figure 75 shows the surface of PZT plates after firing the coating in the furnace; a

smooth crack-free coating developed on the surface of the sample. It should be mentioned

that during the dip coating process, excessive accumulation of the sol occurred at the bottom

Page 132: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

115

edge of the sample, and because of this, some cracks were observed at the very end of the

sample; however, the center of the sample was crack-free. The darker areas seen in Figure

75b are pores on the surface of PZT, just covered by the very thin (1-2 m) coating.

Figure 75- Low magnification (a) and high magnification (b) images of the coating on the surface of PZT after

dip coating and after heat treatment of the coating in the furnace (comparing with Figures 73b and c, a smooth

uniform coating has formed on the surface of PZT with the addition of PVA to sol)

Figure 76 shows the cross section of a sample with alumina coating after three

consecutive depositions: there is an evidence of good adhesion between the PZT and the

coating. Furthermore, no cracks were observed in the coating, and there were no pores or

cracks at the interface between the alumina layer and PZT. It should be noted that there

were some cracks in the coating at the bottom end of the sample, where sol accumulation

occurred. As it can be seen from Figure 76b, the sol had good fluidity and surface wettability,

so it was able to fill the holes and grooves on the surface of the PZT plates.

Page 133: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

116

Figure 76- Low magnification (a) and high magnification (b) images of the cross section of the alumina coating.

As it can be seen from (b), the coating had enough fluidity to fill out the pores on the surface of PZT

Leenaars et al. have measured the amount of porosity of Boehmite coatings after

treatment at different temperatures [91]. The porosity and the size of pores are presented in

Table 6, which shows that even after firing as high as 1000C, the coating still contains a

noticeable amount (>40 vol%) of porosity. They also have found that the prolonged heat

treatment for 850 hours at 400C did not change the amount of porosity. Similar results have

also been reported in other works [92], therefore, we also expect the coating to be porous

although we did not measure porosity of the coatings processed in this work.

Page 134: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

117

Table 6- Microstructural charachteristics of alumina coatings a a function of TC [91]

As seen from the high magnification image of the cross-section shown in Figure 77,

the alumina coating after firing is a conglomerate of seemingly separate agglomerates in the

range of about 20-50 nm, with <10 nm particles within the agglomerates. According to the

information provided by the manufacturer of the commercial Boehmite powder used in this

work, each of these particles are actually agglomerations of a few individual crystalline

particles (the average size for agglomerated Boehmite particles is 25 nm and for individual

crystallites is 4.5 nm [57] ). The important point which can be taken from Figure 77 is that

the coating is porous, with inter-agglomerate pore sizes < 100 nm. As the alumina coating is

a meso-porous medium, we expect that it will not prevent the access of molecular hydrogen

(H2) to the PZT surface; however it may prohibit or decrease the diffusion of atomic

Page 135: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

118

hydrogen (H). To further understand the microstructure of the coating after the heat

treatment, XRD analysis was performed to identify what phases were present.

Figure 77- High resolution image of the cross section through the alumina coating processed at 450C in air for

5 hours

Figure 78 shows the transformation sequences of Boehmite with the firing

temperature [93], suggesting that after heat treatment at 450C, the structure of the coating is

-alumina. However, XRD analysis on PZT plates with thin alumina coatings (< 5 um), did

not detected any alumina, and only PZT peaks were identified. Consequently, to understand

the crystal structure of the coatings, 10 g bulk sample of Boehmite powder was heat-treated

at 450C in air for 5 hours (i.e. following the same heat treatment procedure used for the

coatings) and then the crystal structure was investigated by XRD, Figure 79. The peaks are

Page 136: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

119

identified as -alumina [94] and they are not very sharp, which indicates a low degree of

crystallinity and a fine particle size distribution.

Figure 78- Transformation sequence of the different aluminum hydroxides with temperature (adapted from

[93]).

Figure 79- XRD results for the as-received boehmite powder and after heat treatment at 450C for 5 hr

It should be pointed out here that the -alumina has a spinel structure, but the

chemical formula for -alumina containing hydrogen (H) has been the subject of debate [95].

10 20 30 40 50 60 70 80 90

2

as-received

after treatment(111)

(220)

(311)(222)

(400)

(511)

(440)

(444)

Page 137: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

120

A recent theoretical study by Sohlberg et al. has shown that the -alumina can exist over a

range of H content and the chemical formula for -alumina can be presented as H3mAl2-mO3

[95]. This theoretical chemical formula for -alumina has been confirmed by the available

experimental data for the -alumina structure [95]. It should be pointed out that there are

different types of hydrogen atoms positions in the structure of the -alumina particles [95].

H atoms present in the bulk of the -alumina particles occupy octahedral and tetrahedral

sites. Additionally hydrogen atoms can be present on the surface of the -alumina particles,

without specific preference towards lattice positions [95]. These hydrogen atoms have

different mobilities and thermal stabilities in the structure of -alumina, and therefore they

play different roles in the properties of -alumina [95]. According to the elemental

compositions by EDX shown in Table 7, the Al/O ratio (0.75) in the coating is lower than the

theoretical ratio Al/O in Al2O3 (1.12). The lower value of Al/O ratio could be because of the

OH bands in the structure of the -alumina.

Table 7- The EDX analysis for the -alumina coating (high concentration of Au is due to the gold coating on the

sample for SEM analysis)

Element O Na Al Cl Ti Au

Concentration (wt%) 43.59 1.76 32.86 0.96 1.83 21.27

5.4.2 Hydrogen resistivity of alumina-coated PZT

To assess the hydrogen resistivity the coated PZT, water electrolysis technique was

used. Thin (10 nm) Au-Pd electrodes were sputtered over the alumina-coated PZT plates, as

well as on as-received PZT, for reference tests. Figure 80 shows the cross section of the

reference sample, without alumina coating, after water electrolysis at room temperature and

Page 138: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

121

the following test parameters: I= 100 mA/cm2, V= 6-10 V, time= 24 hours, 0.1 M NaOH

solution; a corroded layer has formed close to the top surface of the sample (the metallic

electrode cannot be seen in this figure because it is only about 10 nm thick). It can be also

seen that the grain boundaries of PZT close to the electrode are extensively corroded, which

is not the case in the center of the section far from the electrode.

Figure 80- The cross section of the sample with Au-Pd electrodes and after 24 hours water electrolysis. The

thickness of the corroded layer is about 100 microns

Figure 81 shows the cross section of the sputtered sample with alumina coating after

48 hours of water electrolysis (and all other parameters same as these used to produce the

sample in Figure 77). The effect is quite dramatic - the corrosion beneath the electrode is

limited to < 5 m surface film, which suggests that the porous alumina layer deposited by the

sol-gel technique can act as an effective hydrogen barrier layer.

Au-Pd electrode

PZT

Page 139: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

122

Figure 81- The cross section of the sample with Au-Pd electrodes and alumina coating and after 48 hours water

electrolysis

Because the damage to PZT is due to the diffusion of hydrogen atoms (H) into PZT [9,

14, 18-21], we can conclude that the alumina coating has blocked or decreased the diffusion

of hydrogen atoms from the metallic electrode to the surface of PZT. However, as it was

observed in Figure 77 (and confirmed through [91]), the coating is 40 vol% porous, and

thus molecular hydrogen could easily pass through it. Therefore, the question which may

arise is how this coating decreases the damage to PZT, despite allowing H2 access to PZT

surface. Several hypotheses could be formulated in this regard. First consider the possibilities

of the reaction of hydrogen atoms with the coating (refer to Figure 82 for the schematic

illustration of this possibility). Joubert et al. have investigated the reaction of - alumina

dehydrated at 500C with hydrogen and concluded that hydrogen can react and be absorbed

Au-Pd electrode

PZT

coating

Page 140: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

123

on the surface of the - alumina particles at temperatures as low as 25C [96], due to the

defective surface structure of -alumina [96]. On the other hand, Yu et al. have investigated

the surface diffusion of hydrogen atoms on the surface of - alumina and they observed

noticeable surface diffusion at temperatures higher than 250C [97]. Therefore, one

mechanism by which the coating could decrease the amount of hydrogen damage could be

reaction of the -alumina coating with the hydrogen atoms. In other words, - alumina

coating could act like a “sponge” absorbing hydrogen atoms before they reach the surface of

the PZT (Figure 82). It should be noted that the hydrogen atoms absorbed on the surface of

- alumina particles can diffuse along their surfaces and it is anticipated that sooner or later

damage should start to PZT when hydrogen atoms reach the surface of PZT. If this scenario

is true, then we expect that after a prolonged time of water electrolysis, damage to PZT

should occur.

Figure 82- Schematic for the reaction of hydrogen atoms with -alumina particles 1) transformation of

hydrogen atoms to hydrogen molecules which leave the system away from the coating (i.e. as hydrogen bubbles

during water electrolysis), 2) diffusion of hydrogen atoms through the electrode and attachment to -alumina

particles, followed by surface and bulk diffusion through -alumina towards PZT

Au-Pd electrode

PZT

coating

H

H2

HH

H2

HH

H

H

H

H1

2

ɣ-Al2O3

H

electrode

coating

Page 141: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

124

To examine this theory, we made a thinner alumina coating (1 m thick, produced

through single-dip-coating process, Figure 83), and extended the time of water electrolysis to

144 hours (6 days), at conditions same as before (I= 100 mA/cm2, V= 6-10 V, room

temperature, 0.1 M NaOH solution).

Figure 83- An image of the cross section of PZT with alumina coating on top

The microstructure of PZT after water electrolysis for these conditions (Figure 84)

shows that the PZT is somewhat damaged in regions close to the electrode (to the depth of

about 10-20 m). Therefore, it may be concluded that the reaction of the alumina coating

with the hydrogen is a possible mechanism by which the presence of the coating decreases

and delays the hydrogen damage. However, another explanation to be considered is the

transformation of hydrogen atoms to hydrogen molecules after leaving the metallic

electrode, schematically shown in Figure 85. As palladium is well-known for its hydrogen

catalytic activity, hydrogen atoms (H) could easily transform into hydrogen molecules (H2)

on the surface of the electrode at the interface between the electrode and coating (Figure 85).

Page 142: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

125

Figure 84- The cross section of the sample with Au-Pd electrodes and thin alumina coating and after 144 hours

water electrolysis

Figure 85- Schematic image for the combination of hydrogen atoms at the interface of metallic electrode with

- alumina

Au-Pd electrode

PZT

coating

Au-Pd electrode

PZT

coating

H H

H H

H2

HH HH

H2

HH

H2

1

2

ɣ-Al2O3

electrode

coating

Page 143: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

126

Such hydrogen molecules released on the interface of the alumina coating with the

electrode can further diffuse through the pores of the coating and reach the surface of PZT.

Therefore, it may seem that damage could occur to PZT as alumina coating cannot block the

hydrogen molecules. However, as proposed in [9], and confirmed in this work (Chapter 5-1-

1), hydrogen molecule itself cannot damage the PZT below 400C [9]. That is because

hydrogen molecules cannot dissociate on the surface of PZT at low temperatures [9].

It might be concluded therefore that the transformation of hydrogen atoms to

hydrogen molecule on Pd surface, and physical separation of the PZT surface from the

electrode surface (by the porous alumina) is likely another mechanism by which the alumina

coating (despite its porosity) decreases the amount of hydrogen damage of PZT.

An additional point which should be mentioned here is that the metallic electrode

could be porous. If this is the case, then water could diffuse inside the alumina coating

during water electrolysis and thus fraction of the hydrogen molecules released on the

electrode surface dissolve in the water (the solubility of hydrogen gas in the water at room

temperature is 0.8 mole/liter [99]). Therefore hydrogen molecules would be in contact with

the surface of PZT in water, which is also the case in the samples without alumina coating.

However, because the damage was demonstrated to be considerably lower in the coated PZT

samples, we conclude that the degradation during the water electrolysis is due to the

hydrogen atoms that diffuse from the metallic electrode into PZT, and not to the presence of

the molecular hydrogen in water.

Page 144: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

127

The conclusion that can be drawn from the above experiment is that as far as there is

no interaction between atomic hydrogen (H) and PZT, no damage would occur to PZT. If the

direct contact between the metallic electrode and PZT can be diminished or decreased, such

that no hydrogen atoms will be in contact with PZT, then and no damage would occur to

PZT. Through such mechanism, even a porous coating between the electrode and PZT can

noticeably decrease the amount of damage. It is possible that by applying this method in the

manufacturing of PZT actuators, even with other types of porous coatings, the deleterious

effect of hydrogen would be greatly diminished.

It should be however remembered that a thin insulating surface layer might degrade

functionality and performance of electrode/PZT/electrode assembly [100]. Such layer would

act as a capacitor in series with the PZT (Figure 86), and as a result the externally applied

voltage would be distributed across the sample inversely proportionally to the capacitance of

each layer. Therefore, higher voltages would be needed to drive the PZT actuator. To solve

these issues two solutions could be considered: (1) make the coating layer as thin as possible,

and/or (2) make the coating layer from materials with dielectric constant higher than PZT.

In both these cases, the capacitance of the coating layer will be higher than the capacitance

of the PZT layer, and therefore the voltage across the PZT layer would be almost the same as

the applied voltage to the whole assembly.

Page 145: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

128

Figure 86-Equivalent electrical circuit for PZT and PZT with coatings

There is very recent research supporting the idea of introducing a coating between

the electrode and PZT [100]; this work showed that a thin alumina layer between the PZT

and electrode will not affect the functionality of the PZT capacitors. The reason for this was

not clear; however it was suggested that a thin alumina layer between the electrode and PZT

will not act as an insulator and it may act as a resistor [100]. That is because when a voltage is

applied across the sample with alumina coating, a considerable fraction of the applied voltage

would be across the thin alumina layer, whereas the layer would become conductive under

high voltages by either Schottky emission or thermionic field emission [100].

PZT PZT coating

C PZT C PZTC coat C coat

Page 146: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

129

6 Conclusions

High-pressure hydrogen treatment

The microstructural and capacitance changes in the PZT ceramics exposed to high-pressure

(10 MPa) hydrogen atmosphere at 100C were investigated in this part of the work. For bare

PZT, no noticeable damage was observed to the PZT structure after hydrogen treatment for

up to 1200 hours. The grain boundaries of PZT were corroded only in the regions just below

the surface (about 10 µm deep) of the samples. In samples with Ag electrodes, the presence of

metallic electrodes greatly increases the damaging effects of hydrogen on PZT. The structural

degradation observed in the samples with Ag electrodes consisted of the development of a

very porous layer adjacent to the electrodes on the surface, and the detachment of the

electrodes from PZT. It was proposed that the hydrogen spillover is the responsible

mechanism for the formation of the porous electrode on the surface of the samples. In PZT

samples with Ag/Pd electrodes, the PZT damage noticeably increased compared to Ag-only

electrode. It is therefore suggested to decrease the amount of Pd in the Ag/Pd electrodes to

increase the resistance of the actuators to hydrogen damage. No considerable changes were

observed in the dielectric constant of PZT after 200 hours hydrogen treatment at 100°C.

High-temperature hydrogen treatment

The kinetics of the PZT structural modifications due to hydrogen exposure was

investigated by online monitoring of the electrical properties of PZT above Curie

Page 147: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

130

temperature, up to 650C. Considerable changes were observed in the microstructure of the

PZT samples hydrogen-treated at high temperatures, including the detachment of single PZT

grains from the surface, as well as reduction to metallic lead on the surface of the samples. It

was found that the changes in PZT exposed to high-temperature H2 can be described by a

simple nucleation and growth model. Assuming that the changes are controlled by protons

diffusion, the resulting activation energy for the diffusion of protons in cubic PZT was

determined to be 0.440.09 eV.

The dielectric spectroscopy study of PZT samples shows that even after the high

temperature hydrogen treatment, Ti-O and Zr-O dipoles are still present inside the PZT. The

results show two relaxation peaks in the dissipation factor curve of the hydrogen-treated

PZT. While the first peak indicates that the kinetics obeys the classical Arrhenius law, with

the activation energy of 0.664 eV, the second peak indicates the presence of unusual

relaxation kinetics: the relaxation time increases with increasing temperature. This non-

monotonic relaxation kinetics can be attributed to the defects that hydrogen has produced

inside the PZT, or it can be due to the Maxwell-Wagner polarization mechanism.

Water electrolysis treatment

The interaction of hydrogen with PZT in the tetragonal phase was investigated using

the water electrolysis technique. Development of a hydrogen-affected (corroded) layer

adjacent to the electrode functioning as the cathode was observed during water electrolysis.

Page 148: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

131

The thickness of the corroded layer was used to calculate the diffusion coefficient of

hydrogen atoms in PZT, and the value obtained was 9×10-11 cm2/sec. A composite model was

proposed for the microstructure of PZT affected by hydrogen generated during water

electrolysis, and changes of the electrical properties of PZT are linked to the model. The

Maxwell-Wagner polarization mechanism was proposed to be responsible for the changes in

the dielectric properties of PZT after hydrogen charging. Although this polarization

mechanism has been ignored in previous works by other researchers, we believe it is

responsible for the variation in electrical properties of other oxide ceramics during water

electrolysis as well. Furthermore, the results indicate that after aging, the resistivity and

high-frequency dielectric constant of PZT decrease. The decrease in capacitance is expected

to be due to [OH]– dipoles hindering the movement of PZT domains.

Alumina coatings for PZT protection from hydrogen

In this part of the research we have investigated the possibility of protecting PZT

from hydrogen damage by coating it with alumina using the sol gel technique; subsequently

we have assessed the hydrogen resistance of the coated PZT. The results show that the

quality of the alumina coatings obtained with pure boehmite sol was not very good, i.e.

cracks and coating detachment were observed. However, the addition of the poly-vinyl

alcohol (PVA) to the boehmite sol considerably enhanced the quality of the final coating;

neither cracks nor detachment were observed. The hydrogen resistance of the alumina

coated PZT was investigated using the water electrolysis technique, and the results have

Page 149: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

132

shown that the alumina coatings noticeably decrease the level of hydrogen damage to PZT.

It is anticipated that the main contributor to the decreased PZT damage is the physical

separation of the metallic electrode from PZT by the coatings. The insulation of the PZT

from the electrode leads to the re-combination of hydrogen atoms into molecules on the

electrode surface and within the pores of the coating, which effectively prevents access of

the damaging atomic hydrogen to the surface of PZT.

Impacts of the work

The results of the study of long-term high-pressure gaseous hydrogen exposure of

PZT could be beneficial to the design of the modern electronic fuel injectors that use PZT

actuators for valve opening, instead of the conventional solenoid technology. The results

show that the metallic electrode has considerable effects on the level of hydrogen damage,

and it was suggested that Ag/Pd electrodes should be replaced with pure Ag electrodes in

making such actuators. Kinetics of the hydrogen damage to PZT was also investigated in the

present work. The results could be used for the prediction of the degradation caused by the

hydrogen treatment. The results could be used for the re-design of the hydrogen treatment

process of FeRAMs. We also have shown that even a porous separation layer between an

electrode and PZT acts towards decreasing the PZT hydrogen damage. The results of this

part of the research could be further used to better understand the mechanism of PZT

degradation by hydrogen, and to design new methods for decreasing the hydrogen damage to

PZT and other ceramics.

Page 150: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

133

7 Future Work

We suggest that the technique we used in this work for studying the kinetics of

interactions between hydrogen and PZT can be also used for studying the kinetics of the

interaction between hydrogen and other oxides, in particular oxides designated as the

possible fast proton conductors replacing polymeric proton conductors in fuel cells. These

oxides include BaCeO3, BaZrO3, and SrCeO3. The proposed nucleation and growth model for

the structural changes in PZT could be also valid for other oxides, so a general model could

be evaluated for studying the kinetics of interactions between hydrogen and oxides in

hydrogen atmosphere.

The results of this work show that during the exposure of PZT to hydrogen, the

capacitance of PZT capacitors changed, and these changes were due to the diffusion of

protons into PZT. Further work should be done to build upon this observation to develop a

quantitative relationship between the concentration of protons inside PZT and the

capacitance increase. Such quantitative relationship could be generalized for other proton

conductor oxides as well. The quantitative relationship between the capacitance and protons

content inside oxides could be further developed into a method to evaluate the amount of

protons inside the oxides, based on capacitance changes measurements.

Page 151: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

134

Unusually large capacitance values were observed in PZT capacitors in hydrogen

atmosphere at temperatures of 600-650C (unfortunately due to the limitations of the

instruments, the maximum number we were able to read was approximately 100 F).

Further investigation is therefore needed to determine the polarization mechanisms

responsible for this high capacitance observed in the hydrogen-treated PZT capacitors. The

research in this area might lead to the development of “high-temperature” super-capacitors.

The very high capacitance in PZT capacitors could be then used for many practical

applications, such as energy storage.

According to the dielectric spectroscopy results, hydrogen forms dipoles inside PZT.

The nature of such dipoles should be better understood. In this regard, PZT with different

amounts of dopants (Nb) should be prepared, and it should be investigated whether there is

any correlation between the amount of dopants inside PZT and the intensity of the

relaxation peaks. The important point about dipoles formed with Nb is that they may

prohibit the switching of Ti-O or Zr-O dipoles. Therefore, the interactions between the

dipoles that form with Nb and the Ti-O or Zr-O dipoles should be further investigated,

maybe using computer modeling techniques.

We have proposed that a porous layer separating metallic electrode from PZT should

be sufficient to substantially decrease the hydrogen damage of PZT. This theory should be

further tested with various porous coatings other than -alumina, and the mechanisms

Page 152: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

135

responsible for decreased hydrogen damage of PZT should be identified. Depending on the

electrical properties of the insulator layer between PZT and electrode, the functionality and

performance of electrode/PZT/electrode assembly could be degraded. The role of the coating

between the electrode and PZT on the performance of electrode/PZT/electrode assembly

should be evaluated. Insulators with dielectric constant higher or equal to the dielectric

constant of PZT could be explored as the coating materials, to preserve the performance of

electrode/PZT/electrode assembly.

Page 153: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

136

References

[1] R.E. Jones, P.D. Maniar, R. Moazzami, P. Zurcher, J.Z. Witowski, Y.T. Lii, P. Chu, S.J.

Gillespie, Ferroelectric non-volatile memories for low-voltage, low-power applications, Thin

Solid Films, 1995, 270, 584-588

[2] R. Cuppens, P.K. Larsen, G.A.C.M. Spierings, Ferroelectrics for non-volatile memories,

Microelectron. Eng., 1992, 19, 245-252

[3] S. Pitman, K. Alvine, Hydrogen materials compatibility for the H-ICE, Pacific Northwest

National Laboratory, 4/09/2010

[4] S. Kasap, P. Capper, Springer handbook of electronic and photonic materials, 2007,

Springer

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

ferroelectric phase in the Pb(Zr1−xTix)O3 solid solution Appl Phys Lett, 74 (1999), pp. 2059–

2061

[6] W.M. Kriven, Displacive transformations and their applications in structural ceramics, J.

Phys. IV France, 1995, 05, 101-110

[7] D. Damjanovic, Ferroelectric, dielectric and piezoelectric properties of ferroelectric thin

films and ceramics, Rep. Prog. Phys., 1998, 61, 1267-1324

[8] H. Zheng, I.M. Reaney, W.E. Lee, N. Jones, H. Thomas, Effects of strontium substitution

in Nb-doped PZT ceramics, J. Eur. Ceram. Soc., 2001, 21(10-11), 1371-1375

[9] K. Kushida-Abdelghafar, H. Miki, K. Torii, Y. Fujisaki, Electrode-induced degradation of

Pb(ZrxTi1-x)O3 (PZT) polarization hysteresis characteristics in Pt/PZT/Pt ferroelectric thin-

film capacitors, Appl. Phys. Lett., 1996, 69, 3188-3190

[10] R. Ramesh, S. Aggarwal, O. Auciello, Science and technology of ferroelectric films and

heterostructures for non-volatile ferroelectric memories, Mater. Sci. Eng. R, 2001, 32, 191-

236

[11] Y. Wang, W. Y. Chu, L. J. Qiao, Y. J. Su, Hydrogen-induced delayed fracture of PZT

ceramics during dynamic charging under constant load, Mater. Sci. Eng. B, 2003, 98 (1), 1-5

Page 154: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

137

[12] X. Peng, Y. J. Su, K. W. Gao, , L. J. Qiao, W. Y. Chu, Hydrogen fissure in PZT

ferroelectric ceramic, Mater. Letter., 2004, 58 (15), 2073-2075

[13] X. H. Guoz, S. Q. Shiw, L. J. Qiao, Simulation of hydrogen diffusion and initiation of

hydrogen-induced cracking in PZT ferroelectric ceramics using a phase field model, J. Am.

Ceram. Soc., 2007, 90 (9), 2868-2872

[14] H. K. Yuh, E. Yoon, Y. T. Lee, Y. W. Park, S. I. Lee, Effects of Pb/Pt top electrode on

hydrogen-induced degradation in Pb(Zr,Ti)O3, Jpn. J. Appl. Phys., 2002, 41, 42-46

[15] Y. Fujisaki, K. Kushida-Abdelghafar, Y. Shimamoto, H. Miki, The effects of the catalytic

nature of capacitor electrodes on the degradation of ferroelectric Pb(Zr,Ti)O3 thin films

during reductive ambient annealing, J. Appl. Phys., 1997, 82, 341-344

[16] A. R. Krauss, A. Dhote, O. Auciello, J. Im, R. Rarnesh, S. Aggarwal, Studies of hydrogen-

induced degradation processes in PZT and SBT ferroelectric film-based capacitors, Integrated

Ferroelectrics, Proceedings of the 11th International Symposium on Integrated

Ferroelectrics, Colorado Springs, CO, March 7-10, 1999

[17] H. Miki, K. Kushida-Abdelghafar, K. Torii, Y. Fujisaki, Hydrogen-related degradation

and recovery phenomena in Pb(Zr,Ti)O3 capacitors with a platinum electrode, Jpn. J. Appl.

Phys., 1997, 36, 1132-1135

[18] H. Masumoto, A. Kojima, T. Iijima, T. Goto, Preparation of Ag-alloy top-electrode for

ferroelectric Pb(Zr,Ti)O3 films under various atmospheres, Jpn. J. Appl. Phys., 2002, 41,

6882-6885

[19] Y. Shimamoto, K. Kushida-Abdelghafar, H. Miki, Y. Fujisaki, H2 damage of ferroelectric

Pb(Zr,Ti)O3 thin-film capacitors- the role of catalytic and adsorptive activity of the top

electrode, Appl. Phys. Lett., 1997, 70 (23), 3096-3097

[20] K. Kushida-Abdelghafar, M. Hiratani, Y. Fujisaki, Post-annealing effects on

antireduction characteristics of IrO2/Pb(ZrxTi1-x)O3 /Pt ferroelectric capacitors, J. Appl. Phys.,

1999, 85 (2), 1069-1074

[21] J. P. Han, T. P. Ma, Electrode dependence of hydrogen-induced degradation in

ferroelectric Pb(Zr,Ti)O3 and SrBi2Ta2O9 thin films, Appl. Phys. Lett., 1997, 71 (9), 1267-1269

[22] Y. Fukai, The Metal Hydrogen Systems- Basic Bulk Properties, Springer Series in

Materials Science Vol. 21 (Second revision), 2005

Page 155: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

138

[23] J. C. Moreno-Pirajan, Thermodynamics - Interaction Studies - Solids, Liquids and Gases,

InTech, 2011

[24] http://www.rebresearch.com/H2sol2.htm

[25] M. Pasturel, R. J. Wijngaarden, W. Lohstroh, H. Schreuders, M. Slaman, B. Dam, and

R. Griessen, Influence of the chemical potential on the hydrogen sorption kinetics of

Mg2Ni/TM/Pd (TM=transition metal) trilayers, Chem. Mater., 2007, 19, 624-633

[26] Yu. I. Zvezdin, Yu. I. Belyakov, Hydrogen permeability of some transition metals of

group 1 of the periodic system, Fiziko-Khimicheskaya Mekhanika Materialov., 1967, 3 (3),

344-351

[27] S. F. J. Cox, J. S. Lord, S. P. Cottrell, J. M. Gil, H. V. Alberto, A. Keren, D. Prabhakaran,

R. Scheuermann, A. Stoykov, Oxide muonics: I. modelling the electrical activity of hydrogen

in semiconducting oxides, J. Phys. Condens. Matter., 2006, 18, 1061-1078

[28] R. Hempelmann, Hydrogen diffusion mechanism in proton conducting oxides, Physica

B, 1996, 226, 72-77

[29] K. Xiong, J. Robertson, Hydrogen-induced defects and degradation in oxide

ferroelectrics, Appl. Phys. Lett., 2004, 85(13), 2577-2579

[30] C. H. Park, D. J. Chadi, Effect of interstitial hydrogen impurities on ferroelectric

polarization in PbTiO3, Phys. Rev. Let., 2000, 84 (20), 4717-4720

[31] S. Aggarwal, S. R. Perusse, C. W. Tipton, R. Ramesh, H. D. Drew, T. Venkatesan, D. B.

Romero, V. B. Podobedov, A. Weber, Effect of hydrogen on Pb(Zr,Ti)O3-based ferroelectric

capacitors, Appl. Phys. Lett., 1998, 73 (14), 1973-1975

[32] K.D. Kreuer, Aspects of the formation and mobility of protonic charge carriers and the

stability of perovskite-type oxides, Solid State Ionics, 1999, 125, 285-302

[33] W. Chen, L. Li, Y. Wang, Z. Gui, Effects of electrochemical hydrogen charging on lead-

based relaxor ferroelectric multilayer ceramic capacitors, J. Mater. Res., 1998, 13 (5), 1110-

1112

[34] S. Takatani, K. Kushida-Abdelghafar, H. Miki, Effect of H2 annealing on a Pt/PbZrxTi1-

xO3 interface studied by X-ray photoelectron spectroscopy, Jpn. J. Appl. Phys., 1997, 36, 435-

438

Page 156: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

139

[35] Y. Shimakawa, Y. Kubo, Degradation of ferroelectric Pb(Zr,Ti)O3 under reducing

conditions, Appl. Phys. Lett., 2000, 77 (16), 2590-2592

[36] N. Ikarashi, Analytical transmission electron microscopy of hydrogen-induced

degradation in ferroelectric Pb(Zr, Ti)O3 on a Pt electrode, Appl. Phys. Lett., 1998, 73, 1955-

1957

[37] R. Freer, Self-diffusion and impurity diffusion in oxides, J. Mater. Res., 1980, 15 (4), 803-

824

[38] F. Kremer, A. Schönhals, Broadband dielectric spectroscopy, Springer, Berlin, Germany,

2003

[39] S. Havriliak, S. Negami, A complex plane representation of dielectric and mechanical

relaxation processes in some polymers, Polymer, 1967, 8, 161-210.

[40] O. Kamishima, Y. Abe, J. Kawamura, T. Hattori, Dielectric relaxation in Yb-doped

SrZrO3, J. Phys., Condens. Matter, 2004, 16, 4971-4981

[41 ] http://rb-

kwin.bosch.com/en/powerconsumptionemissions/dieselsysteme/dieselsystem/passenger-

car/technology/injection_systems/commonrailsystem/crs_pkw_types.html

[42] A. Scharf, Piezo-actuators in fuel injection systems, EPCOS technical articles, 2006,

(www. epcos.com)

[43] K. J. Alvine, V. Shutthanandan, W. D. Bennett, C. C. Bonham, D. Skorski, S. G. Pitman,

M. E. Dahl, C. H. Henager, High-pressure hydrogen materials compatibility of piezoelectric

films, Appl. Phys. Lett., 2010, (97), 221911-221913

[44] Y. Park, J. K. Lee, I. Chung, J. Y. Lee, Effect of oxygen plasma treatment on hydrogen-

damaged Pt/Pb(ZrxTi1-x)O3/Pt ferroelectric thin-film capacitor, Jpn. J. Appl. Phys., 1999, 38,

577-579

[45] T. Sakoda, T. S. Moise, S. R. Summerfelt, L. Colombo, G. Xing, S. R. Gilbert, A. L. S.

Loke, S. Ma, R. Kavari, L. A. Wills, J. Amano, Hydrogen-robust submicron

IrOx/Pb(Zr,Ti)O3/Ir capacitors for embedded ferroelectric memory, Jpn. J. Appl. Phys., 2001,

40, 2911-2916

[46] T. Saito, T. Tsuji, K. Izumi, Y. Hirota, N. Okamoto, K. Kondo, T. Yoshimura, N.

Fujimura, A. Kitajima, A. Oshima, Fabrication of robust PbLa(Zr,Ti)O3 capacitor structures

Page 157: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

140

using insulating oxide encapsulation layers for FeRAM integration, Electron. Lett., 2011, 47

(8), 486-487

[47] T. S. Bjørheim, A. Kuwabara, I. Ahmed, R. Haugsrud, S. Stølen, T. Norby, A combined

conductivity and DFT study of protons in PbZrO3 and alkaline earth zirconate perovskites,

Solid State Ionics, 2010, 181, 130-137

[48] J43 advanced hydrogen injector, technical opportunities and challenges, Westport Innovations INC (a private presentation)

[49] Ian D. Raistrick, Application of impedance spectroscopy to materials science, Ann. Rev.

Mater. Sci., 1986, 16, pages 343-70

[50] R. N. Lyerl, H. W. Pickering, Mechanism and kinetics of electrochemical hydrogen

entry and degradation of metallic systems, Annu. Rev. Mater. Sci., 1990, 20, pages 299-338

[51] T.P. Perng, J.K. Wu, A brief review note on mechanisms of hydrogen entry into metals,

Materials Letters, 2003, 57, 3437– 3438

[52] J.K. Wu, Electrochemical method for studying hydrogen in iron, nickel and palladium,

Int. J. Hydrogen Energy 17 (1992) 917– 921

[53] M.H. Armbruster, The Solubility of hydrogen at low pressure in iron, nickel and certain

steels at 400 to 600°, J. Am. Chem. Soc. 65 (1943) 1043– 1054.

[54] M. J. Danielson, Use of the Devanathan–Stachurski cell to measure hydrogen

permeation in aluminum alloys, Corros. Sci., 2002, 44 (4), 829–840

[55] L. E. Scriven, Physics and applications of dip coating and spin coating, MRS Proceedings

1988, 121, 717-730

[56] L. Landau, B. Levich, Dragging of a liquid by a moving plate, Acta Physicochim., URSS,

1942, 17, 42-54

[57] http://www.sasoltechdata.com/tds/DISPERAL_DISPAL.pdf

[58] K. Wefers, C. Misra, Oxides & hydroxides of aluminum, Technical Paper 19, Alcoa

Laboratories, 1987

[59] M. Kumagai, G. L. Messing, Controlled transformation and sintering of a boehmite sol-

gel by α-alumina seeding, J. Am. Ceram. Soc., 1985, 68 (9), 500-505

Page 158: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

141

[60] G. Ruhi, O.P. Modi, A.S.K. Sinha, I.B. Singh, Effect of sintering temperatures on

corrosion and wear properties of sol-gel alumina coatings on surface pre-treated mild steel,

Corros. Sci., 2008, 50, 639-649

[61] B.D. Cullity, Elements of X-Ray Diffraction, 1956

[62] E. Baumgarten, C. L. Wagner, R. Wagner, H-D exchange and hydrogen spillover in the

Pd/SiO2; system: kinetics and detection of hydrogen atoms, J. Mol. Catal., 1989, 50, 153-165

[63] E. Baumgarten, E. Denecke, Hydrogen spillover in the system Pt/Al2O3: I. fundamental

observations, J. Catal., 1985, 95, 296-299

[64] J. E. Benson, H.W. Kohn, M. Boudart, On the reduction of tungsten trioxide accelerated

by platinum and water, J. Catal., 1966, 5, 307-313

[65] T. B. Flanagan, W. A. Oates, The palladium-hydrogen system, Annu. Rev. Mater. Sci.,

1991, 21, 269-304

[66] B. Jaffe, W. R. Cook, H. Jaffe, Piezoelectric ceramics, Academic Press, London, 1971

[67] A.C. Rastogi, S.R. Darvish, P.K. Bhatnagar, Phase evolution of electron-beam evaporated

Pb(Zr,Ti)O3 thin films, Mater. Chem. Phys., 2002, 73, 135–143

[68] A. Khawam, D. R. Flanagan, Solid-state kinetic models: basics and mathematical

fundamentals, J. Phys. Chem. B, 2006, 110, 17315-17328

[69] C.H. Bamford, C.F.H. Tipper, Reactions in solid state, comprehensive chemical kinetics,

Vol. 22, Elsevier, Amsterdam, 1980

[70] K.D. Kreuer, St. Adams, W. Munch, A. Fuchs, U. Klock, J. Maier, Proton conducting

alkaline earth zirconates and titanates for high drain electrochemical applications, Solid State

Ionics, 2001, 145, 295-306

[71] W. Munch, K.D. Kreuer, G. Seifert, J. Maier, Proton diffusion in perovskites: comparison

between BaCeO3, BaZrO3, SrTiO3, and CaTiO3 using quantum molecular dynamics, Solid

State Ionics, 2000, 136-137, 183-189

[72] C.J.D. Von Grotthuss, "Sur la décomposition de l'eau et des corps qu'elle tient en

dissolution à l'aide de l'électricité galvanique", Ann. Chim., 1806, 58, 54-73

Page 159: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

142

[73] I. Animitsa, A. Nieman, S. Titova, N. Kochetova, E. Isaeva, A. Sharafutdinov, N.

Timofeeva, Ph. Colomban, Phase relations during water incorporation in the oxygen and

proton conductor Sr6Ta2O11, Solid State Ionics, 2003, 156, 95-102

[74] T. Schober, J. Friedrich, F. Krug, Phase transition in the oxygen and proton conductor

Ba2In2O5 in humid atmospheres below 300C, Solid State Ionics, 1997, 99, 9-13

[75] W. Fischer, G. Reck, T. Schober, Structural transformation of the oxygen and proton

conductor Ba2In2O5 in humid air: an in-situ X-ray powder diffraction study, Solid State

Ionics, 1999, 116, 211-215

[76] T. Hashimoto, Y. Inagaki, A. Kishi, M. Dokiya, Absorption and secession of H2O and

CO2 on Ba2In2O5 and their effects on crystal structure, Solid State Ionics, 2000, 128, 227-231

[77] P. Lunkenheimer, S. Krohns, S. Riegg, S.G. Ebbinghaus, A. Reller, A. Loidl, Colossal

dielectric constants in transition-metal oxides, Eur. Phys. J. Spec. Top., 2010, 180, 61-89

[78] L. Frunza, H. Kosslick, S. Frunza, R. Fricke, A. Schönhals, Molecular dynamics of 4-n-

octyl-4′-cyanobiphenyl in partially filled nanoporous SBA-type molecular sieves, J. Non-

Cryst. Solids, 2006, 90(1-6), 259-270

[79] L. Frunza, H. Kosslick, S. Frunza, R. Fricke, A. Schönhals, Rotational fluctuations of

water inside the nanopores of SBA-type molecular sieves, J. Phys. Chem. B, 2005, 109 (18),

9154-9159

[80] P. B. Ishai, A. J. Agranat, Y. Feldman, Confinement kinetics in a KTN:Cu crystal, Phys.

Rev. B, 2006, 73, 104104-104111

[81] Y. E. Ryabov, A. Puzenko, Y. Feldman, Nonmonotonic relaxation kinetics of confined

systems, Phys. Rev. B, 2004, 69, 014204-014214

[82] J. Schüller, R. Richert, E. W. Fischer, Dielectric relaxation of liquids at the surface of a

porous glass, Phys. Rev. B, 1995, 52, 15232-15238

[83] H. Y. Huang, W. Y. Chu, Y. J. Su, K. W. Gao, J. X. Li, L. J. Qiao, Hydrogen-induced

semiconductor transformation of lead zirconate titanate ferroelectric ceramics, J. Am. Ceram.

Soc., 2007, 90(7), 2062-2066

[84] K. J. Alvine, M. Vijayakumar, M. E. Bowden, A. L. Schemer-Kohrn, and S. G. Pitman,

Hydrogen diffusion in lead zirconate titanate and barium titanate, J. Appl. Phys., 2012, 112,

043511-043517

Page 160: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

143

[85] Y. Pei, W. Chen, Great diversity in stability of perovskite-type oxides studied by

electrochemical hydrogen charging, Wuhan Univ. J. Nat. Sci., 2011, 16, 255-259

[86] W. P. Chen, Y. Wang, H. L. W. Chan, Hydrogen: a metastable donor in TiO2 single

crystals, Appl. Phys. Lett., 2008, 92, 112907-112910

[87] W. P. Chen, X. F. Zhu, Z. J. Shen, J. Q. Sun, J. Shi, X. G. Qiu, Y. Wang, H. L. W. Chan,

Effects of electrochemical hydrogen charging on electrical properties of WO3 ceramics, J.

Mater. Sci., 2007, 42, 2524-2527

[88] W. Chen, Y. Wang, H. Chan, H. S. Luo, Hydrogen-related dynamic dielectric behavior

of barium titanate single crystals, Appl. Phys. Lett., 2006, 88, 202906-202909

[89] W. Chen, J. Qi, Y. Wang, X. Jiang, H. Chan, Water-induced degradation in

0.91Pb(Zn1/3Nb2/3)O3-0.09PbTiO3 single crystals, J. Appl. Phys., 2004, 95, 5920-5922

[90] A.L. Ahmad, N.F. Idrus, M.R. Othman, Preparation of perovskite alumina ceramic

membrane using sol–gel method, J. Membrane Sci., 2005, 262, 129-137

[91] A. F. M. Leenaars, K. Keizer, A.J. Burggraaf, The preparation and characterization of

alumina membranes with ultra-fine pores, J. Mater. Sci., 1984, 19, 1077-1088

[92] D. S. Maciver, H. H. Tobin, R. T. Barth, Catalytic aluminas I. surface chemistry of eta

and gamma alumina, J. Catal., 1963, 2, 485-497

[93] K. Wefers, C. Misra, Oxides and hydroxides of aluminum, Alcoa Technical Paper No. 19,

Revised; Alcoa Laboratories: Alcoa Center, PA, 1987.

[94] Y. Liu, D. Ma, X. Han, X. Bao, W. Frandsen, D. Wang, D. Su, Hydrothermal synthesis of

microscale boehmite and gamma nanoleaves alumina, Mater. Lett., 2008, 62, 1297-1301

[95] K. Sohlberg, S. J. Pennycook, S. T. Pantelides, Hydrogen and the structure of the

transition aluminas, J. Am. Chem. Soc., 1999, 121, 7493-7499

[96] J. Joubert, A. Salameh, V. Krakoviack, F. Delbecq, P. Sautet, C. Coperet, J. M. Basset,

Heterolytic splitting of H2 and CH4 on -alumina as a structural probe for defect sites, J. Phys.

Chem. B, 2006, 110, 23944-23950

Page 161: Hydrogen Induced Damage of Lead Zirconate Titanate (PZT)

144

[97] C. Y. Yu, B. K. Sea, D. W. Lee, S. J. Park, K. Y. Lee, K. H. Lee, Effect of nickel deposition

on hydrogen permeation behavior of mesoporous -alumina composite membranes, J.

Colloid. Interf. Sci., 2008, 319, 470-476

[98] K. J. Sladek, E. R. Gilliland, R. F. Baddour, Diffusion on surfaces (II): correlation of

diffusivities of physically and chemically adsorbed species, Ind. Eng. Chem. Fundamen.,

1974, 13 (2), 100-105

[99] H. A. Pray, C. E. Schweickert, B. H. Minnich, Solubility of hydrogen, oxygen, nitrogen,

and helium in water, Ind. Eng. Chem. Res., 1952, 44, 1146-1151

[100] A.Q. Jiang , H. J. Lee , G. H. Kim , C. S. Hwang , The inlaid Al2O3 tunnel switch for

ultrathin ferroelectric films, Adv. Mater., 2009, 21, 2870-2875