chapter one - etda

249
The Pennsylvania State University The Graduate School Department of Materials Science and Engineering SYNTHESIS AND COLLOIDAL PROPERTIES OF ANISOTROPIC HYDROTHERMAL BARIUM TITANATE A Thesis in Materials Science and Engineering by Timothy James Yosenick © 2005 Timothy James Yosenick Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy December 2005

Upload: khangminh22

Post on 23-Apr-2023

1 views

Category:

Documents


0 download

TRANSCRIPT

The Pennsylvania State University

The Graduate School

Department of Materials Science and Engineering

SYNTHESIS AND COLLOIDAL PROPERTIES OF ANISOTROPIC

HYDROTHERMAL BARIUM TITANATE

A Thesis in Materials Science and Engineering

by

Timothy James Yosenick

© 2005 Timothy James Yosenick

Submitted in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

December 2005

This thesis of Timothy James Yosenick was reviewed and approved* by the following:

James H. Adair Professor of Materials Science and Engineering Thesis Co-Advisor Co-Chair of Committee Clive A. Randall Professor of Materials Science and Engineering Thesis Co-Advisor Co-Chair of Committee Susan Trolier-McKinstry Professor of Ceramic Science and Engineering Thomas Shrout Professor of Materials Darrell Velegol Associate Professor of Chemical Engineering James Runt Professor of Materials Science and Engineering Associate Head for Graduate Studies

*Signatures are on file in the Graduate School

iii

ABSTRACT Nanoparticles of high dielectric constant materials, especially BaTiO3, are

required to achieve decreased layer thickness in multilayer ceramic capacitors (MLCCs).

Tabular metal nanoparticles can produce thin metal layers with low surface roughness via

electrophoretic deposition (EPD). To achieve similar results with dielectric layers

requires the synthesis and dispersion of tabular BaTiO3 nanoparticles. The goal of this

study was to investigate the deposition of thin BaTiO3 layers using a colloidal process.

The synthesis, interfacial chemistry and colloidal properties of hydrothermal BaTiO3, a

model particle system, was investigated. After characterization of the material system

particulates were deposited to form thin layers using EPD.

In the current study, the synthesis of BaTiO3 has been investigated using a

hydrothermal route. TEM and AFM analyses show that the synthesized particles are

single crystal with a majority of the particle having a <111> zone axis and {111} large

face. The particles have a median thickness of 5.8 ± 3.1 nm and face diameter of 27.1 ±

12.3 nm. Particle growth was likely controlled by the formation of {111} twins and the

synthesis pH which stabilizes the {111} face during growth. With limited growth in the

<111> direction, the particles developed a plate-like morphology. Physical property

characterization shows the powder was suitable for further processing with high purity,

low hydrothermal defect concentration, and controlled stoichiometry. TEM observations

of thermally treated powders indicate that the particles begin to loose the plate-like

morphology by 900 °C.

iv

The aqueous surface chemistry of BaTiO3 is complex and difficult to model using

current models due to the pH dependent dissolution/readsorption of Ba2+ at the surface.

In addition the precipitation of BaCO3 at high pH influences the surface chemistry. In the

current study a model was developed to account for the effect of dissolved Ba2+ as a

function of pH. Three distinct regions in the surface chemistry are observed as a function

of pH. At low pH, the dissolution of Ba2+ results in a TiO2 surface which can be

described using the MUSIC model. As pH increases the affect of dissolved Ba2+

becomes more prominent. The adsorption of Ba2+ onto the TiO2 is observed and can be

modeled using a modified Stern isotherm. In basic environments (>pH 9.5) the

precipitation of BaCO3 on the surface of the BaTiO3 particles requires the use of a

Nernst-Gouy-Stern charging model to described the surface.

The aqueous passivation, dispersion, and doping of nanoscale BaTiO3 powders

was investigated. Passivation BaTiO3 was achieved through the addition of oxalic acid.

The oxalic acid selectively adsorbs onto the particle surface and forms a chemically

stable 2-3 nm layer of barium oxalate. The negative surface charge of the oxalate

effectively passivated the BaTiO3 providing a surface suitable for the use of a cationic

dispersant, polyethylenimine (PEI). Rheological properties indicate the presence of an

oxalate-PEI interaction which can be detrimental to dispersion. With a better

understanding of the aqueous surface chemistry of BaTiO3 the surface chemistry was

manipulated to control the adsorption of aqueous soluble complexes of Co, Nb, and Bi,

three common dopants in the processing of BaTiO3. Surface charge, TEM, and EDS

analysis showed that while in suspension the dopants selectively absorbed onto the

particle surface forming an engineered coating.

v

The electrophoretic deposition of two different BaTiO3 nanoparticle suspensions

was investigated. The effect of solution chemistry on dispersion, deposition kinetics, and

film microstructure is addressed. The conditions necessary for optimum dispersion

results in low deposition rates and poor film adhesion. High dispersant concentration

leads to electrochemical inhibition at the electrode and reduced field drop in the bulk of

solution. Low effective fields in the bulk of the suspension results in low electrophoretic

velocities and reduced deposition kinetics. Strong repulsive interactions between the

particles and electrode lead to poor adhesion for the particles that do deposit. The

addition of an indifferent electrolyte reduces the repulsion and improves adhesion.

However, the indifferent electrolyte reduces the zeta potential of the particles in

suspension, leading to aggregation prior to deposition. Deposited films comprised of

aggregates exhibit inhomogeneous microstructures.

vi

TABLE OF CONTENTS

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xvii

ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .xix

Chapter One: Research Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

Chapter Two: A Literature Review of the Synthesis, Dispersion, Doping

and Electrical Properties of Barium Titanate Materials . . . . . . . . . . . . . . . . . . . 4

2.1: Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.2: Synthesis of Nanoscale BaTiO3 Powder . . . . . . . . . . . . . . . . . . . . . . . . 6

2.3: Surface Chemistry and Dispersion of BaTiO3 . . . . . . . . . . . . . . . . . . . 16

2.4: Doping and Microstructure of Sintered BaTiO3 … . . . . . . . . . . . . . . . . 21

2.4.1: Core-shell structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.4.2: Doping for Base Metal Electrodes . . . . . . . . . . . . . . . . . . . . 25

2.4.3: Doping Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

2.5: Size Effects and Electrical Properties of Nanoscale BaTiO3 Materials . . . 31

2.5.1: BaTiO3 Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

2.5.2: Bulk BaTiO3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

2.5.3: BaTiO3 Thin Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

2.5.4: Electrode-Dielectric Interactions . . . . . . . . . . . . . . . . . . . . . 37

2.6: Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

vii

Chapter Three: Synthesis of Nanotabular Barium Titanate via a Hydrothermal

Route . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

3.1: Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

3.2: Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.3: Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

3.3.1: BaTiO3 Particle Morphology and Growth . . . . . . . . . . . . . . . 58

3.3.2: Characterization of Physical Properties . . . . . . . . . . . . . . . . . 69

3.3.3: Morphological Evolution as a Function of Temperature . . . . . . 72

3.4: Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

Chapter Four: Aqueous Surface Chemistry of Hydrothermally Derived BaTiO3

Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

4.1: Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

4.2: Experimental Observations of BaTiO3 Surface Charging in an Aqueous

Environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

4.2.1: Acidic pH – Amorphous TiO2 Surface . . . . . . . . . . . . . . . . . 84

4.2.2: Neutral pH – Ba2+ Adsorption . . . . . . . . . . . . . . . . . . . . . . . 86

4.2.3: Basic pH – BaCO3 Formation . . . . . . . . . . . . . . . . . . . . . . . 87

4.3: Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

4.4: Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

4.4.1: Low pH – Amorphous TiO2 Surface . . . . . . . . . . . . . . . . . . . 90

4.4.1.1: Determination of BaTiO3 Surface Groups . . . . . . . . . 90

viii

4.4.1.2: Combination of MUSIC model and Gouy-Chapman

Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

4.4.2: Neutral pH – Ba2+ Adsorption . . . . . . . . . . . . . . . . . . . . . . . 99

4.4.3: High pH – BaCO3 Formation . . . . . . . . . . . . . . . . . . . . . . .101

4.4.4: Comparison of Experimental and Theoretical Calculations . . . .102

4.5: Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .113

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .116

Chapter Five: Passivation, Dispersion, and Aqueous Solution Doping of

Platelet BaTiO3 Powder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .120

5.1: Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .120

5.2: Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .125

5.3: Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .128

5.3.1: Passivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .128

5.3.2: Dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .134

5.3.3: Doping of Nanotabular BaTiO3 . . . . . . . . . . . . . . . . . . . . . .143

5.4: Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .152

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .154

Chapter Six: Electrophoretic Deposition of Hydrothermally Derived Barium

Titanate Tabular Nanoparticles with a Cationic Dispersant . . . . . . . . . . . . . . . .159

6.1: Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .159

6.2: Theoretical Background – Mechanisms of EPD . . . . . . . . . . . . . . . . .161

6.3: Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .163

6.4: Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .167

ix

6.4.1: Dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .167

6.4.2: EPD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .173

6.4.2.1: Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .173

6.4.2.2: Adhesion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .186

6.4.3: Film Microstructure . . . . . . . . . . . . . . . . . . . . . . . . . . . . .193

6.5: Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .199

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .200

Chapter Seven: Conclusions and Suggest Work

7.1: Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .205

7.2: Suggested Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .207

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .210

Appendix A: Algorithm for the Determination of Surface Potential Using the

MUSIC Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .211

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .214

Appendix B: Dispersion of Solution Based Doped BaTiO3 Platelets for

Electrophoretic Deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .215

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .222

Appendix C: Stabil Calculation for Heterogeneous Coagulation . . . . . . . . . . . .223

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .229

x

LIST OF FIGURES 2.1 Schematic of a typical MLCC showing the three materials systems used in the

fabrication of a MLCC: (1) the dielectric, (2) internal electrode, and (3) termination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.2 Ideal solubility diagram for BaTiO3 in an aqueous environment with CO2 showing that BaTiO3 is not the thermodynamically stable form of barium in

water, from Bendale et al . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.3 Plot of the dielectric constant of BaTiO3 versus temperature for single crystal

BaTiO3 taken from Merz. Three distinct peaks in the dielectric constant are observed. The three peaks coincide with the three phase transitions in BaTiO3: rhombohedral to orthorhombic (-90 °C), orthorhombic to tetragonal (0 °C),

and tetragonal to cubic (130 °C) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.4 The superposition of individual transitions results in a broad diffuse transition,

which is more stable with change in temperature . . . . . . . . . . . . . . . . . . . . 26 2.5 Plot showing the dependence of the dielectric constant of BaTiO3 with grain size in both bulk ceramics (●) and thin films (Δ) . . . . . . . . . . . . . . . . . . . . 32 3.1 Flow diagram for the hydrothermal synthesis of platelet BaTiO3. The starting

solution is 500 mL total of a 1M solution, and has an approximate yield of 120g of powder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56 3.2 X-ray diffraction pattern for as-synthesized powder and powder heat treated to

1000 °C. The as-synthesized powder is pseudo-cubic due to the presence of hydrothermal defects in the lattice. After heat treatment at 1000 °C peak

splitting is observable in the (200)/(002) peak (see insert graph) and indicates the material has converted to the tetragonal form of BaTiO3. Note: * Cubic

BaTiO3 peaks JCPDS Card: 31-0174 and + Tetragonal BaTiO3 peaks JCPDS Card: 79-2264 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

3.3 TEM micrograph (a) and associated selected area electron diffraction pattern (b). The single crystal diffraction pattern shows that the particles are single

crystals with <111> zone axis. The absence of diffraction from (111) in the partial ring diffraction pattern (c) from a cluster of particle shows a majority

of the particles in the cluster show texture . . . . . . . . . . . . . . . . . . . . . . . . . 60 3.4 AFM cross-sectional image of the BaTiO3 particle on an atomically flat cleaved mica substrate. The particles have a plate-like morphology with a thickness of 7.9 nm and face diameter of 46.9 nm . . . . . . . . . . . . . . . 62

xi

3.5 Thickness and face diameter size distributions for the hydrothermal BaTiO3 platelets. The distributions were calculated using the AFM offline software

and image analysis software. Both of the distributions are based on a total of 214 particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 3.6 Schematic of BaTiO3 platelets formed via multiple {111} twin formation. After Schmelz and Thomann . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 3.7 Schematic representations of (a) (100) plane, (b) (110) plane and (c) (111) plane rendered using Atoms for Windows. Each figure shows the oxygen coordination of titanium in each plane. The geometry of each plane was used to calculate the Ti planar density and surface OH density . . . . . . . . 66 3.8 Weight loss curve for platelet, commercial powder A, and commercial powder B. The weight loss from 300 to 500 °C is due to the removal of hydroxyl defects, whereas the weight loss at higher temperatures is the removal of BaCO3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 3.9 TGA and dTGA platelet powder shows the presence of four reactions occurring at 350, 425, 660, and 760 °C . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 3.10 Series of TEM images showing the morphological evolution of the platelet particles as a function of temperature: (a) 25 °C – as-synthesized, (b) 375°C, (c) 450 °C, (d) 700 °C, (e) 800 °C, (f) 900 °C, (g) 1000 °C, and (h) 1100 °C. Neck formation is observable at 800 °C with morphological changes

occurring by 900 °C. At 375 °C hydrothermal defects have begun to coalesce and are not removed until 1000 °C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 4.1 Ideal solubility diagram for BaTiO3-H2O-CO2 system from Bendale et al. Ba2+ dissolution is favored at low pH. As pH increases, Ba2+ solubility decreases until the precipitation of BaCO3 is favored. The TEM image of a

BaTiO3 particle treated in water at pH 6.5 show the presence of an amorphous TiO2 surface layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 4.2 α values for four different surface groups on BaTiO3 as a function of suspension pH. α represents the degree of protonation of all of the specific

surface sites present on the surface. For example, below pH 13 all of the …O-1 groups have an associated proton, but above pH 14 all of the groups are deprotonated. Over the entire pH range one of the four reactions is controlling the surface charging of BaTiO3 . . . . . . . . . . . . . . . . . . . . . . . . 97 4.3 Dissolution data for aqueous BaTiO3 suspensions with increasing solid loadings. As expected Ba2+ dissolution is minimized at high pH and increases with increasing amount of surface area present in suspension. Data from Chodelka . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .104

4.4 Dissolution data for BaTiO3 as function of solution pH normalized for surface area present in solution. A good linear fit of observed and the empirical equation was used to determine the concentration of dissolved Ba2+ in the modeling of surface chemistry of BaTiO3. r2 = 0.937 . . . . . . . . .105 4.5 Zeta potential of aqueous BaTiO3 suspension (40 m2/L) showing the three

different regions of surface charge in BaTiO3. Region I controlled by a native TiO2 surface. The increase in Region II is due to the adsorption of Ba2+

(aq) onto the TiO2 surface. The decrease in Region III is due to precipitation of

BaCO3 on the BaTiO3 surface. r2 = 0.884 . . . . . . . . . . . . . . . . . . . . . . . . .106 4.6 Schematic showing the evolution of the surface charging mechanism as a function of solution pH for BaTiO3 in an aqueous environment. At low pH Ba2+ dissolution leads to a TiO2 surface. As the pH increases Ba2+

(aq) adsorption results in a deviation from an ideal TiO2 surface. In a basic

environment the precipitation of BaCO3 on the surface controls the surface charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .107 4.7 TEM images of the commercial BaTiO3: (a) image showing that the particles are equiaxed, and (b) high resolution image showing lattice fringes a selected

particle with a <011> zone axis and that the surface of the particle is -

terminated by the (100) (0.40 nm) and (011) (0.28 nm) planes . . . . . . . . . . .111 4.8 Plot of zeta potential versus pH for 1wt% suspensions of platelet and equiaxed particles in a 95/5 ethanol/water solvent mixture. Suspensions were

prepared in the solvent mixture to limit Ba2+ dissolution. However, a small amount of dissolved Ba2+ is present at pH greater than pH 8 necessitating the inclusion of Ba2+ adsorption at high pH to account for low negative zeta

potential values and pH greater than pH 10 . . . . . . . . . . . . . . . . . . . . . . . .112 5.1 TEM image of an oxalate passivated BaTiO3 particle. Treatment with oxalic acid results in a 2 nm thickness surface layer of barium oxalate which inhibits the surface from degradation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .129 5.2 Zeta potential of 1wt% suspension of nanoplatelet BaTiO3 in water with

increasing amounts of oxalic acid as a function of pH. Full surface passivation is achieved by an oxalic acid concentration of 3x10-3 M (3.75w/w). A further increase in the oxalic acid concentration results in an increase in the

magnitude of the zeta potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .130 5.3 Plot of soluble Ba2+

(aq) concentration as a function of suspension pH and oxalic acid concentration for the platelet BaTiO3. An oxalic acid concentration of 5x10-2 M yields the best surface passivation yet lower

concentrations are acceptable as long as the solution pH remains greater

xii

than pH 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .132

xiii

5.4 At acidic pH values Ti forms a soluble complex with oxalic acid. At high oxalic acid concentration Ti dissolution from the BaTO3 surface is unacceptable. However, at an oxalic acid concentration of 3x10-3 M at pH values greater than pH 5 the Ti4+ concentration in solution is negligible. (--) at 10-7 M indicates limit of detection for Ti4+with ICP-ES. Lines are trend lines only . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .133 5.5 The addition of PEI to oxalic acid passivated BaTiO3 suspensions results in a

positive zeta potential due to the adsorption of PEI on the barium oxalate surface. In addition to large positive zeta potential values PEI adds a steric

hindrance to aids in particle dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . .136 5.6 The shear stress and viscosity of 10vol% suspensions with 2.5w/w PEI shows that increasing the oxalic acid concentration results in a deviation from

Newtonian behavior due to the interaction of PEI and oxalic acid to form a gel network of amine oxalate. The linear regions of (a) were fit with Bingham’s law and the y-intercept was reported as τB, the yield point. All

suspensions measured were at pH 7 to maintain an approximate zeta potential of approximately +25mV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .137 5.7 Apparent viscosity (a) and yield point (b) values as a function of oxalate concentration for 10vol% suspensions with varying concentration of PEI. The suspensions containing excess oxalic acid exhibit increased viscosity and yield point due to the formation of amine oxalate gel network. All suspensions were measured at pH 7 to maintain a constant zeta potential of +25 mV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .141 5.8 Zeta potential of doped platelet particle as a function of dopant concentration. As Co is added to the suspension a reaction with excess oxalic acid occurs to form a cobalt oxalate surface. When PEI is added, the surface becomes positive and is suitable for the adsorption of the negatively charge Nb and Bi

complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .144 5.9 Schematic representation of the doping processing. First oxalic acid and cobalt are added forming an oxalate surface layer which passivates the surface. PEI is added to disperse the particles and provide a positive surface

charge for the adsorption of the Nb and Bi which are added in the final step . .146 5.10 TEM images showing the morphological evolution of the particle surface at each step of the doping process: (a) as synthesized particle, (b) oxalic acid

passivated particle, the insert shows a 2-3 nm surface layer of barium oxalate, (c) Co doped particle, (d) Co, Nb doped particle, and (e) fully doped particle

which shows the addition of Bi results in 1-2 nm deposits on the surface of the particle. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .147

xiv

5.11 EDS spectrum of a cluster of doped platelet particles. The spectrum shows the presence of the three dopants Co, Nb, and Bi. The C, Cu, and Fe present are due to contamination from either the TEM sample grid or TEM instrument. The insert detail is added to show the presence of Nb and Bi because the signal to noise ratio at lower energies is too small to indicate the

presence of the dopants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .148 5.12 XRF data for doped platelet powder samples showing that the actual concentrations of the CoO, Nb2O5, and Bi2O3 deviate from the prepared concentrations expect Sample 1, which is only doped with 5wt% Bi2O3 . . . . .151 6.1 Zeta potential of equiaxed BaTiO3 powder shows that the zeta potential decreases as the HCl concentration increases due to increased ionic strength in solution. Although high zeta potential values are observed, suspensions

prepared by electrostatic dispersion were not stable and therefore not suitable for deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .170 6.2 Zeta potential of equiaxed BaTiO3 as a function of PEI concentration for differing concentrations of oxalic acid with increasingly negative surface charge. The negative surface charge is suitable for the adsorption of a cationic dispersant, PEI. As the PEI concentration increases the sign of the

surface charge reverses. A PEI concentration of 2w/w results in zeta potential values of approximately 80 mV. All suspensions prepared had an HCl concentration of 10-3 M . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .172

6.3 Particle size distribution for the HOx/PEI dispersed platelet particles shows that a PEI concentration of only 0.25w/w results in the best dispersion with a

median particle size, D(50), of 16.6 nm. Low PEI concentrations do not provide enough surface charge for good dispersion while high PEI concentrations can result in bridging flocculation which degrades dispersion. All suspensions prepared had 2w/w HOx and 10-3 M HCl . . . . . . . . . . . . . .174 6.4 (a) Ideal equivalent circuit for the EPD cell. Rexp and Ccell are the experimental setup resistance and capacitance of the EPD cell, respectively. Both are dependent on the experimental setup and remain constant. Rsol is the solution resistance, Cdl is the capacitance of the electrode double layer, and Rtran is the electron transfer resistance of electrochemical reactions. (b)

Schematic representing the ideal Cole-Cole plot for the equivalent circuit . . . .176 6.5 Cyclic voltamagramm for three ethanol solutions containing HCl and PEI. The addition of 1mM HCl shows the evidence of two electrochemical reactions that occur at the cathode both of which have a profound effect on the pH of the solution near the cathode. The presence of PEI inhibits the

electrochemical reactions by adsorbing onto the electrode and increasing the electron transfer resistance at the cathode . . . . . . . . . . . . . . . . . . . . . . . . .178

xv

6.6 Cole-Cole plot for a 10-3 M HCl solution in ethanol. The center of the second

semi-circle is depressed below the x-axis indicating the Cdl is not an ideal capacitor but a constant phase element, which is due the roughness of the electrode surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .179

6.7 Schematic showing the effect of excess PEI on the electrochemistry of deposition. When excess PEI is present it contaminates the electrode and inhibits electrochemical reactions and particle deposition. At areas of the electrode unaffected by PEI electrochemical reactions occurs and particle deposition occurs. If the PEI concentration is too large the entire electrode can be contaminated and deposition is completely inhibited . . . . . . . . . . . . .183 6.8 Deposition current and rate as a function of PEI concentration used for dispersion. As the PEI concentration increases the current and deposition rate

decreases due to the presence of unabsorbed PEI on solution. During deposition the excess PEI absorbs onto the cathode and inhibits electrochemical reactions decreasing the current . . . . . . . . . . . . . . . . . . . .185 6.9 Schematic showing the process of charge neutralization. Water is reduced at the cathode and the pH increases due to production of hydroxyl groups. The

increased pH results in the PEI losing charge and desorbing from the BaTiO3 particle surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .188

6.10 Zeta potential of platelet BaTiO3 and Pt electrode in pure ethanol as a function of LiCl concentration. The addition of an indifferent electrolyte lowers the zeta potential and therefore the repulsive interactions between the

depositing particles and the electrode. Without the addition of the LiCl the large repulsive interactions between the particles and electrode lead to a lack of adhesion of the particles on the electrode . . . . . . . . . . . . . . . . . . . . . . .190 6.11 Interaction energy curves for the platelet BaTiO3 and Pt electrode. The curves

were calculated using Stabil and the physical constants in Table 6.3. The addition of LiCl, as expected, lowers the repulsive interaction between the particles and electrodes. However, the addition of ≥ 1mM LiCl results in a small repulsive interaction which is not suitable for good dispersion . . . . . . .191 6.12 The conductivity and deposition rate of suspensions with LiCl added are highly dependent on the concentration of LiCl. Although adding LiCl improves film adhesion it results in decreased deposition kinetics . . . . . . . . .194 6.13 (a) TEM image of a film cross-section and (b) AFM deflection image of top

surface of an EPD film of equiaxed BaTiO3 particles. The film has a thickness of 613 nm with a surface roughness of 106 nm . . . . . . . . . . . . . . .195

xvi

6.14 AFM deflection image of EPD film deposited from a platelet BaTiO3 suspension. Electrode surface coverage is incomplete and the film appears to be comprised of aggregates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .197 6.15 XRD analysis of deposited platelet powders showing that crystallographic texture did not develop during deposition. The lines represent the peak positions and relative intensity for tetragonal BaTiO3 powder (JCPDS 79-2264). Increasing the voltage is expected to: (1) increase the layer thickness, and (2) provide a higher driving force for the flat laydown on

platelet particles, but no improvement in texture is observed as the deposition voltage increases. The presence of particle aggregates as seen in AFM images

the reason for the lack of texture. Note: * The diminishing peak at 26º 2θ is due the underlying Pt/Mylar substrate used as the electrode . . . . . . . . . . . . .198 A.1 Simple flow diagram showing the steps necessary to calculate the surface potential as a function of solution pH using the MUSIC model . . . . . . . . . . .212 B.1 The zeta potential of doped and undoped BaTiO3 suspensions as a function of PEI concentration. The PEI provides surface charge as well as adsorption sites for the ionic dopants. When the dopants adsorb, surface sites are neutralized and the surface charge decreases lowering the zeta potential.

Decreasing the pH increases the zeta potential, but it is not sufficient to create stable dispersions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .217 B.2 Interaction energy curves for doped and undoped BaTiO3 showing that a repulsive energy barrier does not exist for the doped suspensions. This is due to the reduction of the zeta potential and increase in the ionic strength as the

dopants are added to the suspension. The interaction energy curves were generated using Stabil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .219

B.3 Particle size distribution for doped and undoped BaTiO3 suspensions at pH 7. As expected from the zeta potential and interaction results the doped suspension is highly aggregated . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .220 C.1 Interaction energy curves for platelet BaTiO3 and Pt electrode. The curves were calculated using Stabil and the physical constants in Table C.1 . . . . . . .228

xvii

LIST OF TABLES

2.1 List of common techniques and their characteristics used in the synthesis of

nanoscale BaTiO3. Taken from Adair and Suvaci . . . . . . . . . . . . . . . . . . . 7 2.2 Electronic Industry Alliance (EIA) of United States codes for allowable

capacitance change and temperature ranges for capacitors . . . . . . . . . . . . . . 23 2.3 List of common dopants added to BaTiO3 in the processing of MLCCs (compiled from Jaffe et al., Tsur et al., Hennings et al., and Lee et al.) . . . . . . 24 3.1 Planar density of Ti and surface hydroxide for low index planes in cubic BaTiO3 (a = 4 Å) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 3.2 Physical properties of platelet, commercial powder A, and commercial powder B hydrothermally derived BaTiO3 powders . . . . . . . . . . . . . . . . . . 70 4.1 Possible surface groups of Ba2+ depleted BaTiO3 with the associated protonation reactions and calculated log K values. The reactions in bold are the only reactions that occur in the normal pH range (1-14) and those used in the calculation of the surface charge . . . . . . . . . . . . . . . . . . . . . 94 4.2 ICP-ES results for the dissolution of Ba2+ in 95/5 ethanol/water mixtures. The data shows that dissolution is limited until pH 8 when a maximum concentration of 10-3 M is observed. Because of the dissolution it is necessary to account for the adsorption of Ba2+

(aq) in model the surface charge in the solvent mixture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .109 4.3 Site densities of Ti for the three low index planes of BaTiO3 based on the

structure of BaTiO3 and the values used in the modeling of the surface in a ethanol/water solvent mixture. The difference between the actual and model values is due the model not accounting for the potential drop in the IHP. The normalized values show that the model is in good agreement with the actual surface with respect to the relative density for each plane . . . . . . . . . . . . . . .114

5.1 Rheological properties for 10vol% BaTiO3 suspensions prepared with varying

amounts of oxalic acid and PEI. The addition of excess oxalic acid results in increased viscosity and yield points due to the formation of a gel network of amine oxalate that increases interparticle interactions. All suspensions were measured at pH 7 in order to maintain an approximate zeta potential of +25

mV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .140

xviii

5.2 XRF data for doped platelet powder samples. The data shows that as the CoO and Nb2O5 concentration increases the actual concentration deviates more from the prepared concentration due to competitive adsorption of the dopants on the surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .149 6.1 Concentration of Ba2+ in pure ethanol as determined by direct couple plasma

emission spectroscopy. Results show the dissolution of Ba2+ in pure ethanol is not prevalent and therefore surface passivation is not necessary prior to

dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .168 6.2 List of EPD cell variables with increasing PEI concentration. The solution

resistance decreases as PEI increases because the proton mobility is increased. The transfer resistance increases in the presence of small concentration of PEI.

In a solution of 0.01wt% PEI 98.6% of the voltage drop occurs at the electrode -solution interface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .181 6.3 List of physical constants used in the calculation of the interaction between

BaTiO3 and Pt in pure ethanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .192

C.1 List of physical constants used in the calculation of the interaction between

BaTiO3 and Pt in pure ethanol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .224

xix

ACKNOWLEDGEMENTS I must first thank my thesis advisor Dr. James Adair for his guidance and patience

over the past five and half years while at Penn State. I would also like to acknowledge

Dr. Clive Randall who graciously served as my co-advisor and provided an additional

point of view in my scientific education. I thank the three other members of my thesis

committee: Dr. Tom Shrout, Dr. Susan Trolier-McKinstry, and Dr. Darrell Velegol for

their guidance and helpful suggestions.

All members of the Adair Research Group through the years have been extremely

helpful and made coming to work everyday worth it. I must specifically thank Dr.

Jennifer Nelson, Dr. Rajneesh Kumar, and Ms. Ying Yuan for their help running the

TEM. Any TEM images that appear in this thesis are due to their help. It would not have

been possible to finish this thesis without the help of the staff of the MRL in performing

many task and experiments. I would like to thank TDK Corp. for the opportunity to visit

Japan for four and half months, and allowing me to work in the materials research center

in Narita.

My friends, both at Penn State and home, have been supportive and provided the

necessary diversions through the years. I hope their support continues, and I will

continue to rely on them for temporary diversions from time to time. The support of my

family is indescribable and immense as it has always been, and I know it always will be.

CHAPTER ONE

Research Objectives The objective of this research is to understand the role of particle morphology in

the processing of hydrothermal BaTiO3 for capacitive applications. To accomplish this

goal, tabular nanoparticles were synthesized via a hydrothermal route and processed in

parallel with commercially available nominally 90 nm equiaxed hydrothermal BaTiO3

powder. Further objectives of this thesis included an investigation of nanoparticle

synthesis, modeling of the aqueous surface chemistry of BaTiO3, characterization of the

aqueous stability of the synthesized nanoparticles, aqueous solution based doping of

nanoparticles, and a brief investigation of the electrophoretic deposition (EPD) of thin

BaTiO3 films.

In the literature review (Chapter Two) an understanding of the problems and

challenges in achieving the outlined goals were developed. The thesis chapters are

arranged in such a way as to investigate the processing of the material system using a

bottom-up approach, beginning with the synthesis of nanotabular BaTiO3 ceramics. The

following is a summary of the research objective of each experimental chapter:

The synthesis of nanotabular BaTiO3 via a hydrothermal route is addressed in

Chapter Three. Emphasis is place on the synthesis of nanotabular BaTiO3 particles with

a high yield using a hydrothermal method. An examination and discussion of the

development of particle morphology during synthesis based on solution pH and crystal

chemistry is presented. The physical properties of nanotabular powder are compared to

1

those of commercially available hydrothermal BaTiO3 powders. Finally, the

morphological evolution of the platelet particles upon heating is discussed.

A model for the complex aqueous surface chemistry of BaTiO3 was developed in

Chapter Four. The model is based on current surface charging models, but includes

aspects which account for the dissolution/adsorption of Ba2+ from the particle surface and

the precipitation of BaCO3 on the particle surface at high pH. The model illustrates the

complex nature of the BaTiO3 surface in an aqueous environment at different solution pH

values. This underlines the importance of addressing the stability of BaTiO3 during

aqueous processing, which is discussed in Chapter Five.

Chapter Five focuses on the aqueous dispersion and doping of the hydrothermally

synthesized nanoscale BaTiO3 powders. The passivation of the BaTiO3 surface was

investigated using oxalic acid at room temperature as a function of solution pH. A

dispersion scheme based on electrosteric stabilization is presented and characterized

using zeta potential, quasi-elastic light scattering and viscosity measurements. Control of

the surface chemistry was used to dope the BaTiO3 while in suspension. Co, Nb and Bi

were selectively adsorbed to the particle surface to create engineered coatings on the

particle eliminating the need for a solid-state doping technique.

The electrophoretic deposition of BaTiO3 nanoparticles is presented in Chapter

Six. A preliminary understanding of the effect of solution chemistry (i.e. dispersant

concentration, conductivity, ionic strength, etc…) on the kinetics of deposition was

investigated. Electrochemical reactions that occur during electrophoretic deposition were

characterized and the effect of solution chemistry on the inhibition of electrochemical

reactions is discussed. Finally the physical properties of deposited films using AFM,

2

TEM, and XRD were characterized, and the effect of solution chemistry on the

microstructure is discussed.

3

CHAPTER TWO

A Literature Review of the Synthesis, Dispersion, Doping and Electrical Properties

of Barium Titanate Materials

2.1 Introduction

The synthesis and processing of bulk BaTiO3 powders is of interest because of the

great significance that BaTiO3 has in the electronics industry. For over 60 years BaTiO3

has been the primary material in the development of multilayer ceramic capacitors

(MLCC).1 MLCCs currently play an important role in cellular and computer electronics.

Not only are demands placed on the physical properties of the material, but also the limits

to which current processing techniques can produce thinner layers.

Figure 2.1 is a schematic of a typical MLCC. There are three material systems

that make up an MLCC: (1) the dielectric, (2) internal electrodes, and (3) end

termination. Each material system plays an important role in the processing, physical

properties, and end use of the MLCC. The current work focuses on the synthesis and

processing of BaTiO3, the dielectric material. An important aspect of the processing of

BaTiO3 is the interaction of the BaTiO3 with the internal electrode, which in recent years

has switched from Ag/Pd alloys to Ni. The use of a base-metal electrode system

introduces unique processing issues that will be discussed later.

The volumetric capacitance of a MLCC is dictated by the thickness of the active

dielectric layer in the capacitor.1 The ability to produce thinner layers makes it possible

to replace a capacitor with a new capacitor of equivalent capacitance with a smaller

4

Figu

re 2

.1.

Sche

mat

ic o

f a ty

pica

l MLC

C sh

owin

g th

e th

ree

mat

eria

l sys

tem

suse

d in

the

fabr

icat

ion

of a

MLC

C:

(1) t

he

diel

ectri

c, (2

) int

erna

l ele

ctro

de, a

nd (3

) ter

min

atio

n.

Die

lect

ric(B

aTiO

3)In

tern

al E

lect

rode

(Ag/

Pd, N

i)

Term

inat

ion

5

footprint. Thinner layers also lead to a savings in weight and cost, and ultimately the

replacement of other capacitor technologies such as polymer films and low capacitance

electrolytics. The current standard for ceramic layer thickness by tape casting is 1 μm

and is partially limited by particle size. Nanoparticles are an important material

component required to achieve thinner layers.

2.2 Synthesis of Nanoscale BaTiO3 Powder

Later in this work the hydrothermal synthesis of anisotropic nanoscale BaTiO3

particles is presented. There are several different synthesis routes for nanoscale BaTiO3

in the literature ranging from solid-state to wet chemical methods.2-17 To understand the

advantages and disadvantages of each route, a review of the common methods is

presented. Recent reviews by Adair and Suvaci17 and Pithan et al.18 provide good

overviews of the different synthesis routes. One of the main disadvantages of several of

the routes is the necessary high-temperature calcination step which leads to the formation

of hard agglomerates. Table 2.1 is a list of the most common synthesis routes used to

produce nanoscale BaTiO3 powders and their characteristics.

Solid-state synthesis is one of the traditional methods for the synthesis of BaTiO3

powder.2, 19-22 The solid-state route is well studied and relatively inexpensive which

makes it an ideal choice for commercialization. In solid-state synthesis BaCO3 and TiO2

are calcined at high temperature. The high temperature calcination is necessary for the

decomposition of BaCO3 and the diffusion of Ba into the lattice of the TiO2 particles.

Hennings synthesized 400 nm BaTiO3 using a conventional solid-state method where

BaCO3 reacted with TiO2 at temperatures ranging from 780 to 900 °C.2 To achieve the

6

Met

hod

Par

ticle

Siz

eIm

purit

ies

Adva

ntag

esD

isad

vant

ages

Mix

ed O

xide

400n

m to

100

’s μ

mLa

rge

quan

titie

s of

im

purit

ies

due

to s

tarti

ng

mat

eria

ls a

nd m

illing

m

etho

d.

•Eas

y pr

oces

s to

per

form

on

a la

rge

scal

e.•R

elat

ivel

y ch

eap

star

ting

mat

eria

ls

•Hig

h im

purit

y le

vels

•Hig

h ca

lcin

atio

n te

mpe

ratu

res

•Lar

ge a

mou

nts

of a

ggre

gatio

n le

adin

g to

la

rge

parti

cle

size

s.•M

illing

usu

ally

requ

ired.

•Poo

r sto

ichi

omet

ric c

ontro

l fro

m p

artic

le

to p

artic

le.

Cop

reci

pita

tion

10 n

m –

10’s

μm

Chl

orid

e an

d ot

her

impu

ritie

s pr

esen

t fro

m

star

ting

mat

eria

ls.

Con

tam

inat

ion

if m

illing

is

requ

ired.

•Low

impu

rity

leve

ls.

•Low

reac

tion

tem

pera

ture

s.•S

toic

hiom

etric

mix

ing

appr

oach

es

atom

ic le

vel.

•Usu

ally

requ

ires

a m

illing

trea

tmen

t to

obta

in d

esire

d pa

rticl

e si

ze.

•Mor

e tim

e co

nsum

ing

than

mix

ed o

xide

m

etho

d.•T

edio

us w

ashi

ng re

quire

d to

rem

ove

chlo

ride

ions

.

Sol

-Gel

5-10

0 nm

Min

imal

con

tam

inan

ts

from

org

anic

pre

curs

ors.

Sm

all a

mou

nts

of S

i co

ntam

inat

ion

from

gl

assw

are.

•Ver

y lo

w im

purit

y le

vels

.•S

toic

hiom

etric

on

the

atom

ic le

vel.

•Low

pro

cess

ing

tem

pera

ture

s (2

0-65

0°C

).

•Rel

ativ

ely

expe

nsiv

e st

artin

g m

ater

ials

.•L

ow te

mpe

ratu

re m

etho

ds a

re g

ener

ally

tim

e co

nsum

ing

with

low

pro

duct

yie

ld.

Vap

or P

hase

20 n

m -

mic

ron

leve

lS

mal

l lev

els

of

cont

amin

atio

n fro

m

star

ting

mat

eria

ls.

•Low

pro

cess

ing

tem

pera

ture

s (1

00-

~800

°C).

•Eas

y to

pro

duce

nan

osiz

ed p

artic

les.

•Som

e pr

ecur

sor m

ater

ials

are

cos

tly.

•Col

lect

ion

with

out a

ggre

gatio

n is

diff

icul

t.•S

toic

hiom

etry

con

trol

can

be d

iffic

ult.

Hyd

roth

erm

al3

nm –

mic

ron

leve

lS

mal

l lev

els

of

cont

amin

atio

n fro

m

star

ting

mat

eria

ls a

nd

reac

tion

vess

el.

Hyd

roth

erm

al (O

H)

defe

cts

due

to a

queo

us

synt

hesi

s

•Low

pro

cess

ing

tem

pera

ture

s (6

0-50

0°C

).•P

artic

les

are

form

ed in

sol

utio

n gi

ving

pot

entia

l con

trol o

ver

aggl

omer

atio

n.•H

igh

purit

y an

d at

omic

sca

le

stoi

chio

met

ry.

•Par

ticle

mor

phol

ogy

easi

ly

cont

rolle

d.

•Som

e pr

ecur

sor m

ater

ials

are

cos

tly.

•Rec

over

y fro

m s

uspe

nsio

ns w

ithou

t ag

glom

erat

ion.

•Red

ispe

rsio

n of

agg

lom

erat

es.

Tab

le 2

.1.

List

of c

omm

on te

chni

ques

and

thei

r cha

ract

eris

tics u

sed

in th

e sy

nthe

sis o

f nan

osca

le B

aTiO

3. Ta

ken

from

Ada

ir an

d Su

vaci

17

7

small particle size, Hennings used a TiO2 powder with a starting size of 200 nm. With

solid-state synthesis the high calcination temperature leads to agglomeration and requires

milling the powder to realize the primary particle size. A second disadvantage of the

solid-state technique is that the size and morphology of the synthesized particles are

limited by the size of the starting TiO2 particles.

Synthesis routes based on the coprecipitation of complex metal salts remain one

of the most widely used commercial routes for the synthesis of BaTiO3. The Clabaugh

process is a wet chemical technique where solutions of BaCl2, TiCl4 and oxalic acid are

mixed and particles of barium titanyl oxalate (BTO) precipitate.23 After synthesis the

particles are calcined to drive off the oxalate and form BaTiO3. An advantage of BTO is

that there is little change in particle size during the conversion from BTO to BaTiO3 if the

calcination step is properly controlled. In addition, because the Clabaugh process is

primarily a solution-based synthesis route, good mixing and near atomic scale

homogeneity are possible. However, there are two critical issues associated with the

Clabaugh process: (1) Oswald ripening during synthesis and (2) agglomeration and

crystallite growth during calcination. To overcome the two issues different researchers

have used different approaches to correct the problems at each step in the process.

A modified Clabaugh process has been studied by Kimel et al.5 and Szepesi.5, 24

In the modified process, a small-volume high-shear mixing chamber is used to create

turbulent fluid flow which permits particle nucleation while limiting particle growth.

After precipitation the particles are directly injected into a quenching solution which

coats the particle surface to inhibit Oswald ripening. This method has produced BTO

particles as small as 10 nm.

8

Yamaura et al.25 and Park et al.26 used alcohol based oxalic acid solution in the

synthesis of BTO. The BTO exhibits a lower solubility in alcoholic solution compared to

aqueous solution and therefore growth by Oswald ripening is limited. No notable growth

was observed for BTO particles prepared in alcoholic environments. However, since the

solubility of Ba and Ti in alcohol solution is incongruent it was difficult to precipitate

homogenous BTO powders.

To understand particle evolution during calcination it is important to understand

the decomposition reactions of BTO. Under isothermal conditions during calcination

BTO is believed to decompose by the following reactions27, 28,

)(3342222242 2.)(2)(4 gCOCOOCOTiBaOCBaTiO + [2.1]

)()(23233342222 33.)(2 gg COCOBaTiOTiOBaCOCOOCOTiBa ++++ [2.2]

but the following reaction is also possible,

)(2)(23)(2242 221)( ggg COCOTiOBaCOOOCBaTiO ++++ [2.3]

Independent of the decomposition reaction, BaCO3 and TiO2 must react to form BaTiO3,

)(2323 gCOBaTiOTiOBaCO ++ [2.4]

The reaction in Equation 2.4 is believed to lead to agglomeration and crystallite growth

during calcination. Wada et al. developed a 2-step calcination of BTO with one step

being performed under vacuum.29, 30 Powders with particle size ranging from 17 to 100

nm were reported. Another group of researchers found that precise control of the heating

rate is necessary to control the final particle size of the BaTiO3.27, 28, 31 Using an

intermediate heating rate yields the proper control of nucleation rate with limited growth.

At low heating rates the nucleation rate is low and the duration of the reaction in

9

Equation 2.4 is long enough for substantial growth to occur, whereas at high heating rates

completion of the reaction does not occur until higher temperatures, which promotes

growth.32 Under optimum conditions BaTiO3 powder with a particle size ranging from

20-40 nm can be synthesized

Other synthesis methods based on the thermal decomposition of double metal

salts have been presented, but the most common is the Pechini method, or the citrate

method.33, 34 The method is similar to the Clabaugh process except that citric acid is used

instead of oxalic acid to form a complex double metal salt. The decomposition reactions

involved in the Pechini method are more complex than that of the Clabaugh process.

Since both the Clabaugh process and Pechini method are based on carboxylic acids, the

formation of BaCO3 during thermal decomposition is unavoidable.18

Several research groups have used a sol-gel method for the preparation of

nanoscale BaTiO3 powders.6-8, 10, 12, 35, 36 Most sol-gel routes begin with the formation of

non-aqueous sols using high purity Ba and Ti reagents, commonly organo-metallics. The

sols are then converted to gels with the addition of water which forms a hydrolyzed gel

structure. After gelation, the gels are dried and calcined at high temperatures to remove

the chemically bound water and crystallize the amorphous gel. The calcination

temperature is lower than that of solid-state routes, and therefore agglomerates formed

are weaker and easier to reduce during milling. The main disadvantage of sol-gel routes

is that the processes are costly with low yields. To further reduce the particle size and

tailor the particle size distribution, Hempelmann and co-workers performed sol-gel

synthesis in a microemulsion system.37, 38 By using such a system the nucleation and

growth of the particles was confined to the aqueous phase of a water-in-oil

10

microemulsion, which limit the particle size. Using different surfactant systems and

varying the experimental conditions, narrow particle size distributions with mean sizes

ranging from 3 to 16 nm were synthesized.

Recently, a limited research effort has focused on vapor phase synthesis routes for

nanocrystalline BaTiO3.39-41 The synthesis uses vapor phase Ba and Ti sources such as

liquid precursors that are either boiled or have inert gas bubble through them, then the

vapors are then mixed at elevated temperatures and quenched. Because of the high

quenching rates growth of the particle after nucleation is severely limited. It is also

possible to use electron beam evaporation or sputtering of solid precursors to generate the

vapor.39 Particle sizes less then 20 nm have been reported. One of the major issues

during synthesis is controlling the mixing of the vapors and the chemical stoichiometry

(i.e. Ba:Ti ratio) of the particles. The formation of BaCO3 is also a problem if the

atmosphere is not properly controlled.40

Direct wet-chemical synthesis routes based on precipitation have been presented

in the literature. Whether the technique is called low temperature aqueous synthesis

(LTAS), low temperature direct synthesis (LTDS), solvent refluxing, or hydrothermal

synthesis, the basic synthesis steps are similar. Aqueous solutions of Ba and Ti sources

are injected into a high pH solution, and then aged as needed. The Ba and Ti sources,

solution pH, and temperature vary from technique to technique leading to powders with a

variety of physical properties. Work by Nanni and co-workers42-44, Wada et al.10, 45, and

Wang et al.13 focused on routes to directly precipitate BaTiO3 in an aqueous environment

at or near room temperature under ambient pressure. Because the solutions contain large

amounts of Na and Cl it is necessary to thoroughly wash the particles after synthesis.

11

The high solution pH during synthesis also leads to the incorporation of large amounts of

hydroxide defects into the lattice, and since the reaction is open to the ambient

atmosphere the presence of BaCO3 is difficult to eliminate. By adjusting the synthesis

variables (i.e. solution concentration, temperature, etc.) particle size can be varied from

20-900 nm.

Hydrothermal synthesis of BaTiO3 has been the most widely studied of the wet-

chemical routes.11, 14-16, 46-52 Under the optimized synthesis conditions, powders with low

defect concentrations and controlled stoichiometry that require no further processing can

be synthesized, making hydrothermal synthesis an excellent choice for the commercial

synthesis of BaTiO3.17 In hydrothermal synthesis, aqueous solutions of barium and

titanium sources are mixed and sealed in a high temperature-pressure reaction vessel and

heated. Osseo-Asare et al.53 and Lencka and Riman54 studied the thermodynamics of the

hydrothermal formation of BaTiO3 and found that a basic environment is necessary for

BaTiO3 to precipitate, and that pH was dependent on the Ba concentration in the starting

solution.

The hydrothermal synthesis of BaTiO3 is extensively commercialized and

protected by a variety of patents.55-60 The methods invented by Abe et al.55 and Menashi

et al.58 are two of the primary methods used for the commercial synthesis of

hydrothermal BaTiO3. The synthesis steps in each method are similar with differences

arising in the post-synthesis treatments. Abe et al. uses hydroxides of both Ba and Ti as

the source material, which are mixed in an aqueous solution and heat treated. After

synthesis the powder is washed with an acetic acid solution to remove BaCO3. However,

the acid wash leads to Ba dissolution from the particle and a Ba deficient surface.

12

Stoichiometry is controlled by a post-washing treatment with an insoluble Ba metal salt

to adjust to the desired Ba:Ti stoichiometry.

Menashi et al. used an amorphous hydrous Ti-gel, Tiy(OH)x, as the Ti source with

Ba(OH)2 as the Ba source. After synthesis, the particles are washed with a 0.005 to 0.02

M Ba(OH)2 solutions. The use of a Ba-rich wash solution limits Ba dissolution and

eliminates the need to adjust the stoichiometry with a second treatment. Regardless of

the method used to the synthesis, the general reaction for the formation of BaTiO3 during

hydrothermal treatment is,

OHBaTiOOHTiOBa ss 2)(3)(22 2 +++ −+ [2.5]

Two rate-limiting mechanisms have been observed for the hydrothermal synthesis

of BaTiO3: (1) phase boundary and diffusion limited49-51, 61 or (2) nucleation and

growth.11, 15, 47, 48, 52, 62 The difference in formation mechanism is generally dependent on

the phase of the TiO2 source. If the TiO2 is crystalline or of large size, then the TiO2

particles have a low solubility and growth occurs by the reaction of Ba2+ at the surface of

the TiO2 followed by diffusion of Ba2+ into the lattice, eventually leading to the

conversion of the TiO2 to BaTiO3. A secondary effect of this growth mechanism is that

size and morphology are limited by the size and morphology of the starting TiO2

particles.63 Hertl studied the kinetics of hydrothermal synthesis using a crystalline TiO2

source. At low Ba concentrations diffusion of Ba into the lattice of the TiO2 is the rate

limiting step. In contrast, at higher Ba concentrations, the reaction of the Ba with the

surface of TiO2 particles is the rate limiting step in BaTiO3 growth.51

When a highly soluble TiO2 source is used, for example, a Ti-organometallic or

sol-gel derived Ti-hydrous-oxide gel, both the Ba and Ti exhibit high solubility at

13

elevated temperatures and synthesis proceeds by nucleation and growth. To fully

investigate hydrothermal growth under such conditions Kershner et al.52 used TEM to

image particles synthesized using a TiCl4-based gel as the TiO2 source. At all stages of

growth homogenous single crystal BaTiO3 particles were observed. If a surface

reaction/diffusion mechanism was responsible for growth, then at the early stages of

growth, inhomogeneous particles with a TiO2 core and a shell of BaTiO3 are expected;

however, this was not observed. This lack of evidence for a surface reaction/diffusion

mechanism was later confirmed with kinetic studies from Moon et al.11, which led to the

conclusion that a nucleation and growth mechanism controls the growth of hydrothermal

BaTiO3 when a high solubility TiO2 source is used.

The low temperature hydrothermal synthesis of BaTiO3 is of interest because of

the savings of time and energy. At low temperature the interface-diffusion growth

mechanism is kinetically limited. However, the mixing of the Ti and Ba is a problem

when using a Ti-gel precursor. For example, when titanium isopropoxide is mixed with

water at high pH, a TiOy(OH)x gel readily forms. The local structure of the gel is

comprised of Ti-O-Ti bonds. It is necessary to break the Ti-O bonds for complete mixing

of the Ti and Ba.64 Moon et al. modified titanium isopropoxide with acetylacetone which

inhibits the hydrolysis of Ti and the formation of TiOy(OH)x network.65, 66 This results in

the Ti precursor having greater water solubility and permits better mixing of the Ti and

Ba. Using a modified Ti precursor, Moon et al. synthesized BaTiO3 at temperatures as

low as 50 °C with particle sizes ranging from 50 to 350 nm.

Although high pH is necessary for synthesis it also leads to the greatest issue with

hydrothermal powders: hydroxide defects. During synthesis hydroxyl groups are

14

incorporated into the lattice of the particles.67 After synthesis, heat treatment of the

powders is needed to remove the hydroxyl groups from the lattice. Hennings showed that

the removal of the hydroxyl groups is compensated by the generation of oxygen

vacancies in the lattice.67 If a large concentration of hydroxide defects is present, during

heat treatment the oxygen vacancies coalesce to form large pores, which degrade

electrical permittivity and physical properties, crystallinity and density, of the bulk

materials.

In the synthesis of BaTiO3 the quality and physical properties of the powder must

meet high standards. Defects, contamination, and incorrect stoichiometry are all

problems which will affect the densification and sintering of bulk materials. Large

intragranular pores, exaggerated grain growth and secondary phase are all possible if the

physical properties of the powder are not well-controlled.68

An advantage of hydrothermal synthesis is the ability to control particle

morphology. A variety of shapes have been reported, including tubes69, hexapods16, and

platelets70 all in the nanoscale and all by hydrothermal synthesis. By limiting growth in a

specific direction an anisotropic morphology is achieved. Crystal chemistry and the

presence of specific adsorbates affect which crystal habit is favored for growth. Bagwell

found the stable crystal habit in hydrothermally-derived BaTiO3 changed from the {111}

plane to the {100},{110}, and {211} planes with the addition of polymeric additives.71

Since the ferroelectric properties of BaTiO3 are strongly dependent on the

crystallographic orientation of the materials, these developments in morphology control

could possibly lead to an enhancement in the electrical properties of bulk samples

prepared from these powders.

15

2.3 Surface Chemistry and Dispersion of BaTiO3

Colloidal forming techniques, mainly tape casting, are the preferable forming

methods for the dielectric layers in most MLCCs. Currently most tape cast slurries are

based on non-aqueous dispersion of BaTiO3 powder with binders and other organic

additives.1 For financial and environment reasons, aqueous tape casting is of growing

interest. However, many problems with aqueous based tape casting still exist. In

general, foaming, cracking during drying, and inadequate binder systems lead to low

quality tapes with poor mechanical problems. If these issues can be overcome, it will be

necessary to better understand the dispersion and interactions of BaTiO3 in aqueous-

based suspensions. The issues of the colloidal instability and incongruent solubility of

BaTiO3 in an aqueous environment makes the dispersion of BaTiO3 nanoparticles

difficult. A complicated surface chemistry and the resultant interactions in an aqueous

environment lead to a myriad of problems. The aqueous surface chemistry will be

discussed and the inherent problems presented in concert with the relevant studies that

have attempted to address the problems.

The instability of BaTiO3 in water is well-documented.72-78 In acidic

environments the dissolution of BaTiO3 is thermodynamically favorable,

OHTiOBaHBaTiO ss 2)(22

)(3 2 +++ ++ [2.6]

and leaves a TiO2-rich surface.72, 77 Figure 2.2 is a plot of the stability of BaTiO3 as a

function of solution pH generated using OPAL™79, 80 with the thermodynamic data from

Venigalla and Adair74 and Bendale et al.81 accounting for the presence of CO2. Ba2+

solubility is minimized under extremely alkaline conditions, but as the pH decreases an

16

Figu

re 2

.2Id

eal s

olub

ility

dia

gram

for B

aTiO

3in

an

aque

ous e

nviro

nmen

t with

CO

2sh

owin

g th

at B

aTiO

3is

not

the

ther

mod

ynam

ical

ly st

able

form

of b

ariu

m in

wat

er, f

rom

Ben

dale

et a

l.81

17

increase in the solubility occurs. Dissolution experiments show that barium dissolution is

almost instantaneous and reaches steady-state rapidly.78 A difference in observed

concentrations and thermodynamically calculated concentration led to the hypothesis that

at extended times barium diffusion from the lattice to the surface becomes the rate

limiting step in dissolution.

The amount of barium in solution is also dependent on a wide variety of factors.

The stoichiometry of the powder, as expected, affects the dissolved barium

concentrations.78 Chiang and Jean performed a series of experiments on samples with

Ba/Ti ratios ranging from 0.992 to 1.004 and found that Ba-rich powders yield higher

dissolved barium concentrations. The amount of surface area exposed to suspension also

affects barium dissolution.75 Both higher surface areas and higher BaTiO3 solid loading

lead to an increase in barium concentration. If the barium concentration become too high

then specific readsorption of the barium occurs on the BaTiO3 surface.77, 82 This leads to

the commonly observed effect of the suspension isoelectric point (IEP) being dependent

on solids loading.75, 78

Not only must the reaction of BaTiO3 with water be considered, but reactions with

dissolved species, for example CO2, must also be addressed. Water absorbs CO2 from

the atmosphere and forms carbonic acid, H2CO3. Dissolved Ba2+ reacts with CO32- to

form BaCO374. BaCO3 has been noted to be a problem during sintering. During sintering

BaCO3 evolves from the sample at temperatures where closed porosity is present. At

elevated temperatures CO2 gas forms in the closed pores and leads to localized de-

sintering and retrograde densification.83

18

The degradation of BaTiO3 not only affects the suspension properties but also the

microstructure of bulk materials processed in aqueous based suspensions. Anderson

found that milling aqueous suspensions at different pH values affected the amount of

exaggerated grain growth, but was unable to determine if the TiO2-rich surface or

readsorption of barium was the cause of exaggerated grain growth.82 Crampo et al.

found in a series of experiments with pellets prepared from leached powder, that the

presence of BaCO3 is necessary for exaggerated grain growth.84 Two powders samples

were leached, one in a CO2-free environment, and another prepared in ambient

atmosphere. Only the pellets exposed to CO2, with BaCO3 present, exhibited

exaggerated grain growth.

To avoid the problem of aggregation and barium dissolution, Kamiya synthesized

BaTiO3 powder in the presence of a surfactant to limit particle aggregation and

dissolution during nucleation and growth.85 However, this synthesis route has low yields

and is cost prohibitive for the commercial synthesis of BaTiO3. Finding other suitable

dispersants and binder systems for the aqueous processing of BaTiO3 has also been of

recent interest.86-91 Kirby et al. used a comb polymer based on poly(acrylic acid) (PAA)

and poly(ethylene oxide) (PEO) to stabilize aqueous suspension of BaTiO3.92, 93 The

carboxylic acid groups of the PAA absorbed at the Ba sites on the BaTiO3 surface while

the PEO extended into the solution to provide a steric repulsion. Kirby et al., like almost

all other researchers, noted the instability of BaTiO3 in water, but took little or no

precautions to limit barium dissolution. In contrast, those who addressed barium

dissolution only performed analysis at ≥ pH 9, where barium dissolution is

thermodynamically minimized.

19

Paik et al. studied the effect of PAA and poly(vinyl alcohol) (PVA) on the

dissolution of Ba2+ from BaTiO3 in the pH range from 3 to 11.94 Below pH 6 where only

a small fraction of the carboxylic acid groups are deprotonated the PAA strongly

absorbed to the surface and provides a small degree of degradation resistance. As pH

increased the PAA acted as a sink and actively increased the concentration of Ba2+ in

solution. The change in behavior with pH was attributed to the affinity of the PAA for

the BaTiO3 surface as a function of the degree of dissociation of the carboxylic acid

group. When a low degree of dissociation is observed, the PAA is sparingly soluble in

solution and prefers to adsorb to the particle surface. PVA showed little to no effect on

the dissolution of Ba2+ from the BaTiO3 particles.

Surface passivation is a possible route for enhancing the stability of BaTiO3 in an

aqueous environment.80, 95 A passivation agent reacts with the surface of the powder to

form a barrier to dissolution. Vasques et al. discussed the required parameters of a

passivation agent for a similar material, YBa2Cu3O7-x superconductor which are, (1) low

solubility in the solvent, and (2) a good diffusion barrier.96 Although low solubility is

necessary, it is not the only concern. For example, BaCO3 has low solubility in water at

pH 9 to pH 11 but does not prevent the aqueous degradation of YBa2Cu3O7-x

superconductors because the CO32-

(aq) does not passivate all components. However,

solubility can be used as an initial criterion to limit the search for a passivation agent.

20

2.4 Doping and Microstructure of Sintered BaTiO3

2.4.1 Core-shell structure

BaTiO3 exhibits three phase transitions: rhombohedral to orthorhombic,

orthorhombic to tetragonal, and tetragonal to cubic at -90, 0 and 130 °C, respectively.97

Figure 2.3 shows the increased dielectric response of BaTiO3 at the phase transition

temperatures. Unfortunately changes in dielectric constant over a narrow temperature

range are undesirable for electronic components where the temperature is not precisely

regulated. With increasing use of mobile and cellular technologies, capacitors are

expected to work in temperature extremes from the dead of winter (-20 °C) to the height

of summer (40 °C). Table 2.2 is a list of capacitor classifications specified by the

Electronic Industries Alliance (EIA) of the United States for allowable capacitance

variability in specified temperature ranges.98 One of the most temperature stable

specifications is designated as X7R and it is primarily used in mobile and cellular devices

that will be used both indoors and outside. X7R capacitors must maintain capacitance

within ±15% of the room temperature value over a temperature range from -55 to 125 °C.

To achieve the temperature stability of an X7R capacitor, the BaTiO3 must be

doped to flatten out the temperature response of the dielectric constant.99-106 A series of

materials called Curie shifters shift the Curie temperature of BaTiO3. Sr2+, Zr4+, Hf4+,

Sn4+, Nb5+, Ta5+, W6+, Ni2+, Co2+, Fe3+, Ag+, Zn2+ all decrease the Curie temperature (Tc)

of BaTiO3.107 Table 2.3 is a list of common dopants used in the processing of BaTiO3

MLCCs.107-110 The dopants are divided into two categories: (1) those used to tailor the

dielectric temperature response of the material such as Curie shifters, and (2) those added

to improve oxidation resistance of the BaTiO3 for use with base metal electrode systems.

21

Figu

re 2

.3.

Plot

of t

he d

iele

ctric

con

stan

t of B

aTiO

3ve

rsus

tem

pera

ture

for s

ingl

e cr

ysta

l BaT

iO3

take

n fr

om M

erz.

97Th

ree

dist

inct

pea

ks in

die

lect

ric c

onst

ant a

re o

bser

ved.

The

thre

e pe

aks c

oinc

ide

with

the

thre

e ph

ase

trans

ition

s in

BaT

iO3:

rhom

bohe

dral

to o

rthor

hom

bic

(-90

°C),

orth

orho

mbi

c to

tetra

gona

l (0

°C),

and

tetra

gona

l to

cubi

c (1

30 °C

).

22

Tab

le 2

.2El

ectro

nic

Indu

stry

Alli

ance

(EIA

) of t

he U

nite

d St

ate

code

s for

allo

wab

le

capa

cita

nce

chan

ge a

nd te

mpe

ratu

re ra

nges

for c

apac

itors

98

* Cap

acita

nce

chan

ges a

re m

easu

red

as th

e pe

rcen

t cha

nge

from

the

room

tem

pera

ture

cap

acita

nce.

EIA

Cod

eTe

mpe

ratu

re R

ange

(ºC

)E

IA C

ode

Cap

acita

nce

Cha

nge*

(%)

X7

-55

to +

125

D±3

.3X

5-5

5 to

+85

E±4

.7Y5

-30

to +

85F

±7.5

Z5+1

0 to

+85

P±1

0R

±15

S±2

2T

-33/

+22

U-5

6/+2

2V

-82/

+22

23

Table 2.3 List of common dopants added to BaTiO3 in the processing of MLCC’s(compiled from Jaffe et al.107, Tsur et al.108, Hennings109, and Lee et al.110)

Dopant Valence Site TypeCa +2 A NeutralSr +2 A NeutralPb +2 A NeutralTb +3/4 B NeutralZr +4 B NeutralSn +4 B NeutralHf +4 B NeutralMg +2 B AcceptorNi +2 B AcceptorMn +2/3 B AcceptorCo +2/3 B AcceptorYb +2/3 B AcceptorFe +2/3 B AcceptorLu +3 B AcceptorSm +2/3 A DonorEu +2/3 A DonorLa +3 A DonorNd +3 A DonorBi +3 A DonorCd +3 A DonorNb +5 B DonorTa +5 B DonorW +6 B DonorEr +2/3 Amphoteric Acceptor/DonorHo +3 Amphoteric Acceptor/DonorY +3 Amphoteric Acceptor/DonorDy +3 Amphoteric Acceptor/DonorCe +3/4 Amphoteric Acceptor/DonorPr +3/4 Amphoteric Acceptor/DonorGd +3 Amphoteric-Mainly A Acceptor/DonorTm +2/3 Amphoteric-Mainly B Acceptor/Donor

24

Simply doping BaTiO3 with any of the aforementioned dopants is insufficient to achieve

a flat dielectric response; it is also necessary to have a chemically inhomogeneous

microstructure, referred to as a core-shell microstructure, to achieve the nearly

temperature independent dielectric properties.

In a core-shell microstructure, individual grains have a chemical gradient ranging

from undoped BaTiO3 at the center, or core, of the grain, to fully doped BaTiO3 at the

edge of the grain. The chemical gradient leads to a gradient in the Curie temperature

from pure BaTiO3 at the core, to the Tc of doped BaTiO3 in the shell. The effect on the

overall dielectric response of the material can be envisioned as the superposition of the

dielectric response of several different doped BaTiO3 compositions. Figure 2.4 is a

schematic showing the additive effect of several different compositions with sharp

individual transitions on the temperature response. The superposition of the individual

transitions yields a flat, temperature independent response. Although the doping

formulation presented in this thesis is for precious metal electrode systems, a discussion

of the development of base metal systems is presented to provide a perspective on the

current and future direction in the processing of MLCCs.

2.4.2 Doping for Based Metal Electrodes

Traditionally, precious metal electrodes (PME) have been used as the internal

electrode material in MLCCs. The oxidation resistance of the precious metals permits the

sintering of PME-MLCCs in ambient atmosphere. But recently, the rising cost of

precious metals, most notably Pd, has forced the use of base metal electrodes (BME) in

MLCCs to lower manufacturing costs. Ni is the mostly commonly used material in

BME, but recent interest has also focused on Cu.111 For either material, BME-MLCCs

25

Figu

re 2

.4.

The

supe

rpos

ition

of i

ndiv

idua

l tra

nsiti

on re

sults

in a

bro

ad d

iffus

e tra

nsiti

on, w

hich

is m

ore

stab

le w

ith c

hang

es in

te

mpe

ratu

re.

26

must be fired in a reducing atmosphere to prevent the oxidation of the internal electrodes.

The primary disadvantage of BME-MLCCs has been limited lifetime compared to PME-

MLCCs. The use period of the early commercial BME-MLCCs was as little as several

hours.109 The lifetime of current capacitor technology is limited by the degradation rate

of insulation resistance.112 Over time as the resistance decreases, the leakage current

increase leads to enhanced thermal breakdown as resistive heating occurs. In a study on

degradation of polycrystalline perovskite titanate materials, Waser et al. found oxygen

vacancy migration to be one of the primary factors in electrical degradation.112 However,

the work by Waser et al. only focused on degradation mechanisms in the ceramic and did

not address the electrode/ceramic interactions which will be discussed later.

Initial work on sintering in a reducing atmosphere achieved a dramatic decrease in

the insulation resistance of pure BaTiO3 because of the generation of oxygen

vacancies109,

OxHVoBaTiOxHBaTiO xx 2323 ][ +→+ − [2.7]

which are compensated by electrons,

eVoVo ′+→ •• 2 [2.8]

To overcome the problem of decreased resistance, acceptor doping was

attempted.113-115 Acceptors present on the Ti site were noted to be excellent electron

traps. However, the acceptor defects in the lattice are also compensated by the generation

of oxygen vacancies. For example, a Mn2+ substitution on the Ti-site109,

••++′′→ OOTi VOnMMnO [2.9]

27

generates one oxygen vacancy for each Mn2+ substituted. At room temperature oxygen

vacancies are highly mobile under an applied electric field, and the high ionic

conductivity leads to reduced lifetime.112, 116

A post sintering anneal in oxygen has been used to oxidize the acceptors to

eliminate oxygen vacancies115,

•••• +′→++′′ OTiOTi VnMOVnM 22122 2 [2.10]

Re-oxidation improved lifetime, but a large concentration of oxygen vacancies still

existed and BME capacitors still exhibited limited lifetime compared to air fired precious

metal electrode capacitors.114

Lifetime was further improved by doping with donor-acceptor complexes.109, 117-

119 An example of this is the co-doping of Mn2+ with W6+, both dopants are B-site

substitutions and when the defects are closely associated they form a neutral donor-

acceptor complex,

0}{ •••• ′′→+′′ TiTiTiTi WnMWnM [2.11]

The donor dopants compensate the acceptor dopants and oxygen vacancies are eliminated

to improve lifetime. The close association of the donor-acceptor complexes reduces the

mobility of the ionic defects under an applied field further improving lifetime.109, 117

Doping with rare earth (RE) elements, specifically, Dy and Ho, also resulted in

improved degradation characteristics.109, 110, 118, 120, 121 Initial experiments suggested that

RE occupy the A- and B-site simultaneously and acts as self-compensating acceptor-

donor complexes similar to the co-doping process. Tsur et al. performed a analysis of

rare earth doping in BaTiO3 and found that the stable valence and amphoteric nature of

28

specific RE dopants, were the reason for the improved lifetime.108 Ho was noted to be

most stable amphoteric dopant.

2.4.3 Doping Methodology

Whether it is a PME or BME system current, dopant methods are based on solid-

state methods where dopant oxide particles are added to a large volume of matrix

particles.122-124 The mixture is typically ball milled to try to blend dopants and to

homogenize the mixture as much as possible. A liquid sintering aid is typically added to

help distribute the dopants and aid in the formation of the core-shell structure. Many

authors have stated that the presence of the liquid is necessary for the development of a

core-shell microstructure.99, 100, 125-127 The observed microstructure is attributed to a

dissolution-precipitation mechanism during sintering. However, Chazono and Kishi

observed the development of a core-shell structure at temperatures well-below the

melting point of the liquid.128 Chazono and Kishi proposed that when the dopants exhibit

a low solubility in the liquid or if no liquid is present, diffusion can develop the core-shell

structure, but the high temperatures and long times needed for diffusion make the use of a

liquid advantageous.

The process of solid-state doping has been studied by Wiseman in the doping of

ZnO for the fabrication of varistors. It showed that if dopants are present as particulates

that chemically homogeneity is difficult or nearly impossible.129 Although a liquid-phase

sintering aid is present, high temperatures and long sintering times are still necessary for

the solubility, distribution, and diffusion of the dopants to remove the chemical

inhomogenieties that exist due to doping with particulates.

29

A chemical approach is an alternative method of doping in which the dopant is

added in an ionic or molecular form. Doping then occurs by the adsorption or

precipitation of the dopants on the particle surface which then creates a homogeneous

dopant layer on the particle. Due to the increase in chemical homogeneity in the bulk

samples, sintering times and temperatures can be reduced. Both approaches, ionic and

molecular, have been shown to be successful in the processing of highly engineered

materials such as ZnO-based varistors130 and BaTiO3-based dielectrics.122-124, 131-133

Although an aqueous approach for BaTiO3 has been previously outlined, no attempt has

been made to passivate to protect the surface from degradation. For solution based

doping in aqueous suspension the chemical stability of BaTiO3 must first be addressed.

Fernandez et al. studied the effect of the doping technique on the electrical

properties of BaTiO3 doped with Co and Nb.133 Three different techniques were used: (1)

doping during powder synthesis, (2) conventional solid-state doping, and (3) a chemical

approach. The first technique resulted in a homogenous composition which exhibited a

single sharp peak in the dielectric response. The other two techniques yielded materials

which meet X7R standards and have core-shell microstructures. Fernandez et al. found a

significant microstructural difference between solid-state and chemical doping. Solid-

state doping led to local Co- and Nb-rich regions whereas chemical doping resulted in a

homogenous intergranular structure with chemically inhomogeneous grains. However,

both methods result in an inhomogeneous chemical composition with significant

dispersion in transition temperatures that produces an overall diffuse dielectric response.

30

2.5 Size Effects and Electrical Properties of Nanoscale BaTiO3 Materials

Reducing the layer thickness of a MLCC has many advantages, but the primary

disadvantage of reduced layer thickness is reduced dielectric constant. The effect of size

on the dielectric constant of BaTiO3 dielectrics is well known and documented.134-139

Figure 2.5 is a plot of the room temperature dielectric constant as a function of grain size

for undoped BaTiO3 from Shaw et al.140 where there is a maximum in the dielectric

constant at a grain size of approximately 1 μm. The use of nanoparticles in the current

work is expected to increase the importance of size effect phenomena. Therefore, a

review of current theories on size effects is presented.

2.5.1 BaTiO3 Particles

A wide variety of research has been done to understand the size effect in bulk

BaTiO3.30, 138, 141-143 Over the past 10 years it has been shown that the size effect in

BaTiO3 is highly dependent on the state of the materials. There exists a difference in the

properties of particles and bulk materials due to a difference in the boundary conditions at

either the particle surface or grain boundary. It has been theorized that the stress-state,

presence of secondary phases, and electrical boundary conditions all affect the intrinsic

ferroelectric properties as the size of BaTiO3 grains are reduced.

Theoretical calculations based on free energy arguments of the intrinsic size limit

of BaTiO3 have been presented in the literature. 144 Using Landau-Devonshire theory

Wang et al. calculated the critical limit of ferroelectric behavior in BaTiO3.

Assumptions for the calculation include that the ferroelectric-paraelectric transition was a

first order transition regardless of the size and boundary conditions. A critical size of 44

nm was calculated. Using Landau theory, but accounting for depolarization fields and

31

Figu

re 2

.5.

Plot

show

ing

the

depe

nden

ce o

f the

die

lect

ric c

onst

ant o

f BaT

iO3

with

gra

in si

ze in

bot

h bu

lk c

eram

ics (●)

and

thin

fil

ms (Δ)

take

n fr

om S

haw

et a

l.137

32

the presence of a Schottky barrier at the particle edges due to space charge Shih et al.

calculated an intrinsic limit for BaTiO3 of various sizes.145 The intrinsic limit of

ferroelectric behavior was shown to be dependent on the thickness of the Schottky layer;

with thinner Schottky barrier resulting in a lower critical size limit. Recently, Junquera

and Ghosez used a model system of epitaxial BaTiO3 on SrRuO3 electrode for first

principle calculation of the intrinsic loss of ferroelectric behavior.146 A limit of 24 Å or

six unit cells was calculated.

Begg et al. used X-ray diffraction (XRD) and differential scanning calorimetery

(DSC) to investigate the phase and phase transition for hydrothermal and Clabaugh-

derived BaTiO3 particles.147 A size limit of 190 nm was the limit of the tetragonal phase

of BaTiO3 to temperatures as low as 80 K. However, little to no shift was observed in the

TC of the particles as a function of c/a ratio for tetragonal materials. The intrinsic limit

was theorized to be due to surface energy constraints on the system. Li et al. used a

similar approach, but accounted for the degree of agglomeration in the synthesized

powders.148 Li et al. found 20 - 30 nm to be the limit of the tetragonal phase in BaTiO3

particles. However, it was noted that only agglomerated particles were able to maintain

the tetragonal phase to such small sizes. Li et al. concluded that the depolarization

energy in particles with the tetragonal phase was minimized when the particles

agglomerated, and if the particle remained unagglomerated the cubic-tetragonal phase

transition occurred to minimize the depolarization energy. Wada et al. recently

developed a technique to measure the dielectric constant of particles in a slurry.30, 149 The

dielectric constant of a highly loaded suspension was measured. Using mixing rules the

dielectric constant of the particles was calculated. Particles with sizes ranging from 17 to

33

1000 nm were synthesized and measured. A size of 140 nm yielded the maximum

dielectric response with the loss of ferroelectricity taking place from 17 to 40 nm. Wada

et al. did not characterize the state of dispersion of the particle suspension and therefore

the effect of agglomeration on the intrinsic loss of ferrolectricity was not discussed.

McCauley et al.141and Randall et al.150used a glass ceramic system to precipitate

out and control the size of BaTiO3 particles dispersed in a glass matrix. BaTiO3

crystallite size of 20 – 80 nm was observed. Through extrapolation of the experimental

data a critical size for ferroelectric behavior of 17 nm was determined. A substantial

decrease of the TC and broadening of the curve as crystallite size decreases was observed.

In addition, it was believed that the phase transition had become a second order phase

transition. The electrical boundary conditions imposed on the BaTiO3 crystallites by the

surrounding glass matrix influenced the distribution of the spontaneous polarization

within the crystallites. This distribution results in a quasi-paraelectric shell which

exhibits temperature independent dielectric response leading to shifts in Tc and

broadening of the dielectric temperature response.

2.5.2 Bulk BaTiO3

All of the previously described work focused on understanding the critical limit

and size effects in discreet particle systems. However, most BaTiO3 based materials are

used in a bulk state where the boundary conditions change significantly. In work by Frey

et al. sol-gel derived bulk BaTiO3 samples showed no shift in the TC for samples with

grain sizes as small as 35 nm.142, 151 However, a substantial decrease in the maximum

dielectric constant was noted. TEM analysis showed the presence of an 8 Å thick grain

boundary region, which was believed to have a low dielectric constant. It was theorized

34

that the low dielectric constant grain boundary region acted in series or parallel with the

grain decreasing the effective dielectric constant of the bulk samples. Using a diphasic

brick-wall model proposed by Payne and Cross152 and assuming a dielectric constant of

130 for the grain boundary region, good agreement between the model and experimental

results were obtained.

Ragulya et al. also used the brick-wall approach to model the dielectric constant

of BaTiO3 samples sintered under different conditions.153, 154 Ragulya and co-workers

used rate controlled sintering (RCS) and hot-pressing to yield dense BaTiO3 samples with

grains sizes from 100 to 450 nm. Samples sintered under different conditions which had

similar grain sizes (130 nm) exhibited different maximum dielectric constants. Samples

sintered using RCS exhibited an increase of 2500 in the maximum dielectric constant

over the hot-pressed sample. TEM microstructural analysis showed that samples sintered

with RCS had low angle grain boundaries and an overall thinner grain boundary region

compared to the hot-pressed samples. When the brick-wall model was applied the two

different samples the difference in the thickness of the grain boundary regions explained

the decrease in the dielectric constant for the hot-pressed samples.

2.5.3 BaTiO3 Thin Films

The dielectric properties of ferroelectric thin films show size effects similar to

bulk ferroelectric samples. However, the underlying substrate can lead to effects that are

only an issue for thin films. Shaw et al. noted that intrinsic size and thickness effects in

ferroelectric thin film can be difficult to separate because factors, such as grain size, can

be dependent on the film thickness.140 Thin layers in MLCCs are different than those

deposited on rigid, single crystal substrates. For example, when BaTiO3 is deposited on a

35

single crystal Si wafer, there is a mismatch between the lattice constant of the BaTiO3

and the wafer. In non-epitaxial films this lattice mismatch leads to a strain, α, which has

been shown to shift the Tc of ferroelectric thin films.

Due to differences in processing approaches, the thin layers in MLCCs do not

exhibit similar stress states as those of other thin films. However, the inherent

differences in the shrinkage, thermal expansion, and lattice mismatch between the

dielectric and electrode materials led to residual stresses in MLCCs. Shin et al. and Park

used Vicker’s microhardness to measure the residual stress in BaTiO3 MLCCs.155, 156

Both researchers found that the stress was complex and was dependent on the geometry

of the MLCC. However, two general stress states were found: (1) stresses parallel to the

electrodes were compressive whereas, (2) stresses perpendicular to the electrode were

tensile. No comprehensive study has been performed on the effect of residual stress in

MLCCs on the dielectric properties, but the work by Shin et al. and Park show that the

stress state is complex and non-trivial and it is therefore assumed that the dielectric

properties of the BaTiO3 are affected.

In addition to the presence of residual stress, pure geometric constraints also

affect the electrical properties of thin films. When the grain size is on the order of the

film thickness, a brick wall model no longer accurately represents the connectivity of

secondary phases in the microstructure. Depending on the microstructure either series or

parallel conductivity of secondary phases is possible. Waser has shown how the

differences in connectivity can have a profound effect on the dielectric properties of thin

films.157 Using simple series and parallel mixing rules the effect of secondary phases on

the dielectric properties of thin film is easily demonstrated. When secondary phase are

36

present at grain boundaries which are perpendicular to the electrode the capacitance of

grains and secondary phase have an additive effect. In contrast, secondary phases which

are parallel to the electrode have a diluting effect on the overall dielectric constant of the

film.

2.5.4 Electrode-Dielectric Interactions

The previous discussion focused on the degradation mechanism in the ceramic

material due to doping and presence of oxygen vacancies. In the current section

degradation and diminished dielectric properties will be addressed in the context of

electrode/dielectric interactions. Ag/Pd alloys and Ni are the most commonly used

electrode materials in MLCCs, but recently Cu has begun to be used more frequently.158

The reaction of BaTiO3 with electrode materials has been investigated to better

understand the ceramic/metal interface. Each electrode material presents unique issues in

the degradation of BaTiO3 based MLCCs.

In PME-MLCCs Ag and Pd exhibit a low solubility in BaTiO3 and do not oxidize

to form a secondary phase,159 but Ag has a high mobility at elevated temperature and can

precipitate out in the matrix.160, 161 Electromigration of Ag in MLCCs is a common

problem; under an electric field Ag migrates, typically via the grain boundaries, in

dendritic patterns resulting it conductive pathways in the dielectric layers.162 The growth

of the dendrite reduces the effective interelectrode distance, and at the end of the dendrite

the electric field is concentrated because of the sharp tip. Breakdown then occurs

between the dendrite and the opposite electrode because the local electric field or current

is large enough to induce intrinsic or thermal breakdown. As the dielectric layer

37

thickness is reduced in MLCCs the degradation rate is expected to increase because the

high mobility of Ag will lead to rapid dendrite formation.

While Ni is not as mobile as Ag with respect to migration, problems persist that

must be overcome. Studies on the interduffision of Ni and the dielectric layer during the

sintering in reducing atmosphere show that Ni readily diffuses into the dielectric layer.163,

164 Samples sintered at 1250 °C in a pO2 of 10-10 to 10-13 atm showed Ni diffusion up to

500 nm into the dielectric layer. No intermediate or secondary phases were noted at the

electrode-dielectric interface. The substitution of Ni into BaTiO3 is a problem because it

lowers the Curie point of BaTiO3 and increases the oxygen vacancy concentration. If

sintering temperatures were higher and time were longer, the solubility limit of Ni in

BaTiO3 would be reached, and when Ni is above the solubility limit NiO will form.165 If

present, NiO can act as a low K secondary phase which would decrease the dielectric

constant.

Ni diffusion is due to a high chemical potential gradient at the interface, and the

stability of NiO at elevated temperatures. Even in samples fired in low pO2 the driving

force for Ni oxidation is still large. Therefore, the Ni will change its oxidation state. To

overcome the problems of Ni diffusion lower sintering temperatures, shorter dwell times,

and low pO2 values have been suggested.

A complete analysis of BaTiO3-Cu system has not yet been performed. However,

Song and Randall exhibited the feasibility of a Cu-electrode of co-fired BaTiO3 X7R

MLCC.158 To lower the sintering temperature a ZnO-B2O3 flux was added. An 80 nF

prototype MLCC with Cu electrode and 16 active dielectric layers was fabricated.

SEM/EDS analysis showed there is no interaction between the Cu electrode and BaTiO3

38

dielectric during sintering. But work by Langhammer et al. has shown that Cu can

modify the BaTiO3 crystal structure and stabilize the high temperature hexagonal phase at

room temperature if the sintering temperature is too high.166 Even though most of these

effects can be managed, the combined effects will become more pronounced as layer

thickness is further reduced. At reduced grain size and layer thickness the volume

fraction of secondary phases will be greatly increased. Thinner layers will results in

shorter grain boundary paths between electrode and the problem of Ag migration will

result in diminished lifetimes.

2.6 Conclusions

BaTiO3 is the most important dielectric material in the fabrication of MLCCs.

For over 60 years research has been conducted on BaTiO3 over a variety of topics from

synthesis to processing to electrical properties. With the current drive in reduction of

layer thickness it is necessary to understand the impact of all steps in the process on the

final derived properties.

The goal of this research is to understand the impact of tabular BaTiO3

nanoparticles on the properties and deposition of thin BaTiO3 layers. With insight gained

from the literature, the hydrothermal synthesis of anisotropic BaTiO3 nanoparticles is

investigated (Chapter Three). After synthesis, a systematic study of the surface chemistry

was conducted (Chapter Four). From the literature it was determined that a passivation

technique is needed to limit Ba2+ dissolution and the aqueous degradation of BaTiO3

(Chapter Five). Finally, the deposition of BaTiO3 for the buildup of thin dielectric layers

was investigated (Chapter Six).

39

References

1. A.J. Moulson and J.M. Herbert, Electroceramics: Materials, properties, applications, 1st ed. (Chapman & Hall, London, 1990). 2. D.F.K. Hennings, B.S. Schreinemacher, and H. Schreinemacher: Solid-state preparation of BaTiO3-based dielectrics, using ultrafine raw materials. J. Am. Ceram. Soc. 84, (12), 2777 (2001). 3. L. Gao: Preparation of spherical, solid barium titanate particles with uniform composition at high precursor concentration by spray pyrolysis. J. Am. Ceram. Soc. 87, (5), 830 (2004). 4. B.K. Kim, D.Y. Lim, R.E. Riman, J.S. Nho, and S.B. Cho: A new glycothermal process for barium titanate nanoparticle synthesis. J. Am. Ceram. Soc. 86, (10), 1793 (2003). 5. R.A. Kimel, V. Ganine, and J.H. Adair: Double injection synthesis and dispersion of submicrometer barium titanyl oxalate tetrahydrate. J. Am. Ceram. Soc. 84, (5), 1172 (2001). 6. B. Jiang, J.L. Peng, L.A. Bursill, T.L. Ren, P.L. Zhang, and W.L. Zhong: Defect structure and physical properties of barium titanate ultra-fine particles. Phys. B. 291, (1-2), 203 (2000). 7. K.M. Hung, W.D. Yang, and C.C. Huang: Preparation of nanometer-sized barium titanate powders by a sol-precipitation process with surfactants. J. Euro. Ceram. Soc. 23, 1901 (2003). 8. Y. Kobayashi, A. Nishikata, T. Tanase, and M. Konno: Size effect on crystal structures of barium titanate nanoparticles prepared by a sol-gel method. J. Sol-gel Sci.Tech. 29, (1), 49 (2004). 9. Z. Novak, Z. Knez, M. Drofenik, and I. Ban: Preparation of BaTiO3 powders using supercritical CO2 drying of gels. J. Non-Cryst. Solid 285, (1-3), 44 (2001). 10. S. Wada, T. Tsurumi, H. Chikamori, T. Noma, and T. Suzuki: Preparation of nm-sized BaTiO3 crystallites by a LTDS method using a highly concentrated aqueous solution. J. Cryst. Grow. 229, (1), 433 (2001). 11. J. Moon, E. Suvaci, A. Morrone, S.A. Constantino, and J.H. Adair: Formation mechanisms and morphological changes during the hydrothermal synthesis of BaTiO3 particles from a chemically modified, amorphous titanium (hydrous) oxide precursor. J. Euro. Ceram. Soc. 23, (12), 2153 (2003).

40

12. D. Hennings, G. Rosenstein, and H. Schreinemacher: Hydrothermal preparation of barium titanate from barium-titanium acetate gel precursors. J. Euro. Ceram. Soc. 8, 107 (1991). 13. X. Wang, B.I. Lee, M. Hu, E.A. Payzant, and D.A. Blom: Synthesis of nanocrystalline BaTiO3 by solvent refluxing method. J. Mater. Sci. Lett. 22, (7), 557 (2003). 14. P. Gherardi and E. Matijevic: Homogenous precipitation of spherical colloidal barium titanate particles. Colloid Surface 32, (3-4), 257 (1988). 15. P. Pinceloup, C. Courtois, A. Leriche, and B. Thierry: Hydrothermal synthesis of nanometer-sized barium titanate powders: Control of barium/titanium ratio, sintering, and dielectric properties. J. Am. Ceram. Soc. 82, (11), 3049 (1999). 16. Q. Huang and L. Gao: Synthesis and characterization of hexapod-shaped tetragonal-phase barium titanate single crystals. J. Am. Ceram. Soc. 87, (7), 1350 (2004). 17. J.H. Adair and E. Suvaci: Submicron electroceramic powder by hydrothermal synthesis. In Encyclopedia of Materials: Science and Technology, edited by K.H.J. Buschow, R.W. Cahn, M.C. Flemings, B. Ilschner, E.J. Kramer, and S. Mahajan, (Elsevier Science Ltd., 2001) pp 8933. 18. C. Pithan, D.F.K. Hennings, and R. Waser: Progress in the synthesis of nanocrystalline BaTiO3 powders for MLCC. Int. J. Appl. Ceram. Technol. 2, (1), 1 (2005). 19. L.K. Templeton and J.A. Pask: Formation of BaTiO3 from BaCO3 and TiO2, in air and CO2. J. Am. Ceram. Soc. 42, 212 (1959). 20. A. Amin, M.A. Spears, and B.M. Kulwicki: Reaction of anatase and rutile with barium carbonate. J. Am. Ceram. Soc. 66, (10), 733 (1983). 21. A. Beauger, J.C. Mutin, and J.C. Niepce: Synthesis reaction of metatitanate BaTiO3, Part 1: Effect of the gaseous atmosphere upon the thermal evolution of the system BaCO3-TiO2. J. Mater. Sci. 18, 3041 (1983). 22. A. Beauger, J.C. Mutin, and J.C. Niepce: Synthesis reaction of metatitanate BaTiO3, Part 2: Study of solid-state reaction interfaces. J. Mater. Sci. 18, 3543 (1983). 23. W. Clabaugh, E. Swiggard, and R. Gilchrist: Preparation of barium titanyl oxalate tetrahydrate for conversion to barium titanate of high purity. J. Res. Nat. Bur. Stand. 56, (5), 289 (1956).

41

24. C. Szepesi: Synthesis and processing of sub-micron barium titanyl oxalate tetrahydrate. M.S. Thesis, The Pennsylvania State University, University Park, PA, (2004). 25. H. Yamamura, A. Watanabe, S. Shirasaki, Y. Moriyoshi, and M. Tanada: Preparation of barium titanate by oxalate method in ethanol solution. Ceram. Int. 11, 17 (1985). 26. Z.H. Park, H.S. Shin, B.K. Lee, and S.H. Cho: Particle size control of barium titanate prepared from barium titanyl oxalate. J. Am. Ceram. Soc. 80, 1599 (1997). 27. O.O. Vasyl'kiv, A.V. Ragulya, and V.V. Skorokhod: Synthesis and sintering of nanocrystalline barium titanate powder under nonisothermal conditions. II. Phase analysis of the decomposition products of barium titanyl-oxalate and the synthesis of barium titanate. Powder Metallurgy and Metal Ceramics 36, (5-6), 277 (1997). 28. O.O. Vasyl'kiv, A.V. Ragulya, V.P. Klimenko, and V.V. Skorokhod: Synthesis and sintering of nanocrystalline barium titanate powder under nonisothermal conditions. III. Chromatographic analysis of barium titanyl-oxalate gaseous decomposition products. Powder Metallurgy and Metal Ceramics 36, (11-12), 575 (1997). 29. S. Wada, M. Narahara, T. Hoshina, H. Kakemoto, and T. Tsurumi: Preparation of nm-sized BaTiO3 particles using a new 2-step thermal decomposition of barium titanyl oxalate. J. Mater. Sci. 38, 2655 (2003). 30. S. Wada, H. Yasuno, T. Hoshina, S. Nam, H. Kakemoto, and T. Tsurumi: Preparation of nm-sized barium titanate fine particles and their powder dielectric properties. Jpn. J. Appl. Phys. 42, 6188 (2003). 31. A.V. Ragulya, O.O. Vasyl'kiv, and V.V. Skorokhod: Synthesis and sintering of nanocrystalline barium titanate powder under nonisothermal conditions. I. Control of dispersity of barium titanate during its synthesis from barium titanyl oxalate. Powder Metallurgy and Metal Ceramics 36, (3-4), 170 (1997). 32. A.V. Polotai, A.V. Ragulya, T.V. Tomila, and C.A. Randall: The XRD and IR study of the barium titanate nano-powder obtained via oxalate route. Ferroelectrics 298, 243 (2004). 33. M.P. Pechini: Method of preparing lead and alkaline earth titanates and niobates and coating methods using the same to form a capacitor. US Patent # 3,330,697, (1967). 34. B.J. Mulder: Preparation of BaTiO3 and other ceramic powders by co-precipitation of citrates in an alcohol. Amer. Ceram. Soc. Bull. 49, 990 (1970).

42

35. H. Matsuda, M. Kuwabara, K. Yamada, H. Shimooka, and S. Takahashi: Optical adsorption in sol-gel derived crystalline barium titanate fine particles. J. Am. Ceram. Soc. 81, (11), 3010 (1998). 36. H. Shimooka and M. Kuwabara: Preparation of dense BaTiO3 ceramics from sol-gel derived monolithic gels. J. Am. Ceram. Soc. 78, (10), 2849 (1995). 37. H. Herrig and R. Hempelmann: A colloidal approach to nanometre-sized mixed oxide ceramic powders. Mater. Lett. 27, 287 (1996). 38. C. Beck, W. Hartl, and R. Hempelmann: Size-controlled synthesis of nanocrystalline BaTiO3 by a sol-gel type hydrolysis in microemulsion-provided nanoreactors. J. Mater. Res. 13, (11), 3174 (1998). 39. S. Li, J.A. Eastman, and L.J. Thompson: Gas-condensation synthesis of nanocrystalline BaTiO3. Appl. Phys. Lett. 70, (17), 2244 (1997). 40. K. Suzuki and K. Kijima: Well-crystallized barium titanate nanoparticles prepared by plasma chemical vapor deposition. Mater. Lett. 58, 1650 (2004). 41. Y. Itoh, T.O. Kim, M. Shimada, K. Okuyama, and I. Matsui: One-step preparation of BaTiO3 nanoparticles by liquid source chemical vapor deposition. J. Chem. Eng. Jap. 37, (3), 454 (2004). 42. M. Viviani, J. Lemaitre, M.T. Buscagalia, and P. Nanni: Low-temperature aqueous synthesis (LTAS) of BaTiO3: A statistical design of experiment approach. J. Euro. Ceram. Soc. 20, 315 (2000). 43. M. Viviani, M.T. Buscagalia, A. Testino, V. Buscagalia, P. Bowen, and P. Nanni: The influence of concentration on the formation of BaTiO3 by direct reaction of TiCl4 with Ba(OH)2 in aqueous solution. J. Euro. Ceram. Soc. 23, 1383 (2003). 44. A. Testino, M.T. Buscagalia, M. Viviani, V. Buscagalia, and P. Nanni: Synthesis of BaTiO3 particles with tailored sized by precipitation from aqueous solution. J. Am. Ceram. Soc. 87, (1), 79 (2004). 45. S. Wada, H. Chikamori, T. Noma, and T. Suzuki: Synthesis of nm-sized barium titanate crystallites using a new LTDS method and their characterization. J. Mater. Sci. 35, (19), 4857 (2000). 46. S.S. Flaschen: An aqueous synthesis of barium titanate. J. Am. Chem. Soc. 77, 6194 (1955). 47. K. Kiss, J. Magder, M.S. UVukasovich, and R.J. Lockhard: Ferroelectrics of ultrafine particle size: I, Synthesis of titanate powders of ultrafine particle size. J. Am. Ceram. Soc. 49, (6), 291 (1966).

43

48. K.S. Mazdiyasni, R.T. Dolloff, and J.S. Smith: Preparation of high-purity submicron barium titanate powders. J. Am. Ceram. Soc. 52, (10), 523 (1969). 49. W. Ovramenko, L. Shvets, F. Ovcharenko, and B. Kornilovich: Kinetics of hydrothermal synthesis of barium metatitanate. Inorg. Mat. 15, (11), 1560 (1979). 50. T.R.N. Kutty, R. Vivekanandan, and P. Murugaraj: Precipitation of rutile and anatase (TiO2) and their conversion to MTiO3 (M= Ba, Sr, Ca). Mater. Chem. Phys. 19, (6), 533 (1988). 51. W. Hertl: Kinetics of barium-titanate synthesis. J. Am. Ceram. Soc. 71, (10), 879 (1988). 52. J.A. Kerchner, J. Moon, R. Chodelka, A. Morrone, and J.H. Adair: Nucleation and formation mechanisms of hydrothermal derived barium titanate. In Synthesis and Characterization of Advanced Materials. ACS Symposium Series, Vol 681, edited by M.A. Serio, D.M. Gruen, and R. Malhotra, (American Chemical Society, 1998) pp 106. 53. K. Osseo-Asare, F.J. Arriagada, and J.H. Adair: Solubility relationships in the coprecipitation synthesis of barium titanate: Heterogeneous equilibria in the Ba-Ti-C2O4-H2O system. In Ceramic Transactions, Ceramic Powder Science, edited by G.L. Messing, E.R. Fuller, Jr., and H. Hausner, (The American Ceramic Society, 1988) pp 47. 54. M.M. Lencka and R.E. Riman: Thermodynamic modeling of hydrothermal synthesis of ceramic powders. Chem. Mater. 5, (1), 61 (1993). 55. K. Abe, M. Aoki, H. Rikimaru, T. Ito, K. Hidaka, and K. Segawa: Process for producing a composition which includes perovskite compounds. US Patent # 4,643,984, (1987). 56. E. Lilley and R.R. Wusirika: Method for the production of mono-sized barium titanate. US Patent # 4,764,493, (1988). 57. J. Menashi, R.C. Reid, and L. Wagner: Barium titanate powders. US Patent # 4,829,033, (1989). 58. J. Menashi, R.C. Reid, and L. Wagner: Barium titanate based dielectric compositions. US Patent # 4,832,939, (1989). 59. J. Menashi, R.C. Reid, and L. Wagner: Doped barium titanate based compositions. US Patent # 4,863,883, (1989). 60. M.K. Klee and H.W. Brand: Method of manufacturing powdered barium titanate. US Patent # 4,859,448, (1989).

44

61. M. Hu, V. Kurian, E.A. Payzant, C.J. Rawn, and R.D. Hunt: Wet-chemical synthesis of monodisperse barium titanate particles - hydrothermal conversion of TiO2 microspheres to nanocrystalline BaTiO3. Powder Tech. 110, 2 (2000). 62. R.I. Walton, D. Millange, R.A. Smith, T.C. Hansen, and D. O'Hare: Real time observation of the hydrothermal crystallization of barium titanate using in-situ neutron powder diffraction. J. Amer. Chem. Soc. 123, (50), 12547 (2001). 63. D.V. Miller: Synthesis and properties of barium titanate nanocomposites. Ph.D. Thesis, The Pennsylvania State University, University Park, PA, (1991). 64. R. Vivekanandan, S. Philip, and T.R.N. Kutty: Hydrothermal preparation of Ba(Ti,Zr)O3 fine powders. Mater. Res. Bull. 22, (1), 99 (1986). 65. J. Moon, J.A. Kerchner, H.G. Krarup, and J.H. Adair: Hydrothermal synthesis of ferroelectric perovskites from chemically modified titanium isopropoxide and acetate salts. J. Mater. Res. 14, (2), 425 (1999). 66. J. Moon, E. Suvaci, T. Li, S.A. Costantino, and J.H. Adair: Phase development of barium titanate from chemically modified-amorphous titanium (hydrous) oxide precursor. J. Euro. Ceram. Soc. 22, (6), 809 (2002). 67. D.F.K. Hennings, C. Metzmacher, and B.S. Schreinemacher: Defect chemistry and microstructure of hydrothermal barium titanate. J. Am. Ceram. Soc. 84, (1), 179 (2001). 68. J.K. Lee, K.S. Hong, and J.W. Jang: Roles of Ba/Ti ratios in the dielectric properties of BaTiO3 ceramics. J. Am. Ceram. Soc. 84, (9), 2001 (2001). 69. Y.B. Moa, S. Banerjee, and S.B. Wong: Hydrothermal synthesis of perovskite nanotubes. Chem. Comm. 3, 408 (2003). 70. T.J. Yosenick, D.V. Miller, R. Kumar, J.A. Nelson, C.A. Randall, and J.H. Adair: Synthesis of nanotabular barium titanate via a hydrothermal route. J. Mater. Res. 20, (4), 837 (2005). 71. R.B. Bagwell, J. Sindel, and W. Sigmund: Morphological evolution of barium titanate synthesized in water in the presence of polymeric species. J. Mater. Res. 14, (5), 1844 (1999). 72. M.C. Blanco-Lopez, B. Rand, and F.L. Riley: The properties of aqueous phase suspensions of barium titanate. J. Euro. Ceram. Soc. 17, 281 (1997). 73. M.C. Blanco-Lopez, G. Fourlaris, and F.L. Riley: Interactions of barium titanate powders with an aqueous suspending medium. J. Euro. Ceram. Soc. 18, 2183 (1998).

45

74. S. Venigalla and J.H. Adair: Theoretical modeling and experimental verification of electrochemical equilibria in the Ba-Ti-C-H2O system. Chem. Mater. 11, (3), 589 (1999). 75. U. Paik and V.A. Hackley: Influence of solids concentration on the isoelectric point of aqueous barium titanate. J. Am. Ceram. Soc. 83, (10), 2381 (2000). 76. A. Neubrand, R. Linder, and P. Hoffman: Room-temperature solubility behavior of barium titanate in aqueous media. J. Am. Ceram. Soc. 83, (4), 860 (2000). 77. U. Paik, S. Lee, and V.A. Hackley: Influence of barium dissolution on the electrokinetic properties of colloidal BaTiO3 in an aqueous medium. J. Am. Ceram. Soc. 86, (10), 1662 (2003). 78. C.W. Chiang and J.H. Jean: Effects of barium dissolution on dispersing aqueous barium titanate suspensions. Mater. Chem. Phys. 80, (3), 647 (2003). 79. H. Krurap, R. Linhart, and J.H. Adair, Opal, 1.93x; Gainsville, FL, (1997). 80. J.H. Adair and H.G. Krurup: The role of solution chemistry in understanding colloidal stability of ceramic suspensions. In Advances in process measurements for the ceramic industry, edited by A. Jillavenkatesa, and G.Y. Onoda, (The American Ceramic Society: Westerville, OH, 1999) pp 205. 81. P. Bendale, S. Venigalla, J.R. Ambrose, E.D. Verink, and J.H. Adair: Preparation of barium-titanate films at 55-degrees-C by an electrochemical method. J. Am. Ceram. Soc. 76, (10), 2619 (1993). 82. D.A. Anderson, J.H. Adair, D.V. Miller, J.V. Biggers, and T.R. Shrout: Surface chemistry effect on ceramic processing of BaTiO3 powder. In Ceramic Transactions, Ceramic Powder Science, edited by G.L. Messing, E.R. Fuller, Jr., and H. Hausner, (The American Ceramic Society: Westerville, OH, 1988) pp 485. 83. P. Duran, J. Tartaj, and C. Moure: Sintering behavior and microstructural evolution of agglomerated spherical particles of high-purity barium titanate. Ceram. Int. 29, 419 (2003). 84. J.L. Crampo: The role of passivation/dispersion on grain growth in barium titanate ceramics. B.S. Thesis, The Pennsylvania State University, University Park, PA, (2001). 85. H. Kamiya, K. Gomi, Y. Iida, K. Tanaka, T. Yoshiyasu, and T. Kakiuchi: Preparation of highly dispersed ultrafine barium titanate powder by using mircobial-derived surfactant. J. Am. Ceram. Soc. 86, (12), 2011 (2003). 86. A.W.M. de Laat and W.P.T. Derks: Colloidal stabilization of BaTiO3 with poly(vinyl alcohol) in water. Colloid Surface A 71, (2), 147 (1993).

46

87. X. Wang, B. Lee, and L. Mann: Dispersion of barium titanate with polyaspartic acid in aqueous media. Colloid Surface A 202, (1), 71 (2002). 88. Z. Shen, J. Chen, H. Zou, and J. Yun: Dispersion of nanosized aqueous suspensions of barium titanate with ammonium polyacrylate. J. Colloid Interface Sci. 275, (1), 158 (2004). 89. J. Zhao, X. Wang, Z. Gui, and L. Li: Dispersion of barium titanate with poly(acrylic acid-co-maleic acid) in aqueous media. Ceram. Int. 30, (7), 1985 (2004). 90. Y. Song, X. Liu, and J. Chen: The maximum solid loading and viscosity estimation of ultra-fine BaTiO3 aqueous suspensions. Colloid Surface A 247, 27 (2004). 91. K. Hsu, K. Ying, L. Chen, B. Yu, and W. Wei: Dispersion properties of BaTiO3 colloids with amphoteric polyelectrolytes. J. Am. Ceram. Soc. 88, (3), 524 (2005). 92. G.H. Kirby, D.A. Harris, Q. Li, and J.A. Lewis: Poly(acrylic acid)-poly(ethylene oxide) comb polymer effects on BaTiO3 nanoparticle suspension stability. J. Am. Ceram. Soc. 87, (4), 181 (2004). 93. G.H. Kirby, D.A. Harris, Q. Li, and J.A. Lewis: PAA-POE comb polymer dispersants for colloidal processing. Key Eng. Mater. 264-268, 161 (2004). 94. U. Paik, V.A. Hackley, J. Lee, and S. Lee: Effect of poly(acrylic acid) and poly(vinyl alcohol) on the solubility of colloidal BaTiO3 in an aqueous medium. J. Mater. Res. 18, (5), 1266 (2003). 95. J.H. Adair and S.A. Constantino: Ceramic slip composition and method for preparing the same. US Patent # 6,214,756, (2001). 96. R.P. Vasquez, B.D. Hunt, and M.C. Foote: Wet chemical passivation of YBa2Cu3O7-

x. J. Electrochem. Soc. 137, (7), 2344 (1990). 97. W.J. Merz: The electrical and optical behavior of BaTiO3 single-domain crystals. Phys. Rev. 76, (8), 1221 (1949). 98. EIA-198-1-F - Ceramic dielectric capacitors classes I, II, III and IV - Part I: Characteristics and requirements 99. D.F.K. Hennings and G. Rosenstein: Temperature-stable dielectrics based on chemically inhomogeneous BaTiO3. J. Am. Ceram. Soc. 67, (4), 249 (1984). 100. D.F.K. Hennings and B.S. Schreinemacher: Temperature-stable dielectric materials in the system BaTiO3-Nb2O5-Co3O4. J. Euro. Ceram. Soc. 15, (4), 463 (1994).

47

101. M. Kahn: Influence of grain growth on dielectric properties of Nb-doped BaTiO3. J. Am. Ceram. Soc. 54, (9), 455 (1971). 102. B.S. Rawal, M. Kahn, and W.R. Buessem: Grain core-shell structure in barium titanate-based dielectrics. In Advances in Ceramics, Vol. 1, Grain Boundary Phenomena in Electronic Ceramics, edited by L.M. Levinson, (American Ceramic Society, 1981) pp 172. 103. T.R. Armstrong and R.C. Buchanan: Influence of core-shell grains on the internal stress state and permittivity response of zirconia-modified barium titanate. J. Am. Ceram. Soc. 73, (5), 1268 (1990). 104. T.R. Armstrong, K.A. Young, and R.C. Buchanan: Dielectric properties of fluxed barium titanate ceramics with zirconia additives. J. Am. Ceram. Soc. 73, (3), 700 (1990). 105. F. Azough, R. Al-Saffar, and R. Freer: A transmission electron microscope study of commercial X7R-type multilayer ceramic capacitors. J. Euro. Ceram. Soc. 18, 751 (1998). 106. Y. Park and Y.H. Kim: The dielectric temperature characteristics of additives modified barium titanate having core-shell structured ceramics. J. Mater. Res. 10, (11), 2770 (1995). 107. B. Jaffe, W.R. Cook, and H. Jaffe, Piezoelectric Ceramics, 1st ed. (Academic Press, London, 1971). 108. Y. Tsur, T.D. Dunbar, and C.A. Randall: Crystal and defect chemistry of rare earth cations in BaTiO3. J. Electroceram. 7, 25 (2001). 109. D.F.K. Hennings: Dielectric materials for sintering in reducing atmospheres. J. Euro. Ceram. Soc. 21, 1637 (2001). 110. W. Lee, W.A. Groen, H. Schreinemacher, and D.F.K. Hennings: Dsyprosium doped dielectric materials for sintering in reducing atmospheres. J. Electroceram. 5, (1), 31 (2000). 111. Y. Tsur, J.H. Adair, and C.A. Randall: Improving the oxidation resistance of base metal powders. Jpn. J. Appl. Phys. 39, (10), 6004 (2000). 112. R. Waser, T. Baiatu, and K.H. Hardtl: dc electrical degradation of perovskite-type titanates: I, Ceramics. J. Am. Ceram. Soc. 73, (4), 1645 (1990). 113. J.M. Herbert: High permittivity ceramics sintered in hydrogen. Trans. Br. Ceram. Soc. 62, (8), 645 (1963).

48

114. I. Burn and G.H. Maher: High resistivity BaTiO3 ceramics sintered in CO-CO2 atmosphere. J. Mater. Sci. 10, 633 (1975). 115. H.J. Hagemann and D.F.K. Hennings: Reversible weight change of acceptor-doped BaTiO3. J. Am. Ceram. Soc. 64, (10), 590 (1981). 116. R. Waser, T. Baiatu, and K.H. Hardtl: dc electrical degradation of perovskite-type titanates: II, Single crystals. J. Am. Ceram. Soc. 73, (6), 1654 (1990). 117. K. Albertsen, D.F.K. Hennings, and O. Steigelemann: Donor-acceptor charge complex formation in barium titanate ceramics: Role of firing atmosphere. J. Electroceram. 2, (3), 193 (1998). 118. H. Kishi, Y. Mizuno, and H. Chazono: Base-metal electrode-multilayer ceramic capacitors: Past, present and future perspectives. Jpn. J. Appl. Phys. 42, (1), 1 (2003). 119. C. Lee, S. Kang, D. Sinn, and H. Yoo: Co-doping effect of Mn and Y on charge and mass transport properties of BaTiO3. J. Electroceram. 13, 785 (2004). 120. Y. Okino, H. Shizuno, S. Kusumi, and H. Kishi: Dielectric properties of rare-earth-oxide-doped BaTiO3 ceramics fired in reducing atmosphere. Jpn. J. Appl. Phys. 33, (9B), 5393 (1994). 121. Y. Tsur and C.A. Randall: Point defect concentrations in barium titanate revisited. J. Am. Ceram. Soc. 84, (9), 2147 (2001). 122. J.F. Fernandez, A.C. Caballero, P. Duran, and C. Moure: Improving sintering behavior of BaTiO3 by small doping additions. J. Mater. Sci. 31, (4), 975 (1996). 123. A.C. Caballero, J.F. Fernandez, C. Moure, P. Duran, and Y.M. Chiang: Grain growth control and dopant distribution in Zno-doped BaTiO3. J. Am. Ceram. Soc. 81, (4), 939 (1998). 124. A.C. Caballero, M. Villega, J.F. Fernandez, C. Moure, P. Duran, P. Florian, and J.P. Coutures: Reactive sintering of phosphorous coated BaTiO3. J. Euro. Ceram. Soc. 19, (6-7), 979 (1999). 125. H. Lu, J. Bow, and W. Deng: Core-shell structures in ZrO2-modified BaTiO3 ceramic. J. Am. Ceram. Soc. 73, (12), 3562 (1990). 126. C.A. Randall, S.F. Wang, D. Laubscher, J.P. Dougherty, and W. Huenber: Structure property relationships in core-shell BaTiO3-LiF ceramics. J. Mater. Res. 8, (4), 871 (1993). 127. Y. Kuromitsu, S.F. Wang, S. Yoshikawa, and R.E. Newnham: Interactions between barium titanate and binary glasses. J. Am. Ceram. Soc. 77, (2), 493 (1994).

49

128. H. Chazono and H. Kishi: Sintering characteristics in the BaTiO3-Nb2O5-Co3O4 ternary system: II, Stability of the so-called "core-shell" structure. J. Am. Ceram. Soc. 83, (11), 101 (2000). 129. G.H. Wiseman: Advanced manufacturing process for zinc oxide surge arrester disks. Key Eng. Mater. 150, 209 (1998). 130. S. Ural: Aggregate breakdown and aqueous processing of zinc oxide varistors. M.S. Thesis, The Pennsylvania State University, University Park, PA, (2003). 131. X. Liu: Structure-property relationships in submicron X7R dielectric materials. M.S. Thesis, The Pennsylvania State University, University Park, PA, (1999). 132. S.A. Bruno: Ceramic dielectric compositions and method for enhancing dielectric properties. US Patent # 5,082,811, (1992). 133. J.F. Fernandez, P. Duran, and C. Moure: Influence of the doping method on X7R based-dielectric capacitors. Ferroelectrics 127, 47 (1992). 134. R.J. Brandmayr, A.E. Brown, and A.M. Dunlap:Annealing effects on microstructure and dielecrtric properties of hot-pressed, ultrafine grained BaTiO3; ECOM-2614; (1965). 135. W.R. Buessem, L.E. Cross, and A.K. Goswami: Phenomenological theory of high permittivity in fine-grain barium titanate. J. Am. Ceram. Soc. 49, (1), 33 (1966). 136. W.R. Buessem, L.E. Cross, and A.K. Goswami: Effect of two-dimensional pressure on the permittivity of fine- and coarse-grained barium titanate. J. Am. Ceram. Soc. 49, (1), 36 (1966). 137. H.T. Martirena and J.C. Burfoot: Grain-size effects on properties of some ferroelectric ceramics. J. Phys. C 7, 3182 (1974). 138. G. Artl, D.F.K. Hennings, and G. de With: Dielectric properties of fine-grained barium titanate ceramics. J. Appl. Phys. 58, (4), 1619 (1985). 139. L. Mitoseriu, V. Tura, C. Papusoi, T. Osaka, and M. Okuyama: A comparative study of the grain size effects on ferro-para phase transition in barium titanate ceramics. Ferroelectrics 223, 99 (1999). 140. T.M. Shaw, S. Trolier-McKinstry, and P.C. McIntyre: The properties of ferreolectric films at small dimensions. Annu. Rev. Mater. Sci. 30, 263 (2000). 141. D. McCauley, R.E. Newnham, and C.A. Randall: Intrinsic size effects in a barium titanate glass-ceramic. J. Am. Ceram. Soc. 81, (4), 979 (1998).

50

142. M.H. Frey and D.A. Payne: Grain-size effect on structure and phase transformations for barium titanate. Phys. Rev. B. 54, (5), 3158 (1996). 143. Z. Zhao, V. Buscagalia, M. Viviani, M.T. Buscagalia, L. Mitoseriu, A. Testino, M. Nygren, M. Johnsson, and P. Nanni: Grain-size effects on the ferroelectric behavior of dense nanocrystalline BaTiO3 ceramics. Phys. Rev. B. 70, (2), (2004). 144. Y.G. Wang, W.L. Zhong, and P.L. Zhang: Size driven phase transitions in ferroelectric particles. Solid State Comm. 90, (5), 329 (1994). 145. W.Y. Shih, W.H. Shih, and I.A. Aksay: Size dependence of the ferroelectric transition of small BaTiO3 particles: Effect of depolarization. Phys. Rev. B. 50, (21), 15575 (1994). 146. J. Junquera and P. Ghosez: Critical thickness for ferroelectricity in perovskite ultrathin films. Nature 422, 506 (2003). 147. B.D. Begg, E.R. Vance, and J. Nowotny: Effect of particle size on the room-temperature crystal structure of barium titanate. J. Am. Ceram. Soc. 77, (12), 3186 (1994). 148. X. Li and W. Shin: Size effect in barium titanate particles and clusters. J. Am. Ceram. Soc. 80, (11), 2844 (1997). 149. S. Wada, T. Hoshina, S. Nam, H. Kakemoto, T. Tsurumi, and M. Yashima: Size dependence of dielectric properties for nm-sized barium titanate crystallites and its origins. J. Kor. Phys. Soc. 46, (1), 303 (2005). 150. C.A. Randall, D. McCauley, and D.P. Cann: Finite size effects in BaTiO3 ferroelectric glass ceramic. Ferroelectrics 206-207, 325 (1998). 151. M.H. Frey, Z. Xu, P. Han, and D. Payne: The role of interfaces on an apparent grain size effect on the dielectric properties for ferroelectric barium titanate ceramics. Ferroelectrics 206-207, 337 (1998). 152. D.A. Payne and L.E. Cross: Microstructure-property relations for dielectric ceramics. II. The brick-wall model of the polycrystalline microstructure. In Microstructure and properties of ceramic materials, edited by T.S. Yen, and J.A. Pask, (Beining Science Press, 1984) pp. 153. A.V. Ragulya, V.V. Skorokhod, and A.V. Polotai: Synthesis and sintering of nanocrystalline barium titanate powder under nonisothermal conditions. VI. Structure, grain boundaries, and dielectric properties of barium titanate obtained by various sintering methods. Powder Metallurgy and Metal Ceramics 40, (1-2), 25 (2001).

51

154. A.V. Ragulya and A.V. Polotai: Non-isothermal sintering of barium titanate nano-powders of different origination. Ferroelectrics 254, 41 (2001). 155. D. Park, Y. Jung, and U. Paik: Evaluation of residual stress in BaTiO3-based Ni-MLCC with X7R characteristics. J. Mater. Sci. 15, 253 (2004). 156. Y. Shin, K. Kang, Y. Jung, J. Yeo, S. Lee, and U. Paik: Internal stresses in BaTiO3/Ni MLCCs. J. Euro. Ceram. Soc. 23, 1427 (2003). 157. R. Waser and O. Lohse: Electrical characterization of ferroelectric, paraelectric, and superparaelectric thin films. Inter. Ferro. 21, (1-4), 27 (1998). 158. T.H. Song and C.A. Randall: Copper cofire X7R dielectrics and multilayer capacitors based on zinc borate fluxed barium titanate ceramic. J. Electroceram. 10, 39 (2003). 159. S. Shin and W.H. Tuan: Solubility of silver and palladium in BaTiO3. J. Am. Ceram. Soc. 87, (3), 401 (2004). 160. C.Y. Chen and W.H. Tuan: Evaporation of Ag during co-firing with BaTiO3. J. Am. Ceram. Soc. 83, (7), 1693 (2000). 161. C.Y. Chen and W.H. Tuan: Effect of silver on the sintering and grain growth behavior of barium titanate. J. Am. Ceram. Soc. 83, (12), 2988 (2000). 162. S.J. Krubien: Tutorial: Electrolytic models for metallic electromigration failure mechanisms. IEEE Trans. Reliab. 44, (4), 539 (1995). 163. Y. Wang, L. Li, J. Qi, Z. Ma, J. Cao, and Z. Gui: Nickel diffusion in base-metal-electrode MLCCs. Mater. Sci. and Eng. B99, 378 (2003). 164. Z. Gui, Y.L. Wang, and L.T. Li: Study on the interduffision in base-metal-electrode MLCC's. Ceram. Int. 30, 1275 (2004). 165. W.H. Tzing and W.H. Tuan: Effect of NiO additions on the sintering and grain growth behavior of BaTiO3. Ceram. Int. 25, 69 (1999). 166. H.L. Langhammer, T. Muller, R. Bottcher, and H.P. Abicht: Crystal structure and related properties of copper-doped barium titanate ceramics. Solid State Sci. 5, 965 (2003).

52

CHAPTER THREE

Synthesis of Nanotabular Barium Titanate via a Hydrothermal Route 3.1 Introduction

With the recent demands placed on cellular and mobile technologies the need for

nanoparticles of highly engineered materials has increased. The high volume, low cost,

and superior properties of passive electronic components requires that precision powders

for electronic component are inexpensive, of high quality, and can be produced at high

yields. Control of chemical and hydrothermal defects in the powders is of great

importance in controlling the properties of multilayer ceramic capacitors (MLCC’s) made

from the powder. Therefore, the synthesis of high quality powders of high dielectric

properties, especially perovskite structured materials, has been of specific interest. The

dependence of volumetric capacitance of a MLCC on the thickness of the active layer is

well-documented.1 Nanoparticles and their assembly provide potential to reduce layer

thickness below the current standard of 1 μm.

Much work on a variety of synthesis routes for the synthesis of BaTiO3

nanoparticle has occurred recently. Solid-state carbonate reactions2, a modified Clabaugh

process3, and low temperature direct synthesis (LTDS)4 and several other routes5-7 have

all been investigated. The first two routes are common commercial methods used to

produce powders, however each require further processing, either milling or calcination,

to produce nanoscale BaTiO3, while powders produced by LTDS have a high hydroxyl

defect concentration and a low Ba/Ti ratio. Hydrothermal synthesis is a common method

53

that produces nanosized powders. Under the correct synthesis conditions powders with

low defect concentrations and controlled stoichiometry that requires no further processing

can be produced, making hydrothermal synthesis an excellent choice for the commercial

synthesis of BaTiO3.8

Little to no work on the synthesis of BaTiO3 nanoparticles has focused on

morphology control during synthesis. Instead focus has been on the synthesis of

spherical nanoparticles. Tabular nanoparticles are particles that have a plate-like

morphology with a thickness in the nanoscale and an aspect ratio as high as 100:1.

Nanotabular particles have many advantages over spherical nanoparticles for the laydown

of thin films. For example work by Yener et al. on the electrophoretic deposition of

Ag/Pd nanoparticles has shown improved thickness control and surface roughness of thin

metal layers when the layers are comprised of tabular nanoparticles instead of spherical

nanoparticles.9 The thickness in the nanoscale allows for the laydown of thin layers

with low surface roughness. In addition, the large face area allows for adsorption of a

polyelectrolyte dispersants in high concentrations yielding high surface charge enabling

the creation of stable suspensions.10

A few researchers have shown that morphology control of perovskite materials

during hydrothermal synthesis is possible, but that work yielded micron sized particles.

Work by Moon et al.11 and Cho et al.12 have focused on the synthesis of PbTiO3 and PZT

with varying morphologies. Bagwell controlled the morphology of BaTiO3 by the

addition of polymeric species during synthesis.13 The polymer adsorption varied among

different crystallographic planes leading to different growth rate for each plane. Under

highly alkaline conditions (pH 14) Zhao et al.14 synthesized BaTiO3 octahedra in which

54

the {111} face is the stable habit. At elevated temperatures, 225 °C, Miller synthesized

BaTiO3 particles that resemble hexagonal platelets with the {111} face as the basal

plane.15 Thus, the synthesis scheme of Miller is of interest because of the reported plate-

like nature of the materials. However, there was no report of particle thickness or other

physical properties in the Miller work. Therefore, the current study has verified the

Miller synthesis and provides a more complete characterization.

3.2 Materials and Methods

All chemicals were reagent grade and used without further purification. Titanium

isopropoxide (TI) (97%, Aldrich Chemical Company, Milwaukee, WI) and barium

hydroxide octahydrate (BHO) (98+%, Aldrich Chemical Company, Milwaukee, WI)

were used as the titanium and barium sources, respectively. Figure 3.1 is a flow diagram

of the procedure used in the synthesis. A 500 ml solution of a 1M TI and BHO was

prepared by adding the appropriate amount of BHO to CO2-free DI water and stirring for

10 minutes. CO2-free DI water had been previously prepared by boiling with flowing

argon to remove adsorbed CO2. The removal of CO2 is necessary to limit the formation

of BaCO3 during synthesis. Next, TI was added and the solution was stirred for an

additional 30 minutes. The BHO and TI were mixed in equimolar quantities to obtain

stoichiometric BaTiO3.

The solution was placed in a 1 L reaction vessel (Parr Instrument Company,

Moline, IL) with the stirring speed set at 60 rpm. The thermal treatment consisted of a 2

hour ramp to 225 °C with a 5 hour hold followed by cooling to room temperature. The

pH of the solution measured before and after synthesis and was pH 13.1 and pH 12.7,

55

Barium Hydroxide Octahydrate

Ba(OH)2●8H2O (157.74g)

Titanium IsopropoxideTi[OCH(CH3)2]4

(142.13g)

Add CO2 Free DI H2O(278.8mL)

Combine Equimolar Amountsof Precursors (500 of 1M

BaTiO3 solution) & Stir for 30min

Hydrothermal Treatment

225 ºC for 5hrs

Figure 3.1. Flow diagram for the hydrothermal synthesis of platelet BaTiO3. The starting solution is 500 mL total of a 1M solution, and has an approximate yield of 120g of powder

56

respectively. To investigate the morphological evolution of the particles as a function of

temperature after synthesis powder samples were heat treated at 10 °C/min on Pt foil

placed in an alumina crucible. Samples were heated to 375, 450, 700, 800, 900, 1000,

and 1100 °C removed at temperature and quenched in air.

Particle morphology and size were determined using atomic force microscopy

(AFM) (Multimode IIIa, Digital Instruments, Santa Barbara, CA) and transmission

electron microscopy (TEM) (2010F, JEOL, Japan). For AFM analysis dilute suspensions

were prepared and a drop was placed on atomically cleaved mica substrate. TEM

analysis was performed using holey-carbon film on Cu grids (Electron Microscopy

Sciences, Fort Washington, PA) as the sample holder. A single drop of dilute suspension

was placed on each TEM grid. Size distributions of both the thickness and face diameter

were calculated on a number basis using the offline AFM software (Nanoscope III

Version 5.12r3, Digital Instruments, Santa Barbara, CA) and image analysis software

(Scion Image Beta 4.0.2, Scion Corporation, Fredrick, MD).

Phases present and other physical properties were determined using a variety of

characterization techniques. X-ray diffraction (XRD) (Scintag Pad V, Thermo-ARL,

Dearborn, MI) was used to determine the solid phases present. X-ray fluorescence (XRF)

(1600/10, Phillips, Netherlands) was used to determine the major and minor constituents

of the as-synthesized powder. Density measurements were performed using helium

pycnometery (Multivolume Pycnometer 1305, Micromeritics, Norcross, GA). The

weight loss of the powder up to 1000 °C was analyzed using thermogravimetric analysis

(TGA) (TGA 2050 Thermogravimetric Analyzer, TA Instruments Inc., New Castle, DE).

57

Particle surface area was measured using BET gas adsorption (Gemini, Micromeritics,

Norcross, GA).

Schematic representations of the (100), (110), and (111) planes of BaTiO3 were

calculated and rendered using Atoms for Windows (Version 3.2, Atoms Software,

Kingsport, TN). A cubic crystal structure with m3m space group with a lattice parameter

of 4 Å was assumed. The planar density of both Ti and surface OH were then calculated

using the model planar surfaces. In the current study calculations based on a tetragonal

crystal structure (space group P4mm) were omitted because of the small difference in the

a and c lattice parameters in the tetragonal structure.

3.3 Results and Discussion

3.3.1 BaTiO3 Particle Morphology and Growth

Figure 3.2 is the X-ray diffraction pattern for the as-synthesized powder and

powder heat-treated to 1000 °C. The as-synthesized powder is the pseudo-cubic phase of

the perovskite crystal structure, due to the stain on the lattice introduced by the presence

of hydroxide defects in the lattice.16 Upon heat treatment to 1000 °C, where the

hydroxide defects are no longer present, the phase is the tetragonal perovskite crystal

structure. Figure 3.3a is a TEM image of the as-synthesized BaTiO3 particles. The

associated selected area electron diffraction (SAED) pattern, Figure 3.3b, indicates that

the particles are single crystal with a <111> zone axis parallel to the surface normal.

Figure 3.3c is a partial ring diffraction pattern from a cluster of approximately 100

particles. The pattern shows the presence of rings for diffraction from {100}, {110}, and

{200} planes. Absent from the pattern, between the {110} and {200} rings, is the

58

Figu

re 3

.2.

X-r

ay d

iffra

ctio

n pa

ttern

for a

s-sy

nthe

size

d po

wde

r and

pow

der h

eat t

reat

ed to

100

0 °C

. Th

e as

-syn

thes

ized

pow

der i

s ps

eudo

-cub

ic d

ue to

the

pres

ence

of h

ydro

ther

mal

def

ects

in th

e la

ttice

. A

fter h

eat t

reat

men

t at 1

000

°C p

eak

split

ting

is

obse

rvab

le in

the

(200

)/(00

2) p

eak

(see

inse

rt gr

aph)

and

indi

cate

s the

mat

eria

l has

con

verte

d to

the

tetra

gona

l for

m o

f BaT

iO3.

Not

e: *

Cub

ic B

aTiO

3pe

aks J

CPD

S C

ard:

31-

0174

and

+Te

trago

nal B

aTiO

3pe

aks J

CPD

S C

ard:

79-

2264

59

Figu

re 3

.3a,

b, a

nd c

.TE

M m

icro

grap

h (a

) and

ass

ocia

ted

sele

cted

are

a el

ectro

n di

ffra

ctio

n pa

ttern

(b).

The

sing

le c

ryst

al

diff

ract

ion

patte

rn sh

ows t

hat t

he p

artic

le is

sing

le c

ryst

als w

ith <

111>

zon

e ax

is.

The

abse

nce

of d

iffra

ctio

n fr

om (1

11) i

n th

e pa

rtial

ring

diff

ract

ion

patte

rn (c

) fro

m a

clu

ster

of p

artic

less

how

s a m

ajor

ity o

f the

par

ticle

s in

the

clus

ter s

how

text

ure.

0.35

Å-1

0.25

Å-1

60

expected ring from {111} diffraction. The absence of {111} diffraction indicates that a

majority of that particles in the cluster show texture and have a <111> zone axis

alignment. A few spots from {111} diffraction are present but this is probably due to

particles overlap and misalignment with the electron beam which occurred during the

drying process in the TEM grid.

Figure 3.4 is an AFM cross-section image of the BaTiO3 particles showing that

the particles have plate-like morphology. The thickness and face diameter distribution of

the synthesized particles was calculated used a procedure developed by Yuan and co-

workers.17 The AFM offline software is used to create uniform bins of constant height

which are represented by a distinct color in the AFM image. Any particle with a

thickness which lies within the bin is easily recognized by its color and counted

accordingly. For face diameter measurements the AFM image is converted into a black

and white image, with the area of each particle calculated using image analysis software.

After the area is calculated the face is assumed to be circular and the diameter is

calculated. Figure 3.5 is the log-normal distribution calculated on a number basis for a

total of 214 particles. The cumulative distribution was calculated and the median value

for both the thickness and face diameter was determined to be 5.8 ± 3.1 nm and 27.1 ±

12.3 nm, respectively.

At high pH when titanium isopropoxide is added to water it decomposes to form a

hydrous titania-gel and isopropyl alcohol (IPA) based on the reaction:

Ti[OCH(CH3)2]4 + 4H2O Ti(OH)4 + 4(CH3)2CHOH [3.1]

In the concentrations used to create the 1M solution for the hydrothermal synthesis the

amount of IPA created is considerable, 28vol% of the solution. It was possible during

61

Figu

re 3

.4.

AFM

cro

ss-s

ectio

nal i

mag

e of

BaT

iO3

parti

cles

on

an a

tom

ical

ly fl

at c

leav

ed m

ica

subs

trate

. Th

e pa

rticl

es h

ave

a pl

ate-

like

mor

phol

ogy

with

a th

ickn

ess o

f 7.9

nm

and

face

dia

met

er o

f 46.

9 nm

.

62

Figu

re 3

.5.

Thic

knes

s and

face

dia

met

er si

ze d

istri

butio

ns fo

r the

hyd

roth

erm

al B

aTiO

3pl

atel

ets.

The

dis

tribu

tions

wer

e ca

lcul

ated

usi

ng th

e A

FM o

fflin

e so

ftwar

e an

d im

age

anal

ysis

softw

are.

Bot

h of

the

dist

ribut

ions

are

bas

ed o

n a

tota

l of 2

14

parti

cles

.

63

synthesis that IPA adsorption on the particle surface controls morphology, in a manner

similar to that of polymeric adsorption observed by Bagwell and others.13, 18 A fractional

distillation of the titania gel/IPA solution was carried out to remove the IPA produced by

the hydrolysis of the titanium isopropoxide. The volume of IPA removed by distillation

was measured and CO2-free DI water added in the same volume to ensure the

concentration of the starting solution was identical to all previous experiments. The new

starting solution contained only Ba(OH)2, Ti(OH)4, and CO2-free DI water. The powder

produced was analyzed using TEM and AFM and shows a morphology and dimensions

similar to the other powder synthesized. This critical experiment confirms that IPA

during synthesis does not affect the growth of the particles. Only the high pH, feedstock

concentration, and high packing density of the (111) plane are the primary variables

controlling the morphological evolution of the particles.

With a cubic symmetry during growth the plate-like morphology of the BaTiO3 is

best modeled after (111) double twins proposed by Schmelz and Thomann.19 Schmelz

and Thomann observed the formation of (111) twins in bulk BaTiO3 samples with excess

TiO2 heat treated under sub-eutectic conditions. The high atomic packing density, high

barrier to nucleation, and slow growth of the (111) plane lead to (111) twin formation in

the bulk BaTiO3. The morphology of (111) double twins, shown schematically in Figure

3.6, is a hexagonal shaped platelet similar to the hydrothermal BaTiO3 platelets.

Based on the periodic bond chain (PBC) model by Hartman and Perdok20, Tani et.

al. states the Ti-O bond array is the PBC for the perovskite structure.21 Therefore Ti and

not Ba will be the growth limiting species for the hydrothermal synthesis of BaTiO3.

Figure 3.7 is a schematic representations of the (100), (110), and (111) planes for cubic

64

Figu

re 3

.6.

Sche

mat

ic o

f BaT

iO3

plat

elet

s for

med

via

mul

tiple

{11

1} tw

in fo

rmat

ion.

Afte

r Sch

mel

z an

d Th

oman

n.20

65

Figu

re 3

.7.

Sche

mat

ic re

pres

enta

tions

of (

a) (1

00) p

lane

, (b)

(110

) pla

ne a

nd (c

) (11

1) p

lane

rend

ered

usi

ng A

tom

s for

Win

dow

s.

Each

figu

re sh

ows t

he o

xyge

n co

ordi

natio

n of

tita

nium

in e

ach

plan

e. T

he g

eom

etry

of e

ach

plan

e w

as u

sed

to c

alcu

late

the

Ti

plan

ar d

ensi

ty a

nd su

rfac

e O

H d

ensi

ty.

66

BaTiO3 generated using Atoms for Windows©. Each figure shows the position and

coordination of the Ti in each of the three planes. Table 3.1 is a tabulation of the Ti site

density for each plane and more important the surface hydroxide concentration for each

plane. The Ti surface sites are energetically unstable and at the high pH during synthesis

will be terminated by hydroxyl groups. Surface Ti on the (100) plane are coordinated by

five oxygen while Ti on the (111) surface are only coordinated by three oxygen.

Therefore each surface Ti on the (100) plane will only be able to react with one hydroxyl

group while Ti on the (111) surface can react with three hydroxyl groups resulting in a

increase hydroxide concentration for the (111) plane. Lencka and Riman22 showed that,

at greater than pH 4, Ti(OH)4(aq) is the stable Ti species during the hydrothermal synthesis

of BaTiO3 and that synthesis is a competition between BaTiO3 and Ti(OH)4(aq) formation.

During hydrothermal synthesis, it is speculated that the growth of any new layer a

competition between BaTiO3 and Ti(OH)x(surf) growth. Under these assumptions, the

crystal plane with the highest hydroxide concentration should be the stable facet, which is

the (111) plane. Observations in the hydrothermal synthesis of BaTiO3 by Zhao at pH 14

support these conclusions.14

(111) twin formation during BaTiO3 synthesis is well-known with an example

being “butterfly’ twins observed in BaTiO3 synthesized by the Remeika method.23

“Butterfly” twins are specific twins where the twin plane is the (111) plane with the (100)

plane being the stable crystal habit. In BaTiO3 (which posses a high-temperature

hexagonal phase) (111) twin can be envisioned as a stack fault of hexagonal phase.

Nielsen et al. theorized that the ease of stacking fault formation is the cause of (111)

twinning in BaTiO3.24 With the high likelihood of (111) twin formation and the limited

67

Pla

neTi

/nm

2S

urfa

ce O

H/n

m2

(100

)6.

256.

25(1

10)

4.40

8.80

(111

)3.

6010

.80

Tab

le 3

.1.

Plan

ar d

ensi

ty o

f Ti a

nd su

rfac

e hy

drox

ide

for l

ow in

dex

plan

es in

cub

ic B

aTiO

3(a

= 4

Å).

68

growth in the [111] direction during hydrothermal synthesis crystals with (111) twins and

(111) specific habit are expected. A TEM investigation should be conducted to look for

the presence of (111) twins.

3.3.2 Characterization of Physical Properties

The physical properties of the platelet BaTiO3 was characterized in addition to

two commercial hydrothermal powders to provide standards. Table 3.2 is a comparison

of the physical properties of the uncalcined platelet BaTiO3 and the two commercial as-

received powders. All of the powders have BET surface areas greater than 10 m2/g

which yield equivalent spherical diameters of less than 100 nm. The presence of the

pseudo-cubic perovskite phase in all of the powders indicates that hydrothermal defects

are present in the particles. Figure 3.8 is a TGA curve for the three powders heated to

1000 °C. The platelet BaTiO3 powder experiences two weight losses, in the ranges of

300 to 500 °C and 600 to 800 °C. The first weight loss is only 0.5 wt% and is due to

hydrothermal defects25, whereas, the higher temperature weight loss is due to BaCO3.26

Both commercial powders exhibit similar weight loss characteristics however each has a

higher concentration of hydroxide in the lattice; commercial powder A has 1.62wt% loss

and the commercial powder B has 0.75wt% loss. Hydrothermal defects should be

minimized because of intragranular pore formation during sintering.25 Hennings et al.

developed a defect chemistry model for undoped BaTiO3 based on pycnometery and X-

ray density calculations.25 Charge irregularities due to hydroxide defects in the lattice are

compensated by Ba and Ti vacancies. Upon heating, defects coalesce to form

intragranular pores. The presence of pores in the final fired microstructure degrades the

69

Pow

der

Sur

face

Are

a (m

2 /g)

OH

Def

ect C

once

ntra

tion

(wt%

)D

ensi

ty (g

/cm

3 )B

a/Ti

by

XRF

Pha

se b

y XR

DS

hape

Asp

ect R

atio

Pla

tele

t10

.5 ±

0.3

90.

505.

87 ±

0.0

21.

006 ±

0.00

05P

seud

o-cu

bic

Pla

tele

t4.

7P

owde

r A11

.9 ±

0.6

61.

625.

66 ±

0.0

41.

003 ±

0.00

03P

seud

o-cu

bic

Sph

eric

al

1P

owde

r B15

.6 ±

0.0

90.

755.

81 ±

0.03

0.98

9 ±

0.00

07P

seud

o-cu

bic

Sph

eric

al1

Tab

le 3

.2.P

hysi

cal p

rope

rties

of p

late

let,

com

mer

cial

pow

der A

, and

com

mer

cial

pow

der B

hyd

roth

erm

ally

der

ived

BaT

iO3

pow

ders

.

Not

e: T

he e

rror

bar

s pre

sent

repr

esen

t 95%

con

fiden

ce in

terv

al fo

r an

aver

age

of 3

or m

ore

mea

sure

men

tsX

RF

used

to d

eter

min

e B

a/Ti

ratio

(mea

sure

men

t cou

rtesy

of F

erro

Cor

p.)

70

Figu

re 3

.8.

Wei

ght l

oss c

urve

for p

late

let,

com

mer

cial

pow

der A

, and

com

mer

cial

pow

der B

. Th

e w

eigh

t los

s fro

m 3

00 to

500

°C

is d

ue to

the

rem

oval

of h

ydro

xyl d

efec

ts, w

here

as t

he w

eigh

t los

s at h

ighe

r tem

pera

ture

s is t

he re

mov

al o

f BaC

O3.

71

electrical properties. The density measurements correlate well with the TGA results and

suggest that the plate-like particles will develop homogenous dense microstructures.

XRF analysis of the BaTiO3 shows that the platelet BaTiO3 has a Ba/Ti ratio of

1.006 ± 0.0005. Excess barium is potentially problematic for further processing.

Stoichiometric BaTiO3 has limited solubility for excess barium. At the current Ba/Ti

ratio, a moderate amount of Ba2TiO4 is possible upon heat treatment. Ba2TiO4 has been

shown to form as inclusions in BaTiO3 grains and is detrimental to the density and

dielectric properties.27

3.3.3 Morphological Evolution as a Function of Temperature

To investigate the stability of the plate-like morphology of the particles the

material was thermally treated and the particles were characterized. Figure 3.9 shows the

TGA and derivative curve for the platelet powder samples. There are four observed

reactions occurring at 350, 425, 660, and 760 °C. Based on the four reactions, the

powder samples were thermally treated at 375, 450, 700, and 800 °C to investigate the

effect of the weight loss reactions on the structure and morphology of the particles.

Samples were also thermally treated at 900, 1000, and 1100 °C to determine the onset of

sintering and morphological change. Figure 3.10a-h is a series of TEM images of the

thermally treated powder samples.

By 375 °C (Figure 3.10a) the hydrothermal defects in the powder have coalesced

and appear as 10 nm equiaxed defects in the interior of the particles. This is similar to

the observations of Hennings et al. in which point defects in the lattice coalesce to form

defects which were observable up to temperatures of 800 °C.25 The defects are removed

by 1000 °C (Figure 3.10g). The increased temperature compared to Hennings et al.

72

Figu

re 3

.9.

Wei

ght l

oss a

nd d

eriv

ativ

e cu

rve

for p

late

let p

owde

r sho

ws t

he p

rese

nce

of fo

ur re

actio

ns o

ccur

ring

at 3

50, 4

25, 6

60,

and

760

ºC.

73

Figu

re 3

.10a

-h.

Serie

s of T

EM im

ages

show

ing

the

mor

phol

ogic

al e

volu

tion

of th

e pl

atel

et p

artic

les a

s a fu

nctio

n of

tem

pera

ture

: (a

) 25

ºC –

as-s

ynth

esiz

ed, (

b) 3

75 ºC

, (c)

450

ºC, (

d) 7

00 ºC

, (e)

800

ºC, (

f) 9

00 ºC

, (g)

100

0 ºC

, and

(h) 1

100

ºC.

Nec

k fo

rmat

ion

is o

bser

vabl

e at

800

ºC w

ith m

orph

olog

ical

cha

nges

occ

urrin

g by

900

ºC.

At 3

75 ºC

hyd

roth

erm

al d

efec

ts h

ave

begu

n to

coa

lesc

e an

d ar

e no

t rem

oved

unt

il 10

00 ºC

.

74

75

observations is explained by the high heating rate (10°C/min) and the lack of any dwell in

the thermally treated samples.

Solid bridging and the onset of sintering is seen in samples heated to 800 °C

(Figure 3.10e). Changes in the morphology beings by 900 °C (Figure 3.10f) observed by

the loss of well-faceted particles and the rounding of the particles to a more equiaxed

morphology. Complete loss of morphology occurs by 1000 °C (Figure 3.10g). The loss

of morphology at low temperatures (900-1000 °C) suggests that the development of a

texture microstructure would be difficult using conventional sintering. The present

results are for loose powder samples. However, if a consolidation technique which

results in a textured green microstructure was used in combination with a novel multi-

step sintering approach, similar to those proposed by Polotai et al.28 and Chen and

Wang29, a textured fired microstructure could possibly be obtained. The multi-step

sintering approaches use a low temperature isothermal final step to eliminate pores

without grain growth. If the final step were below 1000 °C it is possible that sintering

with limited morphological change could be obtained.

3.4 Conclusions

Nanotabular BaTiO3 particles were synthesized using a hydrothermal route. The

particles are single crystal with a majority having a [111] zone axis, and have a median

thickness of 5.8 ± 3.1 nm and a face diameter of 27.1 ± 12.3 nm, as determine by atomic

force microscopy. Morphology of the particles was shown to be controlled solely by pH

of the solution during synthesis. It is speculated that the high solution pH stabilizes the

{111} face limiting growth in the <111> direction and leading to multiple {111} twin

76

formation during synthesis. With growth limited in the <111> direction the particle

develop a plate-like morphology. The powder has a low concentration (0.5wt %) of

hydrothermal defect which coalesce to form internal defects when the powder is heated to

375°C, as observed by TEM. Solid bridging and the onset of sintering is observed at 800

°C with a complete loss of plate-like morphology by 1000 °C.

77

References

1. A.J. Moulson and J.M. Herbert, Electroceramics: Materials, properties, applications, 1st ed. (Chapman & Hall, London, 1990). 2. D.F.K. Hennings, B.S. Schreinemacher, and H. Schreinemacher: Solid-state preparation of BaTiO3-based dielectrics, using ultrafine raw materials. J. Am. Ceram. Soc. 84, (12), 2777 (2001). 3. R.A. Kimel, V. Ganine, and J.H. Adair: Double injection synthesis and dispersion of submicrometer barium titanyl oxalate tetrahydrate. J. Am. Ceram. Soc. 84, (5), 1172 (2001). 4. S. Wada, T. Tsurumi, H. Chikamori, T. Noma, and T. Suzuki: Preparation of nm-sized BaTiO3 crystallites by a LTDS method using a highly concentrated aqueous solution. J. Cryst. Grow. 229, (1), 433 (2001). 5. B.K. Kim, D.Y. Lim, R.E. Riman, J.S. Nho, and S.B. Cho: A new glycothermal process for barium titanate nanoparticle synthesis. J. Am. Ceram. Soc. 86, (10), 1793 (2003). 6. H. Kamiya, K. Gomi, Y. Iida, K. Tanaka, T. Yoshiyasu, and T. Kakiuchi: Preparation of highly dispersed ultrafine barium titanate powder by using mircobial-derived surfactant. J. Am. Ceram. Soc. 86, (12), 2011 (2003). 7. P. Gherardi and E. Matijevic: Homogenous precipitation of spherical colloidal barium titanate particles. Colloid Surface 32, (3-4), 257 (1988). 8. J.H. Adair and E. Suvaci: Submicron electroceramic powder by hydrothermal synthesis. In Encyclopedia of Materials: Science and Technology, edited by K.H.J. Buschow, R.W. Cahn, M.C. Flemings, B. Ilschner, E.J. Kramer, and S. Mahajan, (Elsevier Science Ltd., 2001) pp 8933. 9. D.O. Yener, T.J. Yosenick, C.A. Randall, and J.H. Adair: Synthesis of nanosized Ag/Pd platelets in self-assembled bilayers, and thin film metallization by electrophoretic depositions. To be submitted, (2005). 10. D.O. Yener, A.H. Carim, and J.H. Adair: submitted to J. Phys. Chem., (2004). 11. J. Moon, M.L. Carasso, H.G. Krurup, J.A. Kerchner, and J.H. Adair: Particle-shape control and formation mechanisms of hydrothermally derived lead titanate. J. Mater. Res. 14, (3), 866 (1999). 12. S.B. Cho, M. Oledzka, and R.E. Riman: Hydrothermal synthesis of acicular lead zirconate titanate (PZT). J. Cryst. Grow. 226, (2-3), 313 (2001).

78

13. R.B. Bagwell, J. Sindel, and W. Sigmund: Morphological evolution of barium titanate synthesized in water in the presence of polymeric species. J. Mater. Res. 14, (5), 1844 (1999). 14. L. Zhao, A.T. Chen, F.F. Lange, and J.S. Speck: Microstructural development of BaTiO3 powders synthesized by aqueous methods. J. Mater. Res. 11, (6), 1325 (1996). 15. D.V. Miller: Synthesis and properties of barium titanate nanocomposites. Ph.D. Thesis, The Pennsylvania State University, University Park, PA, (1991). 16. D.F.K. Hennings and S. Schreinemacher: Characterization of hydrothermal barium titanate. J. Euro. Ceram. Soc. 9, 41 (1992). 17. Y. Yuan, T.J. Yosenick, and J.H. Adair: Unpublished results. (2004). 18. K.M. Hung, W.D. Yang, and C.C. Huang: Preparation of nanometer-sized barium titanate powders by a sol-precipitation process with surfactants. J. Euro. Ceram. Soc. 23, 1901 (2003). 19. H. Schmelz and H. Thomann: Twinning in BaTiO3 ceramics. Ceram. For. Intern. 61, 199 (1984). 20. P. Hartman and W.G. Perdok: On the relations between structure and morphology of crystals. Acta. Cryst. 8, 49 (1955). 21. T. Tani, Z. Xu, and D. Payne, Thin Ferroelectric Films III, edited by E. Meyers, B.A. Tuttle, S.B. Desu, and P.K. Lauser (in Mater. Res. Soc. Symp. Proc., 310,Pittsburgh, PA, 1993), pp 269. 22. M.M. Lencka and R.E. Riman: Thermodynamic modeling of hydrothermal synthesis of ceramic powders. Chem. Mater. 5, (1), 61 (1993). 23. J.P. Remeika: Method of growth barium titanate single crystals. J. Am. Ceram. Soc. 76, 940 (1954). 24. J.W. Nielsen, R.C. Linares, and S.E. Koonce: Genesis of the barium titanate butterfly twin. J. Am. Ceram. Soc. 45, (1), 12 (1962). 25. D.F.K. Hennings, C. Metzmacher, and B.S. Schreinemacher: Defect chemistry and microstructure of hydrothermal barium titanate. J. Am. Ceram. Soc. 84, (1), 179 (2001). 26. S.W.L. Lu, B.I. Lee, and L.A. Mann: Carbonation of barium titanate powders studied by FT-IR technique. Mater. Lett. 43, 102 (2000). 27. J.K. Lee, K.S. Hong, and J.W. Jang: Roles of Ba/Ti ratios in the dielectric properties of BaTiO3 ceramics. J. Am. Ceram. Soc. 84, (9), 2001 (2001).

79

28. A.V. Polotai, K. Breece, E. Dickey, C.A. Randall, and A.V. Ragulya: A novel approach to sintering nanocrystalline barium titanate ceramics. J. Am. Ceram. Soc. 88, (11), 3008 (2005). 29. I.W. Chen and X.H. Wang: Sintering dense nanocrystalline ceramics without final-stage grain growth. Nature 404, 168 (2000).

80

CHAPTER FOUR

Aqueous Surface Chemistry of Hydrothermally Derived BaTiO3 Nanoparticles

4.1 Introduction

Barium titanate is a complex metal oxide consisting of two chemically differing

end members; a relatively insoluble acidic TiO2 and a highly soluble basic BaO. The

nature of BaTiO3 complicates the interfacial chemistry. Figure 4.1 gives the ideal

stability fields for the BaTiO3-H2O-CO2 system in an aqueous environment,1, 2 and a

TEM image of a BaTiO3 particle allowed to equilibrate in aqueous suspension at pH 6.5.

The stability of BaTiO3 is highly pH dependent. As pH decreases, the solubility of the

Ba2+ is dramatically increased. This large change in solubility is not observed in the

dissolution of Ti from the BaTiO3 surface.3 Thus, the BaTiO3 surface dissolves

incongruently, yielding a Ti-rich surface layer.4, 5 The TEM image shows the amorphous

TiO2 surface of the Ba2+ depleted BaTiO3 surface. The presence of the Ti-rich surface is

an issue because during sintering, local Ti-rich regions act to form low temperature

melting phases and promote exaggerated grain growth.6

Figure 4.1 shows that in addition to Ba2+ dissolution, the formation of BaCO3 is

also a concern. CO2 in the atmosphere readily dissolves into water, and at neutral pH

where Ba2+ dissolves from the surface, BaCO3 can form.7, 8 During sintering the high

temperature (>1200°C) decomposition of BaCO3 can lead to decreased density because

of CO2 gas evolution in closed pores.9 The gas exerts a negative sintering pressure which

inhibits and eventually reverses the sintering process. Because of the deleterious effects

81

Figu

re 4

.1.

Idea

l sol

ubili

ty d

iagr

am fo

r BaT

iO3-

H2O

-CO

2sy

stem

from

Ben

dale

et a

l.2B

a2+di

ssol

utio

n is

favo

red

at lo

w p

H.

As

pH in

crea

ses,

Ba2+

solu

bilit

y de

crea

ses u

ntil

the

prec

ipita

tion

of B

aCO

3is

favo

red.

The

TEM

imag

e of

a B

aTiO

3pa

rticl

e tre

ated

in

wat

er a

t pH

6.5

show

s the

pre

senc

e of

an

amor

phou

s TiO

2su

rfac

e la

yer.

Ba2+

depl

eted

am

orph

ous

TiO

2S

urfa

ce

82

of Ba2+ dissolution and BaCO3 formation it is important to understand the changes in the

interfacial chemistry of BaTiO3 as function of solution pH, such that processing methods

can be developed to overcome aforementioned issues.

Current developments in the synthesis of BaTiO3 have resulted in better

morphological control during synthesis; BaTiO3 wires10, tubes11, hexapods12 and

platelets13 have all been recently reported. Each particle morphology exhibits a specific

crystallographic surface habit. The effect of surface crystal habit on zeta potential and

point of zero charge (PZC) has been extensively studied in TiO2 materials.14-19 These

studies verify that particle morphology contributes to the surface chemistry and

electrokinetic properties of these new BaTiO3 particles. In addition to instability of the

BaTiO3 surface in aqueous environments, an understanding of the surface chemistry of

anisotropic BaTiO3 materials and its affects on the dispersion, and downstream

processing is needed because of the advantages that these new BaTiO3 particles are

expected to exhibit toward the manufacturing of electronic components.

There are four objectives in the current study: (1) to describe the surface

chemistry of BaTiO3 using the MUSIC model as a function of the crystal structure and

morphology, (2) use dissolution data to determine the concentration of Ba2+ in solution as

a function of pH and surface area (3) to account for the adsorption of Ba2+ using a

modified Stern isotherm, and (4) the precipitation of BaCO3 on the BaTiO3 surface.

83

4.2 Experimental Observations of BaTiO3 Surface Charging in an Aqueous

Environment

Extensive studies have shown that the interfacial chemistry of BaTiO3 is highly

dependent on the solution pH and the suspension solid loading.4, 5, 20, 21 Theories have

been presented to explain the experimental observations which have been limited to

descriptions of the surface in specific pH ranges.

4.2.1 Acidic pH – Amorphous TiO2 Surface

The instability of the BaTiO3 in an aqueous environment is well-documented. 4, 5,

20, 21 Several studies have shown that in acidic environments that Ba2+ dissolves from the

surface and resides in the solution phase of aqueous suspensions. The TEM image in

Figure 4.1 shows the presence of an amorphous surface layer on a Ba2+ depleted BaTiO3

particle. Blanco-Lopez et al.4 theorized that since rutile is the thermodynamically stable

form of TiO2 under the acidic conditions that the BaTiO3 surface should act “TiO2-like”.

Paik et al.5 experimentally observed that below pH 5 suspensions of BaTiO3 behave

similar to rutile suspensions.

Since the BaTiO3 particle surface is similar to TiO2 in acidic environments it is

possible to use models present in the literature to describe the surface charge of BaTiO3

in acidic environments. The multisite complexation (MUSIC) model is the most

developed of the models and accounts for the crystal chemistry of the material in the

determination of the surface sites that provide surface charging in solution.14, 15 The

MUSIC approach provides a good starting point to investigate the surface chemistry of

BaTiO3 because of the approach has been widely use to describe one of end member of

the BaTiO3 solid solution, TiO2.

84

The MUSIC model was developed and refined by Hiemstra et al.14, 15 and

Machesky et al.22 and is based on the assumption that the oxygen coordination at the

surface is different from that of the lattice. The difference in coordination affects the

valence of the oxygen at the surface and the resulting reaction with protons that produce

surface charge. Each broken surface oxygen-cation bond has a valence defined by

Pauling, who stated that the charge of the centrally coordinated oxygen is distributed

equally over all of the neighboring cations.23 The bond valence is calculated using the

following equation,

CNzv = [4.1]

where z is the charge of oxygen and CN is the oxygen’s coordination number. For

example, in TiO2, O has a valence of 2- and is coordinated by 3 Ti leading to an average

bond valence of 2/3- for each O-Ti bond.

Pauling’s bond valence23 model assumes that the oxygen valence is equally

shared among all of the coordinated bonds. Actually, the bond valences are dependent on

the bond length. For materials with anisotropic crystal structures the difference in bond

valence is important. Several approaches have been used to calculate the difference in

bond valence due to the crystal structure.18, 22 Once the surface site valence is known the

reaction of the surface groups with the solvent are determined. One common approach,

used by Fedkin18, is to use a periodic density method to generate an ideal surface which is

allowed to relax. After the relaxation, the new bond lengths are measured and the new

bond valence is calculated.

The primary limitation of the MUSIC model to describe the behavior of a multi-

component metal oxide is the inability to account for the presence of specific adsorbates

85

or other phases that may be present. In fact, the surface chemistry of BaTiO3 in an

aqueous environment is controlled by the dissolution and specific adsorption of Ba2+ at

low to moderate solution pH (pH < 9).21, 24 The dissolution of Ba2+ would not be a

problem in the modeling of the surface chemistry if the dissolved barium remained in

solution. Resulting in the surface behaving similar to a TiO2 surface.21 However, the

barium does not remain in solution, but in fact specifically absorbs onto the surface.5, 25

Therefore the use of MUSIC model is limited to acidic environments in the description of

the BaTiO3 surface.

4.2.2 Neutral pH – Ba2+ Adsorption

Several researchers have noted the zeta potential and isoelectric point (IEP) of

aqueous BaTiO3 suspensions are dependent on the solid loading and surface area of the

powder.5, 20 The changes in the magnitude of the zeta potential and shift in the IEP were

attributed to the dissolution/adsorption of Ba2+(aq) at the particle surface. This led to the

hypothesis that at intermediate suspension pH the BaTiO3 surface behaves as a TiO2

surface that specifically adsorbs Ba2+(aq).

Malati and Smith26 and Fuerstenau and co-workers27, 28 studied the adsorption of

alkaline earth cations, including Ba2+(aq), on TiO2. Malati and Smith26 studied the

adsorption of Ba2+(aq) on both rutile and anatase at pH 7 and found that adsorption

followed a Langmuir-type isotherm at solution pH values greater than the PZC of the

surface. Jang and Fuerstenau28 performed a more detailed study of the adsorption onto

the surface of rutile. The adsorption of Ba2+(aq) was consistent with a modified Stern

adsorption isotherm. Jang and Fuerstenau confirmed that Ba2+(aq) adsorption was

negligible below the PZC of the surface.

86

Jang and Fuerstenau used electrokinetic and potentiometric titrations to perform a

detailed study of the adsorption of Ba2+ on the surfaces of rutile. Two models were

proposed for the adsorption of Ba2+ at the TiO2 surface: (1) a monodentate complex

where a Ba2+ reacts with a single TiOH(surf) group according to the following reaction,

TiOH(surf) + Ba2+ TiOBa+(surf) + H+ [4.2]

Or, (2) a bidentate complex where a Ba2+ reacts with two TiOH surface groups,

2TiOH + Ba2+ (TiO)2Ba + 2H+ [4.3]

By measuring the ratio of protons released from the surface for each Ba2+ and modeling

the adsorption density as a function of pH it was shown that the bidentate complex is the

predominant adsorption mechanism.

In describing Ba2+ adsorption, Jang and Fuerstenau found that the Stern isotherm

was not valid. A Stern isotherm is a modified Langmuir isotherm that assumes each

adsorption site has a charge equal to the valence of the adsorption ion. One assumption

of the Langmuir isotherm is that the probability of site occupancy is equivalent for all

sites regardless of adsorption density (i.e. that nearest neighbor interactions do not affect

the adsorption of subsequent adsorbates). For adsorption due to electrostatic attraction

this assumption is not valid.29 Levine et al. modified the Stern isotherm to permit for

increased electrostatic repulsion as adsorption density increases.29 In describing Ba2+

adsorption on a rutile surface Jang and Fuerstenau used the Levine modification of the

Stern isotherm.

4.2.3 Basic pH – BaCO3 Formation

In alkaline environments the dissolution of Ba2+ from BaTiO3 is minimal, but the

precipitation of BaCO3 becomes an issue, as shown in Figure 4.1. Above pH 13, BaCO3

87

is the stable phase of Ba with the onset of precipitation depending on the concentration of

total CO2 species in solution. Several studies have shown that BaCO3 readily forms on

the surface of BaTiO3 at high solution pH in aqueous environments.30-34 Thus, the

electrokinetic properties of BaTiO3 in water also dependent on the presence of BaCO3.35

Describing the aqueous surface chemistry of BaTiO3 it is necessary to accommodate for

the dissolution and subsequent adsorption of barium as well as the formation of BaCO3

on the particle surface at elevated pH conditions

The surface chemistry of BaCO3 is controlled by a Nernst-Gouy-Stern mechanism

with the potential determining ions (PDIs) for the system being Ba2+(aq) and CO3

2-(aq).35, 36

In an aqueous solution, if no atmospheric control is maintained, the total dissolved

carbonate concentration is fixed by the ambient pCO2 of the surrounding environment,

and unlike the concentration of Ba2+, is independent of solution pH. With a fixed

carbonate concentration, the charging of the surface primarily depends on the change in

barium concentration as a function of solution pH.

4.3 Materials and Methods Suspensions of a commercially available hydrothermally derived powder (BT-08,

Cabot Performance Materials, Boyertown, PA) were prepared in DI water at four solid

loadings, 0.5, 2, 6, and 20 weight percent (wt%) (40, 160, 480, 1600 m2/L). Solution pH

was adjusted using 0.1 or 0.01 M solutions of nitric acid (70wt%, Fisher Scientific, Fair

Lawn, NJ) or tetraethylammonium hydroxide (TEAOH) (35wt% in water, Aldrich

Chemical Company, Milwaukee, WI) prior to adding the powder. The suspensions were

allowed to equilibrate for 24 hours. Electrophoretic mobility of the suspension was

measured using electrophoretic light scattering (ZetaPALS, Brookhaven Instrument Corp,

88

Holtsville, NY). The suspensions were centrifuged and the supernatant passed through a

0.22 μm filter. The filtered supernatants were analyzed for dissolved Ba and Ti using

indirectly coupled plasma atomic emission spectroscopy (ICP-AES) (PS3000UV,

Leeman Labs, Los Angles, CA) Standards were prepared by serial dilution from 1000

ppm stock solutions (Hi Purity Standard, Charleston, SC).

The zeta potential of 1wt% suspensions of two different BaTiO3 powders, an

anisotropic, hydrothermal platelet powder synthesized using a procedure previously

described13, and a commercial hydrothermal equiaxed powder (BT-01, Sakai Chemical

Company, Osaka, Japan) were measured in a co-solvent of 95wt% ethanol and 5wt% DI

water (95/5 EW) using electrophoretic light scattering. The platelet powder has a median

face diameter of 27.1 ± 12.3 nm and a median thickness of 5.8 ± 3.1 nm with the (111)

plane as the large face. The spherical powder has a BET surface area of 11.9 m2/g and a

pycnometery density of 5.67g/cm3, yielding an equivalent spherical diameter of 88.9 nm.

The pH was measured using an ion-selective field effect transistor (IS-FET) pH probe

(Sentron Hotline Probe, RL Instruments, Manchoung, MA). The probe was calibrated

using aqueous-based NIST-traceable pH standard with nominal values of pH 4, 7, and 10.

The pH was adjusted to pH 3, 5, 7, 9, and 11 using 0.1 M and 1 M solutions of

hydrochloric acid (37%, J.T. Baker, Phillipsburg, NJ) and tetramethylammonium

hydroxide (TMAOH) (Aldrich Chemical Company, Milwaukee, WI) in 95/5 EW. The

pH values of the suspensions prepared in the 95/5 EW were corrected by subtracting the

residual junction potential present due to aqueous based calibration process. The value

for the residual junction potential (1.05 pH units) was obtained from Table 5.7 in

89

Popovych and Tomkins.37 The concentration of Ba and Ti in solution was analyzed using

ICP-ES.

Transmission electron microscopy (TEM) (2010F, JEOL, Japan) was used to

collect image of the Sakai BT-01 powder. TEM analysis was performed using holey

carbon film on Cu grids as sample holders with a single drop of dilute suspension on a

grid.

4.4 Results and Discussion

4.4.1 Acidic pH – Amorphous TiO2 Surface

4.4.1.1 Determination of BaTiO3 Surface Groups

When a metal oxide is present in aqueous suspension the surface of the material is

fully hydrated.38 In the MUSIC model the chemical reactions that provide surface charge

are based on the reactivity of the surface hydroxides, which are anchored at the oxygen

surface sites on the material. The valence and number of orbitals available for proton

uptake therefore are determined by the change of cation coordination of the surface

oxygen compared to oxygen in the lattice. To apply the MUSIC model to any surface it

is first necessary to determine the structure and cation coordination of the lattice and

surface oxygen.

Because TiO2 has been extensively modeled and is a component of the BaTiO3

solid solution TiO2 is a first approximation to use in the modeling of the BaTiO3 surface.

The two polymorphic forms of TiO2, rutile and anatase, are based on building blocks of

TiO6 octahedra that share edges and faces. This results in lattice oxygen being

coordinated by three Ti. In the analysis of the rutile and anatase surfaces Hiemstra et al.

90

identified three surface oxygen sites that are singly (OI(a)), doubly (OII

(a)), or triply (OIII(a))

coordinated with Ti, which have the ability to react with two, one, or zero protons,

respectively.15 In the notation the superscript numerals denote the Ti coordination of the

surface oxygen while the subscript character is included to differentiate between oxygen

surface sites based on the crystal structure of TiO2 and BaTiO3, which will be discussed

later. The ability to react with protons is related to the difference in Ti coordination of

the oxygen on the surface compared to oxygen in the lattice. Missing Ti on the surface

are compensated by the adsorption of protons from solution. Oxygen in the lattice of

rutile and anatase are coordinated by three Ti. Because of this, the three possible surface

groups on rutile and anatase can be viewed as two, one, or zero Ti deficient, where the Ti

deficiency is equal to the number of protons available for uptake by the surface group.

BaTiO3 is also based on TiO6 octahedra which share corners. This difference in

the crystal structure leads to a decrease in Ti coordination from three to two for lattice

oxygen in BaTiO3. The change in Ti coordination leads to singly (OI(b)) and doubly

(OII(b)) coordinated oxygen surface sites. The primary difference between TiO2 and

BaTiO3 is the presence of the basic BaO end member. However, as noted the dissolution

of Ba2+ at low pH leads to a TiO2 surface and therefore it is not necessary to consider the

Ba coordination of the oxygen in the analysis.

To aid in the analysis of the platelet and equiaxed BaTiO3 the powders, it was

assumed that the local crystal structure is cubic and not tetragonal. Therefore it is not

necessary to calculate the difference in bond length of the different surface sites on

tetragonal BaTiO3 because of the simplifying assumption that all of the sites have equal

bond length. For materials in an aqueous solution the surfaces are fully hydrated and

91

surface relaxation occurs. The relaxation produces bond length changes that according to

the MUSIC model, changes the valence on the oxygen sites, and shift the pK value of the

specific site type. With the dissolution of Ba2+, the surface structure of BaTiO3 becomes

amorphous as shown in Figure 4.1. Without full characterization of the amorphous TiO2

surface it is not possible to fully determine the bond length changes. However, as an

approximation is was assumed that the bond length did not change due to the dissolution

of Ba2+. Only OI(b) and OII

(b) sites are expected from the BaTiO3 crystal structure, but to

account for possible relaxations due to Ba2+ dissolution the presence of OI(a), OII

(a), and

OIII(a) sites are included in the analysis.

Due to the presence of multiple surface sites on particle surfaces it is nearly

impossible to experimentally determine the pK value of any particular surface site. In

developing the MUSIC model Hiemstra derived an empirical equation to calculate the log

Kc,m of an arbitrary surface group15,

log K = -A(ΣSMe-O + mSH + n(1 - SH) + V) [4.4]

where A is a constant with a value 21.7, and is based on the linear regression of various

log K values of unconstrained (i.e. free in solution) protonation reactions22, ΣSMe-O is the

sum of bond valence (SMe-O) of the metal-oxygen bonds, SH is the bond valence for an

associated surface proton, m is equal to the number of orbitals available to coordinate to a

proton, and n is equal to the number of filled orbitals, and V is the valence of oxygen, 2-.

Due to hydrogen bonding of water molecules with the ionized surface group, the bond

valence of the hydrogen-oxygen bond is not equal to the charge on the proton. Hiemstra

et al. determined SH to be equal to 0.80.

92

Table 4.1 is a list of all possible surface groups, reactions, and theoretical log K

values, calculated using Equation 4.4, for a TiO2 surface on BaTiO3 predicted by the

MUSIC model. The chemical structure and protonation reactions of the surface groups

are included to illustrate the chemical change to the surface site as a function of solution

pH. Table 4.1 includes surface groups based on oxygen coordinated by 2 and 3 Ti for a

total of 19 possible surface reactions. However, the six reactions shown in bold typeface

are the only reactions that have pK vales in the normal pH range of pH 1-14, and are the

only reactions that were used in the modeling of surface charge development.

Of interest is the differences in the log K values for identical surface sites, for

example, OII(a) in Table 4.1 The difference arises from the number of possible adsorbed

protons for each site. For OII(a) if it is assumed only one proton can be adsorbed, log K

equals 10.2. However, if the adsorption of two protons is assumed, log K decreases to

5.9. Machesky et al. noted in the refinement of the MUSIC model that depending on

surface relaxations and nearest neighbor interactions that the number of possible proton

per site can change.22

4.4.1.2 Combination of MUSIC and Gouy-Chapman Models

The power of the MUSIC model lies in the ability to calculate the pK values for

several different surface reactions based solely on the crystallography of the material.

From the pK values it is possible to predict the PZC of the material. However, to

calculate the surface potential of the particle as a function of solution pH, it is necessary

to describe the diffuse part of the double layer using a Gouy39-Chapman40 model.

Healy and White developed the ionizable surface group model41 to describe

surface chemical reactions in the development of surface charge. The model assumes a

93

Tab

le 4

.1.

Poss

ible

surf

ace

grou

ps o

f Ba2+

depl

eted

BaT

iO3

in a

n ac

idic

env

ironm

ent w

ith th

e as

soci

ated

pro

tona

tion

reac

tions

an

d ca

lcul

ated

log

K v

alue

s. T

he r

eact

ions

in b

old

are

the

only

rea

ctio

ns th

at o

ccur

in th

e no

rmal

pH

ran

ge (1

-14)

and

thos

e us

ed in

the

calc

ulat

ion

of th

e su

rfac

e ch

arge

. P

rimar

y S

urfa

ce G

roup

Che

mic

al S

truct

ure

Pro

tona

tion

Rea

ctio

SM

e-O

mn

log

K

OII

I (a)

...O

0...

O0 +

H+ →

...O

H+1

20

1-4

.3O

II (a)

...O

-2/3

...O

-2/3

+ H

+ → ..

.OH+1

/31.

330

110

.2O

II (a)

...O

H+1

/3...

OH

+1/3

+ H

+ →...

OH

2+4/3

1.33

10

-2.8

OII (a

)...

O-2

/3...

O-2

/3 +

H+ →

...O

H+1/3

1.33

02

5.9

OII (a

)...

OH

+1/3

...O

H+1

/3 +

H+ →

...O

H2+4

/31.

331

1-7

.2O

II (a)

...O

H2+4

/3...

OH

2+4/3

+ H

+ → ..

.OH

3+7/3

1.33

20

-20.

2O

I (a)

...O

-4/3

...O

-4/3

+ H

+ → ..

.OH

-1/3

0.66

02

20.4

OI (a

)...

OH

-1/3

...O

H-1

/3 +

H+ →

...O

H+2

/30.

661

17.

4O

I (a)

...O

H2+2

/3...

OH

2+2/3

+ H

+ → ..

.OH

+4/3

0.66

20

-5.6

OI (a

)...

O-4

/3...

O-4

/3 +

H+ →

...O

H-1

/30.

660

316

.1O

I (a)

...O

H-1

/3...

OH

-1/3

+ H

+ → ..

.OH

2+2/3

0.66

12

3.0

OI (a

)...

OH

2+2/3

...O

H2+2

/3 +

H+ →

...O

H3+5

/30.

662

1-1

0.0

OI (a

)...

OH

3+5/3

...O

H3+5

/3 +

H+ →

...O

H4+8

/30.

663

0-2

3.0

OII (b

)...

O0

...O

0 + H

+ → ..

.OH

+12

01

-4.3

OI (b

)...

O-1

...O

-1 +

H+ →

...O

H0

10

117

.4O

I (b)

...O

H0...

OH

0 + H

+ → ..

.OH

2+11

10

4.3

OI (b

)...

O-1

...O

-1 +

H+ →

...O

H01

02

13.0

OI (b

)...

OH

0...

OH

0 + H

+ →...

OH

2+11

11

0.0

OI (b

)...

OH

2+1...

OH

+1 +

H+ →

...O

H3+2

12

0-1

3.0

(a) a

nd (b

) den

ote

diffe

renc

e in

the

Ti c

oord

inat

ion

of th

e nu

etra

l sur

face

site

bas

ed o

n th

e cr

ysta

l stru

ctur

eS

hade

d re

gion

s de

note

a c

hang

e in

num

ber o

f ava

ilabl

e pr

oton

for u

ptak

e id

entic

al p

rimar

y su

rface

gro

upΣ

SM

e-O =

Bon

d va

lenc

e su

m o

f the

cat

ion-

oxyg

en b

onds

(SM

E-O

)m

= #

of o

rbita

ls a

vaila

ble

to u

ptak

e a

prot

onn

= #

of o

rbita

ls fi

ll w

ith a

pro

ton

3 Ti

Coo

rdin

ated

Oxy

gen

- TiO

2

2 Ti

Coo

rdin

ated

Oxy

gen

- B

aTiO

3

94

mass-action equation between surface sites with protons at the surface provided from the

solution. The dissociation constant of a surface site can be calculated using the law of

mass action for the general surface reaction,

AH(s) A-(s) + H+

(surf) [4.5]

][

][)(

AH

AaK surfH

eq

−+

= [4.6]

where aH+(surf) is the activity of a proton at the surface which is related to the activity of

the proton in the bulk solution via,41

)exp()( kTe

aa sHsurfH

ψ−= ++ [4.7]

where aH+ is the proton activity in the bulk solution, e is the charge on the electron, ψs the

surface charge, k is Boltzmann’s constant, and T is absolute temperature. Knowing the K

values and surface proton concentration it is possible to calculate the fraction (α) of

protonated and unprotonated surface sites by the following equations,

1][1

,, +

= +smc

mc HKα [4.8]

1][][

,

,1, += +

+

+smc

smcmc HK

HKα [4.9]

The value of α ranges from zero to one depending on solution pH.

The charge at the surface due to chemical reactions is counter balanced in the

solution by a Gouy-Chapman diffuse layer to maintain electroneutrality. This approach

has been used in the current work. The advantage of the ionizable surface group model is

that by relating the chemical reaction of the surface site with the solution chemistry,

mainly solution pH and related equilibrium, the surface potential can be calculated over a

95

pH range. A simplified algorithm and flow diagram for using the MUSIC model to

determine the surface potential as a function of solution pH is present in Appendix A.

From the log K values in Table 4.1 it is possible to calculate the α values for the

surface site of BaTiO3 using Equation 4.8. Figure 4.2 shows the α for the four different

TiO2 sites on the BaTiO3 surface. The plot shows the pH dependence of each site and

throughout the pH range from pH 0 to 14 what surface reaction controls the surface

chemistry of the material.

The presence of several surface sites necessitates a determination of the effect of

each site for a given surface. The surface charge density of any surface can be

represented by the superposition of the effect of all separate sites on the surface,

∑=n

imicisis zNe ,ασ [4.10]

where Nsi is the surface density of site type i, and zi is the valence of the site i.

Given the surface charge density the Gouy-Chapman model can be used. To

maintain electroneutrality, the surface charge density in the solution is equal and opposite

that of the surface change density,

σs = -σdl [4.11]

Because the solution is treated as a continuous dielectric with ions as point charges,

Poisson’s equation is the fundamental electrostatic equation governing the system,

r

dl

OS ε

πσπε

ψ4

412 ⋅⎟⎟

⎞⎜⎜⎝

⎛−=∇ [4.12]

Using the Debye-Hückel approximation42 that ψS ≤ 25 mV then Equation 4.18 simplifies

to a linear differential equation with σdl given by,

96

Figu

re 4

.2.α

valu

es fo

r fou

r diff

eren

t sur

face

gro

ups o

n B

aTiO

3as

a fu

nctio

n of

susp

ensi

on p

H. α

repr

esen

ts th

e de

gree

of

prot

onat

ion

of a

ll of

the

spec

ific

surf

ace

site

s pre

sent

on

the

surf

ace.

For

exa

mpl

e, b

elow

pH

13

all o

f the

…O

-1gr

oups

hav

e an

as

soci

ated

pro

ton,

but

abo

ve p

H 1

4 al

l of t

he g

roup

s are

dep

roto

nate

d. O

ver t

he e

ntire

pH

rang

e on

e of

the

four

reac

tions

is

cont

rolli

ng th

e su

rfac

e ch

argi

ng o

f BaT

iO3.

97

⎟⎠⎞

⎜⎝⎛−=

kTe

ekT sor

dl 2sinh

πκεεσ [4.13]

with,

32

108 xkTIeN

or

A

εεπ

κ −= [4.14]

where εr is the dielectric permittivity of the solvent, εo is the permittivity of free space, k

is Boltzmann’s constant, T is absolute temperature, e is the charge on the electron, ψs is

the surface charge, NA is Avagadro’s number, and I is the ionic strength. 1/κ known as

the Debye length of the electrical double layer and is the distance from the Stern plane to

where the potential drops by a factor of 1/e.

Since both the surface and solution charge density are known the surface charge,

ψs, can be solved for. Combining Equations 4.10, 4.11, and 4.13 yields the following

expression,

⎟⎠⎞

⎜⎝⎛=∑ kT

eekTSNe so

n

imicisi 2

sinh2,

ψπκεεα [4.15]

The surface charge of the particle at a specific pH can be calculated by solving 4.15 using

the log K values derived from the MUSIC model to calculate the α values for each type

of surface group to create a self-consistent array of values for a given ψs value.

For uniquely shaped particles, the surface charge as a function of habit is required

to calculate the overall ψs value. Each crystallographic plane has a different combination

of site type and site density at the solid-solution interface and therefore will have a

unique surface charge. To account for the morphology of the particle the calculation of

the total particle surface charge is based on the fractional area of each plane present in

98

solution. Equation 4.15 can be solved for each specific plane present in solution. Then

the surface charge of the overall particle (ψtot) is a weighted sum of the surface charge of

the ith specific habit (ψh,i) and the area fraction (fh,i) of each plane present to solution:

∑=n

iihihtot f ,,ψψ [4.16]

4.4.2 Neutral pH – Ba2+ Adsorption

By combining an ionizable surface group model with a Gouy-Chapman diffuse

layer the intrinsic contribution from the native surface groups to surface charge and

potential can be described. However, previous research shows that the

dissolution/adsorption of Ba2+ influences the surface chemistry at moderate solution pH.

At low pH and solid loading it has been observed that the Ba2+ depleted surface of

BaTiO3 has electrokinetic properties similar to TiO2.21 As the pH increases the zeta

potential deviates from that of a TiO2 surface due to Ba2+ adsorption.5, 25 The similarities

of the specific absorption of Ba2+ on TiO2 therefore make it an ideal system to model the

depleted surface of BaTiO3 with reabsorbed Ba2+.

Malati and Smith found that near the PZC of rutile and anatase the adsorption of

Ba2+ followed a Langmuir isotherm26,

⎥⎦⎤

⎢⎣⎡ Δ−

=RTG

xnN abssii exp [4.17]

where Ni is the number of filled sites per area, ni is the number of possible surface sites

per area, xs is the mole fraction of adsorbate in solution, ΔGabs is the free energy of

absorption, R is the universal gas constant, and T is absolute temperature. The Stern

99

absorption isotherm is a modification to the Langmuir isotherm that is used to calculate

the surface charge density, σo,

io zeN=σ [4.18]

where z is the valence of the absorbate, and e is the charge on the electron.

The problem with the fundamental Stern isotherm is that as the adsorption density

(Ni) increases the energy required to fill the next site increases due to electrostatic

repulsion as surface sites are filled. Levin et al. modified the Stern isotherm to account

for electrostatic repulsion between a ion in the bulk and the surface29,

( )( ) ⎥⎦

⎤⎢⎣⎡ −Φ−

= − RTze

nzeNnfpzeN oo

op

i

poi

oψσ

σ exp1 [4.19]

where,

( )[ ]RTpHHpa PZCoB −′+−=Φ φ [4.20]

where ΦB is the adsorption potential, p is the number of surface sites occupied by one

adsorbed ion, n is the number of ions per volume in the bulk of solution, n

B

o is the number

of water molecules per volume, and f is the activity coefficient of the ions in the bulk

solution. ΦBB is a factor which accounts for the change in chemical potential between an

ion in the bulk and at the surface as the pH varies from the PZC. pH’PZC refers to the

PZC of the surface in the presence of the adsorbed ion, φo and a are constants dependent

on the ion adsorbing, and ψo is the surface potential, and equal to -15.3 and 1.6,

respectively. If p = 1 and f = 1, Equation 4.19 become the Stern isotherm.

Jang and Fuerstenau used Levin’s modified isotherm to model the adsorption of

Ba2+ onto a rutile surface. Because it was found that a bidentate complex is the mode of

100

adsorption for an alkaline-earth cation on rutile, p =2. Since the Ba2+ concentration is

low it was assumed that f = 1 and Equation 4.19 reduced to,

( ) ⎥⎦⎤

⎢⎣⎡ −Φ

=− RT

eaeN

eN ooM

oi

oi ψ

σ

σ 2exp

5.552 2 [4.21]

where aM is the activity of a cation in the bulk.

Fuerstenau and co-workers found the adsorption density (Γδ) of Ba2+ to be

dependent on solution pH, and follows the Stern-Grahame equation,27

⎥⎦⎤

⎢⎣⎡ Δ−

=ΓRTG

rc absexp2δ [4.22]

where r is the radius of the adsorbed hydrated ion, 1.35 x 10-8 cm, and c is the

concentration of adsorbate in solution. When the solution pH is less than the IEP of the

TiO2 surface adsorption was negligible. From the adsorption density (Figure 5 in Ref 28)

the free energy of adsorption, ΔGabs, was calculated using Equation 4.22. It is then

possible to calculate the surface charge density, σo, as a function of Ba2+ concentration

based on the dissolution of BaTiO3 and ΔGabs both of which are dependent on the

solution pH. Using Equation 4.19 ψo can then be calculated as a function of pH.

4.4.3 Basic pH – BaCO3 Formation

Thermodynamic calculations2, 3 and the literature30, 32, 34 shows that as solution pH

increases BaCO3 precipitation occurs. Several researchers have noted that BaCO3

precipitation occurs preferentially on the BaTiO3 particle surface. 30, 32, 34 This results in

the electrokinetic properties of BaTiO3 being similar to BaCO3 in alkaline environments.

The surface chemistry of BaCO3 is regulated by a Nernst-Gouy-Stern charging

101

mechanism where Ba2+ and CO32- are the PDIs.35, 36 The general surface reaction of

barium carbonate is,

BaCO3(s) CO32-

(s) + Ba2+(surf) [4.23]

with an equilibrium constant of,35

9

3

23

101.5][

][2−

==+

xBaCO

COaK surfBa

eq [4.24]

the activity of the surface Ba2+ is related to the bulk activity through,

]exp[][ 22

kTe

Baa sBasurf

ψ−= +

+ [4.25]

Similar to the TiO2 surface site reactions, the α values of the barium carbonate surface

can be calculated using Equation 4.18. The α value is a function of Ba2+, which is in turn

a function of pH due to the changing solubility of the BaTiO3. The α value can then be

input into Equation 4.21 and the surface charge of BaCO3 as a function of solution pH

calculated.

4.4.4 Comparison of Experimental Results and Theoretical Calculations

The surface potential and zeta potential of a surface are not equal due to strongly

adsorbed species inside the shear plane at the particle surface. The experimental results

are based on electrophoretic mobility experiments which measure the charge at the shear

plane and not the particle surface. The current theoretical calculations are based only on

the surface reactions and do not account for the presence of strongly adsorbed species

inside the shear plane. In the current model, there is a region, between the Stern plane

and solid surface, which is assumed not to affect the diffuse outer layer, which is known

be incorrect. In spite of this reservation the model can be used to predict the dependence

of the zeta potential on pH.

102

Prior to modeling the surface charge of BaTiO3 an empirical formula for the

dissolution of Ba2+ from the surface of BaTiO3 as a function of solution pH is required to

generally describe Ba2+(aq) concentration as a function of solution pH and solid loading.

Figure 4.3 shows the Ba2+(aq) concentration as a function of solution pH and solid loading

from Chodelka.43 An increase in Ba2+(aq) concentration with decreasing solution pH and

increasing surface area present in solution is observed. The data was normalized to the

surface area present in solution and plotted as a function of solution pH shown in Figure

4.4. Normalization to surface area collapses the soluble Ba2+(aq) as a function of solution

pH and solid loading over a common line. A good linear fit was obtained and used to

calculate the Ba2+(aq) concentration in solution as a function of solution pH in the

modeling of the surface chemistry of BaTiO3.

Figure 4.5 shows the experimental zeta potential values and theoretical

calculation of the surface charge for a suspension with a solid loading of 40 m2/L. There

are three distinct regions as a function of solution pH: Region I, pH <IEP of TiO2, Region

II, IEP TiO2 < pH < onset of BaCO3 precipitation, and Region III, pH > BaCO3

precipitation. Figure 4.6 is a schematic representation of the interactions at the solid-

solution interface at each region.

In Region I, below the IEP of TiO2, Ba2+ dissolution is thermodynamically

favored, but adsorption is limited due to the electrostatic repulsion between the positively

charged Ba2+(aq) and the TiO2 surface. Therefore the surface behaves as TiO2 and

analysis using the MUSIC model is applicable because only the intrinsic surface sites

based on the crystal chemistry control the surface charge of the material. In applying the

MUSIC model only the presence of OIII(a) and OII

(b) groups in a 1:1 ratio are required to

103

Figu

re 4

.3.D

isso

lutio

n da

ta fo

r aqu

eous

BaT

iO3

susp

ensi

ons w

ith in

crea

sing

solid

load

ings

. A

s exp

ecte

d, B

a2+di

ssol

utio

n is

m

inim

ized

at h

igh

pH a

nd in

crea

ses w

ith in

crea

sing

surf

ace

area

pre

sent

in su

spen

sion

. D

ata

from

Cho

delk

a.43

104

Figu

re 4

.4.

Dis

solu

tion

data

for B

aTiO

3as

func

tion

of so

lutio

n pH

nor

mal

ized

for s

urfa

ce a

rea

pres

ent i

n so

lutio

n. A

goo

d lin

ear

fit o

f the

obs

erve

d an

d th

e em

piric

al e

quat

ion

was

use

d to

det

erm

ine

the

conc

entra

tion

of d

isso

lved

Ba2+

in th

e m

odel

ing

of su

rfac

e ch

emis

try o

f BaT

iO3.

r2=

0.93

7

105

Figu

re 4

.5.

Zeta

pot

entia

l of a

queo

us B

aTiO

3su

spen

sion

(40

m2 /L

) sho

win

g th

e th

ree

diff

eren

t reg

ions

of s

urfa

ce c

harg

e in

B

aTiO

3. R

egio

n I c

ontro

lled

by a

nat

ive

TiO

2su

rfac

e. T

he in

crea

se in

Reg

ion

II is

due

to th

e ad

sorp

tion

ofB

a2+(a

q)on

to th

e su

rfac

e Ti

O2

surf

ace.

The

dec

reas

e in

Reg

ion

III i

s due

to p

reci

pita

tion

ofB

aCO

3on

the

BaT

iO3

surf

ace.

r2

= 0.

884

106

BaT

iO3

TiO

2

O

OH

O

Ba2+

Ba2+

BaT

iO3

TiO

2

OH2+

OH2+

OHB

a2+B

a2+

BaT

iO3

TiO

2

BaCO3

CO32-

BaCO3

Ba2+

CO

32-

Figu

re 4

.6.

Sche

mat

ic sh

owin

g th

e ev

olut

ion

of th

e su

rfac

e ch

argi

ng m

echa

nism

as a

func

tion

of so

lutio

n pH

for B

aTiO

3in

an

aque

ous e

nviro

nmen

t. A

t low

pH

Ba2+

diss

olut

ion

lead

s to

a Ti

O2

surf

ace.

As t

he p

H in

crea

ses B

a2+(a

q)re

adso

rptio

n re

sults

in a

de

viat

ion

from

an

idea

l TiO

2su

rfac

e. I

n a

basi

c en

viro

nmen

t the

pre

cipi

tatio

n of

BaC

O3

on th

e su

rfac

e co

ntro

ls th

e su

rfac

e ch

arge

.

pH 3

pH 1

1

Reg

ion

IR

egio

n II

Reg

ion

III

107

describe the surface potential in Region I. The OII(b) groups are expected from the crystal

chemistry of BaTiO3, however, the OIII(a) groups are only possible if a relaxation occurs

in which the TiO6 octahedra share edges or faces. This relaxation supports the concept

that the dissolution of Ba2+ results in a surface similar to rutile or anatase.

In Region II, the solution pH is above the IEP of TiO2 with electrostatic attraction

between the Ba2+(aq) and TiO2 resulting in adsorption of the Ba2+

(aq) onto the TiO2 surface.

The adsorption of Ba2+(aq) on the TiO2 can be described using the modified Stern isotherm

proposed by Levine et al.29 As solution pH continues to increase the Ba2+(aq)

concentration decreases yet the surface potential continues to increase. The free energy

of adsorption increases due to increase electrostatic attraction between the Ba2+(aq) and

TiO2 surface and overcomes the decrease Ba2+(aq) concentration. The surface potential in

Region II is a balance between the negative TiO2 surface and the influence of adsorbed

Ba2+(aq).

As solution pH increases to Region III, BaCO3 precipitation occurs and the

charging becomes dependent on the surface charging mechanism of BaCO3. BaCO3

develops surface charge via a Nernst-Gouy-Stern mechanism with Ba2+(aq) and CO3

2-(aq)

being the PDI. Therefore, as the Ba2+(aq) concentration continues to decease and the

surface charge becomes more negative.

To address the expected difference in surface chemistry due to differences in

morphology the zeta potential of an equiaxed and platelet powder was measured in a

solvent mixture, 95wt% ethanol/5wt% water (95/5 E/W). The solvent mixture was

chosen to limit Ba2+ dissolution and measure the intrinsic BaTiO3 surface reactions.

Table 4.2 is list of the ICP data for the dissolution of BaTiO3 in 95/5 E/W. Ba2+

108

Sol

utio

n pH

Pla

tele

t (p

pm)

Equ

iaxe

d (p

pm)

27

<.5

4<.

5<.

56

<.5

<.5

8<.

515

810

6566

123

N/A

Tab

le 4

.2.

ICP-

ES re

sults

for t

he d

isso

lutio

n of

Ba2+

in 9

5/5

etha

nol/w

ater

mix

ture

s. T

he d

ata

show

s tha

t dis

solu

tion

is li

mite

d un

til p

H 8

whe

n a

max

imum

con

cent

ratio

n of

10-3

M is

obs

erve

d. B

ecau

se o

f the

dis

solu

tion

it is

nec

essa

ry to

acc

ount

for t

he

adso

rptio

n of

Ba2+

(aq)

in m

odel

ing

the

surf

ace

char

ge in

the

solv

ent m

ixtu

re.

109

dissolution is limited except in alkaline environments, with concentrations limited to less

than 10-3 M. Prior to calculating the surface charge of either the platelet or equiaxed

particle, a determination the area fraction of each specific plane that bounds the particle is

required. For the platelet particles a right hand cylinder geometry was assumed with a

thickness of 5.8 nm and a face diameter of 27.1 nm. The thickness and diameter

dimensions are the medium values reported previously.13 The calculation yielded an area

fraction of 0.7 for the {111} plane. The remaining area fraction was equally distributed

between the {100} and {110} planes. For the determination of the surface structure of

the commercial particles, TEM images were collected. Figure 4.7a is a TEM image

showing that the particles are approximately equiaxed. Figure 4.7b is a high resolution

TEM image showing the lattice fringes of one of the BaTiO3 particles. By measuring the

distances among the planes it was verified that the particles are bounded by the {100}

and {110} planes. This observation is consistent with those of Jiang et al.44 in the

analysis of equiaxed BaTiO3 particles. For the analysis the area of the equiaxed particles

was equally divided between the {100} and {110} planes.

Figure 4.8 shows the zeta potential as a function of solution pH for the platelet

and equiaxed powders in 95/5 E/W. The lines represent the best fit theoretical

calculations based on the surface reactions and log K values (Table 4.1) from the MUSIC

model and the site densities of the surface groups. The points on the plot are

experimentally measured zeta potential values. To compare the model to the

experimental data Ba2+ adsorption at high pH was described using the modified Stern

isotherm. The ICP results in Table 4.2 show that significant dissolved Ba2+ is present in

solution pH greater than pH 8.

110

Figu

re 4

.7a

and

b.TE

M im

ages

of t

he c

omm

erci

al B

aTiO

3: (a

) im

age

show

ing

that

the

parti

cles

are

equ

iaxe

d, a

nd (b

) hig

h re

solu

tion

imag

e sh

owin

g la

ttice

frin

ges o

f a se

lect

ed p

artic

le w

ith a

<01

1> z

one

axes

and

that

the

surf

ace

of th

e pa

rticl

e is

te

rmin

ated

by

the

(100

) (0.

40 n

m) a

nd (0

11) (

0.28

nm

) pla

nes.

100

nm

-

111

Figu

re 4

.8.

Plot

of z

eta

pote

ntia

l ver

sus p

H fo

r 1w

t% su

spen

sion

s of p

late

let a

nd e

quia

xed

parti

cles

in a

95/

5 et

hano

l/wat

er

solv

ent m

ixtu

re. S

uspe

nsio

ns w

ere

prep

ared

in th

e so

lven

t mix

ture

to li

mit

Ba2+

diss

olut

ion.

How

ever

, a sm

all a

mou

nt o

f dis

solv

ed

Ba2+

is p

rese

nt a

t pH

gre

ater

than

pH

8 n

eces

sita

ting

the

incl

usio

n of

Ba2+

adso

rptio

n at

hig

h pH

to a

ccou

nt fo

r low

neg

ativ

e ze

ta

pote

ntia

l val

ues a

t pH

gre

ater

than

pH

10.

112

There is a difference of approximately 1 pH unit in the IEP of the two powders,

but this difference is not significant due to the observed error in the measurements near

the IEP. However, there is a significant difference between the magnitudes of the zeta

potential of the two powders at pH values remote from the IEP. Table 4.3 is a list of the

surface site densities for the Ti terminated surface of the three low index planes of

BaTiO3 based on the crystallography of BaTiO3 and those values used in the theoretical

calculation of the surface charge. The densities used in the calculation are four orders of

magnitude lower than the actual site densities. This is due to the theoretical model not

accounting for the presence of the IHP where there is a substantial potential drop between

the surface where the surface potential is calculated and the shear plane where the zeta

potential is measured. The site densities were normalized to the highest density plane to

compare the relative change of the three low index planes. The normalized site densities

show there is good agreement between the real and model surface. The lower zeta

potential of the platelet particles can be explained by the large area fraction of {111}

plane on the particles. The normalized values show that the {111} plane has a density of

approximately one half of the highest density plane. With the platelet particles having a

large {111} face the zeta potential is expected to be lower than that of the equiaxed

particles.

4.5 Conclusions

The complex nature of the aqueous surface chemistry of BaTiO3 was investigated

using dissolution studies and electrophoretic mobility measurements. A surface charging

model based on current charging theories was modified with an adsorption isotherm to

113

Tab

le 4

.3.

Site

den

sitie

s of T

i for

the

thre

e lo

w in

dex

plan

es o

f BaT

iO3

base

d on

the

stru

ctur

e of

BaT

iO3

and

the

valu

es u

sed

in

the

mod

elin

g of

the

surf

ace

in a

eth

anol

/wat

er so

lven

t mix

ture

. Th

e di

ffer

ence

bet

wee

n th

e ac

tual

and

mod

el v

alue

s is d

ue th

e m

odel

not

acc

ount

ing

for t

he p

oten

tial d

rop

in th

e IH

P. T

he n

orm

aliz

ed v

alue

s sho

w th

at th

e m

odel

is in

goo

d ag

reem

ent w

ith th

eac

tual

surf

ace

with

resp

ects

to th

e re

lativ

e de

nsity

for e

ach

plan

e.

Pla

necm

-2N

orm

aliz

edcm

-2N

orm

aliz

ed(1

00)

6.25

x 1

0151.

005.

60 x

1011

1.00

(110

)4.

40 x

1015

0.70

4.20

x 1

0110.

75(1

11)

3.60

x 1

0150.

582.

50 x

1011

0.45

Act

ual

Mod

el

114

account for the adsorption of Ba2+(aq). The precipitation of BaCO3 on the surface in a

alkaline environment was also included in the model using a Nernst-Gouy-Stern charging

model. The surface charge showed three distinct regions dependent on the solution pH.

The first region is at low pH where the dissolution of Ba2+ results in the surface behaving

similar to TiO2. In this region where only intrinsic surface reactions control the surface

charge the MUSIC model was applied. At intermediate solution pH values the adsorption

of Ba2+ results in a deviation from the ideal TiO2 surface observed at low pH. The

modified Stern isotherm was used to determine the surface charge due to Ba2+(aq)

adsorption. As the pH continues to increase the precipitation of BaCO3 on the surface

results in the BaTiO3 behaving similar to BaCO3. At pH value greater than the isoelectric

point of a TiO2 surface the charging is highly dependent on the concentration of dissolved

Ba2+(aq). Therefore an empirical formula for the concentration of dissolved Ba2+ was

derived as a function of pH and solid loading and used to calculate the concentration of

Ba2+(aq) present in solution.

115

References 1. K. Osseo-Asare, F.J. Arriagada, and J.H. Adair: Solubility relationships in the coprecipitation synthesis of barium titanate: Heterogeneous equilibria in the Ba-Ti-C2O4-H2O system. In Ceramic Transactions, Ceramic Powder Science, edited by G.L. Messing, E.R. Fuller, Jr., and H. Hausner, (The American Ceramic Society, 1988) pp 47. 2. P. Bendale, S. Venigalla, J.R. Ambrose, E.D. Verink, and J.H. Adair: Preparation of barium-titanate films at 55-degrees-C by an electrochemical method. J. Am. Ceram. Soc. 76, (10), 2619 (1993). 3. S. Venigalla and J.H. Adair: Theoretical modeling and experimental verification of electrochemical equilibria in the Ba-Ti-C-H2O system. Chem. Mater. 11, (3), 589 (1999). 4. M.C. Blanco-Lopez, B. Rand, and F.L. Riley: The properties of aqueous phase suspensions of barium titanate. J. Euro. Ceram. Soc. 17, 281 (1997). 5. U. Paik and V.A. Hackley: Influence of solids concentration on the isoelectric point of aqueous barium titanate. J. Am. Ceram. Soc. 83, (10), 2381 (2000). 6. B. Lee, S. Chung, and S.L. Kang: Necessary conditions for the formation of {111} twins in barium titanate. J. Am. Ceram. Soc. 83, (11), 2858 (2000). 7. M.C.B. Lopez, G. Fourlaris, and F.L. Riley: Interaction of barium titanate powders with an aqueous suspending medium. J. Euro. Ceram. Soc. 18, (14), 2183 (1998). 8. J.H. Adair, B.L. Utech, K. Osseo-Asare, and J.P. Dougherty, Solubility and phase stability of barium titanate in aqueous suspension, edited by J.P. Dougherty, and K. Wakino (in Proceedings of the Fifth US-Japan Seminar on Dielectric and Piezoelectric Ceramics, Kyoto, Japan, 1991), pp. 9. P. Duran, J. Tartaj, and C. Moure: Sintering behavior and microstructural evolution of agglomerated spherical particles of high-purity barium titanate. Ceram. Int. 29, 419 (2003). 10. J.J. Urban, W.S. Yun, Q. Gu, and H. Park: Synthesis of single-crystalline perovskite nanorods composed of barium titanate and strontium titanate. J. Am. Chem. Soc. 124, (7), 1186 (2002). 11. Y.B. Moa, S. Banerjee, and S.B. Wong: Hydrothermal synthesis of perovskite nanotubes. Chem. Comm. 3, 408 (2003). 12. Q. Huang and L. Gao: Synthesis and characterization of hexapod-shaped tetragonal-phase barium titanate single crystals. J. Am. Ceram. Soc. 87, (7), 1350 (2004).

116

13. T.J. Yosenick, D.V. Miller, R. Kumar, J.A. Nelson, C.A. Randall, and J.H. Adair: Synthesis of nanotabular barium titanate via a hydrothermal route. J. Mater. Res. 20, (4), 837 (2005). 14. T. Hiemstra, W.H. van Riemsdijk, and G.H. Bolt: Multisite proton adsorption modeling at the solid-solution interface of (hydr)oxides - A new approach. 1. Model description and evaluation of intrinsic reaction constants. J. Colloid Interface Sci. 133, (1), 91 (1989). 15. T. Hiemstra, P. Venema, and W.H. van Riemsdijk: Intrinsic proton affinity of reactive surface groups of metal (hydr)oxides: The bond valence principle. J. Colloid Interface Sci. 184, (2), 680 (1996). 16. P. Fenter, L. Cheng, M.L. Machesky, M.J. Bedzyk, and N.C. Struchio: Electrical double-layer structure at the rutile-water interface as observed in situ small-period x-ray standing waves. J. Colloid Interface Sci. 225, 154 (2000). 17. W. Piasecki: 1pK and 2pK protonation models in the theoretical description of simple ion adsorption at the oxide/electrolyte interface: The analysis of temperature dependence of potentiometric titration curves. J. Colloid Interface Sci. 254, (1), 56 (2002). 18. M. Fedkin, X.Y. Zhou, J.D. Kubicki, A.V. Bandura, S.N. Lvov, M.L. Machesky, and D. Wesolowski: High temperature microelectrophoresis studies of the rutile/aqueous solution interface. Langmuir 19, (9), 3797 (2003). 19. J.P. Fitts, M.L. Machesky, D. Wesolowski, X. Shang, J.D. Kubicki, G.W. Flynn, T.F. Heinz, and K.B. Eisenthal: Second-harmonic generation and theoretical studies of protonation at the water/α-TiO2 (110) interface. Chem. Phys. Lett. 411, 399 (2005). 20. C.W. Chiang and J.H. Jean: Effects of barium dissolution on dispersing aqueous barium titanate suspensions. Mater. Chem. Phys. 80, (3), 647 (2003). 21. U. Paik, S. Lee, and V.A. Hackley: Influence of barium dissolution on the electrokinetic properties of colloidal BaTiO3 in an aqueous medium. J. Am. Ceram. Soc. 86, (10), 1662 (2003). 22. M.L. Machesky, D. Wesolowski, D.A. Palmer, and M.K. Ridley: On the temperature dependence of intrinsic surface protonation equilibrium constants: An extension of the revised MUSIC model. J. Colloid Interface Sci. 239, 314 (2001). 23. L. Pauling: The principles determining the structure of complex ionic crystals. J. Am. Chem. Soc. 51, 1010 (1929). 24. D.A. Anderson, J.H. Adair, D.V. Miller, J.V. Biggers, and T.R. Shrout: Surface chemistry effect on ceramic processing of BaTiO3 powder. In Ceramic Transactions,

117

Ceramic Powder Science, edited by G.L. Messing, E.R. Fuller, Jr., and H. Hausner, (The American Ceramic Society: Westerville, OH, 1988) pp 485. 25. M.C. Blanco-Lopez, B. Rand, and F.L. Riley: The isoelectric point of BaTiO3. J. Euro. Ceram. Soc. 20, 107 (2000). 26. M.A. Malati and A.E. Smith: The adsorption of the alkine earth cations on titanium dioxide. Powder Tech. 22, 279 (1979). 27. D.W. Fuerstenau, D. Manmohan, and S. Raghavan: The adsorption of alkaline-earth metal ions at the rutile/aqueous interface. In Adsorption from aqueous solutions, edited by P.H. Tewari, (Plenum Press, 1981) pp 93. 28. H.M. Jang and D.W. Fuerstenau: The specific adsorption of alkaline-earth cations at the rutile/water interface. Coll. Surf. 21, 235 (1986). 29. S. Levine, G.M. Bell, and D. Calvert: The discreteness-of-charge effect in electrical double layer theory. Cand. J. Chem. 518, 518 (1962). 30. C.C. Hung and R.E. Riman: X-ray photoelectron spectroscopy investigation of hydrothermal and commercial barium titanate powders. In Chemical processing of advanced materials, edited by L.L. Hench, and J.K. West, (Wiley, 1992) pp 603. 31. C. Herard, A. Faivre, and J. Lemaitre: Surface decontamination treatments of undoped BaTiO3 - Part I. Powder and green body properties. J. Euro. Ceram. Soc. 15, (2), 135 (1995). 32. M.C. Blanco-Lopez, G. Fourlaris, B. Rand, and F.L. Riley: Characterization of barium titanate powders: Barium carbonate identification. J. Am. Ceram. Soc. 82, (7), 1777 (1999). 33. M.M. Lencka and R.E. Riman: Thermodynamic modeling of hydrothermal synthesis of ceramic powders. Chem. Mater. 5, (1), 61 (1993). 34. S.W.L. Lu, B.I. Lee, and L.A. Mann: Carbonation of barium titanate powders studied by FT-IR technique. Mater. Lett. 43, 102 (2000). 35. C.C. Li and J.H. Jean: Dissolution and dispersion behavior of barium carbonate in aqueous suspensions. J. Am. Ceram. Soc. 85, (12), 2977 (2002). 36. D.W. Fuerstenau, Pradip, and R. Herrera-Urbina: The surface chemistry of bastnaesite, barite, and calcite in aqueous carbonate solutions. Coll. Surf. 68, 95 (1992). 37. O. Popovych and R.T. Tomkins, Nonaqueous Solution Chemistry, 1st ed. (John Wiley & Sons, New York, 1981).

118

38. G.A. Parks: The isoelectric points of solid oxides, solid hydroxides, and aqueous hydroxo complex systems. Chem. Rev. 65, (2), 177 (1965). 39. G. Gouy: Constitution of the electric charge at the surface of an electrolyte. J. Physique 9, (4), 457 (1910). 40. D.L. Chapman: Theory of electrocapillarity. Phil. Mag. 25, 475 (1913). 41. W. Healy and W.L. White: Ionizable surface group model of aqueous interface. Adv. Coll. Int. Sci. 9, 303 (1978). 42. R.J. Hunter, Zeta potential in colloid science: Principles and applications, 1st ed. (Academic Press, San Diego, CA, 1981). 43. R. Chodelka: The aqueous processing of barium titanate: passivation, dispersion, and binder formulations for multilayer capacitors. PhD Thesis, University of Florida, Gainsville, FL, (1996). 44. B. Jiang, J.L. Peng, and L.A. Bursill: Surface structures and dielectric response of ultrafine BaTiO3 particles. Ferroelectrics 207, (3-4), 445 (1998).

119

CHAPTER FIVE

Passivation, Dispersion, and Aqueous Solution Doping of Platelet BaTiO3 Powder

5.1 Introduction

The high dielectric constant is the primary reason BaTiO3 is the main material

used in the fabrication of multilayer ceramic capacitors (MLCCs). To achieve the

reduction in dielectric thickness, the colloidal processing of BaTiO3 must be addressed.

To achieve the reduction in layer thickness the use of nanoparticles has readily been

implemented. Because of the reduction of the particle size issues related to the surface

chemistry, doping, and particulate mixing will be of greater importance in the processing

of MLCCs.

The surface chemistry of BaTiO3 is complex due to chemical instability and

surface reactions.1-3 Currently, most BaTiO3 materials are used in the fabrication of

MLCCs where tape casting is the main forming method. Non-aqueous particulate

slurries are the basis for most industrial tape casting.4 For financial and environment

reasons aqueous tape casting is of interest. However, many problems with aqueous based

tape casting exist. In general, foaming, cracking during drying, and inadequate binder

systems lead to low quality tapes with poor mechanical properties. To overcome the

problems it is necessary to better understand the dispersion and interactions of BaTiO3 in

aqueous based suspensions.

Many materials, especially BaTiO3, are not thermodynamically stable in water

and undergo chemical reactions with water present.5 The thermodynamically stable form

120

of most metal oxide ceramics in water is not the pure oxide but the metal hydroxide,

hydroxy carbonate, or carbonate.5, 6 For complex oxide materials composed of a solid

solution of two or more metal oxides, such as BaTiO3, the thermodynamics can be

complicated because of heterogeneous reactions.1, 7, 8

Figure 4.1 shows the ideal stability field for the Ba-Ti-O-C system from Bendale

et al.9 The solubility of Ba2+ is dependent on solution pH with solubility decreasing in an

alkaline environment. The dissolution of Ba2+ during the aqueous processing of BaTiO3

is a problem because it results in a Ti-rich surface layer.10, 11 This is an issue because

during sintering, local Ti-rich regions act to form low temperature melting phases and

promote exaggerated grain growth.12

Figure 4.1 shows that in addition to Ba2+ dissolution, the formation of BaCO3 is

also a concern. CO2 in the atmosphere readily dissolves into water, and at neutral pH

where Ba2+ dissolves from the surface, BaCO3 can form.13, 14 During sintering the high

temperature (>1200°C) decomposition of BaCO3 can lead to decreased density because

of CO2 gas evolution in closed pores.15 The gas exerts a negative sintering pressure

which inhibits and eventually reverses the sintering process. Because of the deleterious

effects of BaCO3 on BaTiO3 processing it is essential to limit Ba2+ dissolution and the

subsequent BaCO3 formation.

Chemical surface passivation can be applied to a variety of materials including

structural ceramics16, semiconductors17, superconducting oxides18, and metals.19-21 In all

cases the material is not stable in the intended use environment and the surface has been

chemically modified to provide a barrier against degradation. Degradation can occur by

two mechanisms; the material can degrade by the diffusion of a species into the lattice,

121

typically oxygen or the surface can be chemically unstable and begin to dissolve. The

primary issue with BaTiO3 is the latter and it is necessary to passivate the surface to

inhibit the dissolution of Ba2+ from the surface. Vasques et al.18 discussed the parameters

required for a passivation agent of a similar material, YBa2Cu3O7-x : (1) it must exhibit

low solubility in the solvent, and (2) act as a good diffusion barrier. Although low

solubility is necessary, it is not the only concern. For example, BaCO3 has low solubility

in water but does not prevent the aqueous degradation of YBa2Cu3O7-x superconductors.

However, the solubility can be used as a starting point to limit the search for a passivation

agent.

Once the chemical stability of BaTiO3 has been addressed, it is necessary to focus

on other processing issues, including doping. Doping is necessary because BaTiO3

exhibits three phase transitions: rhombohedral to orthorhombic, orthorhombic to

tetragonal, and tetragonal to cubic at -90, 0 and 130 °C, respectively.22 At each phase

transition there is an associated spike in the dielectric properties of the materials making

it unsuitable for use over a wide temperature range. There is a need to tailor the electric

properties and to flatten the dielectric response of BaTiO3. The Electronic Industries

Alliance (EIA) of the United States has established guidelines for acceptable changes in

capacitance for defined temperature ranges. One of the criteria in selecting a capacitor

for a specific application is the expected range of temperature use.

The X7R class has an allowable ±15% change in capacitance with respect to the

room temperature capacitance over a temperature range from -55 to 125 °C. X7R

capacitors are typically used in applications which require a broad temperature range,

including mobile electronics. To tailor the electrical properties of BaTiO3 and achieve

122

X7R specifications doping the material is required. A series of materials known as Curie

shifters have been identified which shift the Curie temperature (TC) of BaTiO3 and result

in a flat temperature response.23

To achieve the low temperature coefficient of capacitance of an X7R dielectric a

chemically inhomogeneous microstructure is required. One possible route is the

development of a core-shell microstructure. In a core-shell microstructure, individual

grains have a chemical gradient ranging from undoped BaTiO3 in the center of the grain

to fully doped BaTiO3 the edge of the grain. The chemical gradient results in a gradient

of the TC from pure BaTiO3 in the core to the TC of doped BaTiO3 in the shell.

Current solid-state doping methods require the addition of dopant particles to the

matrix particles. The mixture is typically ball milled to incorporate dopants and

homogenize the mixture. The process has been studied by Wiseman in the doping of

ZnO for the fabrication of varistors. If dopants are present as particulates, chemical

homogeneity is nearly impossible to achieve.24 As the dopant particle size meets or

exceeds that of the matrix particle high sintering temperatures and long sintering times

are required for diffusion in the development of a core-shell microstructure. With the

growing use of nanoscale matrix particles an alternate doping method not limited by

particulate mixing is required.

A chemical approach is an alternative method of doping in which the dopant is

added in ionic or molecular form. In the ionic approach, soluble salts are added to

aqueous suspensions, whereas the molecular approach is used typically in non-aqueous

suspensions where the dopants are added as organo-metallics precursor molecules. In

either approach little care has been taken to ensure that dopants selectively adsorb onto

123

particle surfaces. In addition to adsorption or precipitation of the dopants onto the

particle surface, segregation or homogenous precipitation of the unabsorbed dopants can

occur upon drying of the suspension. Irrespective of the location of the dopants, the

resulting powder mixture has a more homogenous composition compared to powders

doped by a solid state approach. Both approaches, ionic and molecular, have been shown

to be successful in the processing of highly engineered materials such as, ZnO-based

varistors25 and BaTiO3-based capacitors.26, 27 Although an aqueous approach for BaTiO3

has been previously outlined27, no attempt was made to passivate the surface and protect

the surface from degradation. It is necessary to protect or passivate the surface from

degradation to limit the processing problems that have been observed due to Ba2+

dissolution.15, 28

Fernandez et al. studied the doping of BaTiO3 using three different methods:

power synthesis doping, solid-state, and chemical.29 Doping during synthesis was not

effective due to a homogenous chemical composition obtained in the final microstructure.

Both the solid-state and chemical approaches were effective in achieving X7R dielectrics,

but yielded differences in the electrical loss due to differences in the distribution of the

dopants.

The goal of the current study are to: (1) investigate the use of oxalic acid as a

passivation agent for the aqueous processing of BaTiO3, (2) address the dispersion of

passivated BaTiO3 with a cationic polyelectrolyte, and (3) use an aqueous solution based

method for the doping of BaTiO3.

124

5.2 Materials and Methods

Dissolution experiments were conducted to determine the chemical stability of

BaTiO3 in an aqueous environment in the presence of a passivating agent and dispersant.

The experiments were conducted using two different BaTiO3 powders; a commercial

equiaxed powder (BT-01, Sakai Chemical Company, Osaka, Japan) and a hydrothermally

derived anisotropic platelet powder. The details of the synthesis and physical properties

of the platelet powder are outlined elsewhere.30 Oxalic acid dihydrate (HOx) (Fisher

Scientific, Fair Lawn, NJ) was used as the passivating agent and polyethylenimine (PEI)

(25,000 Mw, Aldrich Chemical Company, Milwaukee, WI) was used as the dispersant.

PEI was used because it is a cationic polyelectrolyte which develops positive charge at

pH values less than pH 9.31 Solutions of 0.1 M and 1 M tetraethylammonium hydroxide

(TEAOH) (35wt% in water, Aldrich Chemical Company, Milwaukee, WI) and nitric acid

(70w%, J.T. Baker, Phillipsburg, NJ) were used to adjust suspension pH. All suspensions

prepared for dissolution experiments contained 1wt% BaTiO3 powder and varying

amounts of HOx and PEI in CO2-free DI water that had been previously prepared by

boiling with flowing argon to remove adsorbed CO2.

Suspensions were prepared by adding the HOx and/or PEI to DI water and

adjusting the pH to 3, 5, 7, 9 or 11. Suspension pH was measured using ion-selective

field effect transitor (IS-FET) pH probe (Senton Hotline Probe, RL Instruments

Manchoung, MA) calibrated using NIST-traceable standards with nominal pH values of

4, 7, and 10. After pH adjustment the BaTiO3 powder was added and the suspensions

were allowed to equilibrate for 18 to 24 hours. The zeta potentials of the suspensions

were measured using electrophoretic light scattering (ZetaPALS, Brookhaven Instrument

125

Corp, Holtsville, NY). Suspensions were centrifuged at 10,000 rpm for 7 min, and the

supernatant of each suspension was then filtered using a 0.22 μm syringe filter. Ba and

Ti concentrations for each suspension were measured using indirectly coupled plasma

atomic emission spectroscopy (ICP-AES) (PS3000UV, Leeman Labs, Los Angeles, CA).

Standards for ICP-ES were prepared by serial dilution from 1000 ppm standard stock

solutions (Hi Purity Standards, Charleston, SC).

Rheological properties were measured to investigate the particle-particle

interactions in moderately concentrated, 10 volume percent, suspensions. Suspensions

for analysis were chosen from suspensions which exhibited both low Ba2+ and Ti4+

solubility and high zeta potential. Rheological properties were measured using a cone-

plate type rheometer (CSL 100K Auto Gap, Carri-Med, Surrey, England) with shear rates

ranging from 1 to 1000 sec-1. The apparent viscosity was then calculated from the

rheological data by extrapolating the viscosity as a function of shear rate at shear rate

values greater than 700 sec-1 back to zero shear and reporting the y-intercept of the

asymptote as the apparent viscosity. The Bingham yield point was reported to be the

shear stress linearly extrapolated to zero shear rate.

Three different elements, Co, Nb, and Bi were added to dope the powder. Co and

Nb were chosen because X7R dielectrics have previously been reported using Co and

Nb.32, 33 Bi was added as a flux to lower the sintering temperature and promote liquid

phase sintering.34, 35 A solution of 0.5 M cobalt nitrate hexahydrate (98+%, Aldrich

Chemical Company, Milwaukee, WI) served as the cobalt source. A solution of 0.5 M

bismuth ammonium citrate was the bismuth source and was prepared by dissolving

bismuth citrate (94%, Alfa Aesar, Ward Hill, MA) with ammonium hydroxide (30%,

126

Aldrich Chemical Company, Milwaukee, WI) in a molar ratio of 1:3. An aqueous citrate-

peroxide-niobium complex was prepared using niobium ammonium oxalate (H.C. Starck,

Newton, MA) as the niobium source following the method outline by Narendar and

Messing.36 After preparation the solution was concentrated using a rotoevaporator (RE

51, Yamato Scientific Company, Tokyo, Japan), and the niobium concentration in

solution was determined using indirectly couple plasma atomic emission spectroscopy

(ICP-AES)

The dopants were added in a step-wise manner to ensure adsorption onto the

particle surface. First, the BaTiO3 powder was passivated with oxalic acid dihydrate.

The cobalt nitrate was then added followed by the addition of PEI. Finally, the bismuth

and niobium were added and to the suspension while pH maintained at pH 7 using 0.1M

solutions of either nitric acid (69-70%, JT Baker, Phillipsburg, NJ) or

tetraethylammonium hydroxide (35wt% in water, Aldrich Chemical Company,

Milwaukee, WI).

Zeta potential measurements were performed at each step to determine if the

dopants were specifically absorbing on the particle surfaces. Suspensions were prepared

with 0, 1, 2, 5, and 10 wt% of the individual dopant and then the zeta potential was

measured. TEM analysis was performed using holey carbon film on Cu grids as sample

holders with a single drop of dilute suspension placed on each grid. Dopant layer

thickness was imaged and measured using transmission electron microscopy (TEM)

(2010 LaB6, JEOL. Japan). The presence of the dopant was confirmed with the use of

energy dispersion spectroscopy (EDS) (Gatan Inc., Pleasanton, CA).

127

To quantitatively determine the amount of dopant absorbed during doping three

powder samples were prepared with 5wt% Bi2O3 and 0, 2, and 5wt% CoNb2O6. After

doping the suspensions were centrifuged to separate the doped powder for the supernatant

and then dried. The powder was heated to 600 °C in flowing oxygen for 2 hrs. to remove

any organic present. The powder was then analyzed using X-ray fluorescence (XRF)

(1600/10, Phillips, Netherlands) for both the dopant concentration and the Ba to Ti ratio.

5.3 Results and Discussion

5.3.1 Passivation

Previous attempts at passivation haven been made with limited success.37-39

Much of the work has focused on the use of poly(acrylic acid) (PAA) because the

carboxylic acid groups on the PAA complex well with Ba2+.37-39 Oxalic acid was chosen

as a possible passivation agent because it readily forms an insoluble metal salt with

barium.40 Ba2+ attempting to leach from the surface reacts with the oxalate ion to form

barium oxalate which heterogeneously nucleates on the particle surface creating a

passivation layer.5 This is similar to observations by Mandanas et al. in the passivation

of doped ferrite powders41 and yttria-doped zirconia by Kimel and Adair.42 Figure 5.1 is

TEM image showing the presence of a 2-3 nm surface layer on a BaTiO3 platelet particle

treated with oxalic acid.

In addition to TEM observations, zeta potential determination was used to

investigate the effect of oxalic acid on the surface of BaTiO3. Figure 5.2 give the zeta

potential for suspensions with varying amounts of oxalic acid as a function of pH. As the

concentration of oxalic acid increases two important effects are observed: (1) the

128

Figu

re 5

.1.

TEM

imag

e of

an

oxal

ate

pass

ivat

ed B

aTiO

3pa

rticl

e. T

reat

men

t with

oxa

lic a

cid

resu

lts in

a 2

nm

thic

knes

s sur

face

la

yer o

f bar

ium

oxa

late

whi

ch in

hibi

ts th

e su

rfac

e fr

om d

egra

datio

n.

129

Figu

re 5

.2.

Zeta

pot

entia

l of 1

wt%

susp

ensi

on o

f nan

opla

tele

t BaT

iO3

in w

ater

with

incr

easi

ng a

mou

nts o

f oxa

lic a

cid

as a

fu

nctio

n of

pH

. Fu

ll su

rfac

e pa

ssiv

atio

n is

ach

ieve

d by

an

oxal

ic a

cid

conc

entra

tion

of 3

x10-3

M (3

.75w

/w).

A fu

rther

incr

ease

in

the

oxal

ic a

cid

conc

entra

tion

resu

lts in

an

incr

ease

in th

e m

agni

tude

of t

he z

eta

pote

ntia

l.

130

isoelectric point (IEP) shifts or is suppressed, and (2) the zeta potential becomes

independent of pH. A shift in the IEP is conformation of specific adsorption of the oxalic

acid onto the surface of the particle. The pH independent zeta potential is similar to that

of barium oxalate, implying that as the BaTiO3 particle is treated with oxalate a Nernstian

surface of barium oxalate is formed. The zeta potential data along with the TEM

observations show that in the presence of oxalate a barium oxalate surface layer forms on

BaTiO3.

For full passivation it is necessary to have complete surface coverage. When full

surface passivation with barium oxalate is achieved the zeta potential behaves similar to

that of barium oxalate. At an oxalic acid concentration of 10-3 M an isoelectric point

(IEP) still exists around pH 9, indicating that Ba2+ dissolution still occurs. With an

oxalate concentration of 3x10-3 M the IEP has been suppressed and the zeta potential

curve appears similar to the zeta potential curve for that of barium oxalate, verifying full

passivation is achieved.

In the pH range where barium oxalate is sparingly soluble in water, the surface

charging of barium oxalate is regulated by a Nernst-Gouy-Stern surface reaction that is

independent of pH.40 Only the concentration of barium and oxalate ions control the

magnitude of the zeta potential. Only an excess of either barium or oxalate ions present

in solution should further increase the magnitude of the zeta potential. Increasing the

oxalic acid concentration produces an increase in the magnitude of the zeta potential but

also in an increase in the concentration of free oxalate ions in solution. An oxalic acid

concentration of 5x10-2 M provides the largest magnitude zeta potential.

131

Figu

re 5

.3.

Plot

of s

olub

le B

a2+(a

q)co

ncen

tratio

n as

a fu

nctio

n of

susp

ensi

on p

H a

nd o

xalic

aci

d co

ncen

tratio

n fo

r the

pla

tele

t B

aTiO

3. A

n ox

alic

aci

d co

ncen

tratio

n of

5x1

0-2M

yie

lds t

he b

est s

urfa

ce p

assi

vatio

n ye

t low

er c

once

ntra

tions

are

acc

epta

ble

as

long

as t

he so

lutio

n pH

rem

ains

gra

ter t

han

pH 5

.

132

Figu

re 5

.4.

At a

cidi

c pH

val

ues T

i for

ms a

solu

ble

com

plex

with

oxa

lic a

cid.

At h

igh

oxal

ic a

cid

conc

entra

tion

Ti d

isso

lutio

n fr

om th

e B

aTO

3su

rfac

e is

una

ccep

tabl

e. H

owev

er, a

t an

oxal

ic a

cid

conc

entra

tion

of 3

x10-3

M a

t pH

val

ues g

reat

er th

an p

H 5

the

Ti4+

conc

entra

tion

in so

lutio

n is

neg

ligib

le. (

--) a

t 10-7

M in

dica

tes t

he li

mit

of d

etec

tion

for T

i4+w

ith IC

P-ES

. Li

nes a

re tr

end

lines

on

ly.

133

Although it has been shown the oxalic acid treatment results in a barium oxalate

surface layer it is not known if the layer provides passivation. To investigate the

passivation with oxalic acid, ICP-AES analysis of the suspension supernatants was

performed. Analyzes for both Ba and Ti were conducted. Figures 5.3 and 5.4 show the

Ba and Ti concentrations in solution, respectively, as a function of oxalic acid

concentration and suspension pH for the platelet powder. The Ba2+ dissolution data

shows that as oxalic acid concentration increases, passivation is observed and that 5x10-2

M oxalic acid provides the best passivation over the entire pH range studied. However,

the Ti ICP-AES data shows that increasing the oxalic acid concentration results in

increased Ti solubility with Ti solution concentrations as high as 10-2 M. Oxalic acid

forms a readily soluble complex with Ti43, 44 and the reaction of oxalate with Ti degrades

the surface of the BaTiO3 in a manner similar to Ba2+ dissolution.

To use oxalic acid as a passivation agent in the aqueous processing of BaTiO3 it is

necessary to balance the Ba2+ dissolution and increased Ti solubility. An oxalic acid

concentration of 3x10-3 M, which is equivalent to 3.75 w/w oxalic acid for the 1wt%

suspensions prepared, limits both Ba and Ti solubility as long as solution pH is greater

than pH 5. This constraint sets the lower limit for aqueous processing with oxalic acid at

pH 5.

5.3.2 Dispersion

If the magnitude of the zeta potential were large enough to provide stable

dispersion then dispersion could be achieved through the use of oxalic acid alone.

However, because Ti4+ solubility limits the use of high concentrations of oxalic acid the

zeta potential after passivation is only -15mV. Although dependent on many factors, at

134

room temperature a zeta potential of magnitude greater than ±25 to 30 mV is generally

necessary to produce a stable dispersion in a moderate to low ionic strength suspension.45

Therefore a dispersant is required to promote dispersion in BaTiO3 suspensions. An

added advantage of a polyelectrolyte dispersant is that in addition to electrostatic

repulsion a steric component is provided for dispersion. Polyethylenimine was chosen

because it is a cationic polyelectrolyte which develops positive charge at pH values less

than pH 10.31 The positively charged PEI readily absorbs to the negatively charged

oxalate surface, and as the pH is lowered to neutral pH values, the degree of protonation

of the PEI increases, yielding large zeta potentials.41

Zeta potential measurements and ICP-AES were used to analyze the effect of PEI

on the surface chemistry of an oxalate passivated BaTiO3 and determine the optimum

dosage for dispersion. Figure 5.5 shows the zeta potential of platelet BaTiO3 with

varying amounts of PEI as a function of pH. As the PEI concentration increases, the sign

of the zeta potential reverses verifying PEI absorption on particle surfaces. In the pH

range of interest, pH 5-10, PEI concentrations greater than 1 w/w provide zeta potentials

suitable for dispersion (≥ ±25mV). ICP-AES results, not shown, show negligible

variations in either the Ba2+ or Ti4+ concentration as a function of PEI concentration.

Zeta potential measurements provide insight into the stability of particulate

suspensions and the potential for good dispersion. However, rheological properties are

highly dependent on the state-of-dispersion and are controlled by particle-particle

interactions in the suspensions.46 To evaluate the effect of oxalic acid and PEI on the

interparticle interactions the rheological properties of moderately concentrated, 10 vol%,

suspensions of the equiaxed commercial particles were measured. All suspension

135

Figu

re 5

.5.

The

addi

tion

of P

EI to

oxa

lic a

cid

pass

ivat

ed B

aTiO

3su

spen

sion

s res

ults

in a

pos

itive

zet

a po

tent

ial d

ue to

the

adso

rptio

n of

PEI

on

the

bariu

m o

xala

te su

rfac

e. I

n ad

ditio

n to

larg

e po

sitiv

e ze

ta p

oten

tial v

alue

s PEI

add

s a st

eric

hin

dran

ce to

ai

ds in

par

ticle

dis

pers

ion.

136

Figu

re 5

.6a

and

b.Th

e sh

ear s

tress

and

vis

cosi

ty o

f 10v

ol%

susp

ensi

ons w

ith 2

.5w

/w P

EI sh

ows t

hat i

ncre

asin

g th

e ox

alic

aci

d co

ncen

tratio

n re

sults

in a

dev

iatio

n fr

om N

ewto

nian

beh

avio

r due

to th

e in

tera

ctio

n of

PEI

and

oxa

lic a

cid

to fo

rm a

gel

net

wor

kof

am

ine

oxal

ate.

The

line

ar re

gion

s of (

a) w

ere

fit w

ith B

ingh

am’s

law

and

the

y-in

terc

ept w

as re

porte

d as

τB, t

he y

ield

poi

nt. A

ll su

spen

sion

s mea

sure

d w

ere

at p

H 7

to m

aint

ain

an a

ppro

xim

ate

zeta

pot

entia

l of a

ppro

xim

atel

y +2

5mV

.

137

rheology was measured at pH 7 to maintain a zeta potential of approximately +25mV,

and avoid any possible changes in the particle-particle interactions due to changes in the

degree of protonation of the PEI. Figures 5.6a and b show the shear stress and viscosity

as a function of shear rate for suspension with 2.5w/w PEI and increasing dosages of

oxalic acid. The linear regions of the curves on Figure 5.6a were fit using Bingham’s

law,

[5.1] •

+= γηττ B

where τ is the shear stress, τB is the yield point, η is viscosity, and γ is the shear rate. As

the oxalic acid concentration increases the apparent viscosity increases. The increasing

apparent viscosity and non-Newtonian behavior is an indication of stronger particle-

particle interactions. It is likely that the increased interaction is due to the interaction of

amine groups on PEI with the oxalate. An amine-carboxylic acid complex is formed by

the carboxylic acid group transferring a proton to the nitrogen on an amine group.47

Vaidhyanathan et al. were able to make crystalline amine oxalates using oxalic acid and

organic amines with a limited number of amine groups. However, in the current work the

high concentration of amine groups on the PEI allow for the formation of a gel network

of amine oxalate. The formation of a gel network was confirmed by evaluating the

behavior of PEI oxalate mixtures in solution. For the formation of amine oxalate, free

oxalates in solution react with the PEI. Thus, if concentrations of PEI and oxalate are

higher than required for passivation/dispersion an undesirable gel network forms with

loss of rheological control.

Firth and Hunter developed a physical model for the rheological properties of

colloidal suspensions.48 In the elastic-floc model flocs are envisioned as elastic clusters

138

of hard aggregates that deform under shear conditions. The Bingham yield point is

related to the energy required to rupture a floc, which is comprised of two parts: (1)

Estretch, the energy that is dissipated by elastic deformation of the floc, and (2) Ebreak,

energy required to break the flocs due to interaction potential. Table 5.1 summarizes the

apparent viscosity and Bingham yield points for suspensions with various concentrations

of PEI and oxalic acid. Figures 5.7a and b show the apparent viscosity and yield point as

a function of oxalate concentration for 10 vol% suspensions with varying concentrations

of PEI. Increasing the oxalic acid concentration increases the apparent viscosity and

Bingham yield point. Since the zeta potential and local chemistry (i.e. the amine-oxalate

linkage) were maintained at constant values in the current experiments, it can be reliable

assumed that Ebreak is constant. Therefore, the changes in Bingham yield point with

increasing oxalic acid concentration are due to an increased elastic component, Estretch.

The increasing Bingham yield point supports the notion that the oxalic acid and PEI are

forming a gel-like amine-oxalate network, which under applied shear is easily

deformable.

Although zeta potential measurements and ICP-AES indicate that an oxalic acid

concentration of 3.75w/w provides complete surface coverage, the viscosity data suggests

that an oxalic acid concentration of 2.25w/w provides adequate surface coverage. At an

oxalic acid concentration as low as 3w/w, increased viscosities and yield points indicate

the formation of amine oxalates due to the presence of excess oxalic and PEI. With the

reduced oxalic acid concentration (2.25w/w) a loss of oxalate passivation is a concern.

For the experiment conditions 2.25w/w HOx equals approximately 2x10-3 M HOx.

Based on the zeta potential and Ba2+ ICP-AES results, Figures 5.3 and 5.4, 2x10-3 M

139

Tab

le 5

.1.

Rhe

olog

ical

pro

perti

es fo

r 10v

ol%

BaT

iO3

susp

ensi

ons p

repa

red

with

var

ying

am

ount

s of o

xalic

aci

d an

d PE

I. T

he

addi

tion

of e

xces

s oxa

lic a

cid

resu

lts in

incr

ease

d vi

scos

ity a

nd y

ield

poi

nts d

ue to

the

form

atio

n of

a g

el n

etw

ork

of a

min

e ox

alat

e th

at in

crea

ses i

nter

parti

cle

inte

ract

ions

. A

ll su

spen

sion

s wer

e m

easu

red

at p

H 7

in o

rder

to m

aint

ain

an a

ppro

xim

ate

zeta

pot

entia

l of

+25

mV PE

I (w

/w)

2.25

33.

754.

52

η a =

1.0

cP

η a =

4.4

cP

η a =

5.8

cP

η a =

6.2

cP

τ B =

N/A

τ B =

6.9

4 dy

nes/

cm2

τ B =

11.

05 d

ynes

/cm

2τ B

= 1

2.05

dyn

es/c

m2

2.5

η a =

1.4

cP

η a =

2.8

cP

η a =

3.7

cP

η a =

6.3

cP

τ B =

N/A

τ B =

1.5

6 dy

nes/

cm2

τ B =

3.7

4 dy

nes/

cm2

τ B =

12.

06 d

ynes

/cm

2

3η a

= 3

.8 c

Pη a

= 1

.6 c

Pη a

= 4

.2 c

Pη a

= 5

.5 c

P

τ B =

3.1

9 dy

nes/

cm2

τ B =

N/A

τ B =

6.9

6 dy

nes/

cm2

τ B =

10.

53 d

ynes

/cm

2

Oxa

lic A

cid

(w/w

)

Not

e: η

a=

appa

rent

vis

cosi

tyτ B

= B

ingh

am y

ield

poi

ntN

/A =

bel

ow in

stru

men

t mea

sure

men

t lim

it

140

Figu

re 5

.7.

App

aren

t vis

cosi

ty (a

) and

yie

ld p

oint

(b) v

alue

s as a

func

tion

of o

xala

te c

once

ntra

tion

for 1

0vol

% su

spen

sion

with

va

ryin

g co

ncen

tratio

n of

PEI

. Th

e su

spen

sion

con

tain

ing

exce

ss o

xalic

aci

d ex

hibi

t inc

reas

ed v

isco

sity

and

yie

ld p

oint

due

to th

e fo

rmat

ion

of a

min

e ox

alat

e ge

l net

wor

k. A

ll su

spen

sion

s wer

e m

easu

red

at p

H 7

to m

aint

ain

a co

nsta

nt z

eta

pote

ntia

l of +

25 m

V.

141

HOx should provide a minimum of passivation. Any further reduction in the oxalic acid

concentration should be avoided to limit Ba2+ dissolution.

Suspension rheological properties are dependent on the adsorption of dispersants

onto the particle surface.49 The data in Table 5.1 shows that as the oxalic acid

concentration increase, changes in the rheological behavior as a function of PEI are

observed. Not until the PEI concentration is high enough to react with the free oxalate

will surface coverage be complete enough to promote dispersion. This issue is not a

problem in suspensions with 2.25w/w oxalic acid because low concentrations lead to

oxalate present only on particle surfaces.

In the suspensions with 2.25w/w oxalic acid the increasing PEI concentration

should result in an increased viscosity. As the dispersant concentration increases the

surface coverage should increase and yield higher zeta potentials. High zeta potentials

lead to larger inter-particle repulsion accompanied by decreased viscosity and yield point.

The increased viscosity with increasing concentration indicates that there is a possible

change in the mechanism of PEI adsorption as a function of concentration. PEI is a

commonly used dispersant and its adsorption to SiO2, an ideal negatively charged

surface, has been previously studied.50 The conformation of PEI changes depending on

PEI concentration. At low concentrations, PEI assumes a flat unfolded morphology on

the particle surface. As the PEI concentration increases, a folded morphology with an

increased density of loops and tails is observed.51 In its folded conformation, PEI forms a

patchy absorbed layer leaving regions of uncoated surface. At high PEI concentrations,

because of high loop and tail densities and uncoated surface regions, bridging

flocculation has been observed.50 Thus, as PEI concentration increases, the PEI assumes

142

a more folded conformation and bridging flocculation occurs leading to increased

viscosities.

5.3.3 Doping of Nanotabular BaTiO3

With an understanding of the aqueous passivation/dispersion of BaTiO3 it is

possible to manipulate the surface charge to control the adsorption of ionic species onto

the particle surface. Normal doping routes need to be scaled to the nanometer regime for

BaTiO3 nanoparticles. It is believed that the controlled adsorption of dopant ions is a

possible route for the doping of nanoscale BaTiO3. The advantage of a solution based

technique is a much more homogeneous dopant distribution via the ability to create a

coated particle prior to sintering.

Like Ba, Co forms a sparingly soluble oxalate at neutral pH. The surface charge

of cobalt oxalate is also insensitive to pH and only dependent on the concentration of Co

and oxalate in solution.40 The addition of cobalt nitrate to a suspension of oxalate

passivated BaTiO3 powder results in the deposition of cobalt forming a cobalt oxalate

surface on the BaTiO3 particles. Figure 5.8 gives the zeta potential versus dopant

concentration for the three dopants studied: Co, Nb, and Bi. The figure shows that as

cobalt is added the sign of the zeta potential reverses, going from positive to negative

confirming that cobalt oxalate is forming on particle surfaces. At low cobalt

concentrations the particle still maintains a negative surface charge suitable for the

absorption of PEI for dispersion.

The lack of suitable Nb and Bi metal salts that are soluble at neutral pH, make the

precipitation of niobium oxalate or bismuth oxalate on the particle surface difficult.

However, Nb and Bi form soluble complexes with citric acid that are negatively charged

143

Figu

re 5

.8.

Zeta

pot

entia

l of d

oped

pla

tele

t par

ticle

as a

func

tion

of d

opan

t con

cent

ratio

n. A

s Co

is a

dded

to th

e su

spen

sion

a

reac

tion

with

exc

ess o

xalic

aci

d oc

curs

to fo

rm a

cob

alt o

xala

tesu

rfac

e. W

hen

PEI i

s add

ed, t

he su

rfac

e be

com

es p

ositi

ve a

nd is

su

itabl

e fo

r the

ads

orpt

ion

of th

e ne

gativ

ely

char

ged

Nb

and

Bi c

ompl

exes

.

144

at neutral pH.36, 52 Thus, a positive PEI surface is ideal for the adsorption of these

negative complexes. After the addition of the cobalt nitrate forming a negatively charged

cobalt oxalate surface, PEI was added to provide a positive surface charge suitable for the

absorption of the negatively charged Nb and Bi complexes. Zeta potential measurements

(Figure 5.8) confirm that the negative complexes absorb onto the positive PEI surface.

Similar to the addition of Co, increasing the concentration of Nb and Bi reverses the sign

of the surface charge confirming the adsorption of the Nb and Bi complexes onto the PEI

coated particle surfaces.

Figure 5.9 is a schematic showing the doping process. First, Ba2+ dissolves from

the surface and along with Co2+ reacts with the oxalic acid to form a negatively charged

oxalate surface layer. PEI is added to disperse the particles and reverse the sign of the

surface charge. Finally, NbO(O2)Cit3- and Bi(NH3)3Cit- are added and electrostatically

adsorb onto the particle surface. Figures 5.10a-e are TEM micrographs of platelet

BaTiO3 particles at each step of the doping process beginning with an undoped sample.

The difference between the undoped and fully doped particles is obvious. The surface of

the undoped particle shows few features, whereas the fully doped particle has small (1-2

nm) structures on the surface. Figure 5.11 is an EDS spectrum for the fully doped

particle. The minor peaks of Co, Nb, and Bi confirm the dopants are present on particle

surfaces. The other peaks in the spectrum are artifacts from the experimental setup and

instrumentation. C and Cu are from the sample grid used for the sample preparation and

Fe is contamination from the TEM used for EDS analysis.

Table 5.2 summarizes the XRF analysis of three doped powder samples. The

three samples were doped with 5wt% Bi2O3 and 0, 2, or 5 wt% CoNb2O6. To measure

145

Figu

re 5

.9.

Sche

mat

ic re

pres

enta

tion

of th

e do

ping

pro

cess

ing.

Firs

t oxa

lic a

cid

and

coba

lt ar

e ad

ded

to fo

rm a

n ox

alat

e su

rfac

e la

yer w

hich

pas

siva

tes t

he su

rfac

e. P

EI is

add

ed to

dis

pers

e th

e pa

rticl

es a

nd p

rovi

de a

pos

itive

surf

ace

char

ge fo

r the

ads

orpt

ion

of th

e N

b an

d B

i whi

ch a

re a

dded

in th

e fin

al st

ep.

BaT

iO3

Ba2

+ (aq)

+ C

2O42-

(aq)

BaT

iO3

BaC

2O4

(s)

Co2

+ (aq)

+ C

2O42-

(aq)

CoC

2O4

(s)

PE

I

BaT

iO3

+

+ ++

++

+

+

+

++

NbO

(O2)

Cit3

- (aq)

Bi(N

H3)

3Cit- (a

q)

BaT

iO3

+

+ ++

++

+

+

+

+

Bi(N

H3)

3Cit-

++

Bi(N

H3)

3Cit-

- -

-

--

-

--

Dis

solu

tion N

bO(O

2)C

it3-

NbO

(O2)

Cit3

-

146

ab

c

ed

Figu

res 5

.10a

-e.

TEM

imag

es sh

owin

g th

e m

orph

olog

ical

evo

lutio

n of

the

parti

cle

surf

ace

at e

ach

step

of t

he d

opin

g pr

oces

s: (a

) as

-syn

thes

ized

par

ticle

, (b)

oxa

lic a

cid

pass

ivat

ed p

artic

le, t

he in

sert

show

s a 2

-3 n

m su

rfac

e la

yer o

f bar

ium

oxa

late

, (c)

Co

dope

d pa

rticl

e, (d

) Co,

Nb

dope

d pa

rticl

e, a

nd (e

) ful

ly d

oped

par

ticle

whi

ch sh

ows t

he a

dditi

on o

f Bi r

esul

ts in

1-2

nm

dep

osits

on

the

surf

ace

of th

e pa

rticl

e.

147

Figu

re 5

.11.

EDS

spec

trum

of a

clu

ster

of d

oped

pla

tele

t par

ticle

s. T

he sp

ectru

m sh

ows t

he p

rese

nce

of th

e th

ree

dopa

nts C

o, N

b,

and

Bi.

The

C, C

u, a

nd F

e pr

esen

t are

due

to c

onta

min

atio

n fr

omei

ther

the

TEM

sam

ple

grid

or T

EM in

stru

men

t. Th

e in

sert

deta

ilis

add

ed to

show

the

pres

ence

of N

b an

d B

i bec

ause

the

sign

al to

nois

e ra

tio a

t low

er e

nerg

ies i

s too

smal

l ind

icat

e th

e pr

esen

ce o

f th

e do

pant

s.

148

Sam

ple

CoO

Nb 2

O5

Bi 2O

3C

oO:N

b 2O

5C

oON

b 2O

5B

i 2O3

CoO

:Nb 2

O5

10.

000.

005.

00N

/A<0

.01

<0.0

14.

46N

/A2

0.44

1.56

5.00

1:1

0.25

0.72

2.23

1:2.

93

1.10

3.90

5.00

1:1

0.25

1.29

2.23

1:5.

2

Pre

pare

d (w

t%)

Act

ual (

wt%

)

Tab

le 5

.2.

XR

F da

ta fo

r dop

ed p

late

let p

owde

r sam

ples

. Th

e da

ta sh

ows t

hat a

s the

CoO

and

Nb 2

O5

conc

entra

tions

incr

ease

the

actu

al c

once

ntra

tions

dev

iate

s mor

e fr

om th

e pr

epar

ed c

once

ntra

tion

due

to c

ompe

titiv

e ad

sorp

tion

of th

e do

pant

s on

the

surf

ace.

149

only the dopants specifically absorbed onto the powder surface, the doped powder was

centrifuged to separate the powder from the solution containing unadsorbed dopants.

Figures 5.12a, b, and c illustrate the difference in the CoO, Nb2O5, and Bi2O3

concentrations of the prepared and actual samples. A decrease in the actual dopant

concentration compared to the prepared concentration is present. For the sample with

only Bi2O3 (Sample 1) the drop is minimal compared to the other samples. For the other

samples, the measured dopant concentration is lower than the concentration of added

dopant.

The sample doped with only 5wt% Bi2O3 shows only a slight reduction in the

actual concentration to 4.46wt%. However, for samples with up to 10wt% dopant, the

results show that only 37% of the total dopant concentration is achieved. As the

CoNb2O6 concentration increases the concentration of CoO and Bi2O3 stabilize while the

Nb2O5 concentration continues to increase. This result suggests that the interaction

between the particle surface and the Nb complex is stronger than the other dopants. Not

only is the total dopant concentration lower than expected, but the Co:Nb ration is no

longer that of the desired formulation of 1:3.5, but varies from 1:2.9 to 1:5.2

The specific adsorption of the Nb and Bi complexes on the PEI coated surfaces

occurs by electrostatic attraction. Due to the electrostatic attraction the adsorption is

limited to monolayer coverage and therefore the adsorption typically behaves as a

Langmuir isotherm,

)exp(RTG

nxN abss Δ−= [5.2]

where N is the number of filled sited per area, n is the total number of sites per area, xs is

the mole fraction of dopant in solution, ΔGabs is the free energy of adsorption, R is the

150

Figu

re 5

.12a

, b, a

nd c

.X

RF

data

for d

oped

pla

tele

t pow

der s

ampl

es sh

owin

g th

at th

e ac

tual

con

cent

ratio

ns o

f the

CoO

, Nb 2

O5

and

Bi 2O

3de

viat

e fr

om th

e pr

epar

ed c

once

ntra

tions

exp

ecte

d Sa

mpl

e 1

whi

ch is

onl

y do

ped

with

5w

t% B

i 2O3

151

universal gas constant, and T is absolute temperature. With a Langmuir isotherm there

are fixed number of possible absorption sites on the powder surface (i.e. n is constant).

When multiple dopants are introduced to the suspension competitive absorption becomes

a problem. When the total dopant concentration is low (≤ 5wt%, Sample 1) this problem

does not exist because there are sufficient sites for adsorption of all of the dopants, but as

the dopant concentration increases the number of empty sites decreases and dopants must

compete for the available sites.

Increasing the number of possible adsorption sites would help overcome the low

adsorption densities. This is possible through the use of a higher surface area powder or

a highly branched polyelectrolyte with more ionizable side groups. A discussion of the

dispersion of the doped BaTiO3 is presented in Appendix B.

5.4 Conclusions

Oxalic acid was is a suitable passivation agent for the aqueous processing of

BaTiO3. Passivation occurs by the formation of a 2 to 3 nm layer of insoluble barium

oxalate on particle surfaces. Because of the low zeta potential provided with oxalate

passivation, a cationic polyelectrolyte, PEI is used to disperse the BaTiO3. Viscosity

measurements show the ability for excess oxalic acid and PEI to react and form a gel

network of amine oxalate, which is detrimental to rheological properties. When excess

oxalic acid is not present, the concentration of PEI affects the conformation of PEI on the

surface leading to bridging flocculation. The rheological properties show that

passivation/dispersion is achieved with oxalic acid and PEI, but there is a limited

concentration range for both reagents in which good dispersion is possible.

152

Using an aqueous solution-based approach platelet particles were doped with Co,

Nb, and Bi. The surface chemistry of the BaTiO3 particles was manipulated, resulting in

the selective deposition of the dopants on the particle surfaces. The development of a

engineered coating during doping was confirmed by surface charge, TEM, and EDS

analysis. Possible uses of these doped particles include the tape casting or electrophoretic

deposition of thin formulated BaTiO3 layers for capacitive applications.

153

References

1. S. Venigalla and J.H. Adair: Theoretical modeling and experimental verification of electrochemical equilibria in the Ba-Ti-C-H2O system. Chem. Mater. 11, (3), 589 (1999). 2. A. Neubrand, R. Linder, and P. Hoffman: Room-temperature solubility behavior of barium titanate in aqueous media. J. Am. Ceram. Soc. 83, (4), 860 (2000). 3. U. Paik and V.A. Hackley: Influence of solids concentration on the isoelectric point of aqueous barium titanate. J. Am. Ceram. Soc. 83, (10), 2381 (2000). 4. A.J. Moulson and J.M. Herbert, Electroceramics: Materials, properties, applications, 1st ed. (Chapman & Hall, London, 1990). 5. J.H. Adair and H.G. Krarup: The role of solution chemistry in understanding colloidal stability of ceramic suspensions. In Advances in Process Measurements for the Ceramics Industry, edited by A. Jillavenkatesa, and G.Y. Onoda, (American Ceramic Society, 1999) pp 205. 6. C.R. Baes and R.E. Mesmer, The Hydrolysis of Cations, 1st ed. (Wiley, New York, NY, 1976). 7. M.M. Lencka and R.E. Riman: Thermodynamic modeling of hydrothermal synthesis of ceramic powders. Chem. Mater. 5, (1), 61 (1993). 8. K. Osseo-Asare, F.J. Arriagada, and J.H. Adair: Solubility relationships in the coprecipitation synthesis of barium titanate: Heterogeneous equilibria in the Ba-Ti-C2O4-H2O system. In Ceramic Transactions, Ceramic Powder Science, edited by G.L. Messing, E.R. Fuller, Jr., and H. Hausner, (The American Ceramic Society, 1988) pp 47. 9. P. Bendale, S. Venigalla, J.R. Ambrose, E.D. Verink, and J.H. Adair: Preparation of barium-titanate films at 55-degrees-C by an electrochemical method. J. Am. Ceram. Soc. 76, (10), 2619 (1993). 10. M.C. Blanco-Lopez, B. Rand, and F.L. Riley: The properties of aqueous phase suspensions of barium titanate. J. Euro. Ceram. Soc. 17, 281 (1997). 11. U. Paik, S. Lee, and V.A. Hackley: Influence of barium dissolution on the electrokinetic properties of colloidal BaTiO3 in an aqueous medium. J. Am. Ceram. Soc. 86, (10), 1662 (2003). 12. B. Lee, S. Chung, and S.L. Kang: Necessary conditions for the formation of {111} twins in barium titanate. J. Am. Ceram. Soc. 83, (11), 2858 (2000). 13. M.C.B. Lopez, G. Fourlaris, and F.L. Riley: Interaction of barium titanate powders with an aqueous suspending medium. J. Euro. Ceram. Soc. 18, (14), 2183 (1998).

154

14. J.H. Adair, B.L. Utech, K. Osseo-Asare, and J.P. Dougherty, Solubility and phase stability of barium titanate in aqueous suspension, edited by J.P. Dougherty, and K. Wakino (in Proceedings of the Fifth US-Japan Seminar on Dielectric and Piezoelectric Ceramics, Kyoto, Japan, 1991), pp. 15. P. Duran, J. Tartaj, and C. Moure: Sintering behavior and microstructural evolution of agglomerated spherical particles of high-purity barium titanate. Ceram. Int. 29, 419 (2003). 16. Y. Koh, Y. Kong, S. Kim, and H. Kim: Improved low-temperature environmental degradation of yttria-stabilized tetragonal zirconia polycrystals by surface encapsulation. J. Am. Ceram. Soc. 82, (6), 1456 (1999). 17. T. Hanrath and B.A. Korgel: Chemical surface passivation of Ge nanowires. J. Am. Chem. Soc. 126, 15466 (2004). 18. R.P. Vasquez, B.D. Hunt, and M.C. Foote: Wet chemical passivation of YBa2Cu3O7-

x. J. Electrochem. Soc. 137, (7), 2344 (1990). 19. M. Nayak, M. Ando, and N. Murase, Passivation of CdTe nanoparticles by silane coupling agent assisted silica encapsulation, edited by H. Lin, and M. Singh (in 26th Annual Conference on Composites, Advanced Ceramics, Materials, and Structures: B, 23(4),Cocoa Beach, FL, 2002), pp 695. 20. C. Shih, C. Shih, Y. Su, L.H.J. Su, M. Chang, and S. Lin: Effect of surface oxide properties on corrosion resistance of 316L stainless steel for biomedical applications. Corr. Sci. 46, 427 (2004). 21. R.J. Jouet, A.D. Warren, D.M. Rosenberg, and V.J. Bellitto, Surface passivation of bare aluminum nanoparticles using perfluoroalkyl carboxylic acids, edited by R. Armstrong, N. Thadhani, W. Wilson, J. Gilman, and R. Simpson (in Synthesis, Characterization and Properties of Energetic/Reactive Nanomaterials, 800,Boston, MA, 2003), pp 67. 22. W.J. Merz: The electrical and optical behavior of BaTiO3 single-domain crystals. Phys. Rev. 76, (8), 1221 (1949). 23. B. Jaffe, W.R. Cook, and H. Jaffe, Piezoelectric Ceramics, 1st ed. (Academic Press, London, 1971). 24. G.H. Wiseman: Advanced manufacturing process for zinc oxide surge arrester disks. Key Eng. Mater. 150, 209 (1998). 25. S. Ural: Aggregate breakdown and aqueous processing of zinc oxide varistors. M.S. Thesis, The Pennsylvania State University, University Park, PA, (2003).

155

26. X. Liu: Structure-property relationships in submicron X7R dielectric materials. M.S. Thesis, The Pennsylvania State University, University Park, PA, (1999). 27. S.A. Bruno: Ceramic dielectric compositions and method for enhancing dielectric properties. US Patent # 5,082,811, (1992). 28. D.A. Anderson, J.H. Adair, D.V. Miller, J.V. Biggers, and T.R. Shrout: Surface chemistry effect on ceramic processing of BaTiO3 powder. In Ceramic Transactions, Ceramic Powder Science, edited by G.L. Messing, E.R. Fuller, Jr., and H. Hausner, (The American Ceramic Society: Westerville, OH, 1988) pp 485. 29. J.F. Fernandez, P. Duran, and C. Moure: Influence of the doping method on X7R based-dielectric capacitors. Ferroelectrics 127, 47 (1992). 30. T.J. Yosenick, D.V. Miller, R. Kumar, J.A. Nelson, C.A. Randall, and J.H. Adair: Synthesis of nanotabular barium titanate via a hydrothermal route. J. Mater. Res. 20, (4), 837 (2005). 31. D. Horn: Polyethylenimine - Physiochemical properties and applications. In Polymeric Amines and Ammonium Salts, edited by E.J. Goethals, (Pergamon Press, 1979) pp 333. 32. D.F.K. Hennings and B.S. Schreinemacher: Temperature-stable dielectric materials in the system BaTiO3-Nb2O5-Co3O4. J. Euro. Ceram. Soc. 15, (4), 463 (1994). 33. H. Chazono and H. Kishi: Sintering characteristics in BaTiO3-Nb2O5-Co3O4 ternary system: I, Electrical properties and microstructure. J. Am. Ceram. Soc. 82, (10), 2689 (1999). 34. Y. Kuromitsu, S.F. Wang, S. Yoshikawa, and R.E. Newnham: Interactions between barium titanate and binary glasses. J. Am. Ceram. Soc. 77, (2), 493 (1994). 35. I. Burn: Flux-sintered BaTiO3 dielectrics. J. Mater. Sci. 17, 1398 (1982). 36. Y. Narendar and G.L. Messing: Synthesis, decomposition and crystallization characteristics of peroxo-citrato-niobium: An aqueous niobium precursor. Chem. Mater. 9, 580 (1997). 37. G.H. Kirby, D.A. Harris, Q. Li, and J.A. Lewis: PAA-POE comb polymer dispersants for colloidal processing. Key Eng. Mater. 264-268, 161 (2004). 38. G.H. Kirby, D.A. Harris, Q. Li, and J.A. Lewis: Poly(acrylic acid)-poly(ethylene oxide) comb polymer effects on BaTiO3 nanoparticle suspension stability. J. Am. Ceram. Soc. 87, (4), 181 (2004).

156

39. U. Paik, V.A. Hackley, J. Lee, and S. Lee: Effect of poly(acrylic acid) and poly(vinyl alcohol) on the solubility of colloidal BaTiO3 in an aqueous medium. J. Mater. Res. 18, (5), 1266 (2003). 40. A. Hodgkinson, Oxalic acid in biology and medicine, 1st ed. (Academic Press, London, England, 1977). 41. M.M. Mandanas, W. Shaffer, and J.H. Adair: Aqueous processing and stabilization of manganese zinc ferrite powders via a passivation-dispersion approach. J. Am. Ceram. Soc. 85, (9), 2156 (2002). 42. R.A. Kimel and J.H. Adair: Aqueous degradation and chemical passivation of yttria-tetragonally-stabilized zirconia at 25 degrees C. J. Am. Ceram. Soc. 85, (6), 1403 (2002). 43. E.A. Mazurenko and B.I. Nabivants: Oxalto complexes of titanyl. Soviet Progress in Chemistry 33, (1), 86 (1967). 44. C. Boudaren, T. Bataille, J.P. Auffredic, and D. Louer: Synthesis, structure determination from powder diffraction data and thermal behavior of titanium(IV) [Ti2O3(H2O)2](C2O4).H2O. Solid State Sci. 5, 175 (2003). 45. J.H. Adair, E. Suvaci, and J. Sindel: Surface and colloid chemistry of advanced ceramics. In Encyclopedia of Materials: Science and Technology, edited by K.H.J. Buschow, R.W. Cahn, M.C. Flemings, B. Ilschner, E.J. Kramer, and S. Mahajan, (Elsevier Elsvier Science, Ltd., 2001) pp 8996. 46. R.J. Hunter, Zeta potential in colloid science: Principles and applications, 1st ed. (Academic Press, San Diego, CA, 1981). 47. R. Vaidhyanathan, S. Natarajan, and C.N.R. Roa: Synthesis of a hierarchy of zinc oxalate structures from amine oxalates. J. Chem. Soc., Dalton Trans. 5, 699 (2001). 48. B.A. Firth and R.J. Hunter: Flow properties of coagulated colloidal suspensions. III. The elastic flow model. J. Colloid Interface Sci. 57, (2), 266 (1976). 49. R.J. Hunter, R. Matarese, and D.H. Napper: Rheological behavior of polymer flocculated latex suspensions. Coll. Surf. 7, 1 (1983). 50. E. Poptoshev and P.M. Claesson: Forces between glass surfaces in aqueous polyethylenimine solutions. Langmuir 18, 2590 (2002). 51. T. Radeva and I. Petkanchinh: Electric properties and conformation of polyethylenimine at the hematite-aqueous solution interface. J. Colloid Interface Sci. 196, (1), 87 (1997).

157

52. J.P. Sadler and Z. Guo: Metal complexes in medicine: Design and mechanisms of action. Pure Appl. Chem. 70, (4), 863 (1998).

158

CHAPTER SIX

Electrophoretic Deposition of Hydrothermally Derived Barium Titanate Tabular

Nanoparticles with a Cationic Dispersant

6.1 Introduction

Demands on cellular and mobile technologies have motivated the miniaturization

of passive electronic components. Currently, tape casting is used in the fabrication of

passive electronic components with layer thicknesses of 1 μm.1 The increase in

volumetric efficiency of multilayer device with decreasing layer thickness is well

known.2 Therefore, an enhanced thin film processing technique for the lay down of thin

dielectric layers would be an effective way to increase device performance.

Electrophoretic deposition (EPD) is an ideal technique for the deposition of thin metal

and dielectric layers in passive electronic components.3

EPD is a colloidal processing technique in which the driving force for deposition

is an applied electric field. Electrodes are immersed in a particle suspension and an

electric field is applied. The charged particles in suspension migrate towards the

oppositely charged electrode. Once the particles reach the electrode they rearrange and

deposit, forming a uniform coating on the electrode. Films prepared using EPD exhibit

high green density and homogenous microstructures.4 The first observation of EPD was

by Harsanyi in 1927.5 Since then, many researchers have investigated EPD as a

processing route for a variety of complex structures and engineered materials.6-11

159

EPD offers several distinct advantages over other thin film processing and coating

techniques. For example, EPD can be used to deposit films with controlled thickness

ranging from millimeters to nanometers. Recently, it has been shown that EPD can be

used to deposit films as thin as 400 nm from well-dispersed Ag particle suspensions.3 In

addition to excellent thickness control, EPD can be used to deposit a wide variety of

materials. Any metal, ceramic, semiconductor, and organic particle that is chemically

stable and develop a surface charge in a suitable solvent can be deposited using EPD.

EPD films conform to the underlying electrode when depositing, therefore combining

EPD with other materials fabrication techniques (i.e. photolithography) allows for the

deposition of complex structures. Recently, von Both and Hausselt showed the feasibility

of complex three dimensional coatings by using EPD to deposit alumina on three

dimensional electrodes patterned using photolithography.6

The EPD process is relatively simple when compared to other thin film processing

techniques, some of which may require vacuum chambers or atmospheric control.

Because of this fact, EPD is generally less expensive than other techniques. In EPD, film

chemistry is defined during particle synthesis and dispersion. In other techniques, such

as sol-gel, chemical-vapor-deposition, metal oxide chemical-vapor-deposition, and

physical-vapor-deposition, the chemistry is controlled during the process. Therefore,

there are no processing variables during EPD which affect the film chemistry.

With the increased interest in nanoparticles and the drive towards reduced layer

thickness in MLCCs, EPD has become an ideal processing technique for the deposition of

thin BaTiO3 layers. The kinetics and mechanisms of EPD along with the surface

chemistry of the depositing particles are dependent on the chemistry of the solution phase

160

of the suspensions. It is therefore important to investigate the interdependence of

deposition and suspension stability. The current study focuses on the effect of solution

chemistry (i.e. dispersant concentration, ionic strength, conductivity, etc…) on the

kinetics of EPD and evolution of the resultant films.

6.2 Theoretical Background – Mechanisms of EPD

EPD can be dissected into a three step process: (1) dispersion, (2) migration, and

(3) deposition. In colloidal processing the green microstructure of materials is highly

dependent on the state of dispersion prior to consolidation.12-14 Therefore, for EPD to

produce dense homogenous layers a well-dispersed particulate suspension must be

available. Significant research effort has focused on understanding the dispersion of a

variety of materials systems.15 Migration of particles from the bulk solution to the

electrode surface occurs via electrophoresis, which has been well-studied.16-18 During

deposition the particle suspension is destabilized in a controlled manner at the electrode.

Unlike dispersion and electrophoresis, the mechanisms of deposition have not been

thoroughly studied and remain in some doubt. Recently, van Tassel completed an

overview of the mechanisms of electrophoretic deposition and identified 11 different

deposition mechanisms.19 Each mechanism is defined by the process used to destabilize

the suspension and consolidate the particle into a film at the electrode. Each mechanism

is highly dependent on the solution chemistry of the particulate suspension.

The first attempt to model the deposition process was that of Hamaker and

Verwey.20 Hamaker and Verwey observed that suspensions used for EPD produced

sediment layers if allowed to settle. They reasoned that since the EPD process was

161

similar to sedimentation, that an external force causes the particles to overcome the

repulsive interaction and agglomerate. It was also believed that particles at the solution-

film interface exert an extra force on the particles in the film close to the electrode

leading to an increase in density.

Koelmans and Overbeek theorized that increased ionic strength at the electrode,

due to electrochemical reactions, decreased particle zeta potentials, which lead to

agglomeration of the particles, forming a deposit.21 Brown and Salt calculated the

electric field necessary to overcome the repulsive interaction of the particles and their

predictions agreed well with oxide materials.22 They also deposited metal particles via

the addition of aluminum chloride. The aluminum chloride was believed to form

aluminum hydroxide as a polymeric network that percolated throughout the deposit.

Sarkar and Nicholson observed the deposition of particles on a dialysis membrane

located away from the electrode.23 From this observation they claimed that

electrochemical reactions at the electrode have no role in the deposition process. They

theorized that an electrophoretic force similar to that proposed by Hamaker and Verwey,

aids in overcoming the repulsive interactions between particles. The electrophoretic

force, in combination with double layer polarization, due to electrophoresis, permit the

particles to agglomerate. Recently Kershner et al.24 stated electrochemical reactions that

occur at the deposition electrode change the local pH and drive the suspension pH to the

isoelectric point (IEP) of the suspension.

As van Tassel has explained (see Table 2.4 of Ref 19) most or combinations of

these mechanisms are responsible for deposition, but with such a wide range of

mechanisms it is difficult to fully elucidate the deposition process. Added to the

162

complexity is that each mechanism has an underlying effect on the solution chemistry in

a region located near the electrode. Those properties that control the dispersion and

electrophoresis of the particle in bulk of suspension change greatly near the electrode and

without a priori knowledge of the changes, a basic understanding of the deposition

mechanisms is difficult.

6.3 Materials and Methods

Suspensions for EPD were prepared from two different BaTiO3 powders: a

commercially available equiaxed BaTiO3 (BT-01, Sakai Chemical Company, Osaka,

Japan) with a particle size of 88.2 nm, as determined by BET, and a hydrothermally

derived platelet BaTiO3. The platelet powder has a median thickness of 5.8 ± 3.1 nm

with a median diameter of 27.1 ± 12.3 nm and a surface area of 10.5 m2/g. The details of

synthesis and physical properties of the platelet powder have been described previously.25

Electrostatic, electro-steric, and passivation/dispersion schemes were all used in

the preparation of 0.1vol% suspensions of equiaxed powder. Ethanol (200 proof,

Pharmco, Brookfield, CT) was used as the solvent in all dispersion experiments.

Solutions of 0.1 M hydrochloric acid (HCl) (35-37% JT Baker, Phillipsburg, NJ) and

tetraethylammonium hydroxide (TMAOH) (35 wt% solution in methanol, Aldrich

Chemical Company, Milwaukee, WI) were used to adjust pH and to provide surface

charge in the electrostatic dispersion experiments. The pH of the suspensions was

measured using ion-selective field effect transistor (IS-FET) probes (Sentron Hotline

Probe, RL Instruments, Manchoung, MA) which were calibrated using NIST traceable

aqueous pH standards with nominal values of 4, 7, and 10. The electrophoretic mobility

163

of the suspensions was measured using electrophoretic light scattering (ZetaPALS,

Brookhaven Instrument Corp., Holtsville, NY).

The use of ethanol for dispersion results in the Debye length (κ-1) of the electrical

double layer and particle radius (a) having similar length scales. Resulting is

intermediate κa values, which were outside the limits of both the Smoluchowski and

Hückel approximations for the calculation of zeta potential from electrophoretic mobility

data. The zeta potential was calculated from the measured motilities using the approach

outlined by O’Brien and White.26 For the electrosteric dispersion experiments, anionic

ammonium acrylate dispersants (Darvan C and 821A, R.T. Vanderbilt, Norwalk, CT) and

a cationic polyelectrolyte polyethylenimine (PEI) (25,000 MW, Aldrich Chemical

Company, Milwaukee, WI) were used as dispersants.

For the passivation/dispersion experiments oxalic acid dihydrate (HOx) (Fisher

Scientific, Fair Lawn, NJ) was used as a passivation agent. HOx was added in

concentrations of 1, 3, 4 and 5 wt% with respect to solids present (w/w). Prior to the

addition of the powder, the pH of the oxalic acid solution was adjusted to greater than pH

13 with TMAOH. This was done to ensure complete dissociation of the acid so that it

can allow for the reaction of the oxalate ion with the BaTiO3 surface. After passivation,

the HOx/BaTiO3 suspensions were washed twice with 200 proof ethanol using

centrifugation (7500 rpm, 10min) to remove unreacted HOx and TMAOH. After

washing, the particles were redispersed using an acidified (1mM HCl) PEI/ethanol

solution with PEI concentrations ranging from 0.1 to 2 w/w. Suspensions of the

hydrothermally synthesized BaTiO3 were dispersed using the passivation/dispersion

technique approach described above for comparison to the equiaxed BaTiO3 dispersions.

164

Suspensions of 0.1vol% hydrothermal BaTiO3 with 5w/w HOx and 0.1w/w PEI

were prepared using lithium chloride (LiCl) (99%, Aldrich Chemical Company,

Milwaukee, WI), an indifferent electrolyte. Suspensions were prepared with 0, 0.5, 1, 2,

and 5 mM LiCl. Suspensions of 0.1wt% Pt (0.15-.045 micron, Alfa Aesar, Ward Hill,

Ma) were prepared from the filter supernatant of the hydrothermal BaTiO3 suspensions to

ensure that the platelet and Pt suspensions had identical solution chemistries. The zeta

potential of the BaTiO3 and Pt suspensions were measured as described above.

Cyclic voltammetry (CV) was performed on solutions of ethanol with HCl, LiCl,

and PEI, to determine the effect of ionic strength and dispersant concentration on the

electrochemistry of EPD. Experiments were conducted using a computer controlled

potentiostat (SI 1287 Electrochemical Interface, Solartron, UK) with CorrWare software,

a circular Pt working electrode (1.6 mm diameter, MF-2031, Bioanalytical Systems, West

Lafayette, IN), a Pt mesh (25 mm x 25mm, 52 woven mesh, Alfa Aesar, Ward Hill, MA)

counter electrode and a saturated calomel electrode (SCE) reference. Scans were

performed at a rate of 100 mV/sec from +3 to -3 V versus SCE. Impedance spectroscopy

was also performed on solutions of pure ethanol with 10-3 M HCl with varying amounts

of PEI. The electrode system was changed to a parallel plate conductivity probe with Pt

electrodes for impedance spectroscopy. Scans were run from 106 to 10-3 Hz with an

applied voltage 100 mV. Data was analyzed and fit to a user designed equivalent circuit

using commercially available software (ZView, Ver 2.2, Scribner Associate, Inc.,

Southern Pines, NC).

A series of Ba2+ dissolution studies were conducted on the equiaxed powder to

determine the chemical stability of the powder in ethanol. Suspensions of 1wt% BaTiO3

165

in ethanol were prepared by pre-adjusting the pH of the ethanol with 0.1 M HCl solutions

to pH 2.4, 4.5, 6.5, and 8.3 followed by the addition of the powder. Stirred suspensions

were equilibrated for 24 hrs. The suspensions were centrifuged and the supernatant was

filtered through a 0.22 micron syringe filter. Directly coupled plasma – emission

spectroscopy (DCP-ES) (Spectraspan III, Spectrametrics Inc., Andover, MA) was used to

determine the Ba2+ concentration in solution at 455.4 nm. Barium standards were

prepared by serial dilution from NIST traceable stock solutions.

MylarTM sputtered with Pt was used as a substrate for most deposition

experiments. Samples for TEM cross-section were deposited on single crystal Si wafers

sputtered with Ag/Pd. EPD of the nanoplatelet suspensions was performed under

constant voltage conditions in a TeflonTM cell with vertical electrodes spaced 2.5 cm

apart. Voltage was applied with current monitored using a source meter (Model 2410,

Keithley Instruments, Cleveland, OH) with deposition times ranging from 2 to 10 min.

After deposition, samples were dried in air at room temperature for at least 2 hrs. then

samples of known area were cut from the dried films and the mass was measured to

determine the deposition rate. Cross-sectioned samples for TEM analysis were prepared

by the small angle cleavage technique27 (SACT). X-ray diffraction (XRD) (Scintag Pad

V, Thermo-ARL, Dearborn, MI) was performed on films deposited from platelet powder

suspensions. Continuous scans were run from 15-75° 2θ at a rate of 0.2°/sec.

166

6.4 Results and Discussion

6.4.1 Dispersion

Since EPD is a colloidal processing technique the green density and

microstructure of deposited films is dependent on the state of dispersion of the particles

prior to deposition. In addition to affecting the microstructure the agglomerate size in

suspension limits the minimum layer thickness that can be deposited. Therefore a study

of the dispersion of BaTiO3 in a suitable solvent system prior to deposition was required.

Because electrophoresis is necessary for the migration of the particles to the electrode, an

electrostatic component to the dispersion must be present. Electrostatic and electrosteric

dispersion require a protic solvent that can support charge and promote solution-solid

reactions to develop surface charge.

While water is an ideal solvent choice based on dispersion, electrochemical

reactions between the water and the electrodes prevent the selection of water. Water is

readily reduced at the cathode during deposition, leading to the generation of hydrogen

gas bubbles in the deposit.23 Therefore, non-aqueous solvents are typically used for EPD

because of their low water content, but a relatively polar solvent is necessary for good

dispersion. Ethanol is an excellent choice because it has relatively low water content, yet

supports charge well and allows for electrostatic dispersion.

The chemical instability of BaTiO3 in aqueous environments is well-

documented.28-30 It is also known that surface passivation is necessary to limit Ba2+

dissolution during the aqueous processing BaTiO3.31, 32 Because ethanol it still relatively

polar and can contain significant amounts of water it was necessary to determine the

chemical stability of BaTiO3 in ethanol. Table 6.1 is DCP-ES data from dissolution

167

pHB

a2+ C

once

ntra

tion

(ppb

)2.

315

84.

539

56.

515

88.

3<7

9

Tab

le 6

.1.

Con

cent

ratio

n of

Ba2+

in p

ure

etha

nol a

s det

erm

ined

by

dire

ct c

oupl

e pl

asm

a em

issi

on sp

ectro

scop

y. R

esul

ts sh

ow th

e di

ssol

utio

n of

Ba2+

in p

ure

etha

nol i

s not

pre

vale

nt a

nd th

eref

ore

surf

ace

pass

ivat

ion

is n

ot n

eces

sary

prio

r to

disp

ersi

on.

168

experiments of the equiaxed powder in ethanol. The results indicate that Ba2+ dissolution

is not prevalent in ethanol, and that surface passivation of BaTiO3 with oxalic acid is not

required in ethanol. Therefore, an electrostatic approach to dispersion was first

attempted. Figure 6.1 gives the zeta potential as a function of HCl concentration for the

equiaxed powder. Small additions of HCl result in a high positive zeta potential, but as

the HCl concentration increases, the zeta potential decreases. The decrease is due to the

collapse of the electrical double layer as the ionic strength of the solution increases.15

Suspensions prepared using an electrostatic dispersion approach in ethanol were not

stable and therefore not suitable for EPD.

Since an electrostatic approach was not successful in creating stable dispersion, an

electrosteric approach was used. Electrosteric dispersion uses a charged polymer that

absorbs onto the particle surface and provides both surface charge and a steric

repulsion.15 The steric component limits the distance of closest approach of particles

reducing particle to particle responsible for aggregation. Convnetional anionic

polyacrylate-based dispersants were initially used because of the positive surface charge

of BaTiO3, but were insoluble in ethanol, and therefore not suitable for dispersion.

Suspensions were next prepared with a cationic polyelectrolyte (PEI), but PEI did not

adsorb due to the positive surface charge of the native BaTiO3 surface. Again the

suspensions prepared were not stable or suitable for EPD.

With both previous attempts at dispersion unsuccessful, a third approach was

attempted. PEI is still a good choice as a dispersant because it provides a high degree of

surface charge and it is soluble in ethanol; however PEI requires a negative surface

charge for absorption. Therefore, to use PEI it is necessary to condition the BaTiO3 and

169

Figu

re 6

.1.

Zeta

pot

entia

l of e

quia

xed

BaT

iO3

pow

der s

how

s tha

t the

zet

a po

tent

ial d

ecre

ases

as t

he H

Cl c

once

ntra

tion

incr

ease

s du

e to

incr

ease

d io

nic

stre

ngth

in so

lutio

n. A

lthou

gh h

igh

zeta

pote

ntia

l val

ues a

re o

bser

ved,

susp

ensi

ons p

repa

red

by e

lect

rost

atic

di

sper

sion

wer

e no

t sta

ble

and

ther

efor

e no

t sui

tabl

e fo

r dep

ositi

on.

170

create a negative surface. The passivation/dispersion is an approach used previously in

the aqueous dispersions of BaTiO3 that limits Ba2+ dissolution through the use of a

passivation agent.32 Passivation of the BaTiO3 surface is achieved by the addition of

oxalic acid (HOx) to form a thin layer of barium oxalate (BaC2O4) on the surface.

Barium oxalate has a stable negative surface charge suitable for dispersion by the

addition of PEI. In the current work HOx is not used to passivate the surface, because

ICP results show that Ba2+ dissolution is negligible, but to create a stable negative surface

charge for the adsorption of PEI. The zeta potentials as a function of PEI concentration

are shown in Figure 6.2. As the HOx concentration increases, the surface charge

becomes negative and with the addition of the PEI the surface charge becomes positively

charged.

To quantitatively measure the state of dispersion of particulate suspensions, a

figure of merit called the average agglomeration number, AAN(50), can be calculated

using the following equation33,

3

3)50()50(ESD

DVolume

VolumeANN

BET

eringLightScatt == [6.1]

where D(50) is the medium particle of the particle suspension measured by light

scattering, and ESD is the equivalent equiaxed diameter calculated from the BET surface

area. The ANN(50) represents the average number of primary particles in an

agglomerate in suspension and is based on the assumption of ideal equiaxed particles. In

general, ANN(50) ≤ 10 is considered a well-dispersed system.33 For the equiaxed

powder suspensions an ANN(50) of 7.7 was the best value and was achieved with 5 w/w

HOx and 2 w/w PEI. Because of its high degree of dispersion this suspension was used

in the preliminary deposition experiments.

171

Figu

re 6

.2.Z

eta

pote

ntia

l of e

quia

xed

BaT

iO3

as a

func

tion

of P

EI c

once

ntra

tion

for d

iffer

ing

conc

entra

tions

of o

xalic

aci

d w

ith

incr

easi

ngly

neg

ativ

e su

rfac

e ch

arge

. Th

e ne

gativ

e su

rfac

e ch

arge

is su

itabl

e fo

r the

ads

orpt

ion

of a

cat

ioni

c di

sper

sant

, PEI

.A

s th

e PE

I con

cent

ratio

n in

crea

ses t

he si

gn o

f the

surf

ace

char

ge re

vers

es.

A P

EI c

once

ntra

tion

of 2

w/w

resu

lts in

zet

a po

tent

ial

valu

es o

f app

roxi

mat

ely

80 m

V.

All

susp

ensi

ons p

repa

red

had

an H

Cl c

once

ntra

tion

of 1

0-3M

.

172

The same passivation/dispersion approach was used for the dispersion of the

hydrothermal BaTiO3, and yielded similar results, well-dispersed suspension suitable for

EPD. Figure 6.3 summarizes the particle size distributions for the hydrothermal powder

with 5w/w HOx and varying PEI. Improved dispersion is observed in suspensions with

less PEI; the optimum dispersion being in the presence of 0.25 w/w PEI, yielding a D(50)

= 16.6 nm. With an ESD of 97.3 nm, an ANN(50) of 0.005 is calculated. Due to the

anisotropic nature of the particles an analysis of dispersion based on ANN(50) is not

valid.

6.4.2 EPD

6.4.2.1 Kinetics

The kinetics of deposition is rate-limited by the electrophoretic mobility, and

therefore the variables that affect electrophoresis affect the kinetics of deposition.

Electrophoresis is a balance of the forces on a particle due to the applied electric field and

the viscous drag on the particle as it travels through the fluid. The velocity of a particle

during electrophoresis was calculated by Smoluchowski16, Hückel17, and Henry18 and is

equal to:

ηεζECv = [6.2]

where η is the viscosity of the solvent, ζ is the zeta potential of the particle, ε is the

dielectric constant of the solvent, E is the applied field, and C is a constant ranging from

2/3 to 1; with the velocity independent of particle size. The effective electric field is that

of the bulk of the solution and not the net field as determined by the applied voltage

divided by the interelectrode spacing. The effective field is dependent on a variety of

factors including the solution chemistry, specifically the solution conductivity, and the

173

Figu

re 6

.3.

Parti

cle

size

dis

tribu

tion

for t

he H

Ox/

PEI d

ispe

rsed

pla

tele

t par

ticle

s sho

ws t

hat a

PEI

con

cent

ratio

n of

onl

y 0.

25w

/w

resu

lts in

the

best

dis

pers

ion

with

a m

edia

n pa

rticl

e si

ze, D

(50)

, of 1

6.6

nm.

Low

PEI

con

cent

ratio

ns d

o no

t pro

vide

eno

ugh

surf

ace

char

ge fo

r goo

d di

sper

sion

whi

le h

igh

PEI c

once

ntra

tions

can

resu

lt in

brid

ging

floc

cula

tion

whi

ch d

egra

des d

ispe

rsio

n.

All

susp

ensi

ons p

repa

red

had

2w/w

HO

x an

d 10

-3M

HC

l.

174

structure of the electrode-solution interface. To fully investigate the kinetics of EPD, the

electrochemistry and structure of the electrode-solution interface must be understood.

Figures 6.4a and b are a simple circuit diagram for the EPD cell and a schematic

of the corresponding ideal Cole-Cole plot for the circuit. Rexp and Ccell are the resistance

of the experimental setup and the capacitance of the EPD cell, respectively. Rsol is the

solution resistance and is due to the mobile charged species in solution. Cdl is the double

layer capacitance associated with the electrode-solution interface. Rtran is the charge

transfer resistance and is due to the passage of charge from the electrode to the solution.

Under DC conditions the impedances of Ccell and Cdl are infinite and the circuit diagram

reduces to three resistors in series, Rexp, Rsol, and Rtran. Since Rexp is constant and solely

dependent on the experiment setup, Rsol and Rtran are the two variables that affect the

effective electric field in the bulk of suspension. Changes in either solution resistance or

transfer resistance should affect the kinetics of deposition.

In the initial deposition experiments on equiaxed powders, no deposition was

observed for applied voltages as high as 500V. Low current densities (<10 nA/cm2) were

observed during deposition at lower applied voltages (20 V), indicating there was a high

resistance element present in the EPD cell. The solution resistance is easily measured

with a conductivity probe, whereas it is necessary to use a different electrochemical

technique to measure the electron transfer resistance and model the solution-electrode

interface.

To determine the reason for the reduced kinetics, cyclic voltammetry (CV) and

impedance spectroscopy were used to investigate the solution-electrode interface. Since

it is possible for any or all constituents of the solution to interact with the electrode

175

Figu

re 6

.4a

and

b.(a

) Ide

al e

quiv

alen

t circ

uit f

or th

e EP

D c

ell.

Rex

pan

d C

cell

are

the

expe

rimen

tal s

etup

resi

stan

ce a

nd

capa

cita

nce

of th

e EP

D c

ell,

resp

ectiv

ely.

Bot

h ar

e de

pend

ent o

n th

e ex

perim

enta

l set

up a

nd re

mai

n co

nsta

nt.

Rso

lis t

he so

lutio

n re

sist

ance

, Cdl

is th

e ca

paci

tanc

e of

the

elec

trode

dou

ble

laye

r, an

d R

tran

is th

e el

ectro

n tra

nsfe

r res

ista

nce

of e

lect

roch

emic

al

reac

tions

. (b)

Sch

emat

ic re

pres

entin

g th

e id

eal C

ole-

Col

e pl

ot fo

r the

equ

ival

ent c

ircui

t.

176

surface34 it is necessary to analyze samples with individual constituents and combinations

of constituents in solution to deconvolute the intrinsic contributions of each. CV was

performed on solutions of ethanol, ethanol with 10-3 M HCl, and ethanol with 10-3 M HCl

and 1wt% PEI to check for the presence of electrochemical reactions. Figure 6.5 is the

cyclic voltammogram for the three solutions. The solution of pure ethanol shows no

observable reactions. When HCl is added two reactions are observed:

)(222 gHeH −+ + [6.3]

−− ++ OHHeOH g 222 )(22 [6.4]

On the addition of PEI to the solution the two electrochemical reactions in Equations 6.3

and 6.4 are no longer observed. Electrochemical reactions are necessary for charge

transfer between the electrode and solution. Without electrochemical reactions the

transfer resistance increases.35 CV shows that PEI prevents electrochemical reactions,

but since CV cannot quantify the change in the transfer resistance; it is necessary to use

impedance spectroscopy to measure the transfer resistance. Figure 6.6 is a Cole-Cole

plot for a solution of 10-3 M HCl in ethanol. In a circuit diagram, a capacitor and resistor

in parallel give an ideal Debye relaxor that appears as a semi-circle on a Cole-Cole plot.36

However, in Figure 6.6 the center of the second semi-circle is depressed below the x-axis.

On a Cole-Cole plot a depressed semi-circle is due to the replacement of the ideal

capacitor in a Debye relaxor with a constant phase element (CPE).37 The impedance of a

CPE is,

nCPE iCZ

)(1ω

= [6.5]

where C is the capacitance, ω is the angular frequency, and n is the CPE exponent.

177

Figu

re 6

.5.

Cyc

lic v

olta

mag

ram

mfo

r thr

ee e

than

ol so

lutio

ns c

onta

inin

g H

Cl a

nd P

EI.

The

addi

tion

of 1

mM

HC

l sho

ws t

he

evid

ence

of t

wo

elec

troch

emic

al re

actio

ns th

at o

ccur

at t

he c

atho

de b

oth

of w

hich

hav

e a

prof

ound

eff

ect o

n th

e pH

of t

he so

lutio

n ne

ar th

e ca

thod

e. T

he p

rese

nce

of P

EI in

hibi

ts th

e el

ectro

chem

ical

reac

tions

by

adso

rbin

g on

to th

e el

ectro

de a

nd in

crea

sing

the

elec

tron

trans

fer r

esis

tanc

e at

the

cath

ode.

178

Figu

re 6

.6.

Col

e-C

ole

plot

for a

10-3

M H

Cl s

olut

ion

in e

than

ol. T

he c

ente

r of t

he se

cond

sem

i-circ

le is

dep

ress

ed b

elow

the

x-ax

is in

dica

ting

the

Cdl

is n

ot a

n id

eal c

apac

itor b

ut a

con

stan

t pha

se e

lem

ent,

whi

ch is

due

the

roug

hnes

s of t

he e

lect

rode

surf

ace.

179

A CPE is representative of inhomogenieties or non-ideal systems and is due to a variety

of aspects such as a rough electrode surface38, the inhomogeneous adsorption of organics

on the electrode39, or a distribution of reaction rates.40

Table 6.2 is a list of solution and transfer resistances for ethanol solutions with

increasing PEI concentrations. For pure ethanol both the solution resistance and transfer

resistance are large due to a few carriers and electrochemically active species. As

expected, HCl results in a decrease in the values of both. Adding PEI reduces the

solution resistance and results in an initial rise in the transfer resistance follow by a

decrease. The n value is fairly constant between 0.8 and 0.9 indicating a rough electrode

surface.38 A decrease in solution resistance must be due to an increase in carrier density,

valence or mobility. Prior to adding it to solution, PEI is charge neutral. It is believed

that when PEI is added to solution protons in solution associated with the amine groups

of the PEI. Since the concentration and valence of the protons is constant whether in

solution or associated with PEI, the drop in resistance must be due to an increase in

mobility.

When HCl is added to solution the protons associate (i.e. there are no free proton

in solution) with water to form hydronium ions. When HCl is added to ethanol, the

protons react with ethanol to form ethoxonium ions.41 Without water present, proton

mobility occurs by a hopping mechanism where the proton hops from ethanol to ethanol

molecule. However, when water is present in ethanol the formation of hydronium ions is

favored and the proton mobility becomes limited by the electrophoresis of the hydronium

ion. When PEI is added to the solution there is apparent increase in the proton mobility.

The mechanism responsible is not known and further investigation is beyond the scope of

180

Sol

utio

nR

sol (

)C

dl ( μ

F)

nR

tran

s (M

Ω)

Pur

e E

than

ol1.

470

4.15

0.80

7.01

10-3

M H

Cl

0.13

34.

480.

825.

5210

-3M

HC

l w/ 0

.01w

t% P

EI

0.10

35.

580.

897.

2610

-3M

HC

l w/ 0

.05w

t% P

EI

0.05

85.

790.

907.

3410

-3M

HC

l w/ 0

.1w

t% P

EI

0.04

36.

230.

906.

3010

-3M

HC

l w/ 0

.5w

t% P

EI

0.01

46.

770.

904.

9910

-3M

HC

l w/ 1

wt%

PE

I0.

009

5.84

0.88

4.76

Tab

le 6

.2.

List

of E

PD c

ell v

aria

bles

with

incr

easi

ng P

EI c

once

ntra

tion.

The

solu

tion

resi

stan

ce d

ecre

ases

as P

EI in

crea

ses

beca

use

the

prot

on m

obili

ty is

incr

ease

d. T

he tr

ansf

er re

sist

ance

incr

ease

s in

the

pres

ence

of s

mal

l con

cent

ratio

n of

PEI

. In

a

solu

tion

of 0

.01w

t%, P

EI 9

8.6%

of t

he v

olta

ge d

rop

occu

rs a

t the

elec

trode

-sol

utio

n in

terf

ace.

181

the current work. The transfer resistance of the 10-3 M HCl appears large because only

100 mV was applied during the measurement. When a voltage large enough to initiate

electrochemical reactions is applied the transfer resistance will decrease. This is not true

for the samples with PEI because CV shows that PEI inhibits the electrochemical

reactions.

The actual concentration of unabsorbed PEI in solution was not measured.

However, the PEI concentration of the as prepared suspension was 0.01w/o, but from zeta

potential measurement it known that PEI is adsorbed to the surface so the actual

concentration of free PEI must be less. From the data in Table 6.2 the solution resistance

would be approximately 0.1 MΩ with a transfer resistance of 7.3 MΩ, which results in

98.6% of the voltage drop in the cell occurring at the electrode-solution interface.

It is apparent that PEI interacts with the electrode and stops electrochemical

reactions. When any of the solution constituents adsorbs onto the electrode and hinders

or stops electrochemical reactions, it is referred to as inhibition.42, 43 Organic substances

with amine groups have been noted to be effective cathodic inhibitors.44-46 Adsorbed

molecules which inhibit electrochemical reactions create a layer on the electrode surface

that limits the distance of closest approach of the electrochemically active species in

solution.42 Figure 6.7 is a schematic showing the effect of excess PEI on deposition and

how PEI inhibits electrochemical reactions and deposition. For an electrochemical

reaction to occur it is necessary for the electrons at the surface of the electrode to be

transported through the adsorbed layer to reduce the electrochemical species.

In the current case it is believed that excess PEI remaining in the solution phase

after dispersion has a two fold effect on the kinetics of deposition. The PEI migrates

182

Figu

re 6

.7.

Sche

mat

ic sh

owin

g th

e ef

fect

of e

xces

s PEI

on

the

elec

troch

emis

try o

f dep

ositi

on. W

hen

exce

ss P

EI is

pre

sent

it

cont

amin

ates

the

elec

trode

and

inhi

bits

ele

ctro

chem

ical

reac

tions

and

par

ticle

dep

ositi

on.

At a

reas

of t

he e

lect

rode

una

ffec

ted

by

PEI e

lect

roch

emic

al re

actio

ns o

ccur

s and

par

ticle

dep

ositi

on o

ccur

s. If

the

PEI c

once

ntra

tion

is to

o la

rge

the

entir

e el

ectro

deca

n be

con

tam

inat

ed a

nd d

epos

ition

is c

ompl

etel

y in

hibi

ted.

++++

+

+ +

++

++

++++

+

+ +

++

++

++++

+

+ +

++

++

++++

+

+ +

++

++

Pt Electrode

H+

e- e-H

+

H2(

g)

++++

+

+ +

++

++

++++

+

+ +

++

++

++

++

++ +

++

++

++

++

+

++

++

++

++++

+

+ +

++

++

++++

+

+ +

++

++

BaT

iO3

w/P

EI

PEI

e- e-

H+

H+

++++

+

+ +

++

++

Dep

osite

d Pa

rtic

le

++++

+

+ +

++

++

183

towards the cathode, along with the particles, during deposition, and adsorbs to the

electrode inhibiting electrochemical reactions raising the transfer resistance. The

unabsorbed PEI also decreases the solution resistance, and under constant voltage

conditions decreasing the solution resistance and increasing the transfer resistance results

in a drop in the effective field in the bulk of the suspension lowering the electrophoretic

velocity.

To avoid the problem of PEI inhibition, suspensions were prepared with less PEI

and deposited. Deposition was observed in suspensions with 0.5 and 1 w/w PEI. Similar

deposition experiments were conducted with suspension prepared from the platelet

BaTiO3. Suspensions were prepared with 0.1 to 2 w/w PEI and current was monitored

during deposition. Figure 6.8 is a plot of current and deposition rate as a function of PEI

for depositions at a constant applied voltage of 20 V. It is apparent that there is a direct

correlation between current and deposition rate. Again at high PEI concentrations

inhibition is observed, but at lower PEI concentrations the current density increases as

does the deposition rate. Partial inhibition is possible and even small quantities of

unabsorbed PEI will lower the current density.

Adsorption of a charged species from solution typically behaves as a Langmuir

isotherm which is limited to monolayer coverage with a fixed number of adsorption

sites.15 As PEI concentration increases surface coverage increases and PEI remains in

solution due to a lack of suitable adsorption sites. This is based on the fact that the zeta

potential of the suspensions does not continue to increase as the PEI concentration is

increased beyond 1w/w, as seen in Figure 6.2. As the overall PEI concentration is

increased the concentration of unabsorbed PEI increases and leads to enhanced inhibition

184

Figu

re 6

.8.

Dep

ositi

on c

urre

nt a

nd ra

te a

s a fu

nctio

n of

PEI

con

cent

ratio

nus

ed fo

r dis

pers

ion.

As t

he P

EI c

once

ntra

tion

incr

ease

s th

e cu

rren

t and

dep

ositi

on ra

te d

ecre

ases

due

to th

e pr

esen

ce o

funa

bsor

bed

PEI o

n so

lutio

n. D

urin

g de

posi

tion

the

exce

ss P

EI

abso

rbs o

nto

the

cath

ode

and

inhi

bits

ele

ctro

chem

ical

reac

tions

dec

reas

ing

the

curr

ent.

185

of deposition. The problem of inhibition leads to a trade-off in the dispersion and

deposition of the particles. If the PEI concentration is too large then deposition is

inhibited whereas, if the PEI concentration is lowered then dispersion is no longer

optimized and agglomeration will occur. It is therefore necessary to find a balance

between dispersion and inhibition.

6.4.2.2 Adhesion

In the equiaxed powder samples with reduced PEI concentration where deposition

was observed a lack of adhesion was noted. This was seen by the rearrangement of the

particles upon drying. As the drying front moved from the edges of the deposit towards

the center, the particles move with the front. Uneven thickness and density gradients

were visually observed in all the samples. A lack of adhesion was also observed in all of

the samples deposited from the hydrothermal BaTiO3 suspensions. Even suspensions

with as little 0.1 w/w PEI, where there is little to no unabsorbed PEI in solution to

contaminate the electrode, a lack of adhesion was observed.

Several researchers have observed that during EPD particles can approach but not

deposit at the electrode.47-51 The particles amass in a potential well near the electrode,

and a surge of applied field is needed to overcome the potential barrier and cause

deposition.48 The lack of adhesion indicates that the deposition process is not fully

occurring in the samples. Depending on the mechanism of deposition, the lack of

adhesion can be dependent on several factors. But in general there are two reasons why

the particles do not deposit: (1) the electric field gradient near the electrode is not

sufficient for the particle to overcome the repulsive interactions between the particles and

186

electrode, or (2) the zeta potential of the particles has not be reduced enough to lower the

repulsive interaction at the electrode.19

One mechanism, referred to as neutralization, occurs when the electrochemical

reactions at the electrode generate or consume ionic species near the electrode which

substantially change the solution chemistry near the electrode leading to a reduction and

eventual neutralization of the surface charge.19, 52, 53 Neutralization is believed to be a

common mechanism of deposition for electrosterically dispersed system. Figure 6.9 is a

schematic showing the process on charge neutralization for the current BaTiO3/PEI

system. In the current deposition experiments, two electrochemical reactions, Equations

6.2 and 6.3 are proceeding, and each reaction has a profound effect on the solution

chemistry near the cathode, where deposition is occurring. In the first reaction protons

are consumed to generate hydrogen gas, while in the second reaction water is reduced to

produce hydroxyl ions and hydrogen gas. With the reduction in proton concentration and

increase in the hydroxyl ion concentration the solution pH near the cathode is increasing

greatly. PEI has an isoelectric point (IEP) of approximately 9.5-10 and as the pH is

increases near the cathode the charge on the PEI is reduced as the suspension pH

approaches the IEP of PEI. This is a problem because if the PEI is neutralized then the

electrostatic interactions which bind the PEI to the surface will not longer be present and

the PEI will desorb from the surface. This then leaves the desorbed PEI free in solution

to react with the electrode and inhibit electrochemical reactions. In addition, the particle

surface charge in minimized and the electrophoretic force on the particle is reduced.

Another effective means of lowering the zeta potential without complete charge

neutralization is through the addition of an indifferent electrolyte.15 To lower the zeta

187

Figu

re 6

.9.

Sche

mat

ic sh

owin

g th

e pr

oces

s of c

harg

e ne

utra

lizat

ion.

Wat

er is

redu

ced

at th

e ca

thod

e an

d th

e pH

incr

ease

s due

to

prod

uctio

n of

hyd

roxy

l gro

ups.

The

incr

ease

d pH

resu

lts in

the

PEI l

osin

g ch

arge

and

des

orbi

ng fr

om th

e B

aTiO

3pa

rticl

e su

rfac

ePt Electrode

e- e-

H2(

g)

++++

+

+ +

++

++

BaT

iO3

w/ C

harg

ed P

EIH

2OH

2OH

2O

H2O

H2O

OH

-

OH

-

++++

+

+ +

++

++

++++

+

+ +

++

++

OH

-

OH

-

OH

-+++

++

+ +

++

++

OH

-

OH

-

OH

-

H+

+OH

- →H

2O

H2O

H2O

Neu

tral

PEI

188

potential of the suspensions and the electrode surface an indifferent electrolyte, LiCl was

added up to concentrations of 5x10-3 M. The addition of an indifferent electrolyte will

lower the zeta potential of both particles and electrode. To fully investigate the repulsive

interactions between the particles and electrode the zeta potential of both BaTiO3

particles and Pt electrode was measured. To measure the zeta potential of a low surface

area flat sample, it is necessary to use a streaming potential technique.15 For ease of

measurement the zeta potential of Pt particle suspensions were measured and assumed to

behave similarly to a flat Pt surface. Figure 6.10 is a plot of zeta potential for

suspensions of the platelet BaTiO3 and Pt suspensions and as expected, the zeta potential

of both materials decreases as the LiCl concentration increases.

Using DLVO theory it is possible to calculate the interaction energy curves for

the BaTiO3-Pt system.54, 55 The curves will show the presence and magnitude of a

repulsive energy barrier. It is expected that as the LiCl concentration increases and the

zeta potential decreases the repulsive barrier should decrease. A decreased repulsion

reduces the energy needed to bring the particle and electrode in contact, thus improving

adhesion. The interaction energy curves for the interaction between BaTiO3 and Pt are

shown in Figure 6.11. The interaction energy curves were calculated using Stabil56 and

the physical constants57-59 listed in Table 6.3. Appendix C shows the calculation used by

Stabil to calculate the interaction energy curves. For the suspension without LiCl a

maximum repulsion of 50 kT is observed, however the addition of 0.5 mM LiCl lowers

the maximum to 23.6 kT. Van Tassel noted that a potential barrier of 15kT can

reasonably be overcome with the typical voltage gradients that are present during EPD.19

Excellent adhesion was visually observed in all samples containing LiCl, except for the

189

Figu

re 6

.10.

Zeta

pot

entia

l of p

late

let B

aTiO

3an

d Pt

ele

ctro

de in

pur

e et

hano

l as f

unct

ion

of L

iCl c

once

ntra

tion.

The

add

ition

of

an in

diff

eren

t ele

ctro

lyte

low

er th

e ze

ta p

oten

tial a

nd th

eref

ore

the

repu

lsiv

e in

tera

ctio

ns b

etw

een

the

depo

sitin

g pa

rticl

es a

nd th

e el

ectro

de.

With

out t

he a

dditi

on o

f the

LiC

l the

larg

e re

puls

ive

inte

ract

ions

bet

wee

n th

e pa

rticl

es a

nd e

lect

rode

lead

to a

lack

of

adhe

sion

of t

he p

artic

les o

n th

e el

ectro

de.

190

Figu

re 6

.11.

Inte

ract

ion

ener

gy c

urve

s for

the

plat

elet

BaT

iO3

and

Pt e

lect

rode

. Th

e cu

rves

wer

e ca

lcul

ated

usi

ng S

tabi

l55an

d th

e ph

ysic

al c

onst

ants

in T

able

6.3

. Th

e ad

ditio

n of

LiC

l, as

exp

ecte

d, lo

wer

s the

repu

lsiv

e in

tera

ctio

n be

twee

n th

e pa

rticl

es a

nd

elec

trode

s. H

owev

er, t

he a

dditi

on o

f ≥ 1

mM

LiC

l res

ults

in a

smal

l rep

ulsi

ve in

tera

ctio

n w

hich

is n

ot su

itabl

e fo

r goo

d di

sper

sion

.

191

Ham

aker

Con

stan

t of B

aTiO

31,2

A11

= 2

2.9

x 10

-20 J

Ham

aker

Con

stan

t of P

t3,2

A22

= 6

2.7

x10-2

0 JH

amak

er C

onst

ant o

f Eth

anol

4A

33 =

4.2

x 1

0-20 J

Latti

ce C

onst

ant o

f BaT

iO3

a 1 =

3.9

98 Å

Latti

ce C

onst

ant o

f Pt

a 2 =

4.0

28 Å

BaT

iO3 P

artic

le R

adiu

sr 1

= 5

0 nm

Pt P

artic

le R

adiu

sr 2

= 1

000

mm

Pol

ymer

Thi

ckne

ss o

n B

aTiO

3p 1

= 1

.3 n

mP

olym

er T

hick

ness

on

Pt

p 2 =

0 n

m1 A

131 v

alue

for B

aTiO

3 fro

m T

able

4 o

f Ref

57

2 Equ

atio

n 11

.27

from

Ref

58

used

to c

alcu

late

A11

from

A13

1 dat

a3 A

vera

ge v

alue

of A

131 f

or P

t fro

m T

able

4 o

f Ref

59

4 A33

val

ue o

f eth

anol

from

Ref

58

Tab

le 6

.3.

List

of p

hysi

cal c

onst

ants

use

d in

the

calc

ulat

ion

of th

e in

tera

ctio

n be

twee

n B

aTiO

3an

d Pt

in p

ure

etha

nol.

192

sample with 5 mM LiCl because the sample was not stable and not suitable for

deposition.

Adding an indifferent electrolyte decreases the solution resistance and it is

expected that the deposition rate should decrease. The effect of LiCl on the deposition

rate and suspension conductivity is shown in Figure 6.12. A sharp decrease is observed

in the deposition rate, this is due to two effects: (1) the reduction in zeta potential of the

suspension, and (2) the observed increase in the suspension conductivity. The decrease

in zeta potential is directly related to the electrophoretic velocity through Equation 6.2.

Referring to the circuit diagram, Figure 6.4a, an increase in the suspension conductivity

will results in a lower effective field at a constant applied voltage. CV was used to

determine the effect of LiCl on the electrochemistry. The results (not shown) show no

influence of LiCl on electrochemical reactions.

6.4.3 Film microstructure

With the ability to deposit films, it is important to evaluate the microstructure of

the deposited films. With the changes in solution chemistry needed to improve kinetics

and adhesion it is expected the microstructure of the films will be affected. Figures 6.13

a and b are a TEM image of a film cross-section and an AFM image of the top surface of

the film deposited from the equiaxed powder suspensions exhibiting the best adhesion

(5w/w HOx and 0.5w/w PEI). The film has a thickness of 613± 32 nm and a roughness

of 106 nm. With a primary particle size of 88.2 nm the film is comprised of

approximately 7 particle layers. However, with a D(50) = 175 nm in suspension prior to

deposition the film is comprised of 3.5 aggregate particle layers. If the layer thickness

193

Figu

re 6

.12.

The

cond

uctiv

ity a

nd d

epos

ition

rate

of s

uspe

nsio

ns w

ith L

iCl a

dded

are

hig

hly

depe

nden

t on

the

conc

entra

tion

of

LiC

l. A

lthou

gh a

ddin

g Li

Cl i

mpr

oves

film

adh

esio

n it

resu

lts in

decr

ease

d de

posi

tion

kine

tics

194

Figu

re 6

.13a

and

b.

(a) T

EM im

age

of a

film

cro

ss-s

ectio

n an

d (b

) AFM

def

lect

ion

imag

e of

top

surf

ace

of a

n EP

D fi

lm o

f eq

uiax

ed B

aTiO

3pa

rticl

es.

The

film

has

a th

ickn

ess o

f 613

nm

with

a su

rfac

e ro

ughn

ess o

f 106

nm

.

195

was reduced the presence of defects and pinholes would increase. This underscores the

importance of having well-dispersed particle systems for EPD.

The platelet particles provide a unique advantage in EPD because of their shape

anisotropy. Previous work by Yener et al. showed that with well-dispersed suspension of

2 nm metal platelets 15 nm layer by EPD are possible because upon deposition the

platelets assume an orientation where the large face of the particle lays parallel to the

electrode surface.60 The ability to produce thin layers with a large number of particle

layers is one of the attractive aspects of the EPD of platelet particles. For comparison a

613 nm layer of platelet particles (thickness = 5.8 nm) would have 106 particle layers.

Figures 6.14 is an AFM image of a film deposition from a platelet powder suspension

with 5x10-4 M LiCl, the minimum concentration needed to observe film adhesion. It is

apparent that the film is inhomogeneous and that particle aggregates have deposited on

the electrode. The inhomogeneous nature of the film is also assumed to be due to partial

inhibition by PEI similar to the schematic in Figure 6.7.

Since the platelet particles all have the same crystallographic texture, a (111)

large face, a layer comprised of oriented platelet particles should exhibit crystallographic

texture. Figure 6.15 is the XRD patterns for three films deposited at different voltages.

The patterns show no evidence of texture and the relative ratios of the peak intensities are

similar to that of a powder diffraction pattern for BaTiO3 (JCPDS Card: 31-0174).

Combined with the AFM images the XRD results show that the films deposited are

comprised of large particle agglomerates. In addition to improving film adhesion and

reducing deposition kinetics the addition of LiCl affects the particle-particle interactions

leading to aggregation. Similar to the affects of changing PEI concentration there is a

196

Figu

re 6

.14.

AFM

def

lect

ion

imag

e of

EPD

film

dep

osite

d fr

om a

pla

tele

t BaT

iO3

susp

ensi

on.

Elec

trode

surf

ace

cove

rage

is

inco

mpl

ete

and

the

film

app

ears

to b

e co

mpr

ised

of a

ggre

gate

s.

197

Figu

re 6

.15.

XR

D a

naly

sis o

f dep

osite

d pl

atel

et p

owde

rs sh

ow th

at c

ryst

allo

grap

hic

text

ure

did

not d

evel

op d

urin

g de

posi

tion.

Th

e lin

es re

pres

ent t

he p

eak

posi

tions

and

rela

tive

inte

nsity

for c

ubic

BaT

iO3

pow

der (

JCPD

S C

ard:

31-

0174

). In

crea

sing

the

volta

ge is

exp

ecte

d to

: (1)

incr

ease

the

laye

r thi

ckne

ss, a

nd (2

) pro

vide

a h

ighe

r driv

ing

forc

e fo

r the

flat

layd

own

on p

late

let

parti

cles

, but

no

impr

ovem

ent i

n te

xtur

e is

obs

erve

d as

the

depo

sitio

n vo

ltage

incr

ease

s. T

he p

rese

nce

of p

artic

le a

ggre

gate

s as

seen

in A

FM im

ages

is th

e re

ason

for t

he la

ck o

f tex

ture

. N

ote:

* Th

e di

min

ishi

ng p

eak

at 2

6º2θ

is d

ue th

e un

derly

ing

Pt/M

ylar

su

bstra

te u

sed

as th

e el

ectro

de.

198

trade-off in adding an indifferent electrolyte to the solution, while improving adhesion it

reduces kinetics and destabilize the suspension prior to deposition.

6.5 Conclusions

The electrophoretic deposition of BaTiO3 nanoparticle suspensions was

investigated with an emphasis on understanding the interdependence of solution

chemistry on the kinetics of deposition and microstructure of the deposition films. The

solution chemistry is a controlling variable in the development of well-dispersed

suspensions for EPD. The optimum dispersion conditions are inadequate for deposition.

High concentrations of PEI inhibit electrochemical reactions and increase the resistance

of the electrode-solution interface. This results in a significant reduction in the

electrophoretic velocity of the particles, effectively stopping deposition. When the

deposition kinetics are enhanced the repulsive particle-electrode interactions lead to poor

adhesion of the deposited films. The addition of an indifferent electrolyte mitigates the

repulsive interactions and improved adhesion, but lead to particle aggregation prior to

deposition. Films deposited from aggregated suspensions were of low quality and

unstable for further processing.

199

References

1. Taiyo Yuden: Mass Production of World's First 100μF Multilayer Ceramic Capacitor in EIA1206 Size. http://www.yuden.co.jp/e/release/pdf/2005/20050630e.pdf, (2005). 2. A.J. Moulson and J.M. Herbert, Electroceramics: Materials, properties, applications, 1st ed. (Chapman & Hall, London, 1990). 3. N. Ogata, J. van Tassel, and C.A. Randall: Electrode formation by electrophoretic deposition of nanoparticles. Mater. Lett. 48, (1), 7 (2001). 4. J. Tabellion and R. Clasen, Electrophoretic deposition of SiC from aqueous suspensions, edited by A.R. Boccaccini, O. Van der Biest, and J.B. Talbot (in Electrophoretic deposition: Fundamentals and applications, Banaff, Canada, 2002). 5. E. Harsanyi: Method of coating radiant bodies. US Patent # 1,897,902, (1933). 6. H. von Both and J. Hasselt, Ceramic microstructures by electrophoretic deposition of colloidal suspension, edited by A.R. Boccaccini, O. Van der Diest, and J.B. Talbot (in Electrophoretic deposition: Fundamental and applications, Banff, Canada, 2002), pp 78. 7. C. Kaya, A.R. Boccaccini, and K.K. Chawla: Electrophoretic deposition forming of nickel-coated-carbon-fiber-reinforced borosilicate-glass-matrix composites. J. Am. Ceram. Soc. 83, (8), 1885 (2000). 8. E.M. Wong and P.C. Searson: ZnO quantum particle thin films fabricated by electrophoretic deposition. Appl. Phys. Lett. 74, (20), 2939 (1999). 9. P. Sarkar, X. Haung, and P.S. Nicholson: Structural ceramic microlaminates by electrophoretic deposition. J. Am. Ceram. Soc. 75, (10), 2907 (1993). 10. P. Sarkar, X. Haung, and P.S. Nicholson: Zirconia/alumina functionally gradiented composites by electrophoretic deposition technique. J. Am. Ceram. Soc. 76, (4), 1055 (1992). 11. H.S. Maiti, S. Datta, and R.N. Basu: High-Tc superconducting coating on metal substrates by an electrophoretic technique. J. Am. Ceram. Soc. 72, (9), 1733 (1989). 12. M.D. Sacks and G.W. Scheiffele: Properties of silicon suspensions and slip-cast bodies. Ceram. Eng. Sci. Proc. 6, (7-8), 1109 (1985). 13. T. Kimura, Y. Matsuda, M. Oda, and T. Yamaguchi: Effects of agglomerates on the sintering of alpha-Al2O3. Ceram. Int. 13, (1), 27 (1987).

200

14. A.A. Parker, J. Sun, A.M. Ahern, S. T, D. Wilhelmy, G.H. Armstrong, and J.J. Marcinko: Effect of dispersion state on ceramic green body morphology. Poly. Pre. 33, (1), 1210 (1992). 15. R.J. Hunter, Zeta potential in colloid science: Principles and applications, 1st ed. (Academic Press, San Diego, CA, 1981). 16. M. von Smoluchowski: Contribution a la theorie de l'endosmose electrique et de quelques phenomenes correlatifs. Bull. Acad. Sci. Cracovie 8, 182 (1903). 17. E. Huckel: Die kataphorese der kugel. Phys. Z. 25, 204 (1924). 18. D.C. Henry: The cataphoresis of suspended particles. Part I - The equation of cataphoresis. Proc. Roy. Soc. A 133, 106 (1931). 19. J. van Tassel: Electrophoretic deposition: Fundamentals, mechanisms and examples with an in depth examination of the ion depletion effect. Ph.D. Thesis, The Pennsylvania State University, (2004). 20. H.C. Hamaker and E.J.W. Verwey: The role of forces between the particles in electrodeposition and other phenomena. Trans. Faraday Soc. 36, 180 (1940). 21. H. Koelmans and J.T.G. Overbeek: Stability and electrophoretic deposition of suspensions in non-aqueous media. Discuss. Faraday Soc. 18, (52-63), (1954). 22. D.R. Brown and F.W. Salt: The mechanism of electrophoretic deposition. J. Appl. Chem. 15, 40 (1963). 23. P. Sarkar and P.S. Nicholson: Electrophoretic deposition (EPD): Mechanisms, kinetics, and applications to ceramics. J. Am. Ceram. Soc. 79, (8), 1987 (1996). 24. R.J. Kershner, J.W. Bullard, and M.J. Cima: The role of electrochemical reactions during electrophoretic particle deposition. J. Colloid Interface Sci. 278, 146 (2004). 25. T.J. Yosenick, D.V. Miller, R. Kumar, J.A. Nelson, C.A. Randall, and J.H. Adair: Synthesis of nanotabular barium titanate via a hydrothermal route. J. Mater. Res. 20, (4), 837 (2005). 26. R.W. O'Brien and L.R. White: Electrophoretic mobility of a spherical colloidal particle. J. Chem. Soc. Faraday II 74, 1607 (1978). 27. J. McCaffrey: Small-angle cleavage of semiconductors for transmission electron-microscopy. Ultramicroscopy 38, (2), 149 (1991). 28. S. Venigalla and J.H. Adair: Theoretical modeling and experimental verification of electrochemical equilibria in the Ba-Ti-C-H2O system. Chem. Mater. 11, (3), 589 (1999).

201

29. U. Paik and V.A. Hackley: Influence of solids concentration on the isoelectric point of aqueous barium titanate. J. Am. Ceram. Soc. 83, (10), 2381 (2000). 30. A. Neubrand, R. Linder, and P. Hoffman: Room-temperature solubility behavior of barium titanate in aqueous media. J. Am. Ceram. Soc. 83, (4), 860 (2000). 31. U. Paik, S. Lee, and V.A. Hackley: Influence of barium dissolution on the electrokinetic properties of colloidal BaTiO3 in an aqueous medium. J. Am. Ceram. Soc. 86, (10), 1662 (2003). 32. T.J. Yosenick, Y. Yuan, C.A. Randall, and J.H. Adair: Passivation, dispersion, and aqueous solution doping of platelet BaTiO3 powder. To be submitted, (2005). 33. J.H. Adair, A.J. Rose, and L.G. McCoy: Particle size analysis of ceramic powders. In Advances in ceramics, Volume 11, Processing for improved productivity, edited by (The American Ceramic Society, 1984) pp 142. 34. H. Natter and R. Hempelmann: Nanocrystalline copper by pulsed electrodeposition: The effect of organic additives, bath temperature, and pH. J. Phys. Chem. 100, 19525 (1996). 35. R.A. Marcus: Electron transfer at electrodes and in solution: Comparison of theory and experiment. Electrochim. Acta 13, (5), 955 (1968). 36. J.R. Macdonald, Impedance spectroscopy: Emphasizing solid materials and systems, 1st ed. (John Wiley, New York, New York, 1987). 37. F. Berthier, J.P. Diard, and R. Michel: Distinguishability of equivalent circuits containing CPEs. Part I. Theoretical part. J. Electroanal. Chem. 510, 1 (2001). 38. W.H. Mulder, J.H. Sluyters, T. Pajkossy, and I. Nyikos: Tafel current at fractal electrodes. Connection with admittance spectra. J. Electroanal. Chem. 285, 103 (1990). 39. C.A. Schiller and W. Strunz: The evaluation of experimental dielectric data of barrier coating by means of different models. Electrochim. Acta 46, 3619 (2001). 40. C.H. Kim, S.I. Pyun, and J.H. Kim: An investigation of the capacitance dispersion on the fractal carbon electrode with edge and basal orientations. Electrochim. Acta 48, 3455 (2003). 41. R. De Lise, M. Goffredi, and V.T. Liveri: Effect of water on proton migration in alcoholic solvents. Part 4 - Conductance of hydrogen chloride in butan-2-ol and in ethanol at 25 C. J. C. S. Faraday I 72, 436 (1976).

202

42. G. Trabanelli: Corrosion inhibitors. In Corrosion mechanisms, edited by F. Mansfeld, (Marcel Dekker, Inc., 1987) pp 119. 43. M. Stratmann, W. Furbeth, G. Grundmeier, R. Losch, and C.R. Reinartz: Corrosion inhibition by adsorbed organic monolayers. In Corrosion mechanisms in theory and practice, edited by P. Marcus, and J. Oudar, (Marcel Dekker, Inc., 1995) pp 373. 44. M.J. Incorvia and S. Contarini: X-ray photoelectron spectroscopic studies of metal/inhibitor systems: Structure and bonding at the iron/amine interface. J. Electrochem. Soc. 136, (9), 2493 (1989). 45. P. Li, J.Y. Lin, K.L. Tan, and J.Y. Lee: Electrochemical impedance and x-ray photoelectron spectroscopic studies of the inhibition of mild steel corrosion in acids by cyclohexylamine. Electrochim. Acta 42, (4), 605 (1997). 46. H. Ashassi-Sorkhabi and S.A. Nabavi-Amri: Polarization and impedance methods in corrosion inhibition study of carbon steel by amines in petroleum-water mixtures. Electrochim. Acta 47, 2239 (2002). 47. M. Giersig and P. Mulvaney: Formation of two-dimensional gold colloid lattice by electrophoretic deposition. J. Phys. Chem. 97, (24), 6334 (1993). 48. M. Bohmer: In situ observation of 2-dimensional clustering during electrophoretic deposition. Lang. 12, (24), 5747 (1996). 49. Y. Solomentsev, M. Bohmer, and J.L. Anderson: Particle clustering and pattern formation during electrophoretic deposition: A hydrodynamic model. Lang. 13, (23), 6058 (1997). 50. S.A. Guelcher, Y. Solomentsev, and J.L. Anderson: Aggregation of pair particles on electrodes during electrophoretic deposition. Powder Tech. 110, 90 (2000). 51. Y. Solomentsev, S.A. Guelcher, M. Bevan, and J.L. Anderson: Aggregation dynamics for two particles during electrophoretic deposition under steady fields. Lang. 16, (24), 9208 (2000). 52. F. Grillon, D. Fayeulle, and M. Jeandin: Quantitative image analysis of electrophoretic coatings. J. Mater. Sci. Lett. 11, 272 (1992). 53. J. Mizuguchi, K. Sumi, and T. Muchi: A highly stable nonaqueous suspension for electrophoretic deposition of powdered substances. J. Electrochem. Soc. 136, 2724 (1983). 54. B. Darjaguin and L. Landau: Theory of the stability of strongly charged lyophic sol and of the adhesion of strongly charged particles in solution of electrolytes. Acta Physiochim. URSS 14, 633 (1941).

203

55. E.J.W. Verwey, J.T.G. Overbeek, and K. Nes, Theory of the stability of lyophobic collloids. The interaction of sol particles having an electric double layer, 1st ed. (Elsevier Publishing Company, Inc., New York, New York, 1948). 56. Stabil for Windows, 4.5. 57. L. Bergstrom: Hamaker constants of inorganic materials. Adv. Coll. Int. Sci. 70, 125 (1997). 58. I. Israelachvili, Intermolecular & Surface Forces, 2nd ed. (Harcourt Brace & Company, London, 1992). 59. J. Visser: On Hamaker constants: A comparison between Hamaker constants and Lifshitz - van Der Waals constants. Adv. Coll. Int. Sci. 3, 331 (1972). 60. D.O. Yener, T.J. Yosenick, C.A. Randall, and J.H. Adair: Synthesis of nanosized Ag/Pd platelets in self-assembled bilayers, and thin film metallization by electrophoretic depositions. To be submitted, (2005).

204

CHAPTER SEVEN

Conclusions and Suggested Work 7.1 Summary and Conclusions

With the current in size reduction in dielectric layers in MLCCs the use of

nanoparticles is seen as a necessary requirement to achieve the desired layer thickness.

Combined with controlled particle morphology during synthesis nanoparticles provide

unique advantages including enhanced reactivity and textured microstructures. The

current study investigated the individual steps in the processing of hydrothermal

anisotropic BaTiO3 from the synthesis of nanoparticles to the dispersion, doping,

sintering, and electrical properties of the bulk materials. Nanotabular BaTiO3 particles

were synthesized using a hydrothermal route. The particles are single crystal with a

majority of the particles having a [111] zone axis, and a median thickness of 5.8 ± 3.1 nm

and a face diameter of 27.1 ± 12.3 nm, as determined by atomic force microscopy.

Morphology of the particles was shown to be controlled solely by pH of the solution

during synthesis. It was speculated that the high solution pH stabilizes the {111} face

limiting growth in the <111> direction and leading to multiple {111} twin formation

during synthesis. With growth limited in the <111> direction the particle develop a plate-

like morphology. The powder has a low concentration (0.5 wt%) of hydrothermal defects

which coalesce during heating to form defects in the particles. TEM observations of

thermally treated particles show that the particle lose their plate-like morphology by 1000

°C. The Ba/Ti ratio shows that the powder is slightly Ba rich.

205

After synthesis, a model was which described the complex nature of the BaTiO3

surface in an aqueous environment. Three different regions of surface charging were

observed and modeled using different approaches. The Ba2+ depleted surface at low pH

was treated as a TiO2 surface and the MUSIC model was applied. As pH increases the

influence of adsorbed Ba2+(aq) was accounted by using a modified Stern isotherm. At

high pH the precipitation of BaCO3 on the particle surface was modeled with a Nernst-

Gouy-Stern charging mechanism. The inherent complexity of the surface necessitates an

appropriate aqueous processing technique for BaTiO3.

Oxalic acid was shown to have a passivation effect on the BaTiO3 surface. Oxalic

acid forms a barium oxalate layer on the particle, which limits Ba2+ leaching and BaCO3

formation. PEI can be used as an effective dispersant but the state of dispersion is highly

dependent on the ratios and concentrations of oxalic acid and PEI. Viscosity

measurements show the ability for excess oxalic acid and PEI to react and form an

amine-oxalate gel, which can be detrimental to dispersion.

Using an aqueous solution based approach, both the platelet and spherical BaTiO3

particles were doped with a X7R-type formulation based on Co, Nb and Bi. The surface

chemistry of the BaTiO3 particles was manipulated which resulted in the selective

adsorption of the dopants on the particle surface from solution. The development of an

engineered particle coating during doping was confirmed by surface charge, TEM, and

EDS analysis.

The electrophoretic deposition of spherical and platelet powders was studied. The

effect of solution chemistry on the kinetics, electrochemistry and microstructure of

deposited film was studied. The concentration of dispersant (PEI) and indifferent

206

electrolyte (LiCl) were shown the have a profound on the kinetics of deposition and the

microstructure of the films. Unabsorbed PEI in solution contaminates the electrode

during deposition inhibiting electrochemical reactions and slowing deposition. LiCl was

added to reduce the repulsive interactions between the particle and electrode and improve

film adhesion, but it decreased the deposition rate and led to particle aggregation.

Because of aggregation the platelet particle film did not exhibit any microstructural or

crystallographic texture. Films deposited from platelet powder suspensions were

inhomogeneous and the films as appeared to be isolated islands of aggregation, whereas

films deposited from spherical powder suspensions where confluent layers with a

thickness of approximately 600 nm.

7.2 Suggested Work

Since the platelet particle exhibit crystallographic texture it was expected that film

deposited by EPD should be textured. However, XRD and AFM confirms that particle

aggregates lead to untextured inhomogeneous films. The dispersion of the particle

system must be addressed of textured films are to be deposited. One disadvantage of the

suspensions used in the current work is that ethanol was used instead of water. Ethanol

was used to limit the formation of hydrogen bubbles at the cathode during deposition

which can be detrimental to film microstructure. When EPD is carried out in water most

depositions are done at low voltage1 or using a porous membrane2 located in the center of

the EPD cell to eliminate the effect of the electrochemical reaction at the electrode.

While successful deposition is possible each technique has its limitations. Recently,

Uchikoshi et al. have demonstrated the ability to deposit aqueous alumina and titania

207

suspensions using an applied voltage of 30 V with Pd electrodes.3-5 Pd readily forms

PdHx when it reacts with hydrogen.6, 7 By using Pd electrodes Uchikoshi et al. were able

to adsorb any hydrogen generated at the cathode and stop hydrogen bubble formation.

Uchikoshi et al. in the same research was able to develop crystallographic texture

through the use of an applied magnetic field. Many diamagnetic materials with non-

cubic crystal structure exhibit anisotropy in their magnetic susceptibilities.5 To align

these materials it is necessary to apply a magnetic field such that the magnetic force on

the particle is greater than the force due to thermal energy. However, the difference in

susceptibilities between different orientations is so small that a large (10 T) magnetic

field is required to align the materials. Unfortunately, the magnetic susceptibility

anisotropy of BaTiO3 has not been measured, but work by Uyeda8 suggests that

anisotropy exists. Uyeda found that materials with MO6 metal cation octahedra exhibit

anisotropy in the magnetic susceptibility if the octahedra are slightly elongated with

structurally anisotropy. This suggest that the anisotropy of the TiO6 in tetragonal BaTiO3

should result in an anisotropic magnetic susceptibility the can be used to manipulate the

crystallographic orientation of BaTiO3 particles while in suspension during deposition.

TEM observations show that the platelet particles lose the platelet morphology by

1000 °C. Thus, an advantage obtained by using a platelet material would be lost during

high temperature sintering. However, the observations were made on loose powder

samples where the constraints of particle-particle contacts and pore distribution are not a

factor. With the possible development of a textured microstructure using EPD and

applied magnetic fields a complete analysis of the morphology evolution particles in a

208

powder compact is necessary. In addition to the changes in particle morphology, the

evolution of the pore distribution is also of interest.

The use of conventional sintering approach with a high temperature isothermal

hold is expected to result in significant changes in the microstructure. However, a novel

multi-step sintering approach, similar to those proposed by Polotai et al.9 and Chen and

Wang10, would provide an advantage in that the last sintering step is a low temperature

isothermal hold. In the multi-step approach, the initial steps are designed to result a

uniform distribution of sub-critical pores which can be eliminated without substantial

grain growth during the last step. This suggests that if the green structure were of high

density with a uniform pore distribution that sintering could be achieved with a minimal

or no initial step and a long low temperature isothermal hold. If the isothermal hold were

below 1000 °C, then the sintering of the platelet particle without morphology can is a

possibility.

209

Reference

1. A.L. Rogach, N.A. Kotov, D.S. Koktysh, J.W. Ostrander, and G.A. Ragoisha: Electrophoretic deposition of latex-based 3D colloidal photonic crystals: A technique for rapid production of high-quality opals. Chem. Mater. 12, 2721 (2000). 2. J. Tabellion and R. Clasen, Electrophoretic deposition of SiC from aqueous suspensions, edited by A.R. Boccaccini, O. Van der Biest, and J.B. Talbot (in Electrophoretic deposition: Fundamentals and applications, Banaff, Canada, 2002), pp. 3. T. Uchikoshi, T.S. Suzuki, H. Okayuma, and Y. Sakka: Control of crystalline texture in polycrystalline alumina ceramics by electrophoretic deposition in a strong magnetic field. J. Mater. Res. 19, (5), 1487 (2004). 4. T. Uchikoshi, T.S. Suzuki, K. Okuyama, Y. Sakka, and P.S. Nicholson: Electrophoretic deposition of alumina suspension in a strong magnetic field. J. Euro. Ceram. Soc. 24, 225 (2004). 5. T. Uchikoshi, T.S. Suzuki, F. Tang, K. Okuyama, and Y. Sakka: Crystalline-oriented TiO2 fabricated by electrophoretic deposition in a strong magnetic field. Ceram. Int. 30, 1975 (2004). 6. T. Kuji, Y. Matsumura, H. Uchida, and T. Aizawa: Hydrogen adsorption of nanocrystalline palladium. J. Alloys Compd. 330-332, 718 (2002). 7. S. Kishore, J.A. Nelson, J.H. Adair, and P.C. Eklund: Hydrogen storage in spherical and platelet palladium nanoparticles. J. Alloys Compd. 389, 234 (2005). 8. C. Uyeda: Diamagnetic anistropies of oxide materials. Phys. Chem. Minerals 20, 77 (1993). 9. A.V. Polotai, K. Breece, E. Dickey, C.A. Randall, and A.V. Ragulya: A novel approach to sintering nanocrystalline barium titanate ceramics. J. Am. Ceram. Soc. 88, (11), 3008 (2005). 10. I.W. Chen and X.H. Wang: Sintering dense nanocrystalline ceramics without final-stage grain growth. Nature 404, 168 (2000).

210

APPENDIX A

Algorithm for the Determination of Surface Potential Using the MUSIC Model

To calculate the surface potential of a material system using the MUSIC model1-

3it is first necessary to determine the type and characteristics of surface sites present

based on the crystal structure of the material. Figure A.1 is a flow diagram explaining the

process of using the MUSIC model to predict the surface potential as a function of

solution pH. The crystal structure dictates the cation coordination of the lattice oxygen

and the change in coordination when an ideal surface is cleaved. From the cation

coordination it is possible to calculate the cation-oxygen bond valence4 using Equation

4.7 which is used in calculating the valence, Equation 4.10, and log K values, Equation

4.9, for the surface site. Included in a calculation of the site valence is the number of

orbitals available for proton uptake. For example, in rutile, the main lattice group is

Ti3O, if at the surface the Ti coordination is reduced to two the group becomes Ti2O.

This reduction in Ti coordination results in the oxygen being cation deficient and able to

react with a single proton from solution to compensate for the reduction in coordination.

The calculated log K value represents the equilibrium solution pH at which the proton

will either adsorb or desorb. This is the current limit of the MUSIC model. In order to

calculate the surface charge as a function of solution pH is it necessary to employ a

model for the solution side of the interface, typically a Gouy-Chapman model for the

diffuse double layer.

211

Crystal Structure

Surface Sites

Site Valencelog K Values

Surface Charge Density = f(pH)

Surface Potential = f(pH)

Gouy Chapman Model

MUSIC Model

Solution pH

Figure A.1. Simple flow diagram showing the steps necessary to calculate the surface potential as a function of solution pH using the MUSIC model

212

The Gouy-Chapman model is based on electronuetrality between the solid and

solution sides of the interface, and assumes that charge density on the surface is equal and

opposite to the charge density in solution.5, 6 The information provided by the MUSIC

model is used to calculate the surface charge density,

∑=n

iiisis SNe ασ [A.1]

where e is the charge on the electron, Nsi is the site density of any specific site, Si is the

valence of the site, and αi is the degree of protonation of a specific site. Nsi is based on

the crystal structure of the materials, and Si used a value calculated from the MUSIC

model. αi represents the percentage of specific sites are protonated and is calculated

using Equation 4.15, and is dependent on both the log K value of the specific site and the

solution pH. This results in the surface charge density being dependent on the solution

pH. At each pH value it is possible to calculate the surface potential by setting the

surface charge density equal to the solution charge density, Equation 4.22. The result of

this is a series of surface potential values as a function of pH that can be plotted to yield

the familiar surface charge versus pH curves.

An advantage of this technique is that morphology of the particle can be

accounted for if the structure and area fraction (fi) of each specific habit present to

solution is known. Equation A.1 can be calculated for each specific habit and in turn the

surface potential (ψi) for each habit can be calculated. The total surface potential of the

particle can be calculated by using a weighted average to include the contributions of

each specific habit,

∑=n

iiitot fψψ [A.2]

213

References

1. T. Hiemstra, W.H. van Riemsdijk, and G.H. Bolt: Multisite proton adsorption modeling at the solid-solution interface of (hydr)oxides - A new approach. 1. Model desciption and evaluation of intrinsic reaction constants. J. Colloid Interface Sci. 133, (1), 91 (1989). 2. T. Hiemstra, P. Venema, and W.H. van Riemsdijk: Intrinsic proton affinity of reactive surface groups of metal (hydr)oxides: The bond valence principle. J. Colloid Interface Sci. 184, (2), 680 (1996). 3. M.L. Machesky, D. Wesolowski, D.A. Palmer, and M.K. Ridley: On the temperature dependence of intrinsic surface protonation equilibrium constants: An enxtension of the revised MUSIC model. J. Colloid Interface Sci. 239, 314 (2001). 4. L. Pauling: The principles determining the structure of complex ionic crystals. J. Amer. Chem. Soc. 51, 1010 (1929). 5. G. Gouy: Consititution of the electric charge at the surface of an electrolyte. J. Physique 9, (4), 457 (1910). 6. D.L. Chapman: Theory of electrocapillarity. Phil. Mag. 25, 475 (1913).

214

APPENDIX B

Dispersion of Solution Based Doped BaTiO3 Platelets for Electrophoretic Deposition

Chapter Five addresses the solution based doping of BaTiO3 while in suspension.

In colloidal processing the green microstructure is highly dependent on the state of

dispersion prior to consolidation.1-3 Therefore, it is necessary to have well-dispersed

particle suspensions for the electrophoretic deposition (EPD) of thin particulates layers.

In the case of layers deposited from doped BaTiO3 platelets the lack of dispersion made it

difficult to deposit homogenous dense BaTiO3 layer suitable for further processing or

fabrication of a multilayer ceramic capacitor. The dispersion of particulates in

suspension is dependent on a variety of factors, but the most dominant factors is the

particle surface charge which is affected by the physical chemistry of the surface and the

solution chemistry of the solvent phase. The procedure use to dope the BaTiO3 platelets

while in suspension had a profound effect on both the surface chemistry of the particle

and the solution chemistry of the solvent.

A passivation/dispersion approach was used prior to doping. In

passivation/dispersion the degradation of the BaTiO3 is control by the addition of oxalic

acid, and then the BaTiO3 surface is coated with polyethylenimine (PEI) to provide a

large positive surface suitable for electrosteric dispersion. In addition to dispersion the

positive surface charge was use for the electrostatic attraction of the negatively charged

dopant complexes with the surface. See Chapter Five for further discussion of the

passivation, dispersion and doping of the BaTiO3 particles.

215

Figure B.1 shows the zeta potential of undoped and fully doped BaTiO3 platelet

suspensions as a function of PEI concentration. In all cases the plot shows an initial

increase in the zeta potential with increasing PEI concentration, but the zeta potential

values begin to plateau at PEI concentrations of 3 w/w. This is due the particle surface

being fully coated with PEI and the excess PEI remaining in the solution phase. It is

most important to note the substantial decrease in the zeta potential of the surface in the

presence of the dopants. A zeta potential of 11 mV is observed for a fully doped samples

at pH, which is not large enough to provide adequate dispersion. The doping process was

performed at pH 5 in order to increase the positive surface charge and improve

dispersion. Although an increase in the zeta potential was observed it was minimal and

good dispersion was not observed.

The electrostatic adsorption of species on a surface typically follows a Langmuir

isotherm,

⎥⎦⎤

⎢⎣⎡ Δ−

=RTG

nxN abssi exp [B.1]

where Ni is the number of filled adsorption sites per area, n is the number of possible

adsorption sites per area, xs is the mole fraction of absorbate in solution, ΔGabs is the free

energy of adsorption, R is the universal gas constant, and T is absolute temperature. In a

Langmuir it is assumed that all surface sites are equivalent and there is a fixed

concentration (i.e. n is constant). In the dispersion of the BaTiO3 with PEI it is the amine

groups that simultaneously provide surface charge and adsorption sites for the dopants.

As the dopants are added to the suspension they selectively adsorb and complex with the

amine groups with the PEI,

M(Cit)- + NH4+… M(Cit)NH4… [B.2]

216

Figu

re B

.1.

The

zeta

pot

entia

l of d

oped

and

und

oped

BaT

iO3

susp

ensi

ons a

s a fu

nctio

n of

PEI

con

cent

ratio

n. T

he P

EI p

rovi

des

surf

ace

char

ge a

s wel

l as a

dsor

ptio

n si

tes f

or th

e io

nic

dopa

nts.

Whe

n th

e do

pant

s ads

orb

surf

ace

site

s are

neu

traliz

ed a

nd th

esu

rfac

e ch

arge

dec

reas

es lo

wer

ing

the

zeta

pot

entia

l. D

ecre

asin

g th

e pH

incr

ease

s the

zet

a po

tent

ial,

but i

t is n

ot su

ffic

ient

to

crea

te st

able

dis

pers

ions

.

217

This reaction results in charge neutralization of the surface site. With a fixed number of

surface sites the adsorption of the dopants results in a decreased surface charge.

Decreasing the pH will increase the number of protonated amine groups on the PEI

providing more adsorption site and increasing the surface charge. Figure B.1 shows this

to be true but the increase is insufficient to disperse the particles.

In addition to charge neutralization by specific absorption the addition of the

dopant complexes results in an increase in the ionic strength of the suspension. Since the

complexes are ionic each dopant solution contains a large concentration of indifferent

counter-ions, for example NO3-, which increases the ionic strength and shields the

repulsive particle interactions. Figure B.2 is a plot of the interaction energy curves

generated using Stabil.4 For the undoped samples a repulsive maximum of

approximately 6 kT is observed, whereas for the doped samples no energy barrier is

observed. The lack of repulsion explains the lack of dispersion for the doped samples.

However, a maximum of 6 kT is inadequate to provide long term dispersion in the

undoped samples and in a little as an hour after preparation the suspension began to

destabilize. Figure B.3 is a plot of the particle size distribution for the doped and

undoped powder at pH 7. The figure shows the large increase in the particle size as the

powder is doped, which is expected from the zeta potential results and the interaction

energy curves.

Solution based doping was effective in the doping of powder prepared for

traditional powder processing (i.e. granulation and dry pressing) however the problems

with dispersion make the technique less than ideal for colloidal based processing

techniques, especially electrophoretic deposition. It is believed that the problem present

218

Figu

re B

.2.

Inte

ract

ion

ener

gy c

urve

s for

dop

ed a

nd u

ndop

ed B

aTiO

3sh

ow th

at a

repu

lsiv

e en

ergy

bar

rier d

oes n

ot e

xist

for t

he

dope

d su

spen

sion

s. Th

is is

due

to th

e re

duct

ion

of th

e ze

ta p

oten

tial a

nd in

crea

se in

the

ioni

c st

reng

th a

s the

dop

ants

are

add

ed to

th

e su

spen

sion

. Th

e in

tera

ctio

n en

ergy

cur

ves w

ere

gene

rate

d us

ing

Stab

il.4

219

Figu

re B

.3.

Parti

cle

size

dis

tribu

tion

for d

oped

and

und

oped

BaT

iO3

susp

ensi

on a

t pH

7.

As e

xpec

ted

from

the

zeta

pot

entia

l and

in

tera

ctio

n en

ergy

resu

lts th

e do

ped

susp

ensi

on is

hig

hly

aggr

egat

ed.

220

by the technique can be overcome with the use of a rigorous thermal treatment and

deaggregation step. The disadvantage of this approach is the addition of two processing

step which increase processing time, cost, and complexity.

221

References

1. M.D. Sacks, and G.W. Scheiffele: Properties of silicon suspensions and slip-cast bodies. Ceram. Eng. Sci. Proc. 6, (7-8), 1109 (1985). 2. T. Kimura, Y. Matsuda, M. Oda, and T. Yamaguchi: Effects of agglomerates on the sintering of alpha-Al2O3. Ceram. Int. 13, (1), 27 (1987). 3. A.A. Parker, J. Sun, A.M. Ahern, S. T, D. Wilhelmy, G.H. Armstrong, and J.J. Marcinko: Effect of dispersion state on ceramic green body morphology. Poly. Pre. 33, (1), 1210 (1992). 4. Stabil for Windows, 4.5.

222

APPENDIX C

Stabil Calculation for Heterogeneous Coagulation

Stabil1 is a program designed to calculate the total interaction energy curve for

two particles interacting across a known liquid medium based on DLVO theory.2, 3 The

total energy is the superposition of the attraction van der Waals forces and the repulsive

double layer energy. The calculation of both energies requires a complete knowledge of

the particle-solvent system including the surface and physical chemistry of the particles,

and the solution chemistry of solvent.

First a series of user defined values are input which include the zeta potential,

particle size, particle and solvent Hamaker constants, and others. Table C.1 is a list of the

physical constants4-6 used in the calculation of the interaction of BaTiO3 and Pt in

Chapter 7. Next, the ionic strength of the solvent (I) and the Debye-Hückel parameter7

(κ) are calculated from t physical constants and user defined values,

21

2∑==

n

iii zc

I [C.1]

where, n = number of ionic species ci = concentration of ith species in mol/L zi = valance of ith species

2/12

10008

⎟⎟⎠

⎞⎜⎜⎝

⎛=

TkNIe

BroA εεπκ [C.2]

where, e = charge on the electron (1.602 x 10-19 C)

223

Ham

aker

Con

stan

t of B

aTiO

31,2

A11

= 2

2.9

x 10

-20 J

Ham

aker

Con

stan

t of P

t3,2

A22

= 6

2.7

x10-2

0 JH

amak

er C

onst

ant o

f Eth

anol

4A

33 =

4.2

x 1

0-20 J

Latti

ce C

onst

ant o

f BaT

iO3

a 1 =

3.9

98 Å

Latti

ce C

onst

ant o

f Pt

a 2 =

4.0

28 Å

BaT

iO3 P

artic

le R

adiu

sr 1

= 5

0 nm

Pt P

artic

le R

adiu

sr 2

= 1

000

mm

Pol

ymer

Thi

ckne

ss o

n B

aTiO

3p 1

= 1

.3 n

mP

olym

er T

hick

ness

on

Pt

p 2 =

0 n

m1 A

131 v

alue

for B

aTiO

3 fro

m T

able

4 o

f Ref

42 E

quat

ion

11.2

7 fro

m R

ef 5

use

d to

cal

cula

te A

11fro

m3 A

vera

ge v

alue

of A

131 f

or P

t fro

m T

able

4 o

f Ref

64 A

33 v

alue

of e

than

ol fr

om R

ef 5

Tab

le C

.1.

List

of p

hysi

cal c

onst

ants

use

d in

the

calc

ulat

ion

of th

e in

tera

ctio

n be

twee

n B

aTiO

3an

d Pt

in p

ure

etha

nol.

224

NA = Avogadro’s number (6.02 x 1023 ions/mol) εo = permittivity of free space (8.854 x 10-12 F/m) εr = relative permittivity of the solvent kB = Boltzmann Constant (1.381 x 10-23 J/K) T = absolute temperature in K

The separation distance (D) of two stern planes with two overlap double layer is

calculated by integrating the Poisson-Boltzmann equation twice,

∫∑ ⎟

⎟⎠

⎞⎜⎜⎝

⎛⎥⎦

⎤⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛−⎟⎟

⎞⎜⎜⎝

⎛−⎟⎟

⎞⎜⎜⎝

⎛=

=

δψ

ψ ψψεε

π

ψ

2/1

2/1

1

2/1

2/1

exp)(exp8

)(n

i B

i

B

i

ro

B

Tkez

TkxezTk

xdD [C.3]

where, ψs = stern potential ψx = potential at a distance x from the stern plane ψ1/2 = potential at the “midpoint” of two interacting stern planes It is important to note that in the case of the overlap of symmetric double layers,

ψ1/2 is taken to be the midpoint between the two interacting particles. However, in the

case of asymmetric double layers, which is the case for the interaction of BaTiO3 and Pt,

the “midpoint” is actually the distant from the stern plane to the minimum in the

interaction potential of the double layers. Because of this is it necessary to solve for D

from the stern potential of each particle. Once the separation distance of the two stern

planes has been calculated the total separation distance can be calculated,

abstot DDDD ++= 21 [C.4]

where, D1 = distance from the stern plane of particle 1 D2 = distance from the stern plane of particle 2 Dabs = thickness of any absorbed polymer on the particle surfaces Dtot is the distance at which the particle double layers begin to interact and a rise in

potential in the double layers is observed due to the interaction. Then through a series of

225

iterations Stabil calculates the attractive and repulsive energies between the particles

beginning at Dtot and continuing to the distance of closest of approach of the particles

defined by Dabs.

Heterocogulation is the case where dissimilar particle systems are interacting,

which is the case for the present case. In heterocogulation calculations it is necessary to

calculate the contribution from each particle for both the attractive and repulsive

interactions. For heterocogulation Stabil uses a simplified relation to calculate the

repulsive interaction between electrical double layers,

( )( ) ( )( ⎥

⎤⎢⎣

⎡−−+⎟⎟

⎞⎜⎜⎝

⎛−−−+

= DDDaaDV ror κκκψψεε 2exp1ln

exp1exp1ln2)( 2121 ) [C.5]

where, a1 = radius of particle 1 a2 = radius of particle 2 ψ1 = stern potential of particle 1 ψ2 = stern potential of particle 2

The calculation used for non-retarded van der Waals attration is,

⎥⎦

⎤⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛+++

+++

++++

++−=

yxxyxxxyx

yxxyxy

xxyxyA

DVa 2

2

22132 ln2

12)( [C.6]

where, x = D/(a1 + a2) y = a1/a2 (a1>a2) A132 = effective Hamaker constant

The effective Hamaker constant is calculated from the known Hamaker constant of the

two interaction materials and the solvent,

( )( )322311132 AAAAA −−= [C.7]

226

where, A11 = Hamaker constant of particle 1 A22 = Hamaker constant of particle 2 A3 = Hamaker constant of solvent

The total interaction energy of the particles is simply the addition of the attractive and

repulsive energies,

)()()( DVDVDV ratot += [C.8]

Finally, the interaction energies are plotted as a function of separation distance

and from the plot it is possible to determine the maximum repulsive energy and the

distance at which it occurs. Figure A.1 is a series of interaction energy curves calculated

using Stabil for the BaTiO3 and Pt system.

227

Figu

re C

.1.

Inte

ract

ion

ener

gy c

urve

s for

the

plat

elet

BaT

iO3

and

Pt e

lect

rode

. Th

e cu

rves

wer

e ca

lcul

ated

usi

ng S

tabi

land

the

phys

ical

con

stan

ts in

Tab

le C

.1.

228

References

1. Stabil for Windows, 4.5. 2. B. Derjaguin and L. Landau: Theory of the stability of strongly charged lyophobic sols and of the adhesion of strongly charged particles in solution of electrolytes. Acta Physiochim. URSS 14, 633 (1941). 3. E.J.W. Verwey, J.T.G. Overbeek, and K. Nes, Theory of the stability of lyophobic collloids. The interaction of sol particles having an electric double layer, 1st ed. (Elsevier Publishing Company, Inc., New York, New York, 1948). 4. L. Bergstrom: Hamaker constants of inorganic materials. Adv. Coll. Int. Sci. 70, 125 (1997). 5. I. Israelachvili, Intermolecular & Surface Forces, 2nd ed. (Harcourt Brace & Company, London, 1992). 6. J. Visser: On Hamaker constants: A comparison between Hamaker constants and Lifshitz - van Der Waals constants. Adv. Coll. Int. Sci. 3, 331 (1972). 7. R.J. Hunter, Zeta potential in colloid science: Principles and applications, 1st ed. (Academic Press, San Diego, CA, 1981).

229

VITA

Timothy James Yosenick

Timothy James Yosenick was born September 8th 1976 in Winfield, Illinois the

third child of Peter P. and Joan H. Yosenick. In August of 1994, Tim began attending the

University of Illinois where he earned a B.S. in Ceramic Engineering in May of 1998.

While attending the University of Illinois, he worked at the US Naval Research Lab in

Washington D.C. during a Co-op exchange program where he worked on sonar and

actuator applications for the US Navy. After graduation, he worked for a small research

company in Longmont, Colorado called Nanomaterials Research Corporation. Tim

began his graduate studies at the Pennsylvania State University in January of 2000. From

August to December of 2002 he worked as a visiting scientist for TDK Corporation in

their materials research center in Narita, Japan. Tim graduated from the Pennsylvania

State University in December 2005 with a Ph.D. in Materials Science and Engineering.

He took a position with General Electric’s Global Research Center located in Niskayuna,

New York.