169
The Pennsylvania State University The Graduate School Department of Electrical Engineering NANOSTRUCTURED DIELECTRIC FILMS FOR NEXT GENERATION OF ENERGY STORAGE CAPACITORS A Dissertation in Electrical Engineering by Yash Thakur 2017 Yash Thakur Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2017

NANOSTRUCTURED DIELECTRIC FILMS FOR NEXT GENERATION …

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

The Pennsylvania State University

The Graduate School

Department of Electrical Engineering

NANOSTRUCTURED DIELECTRIC FILMS FOR NEXT GENERATION OF

ENERGY STORAGE CAPACITORS

A Dissertation in

Electrical Engineering

by

Yash Thakur

2017 Yash Thakur

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

August 2017

ii

The dissertation of Yash Thakur was reviewed and approved* by the following:

Qiming Zhang

Distinguished Professor of Electrical Engineering

Dissertation Advisor

Chair of Committee

Jerzy Ruzyllo

Distinguished Professor of Electrical Engineering

Noel Chris Giebink

Charles K. Etner Assistant Professor of Electrical Engineering

James Runt

Professor of Polymer Science in Materials Science and Engineering

Michael Lanagan

Professor of Engineering Science and Mechanics

Kultegin Aydin

Professor of Electrical Engineering

Head of the Department of Electrical Engineering

*Signatures are on file in the Graduate School

iii

ABSTRACT

Advances in modern electronics require the development of polymer-based

dielectric materials with high dielectric constant, low dielectric loss, and high thermal

stability. The dielectric theory suggests that weakly-coupled and strongly-dipolar polymers

have the potential to realize a high dielectric constant. The high dipole moment functional

groups and amorphous structure provides strong scattering to the charge carriers, resulting

in low losses even at high electric fields. These polymers also possess a high glass

transition temperature which makes them suitable for high temperature operation. In this

dissertation, the fundamental understanding has been carried forward to design and develop

next generation of capacitors based on nanostructured materials for compact, light-weight,

and reliable electric power systems to address the commercial, consumer, and military

requirements.

We show through combined theoretical and experimental investigations that

nanostructure engineering of a weakly-coupled and strongly-dipolar polymer can result in

a high-energy density polymer with low loss and high operating temperature. Our studies

reveal that disorder in dipolar polymers creates a significantly larger free volume at

temperatures far below the glass transition (Tg), enabling easier reorientation of dipoles in

response to an electric field. The net result is a substantial enhancement in the dielectric

constant while preserving low dielectric loss and very high breakdown field. It is the free

volume effect that leads to a high dielectric constant (K > 5.6) at temperatures below Tg (>

200°C) in meta-phenylene polyurea (meta-PU). It possesses very low loss (high

charge/discharge efficiency) even at high electric fields (> 600 MV/m).

iv

To extend the idea of free volume, we propose a blending approach where two

glassy state dipolar polymers, poly(arylene ether urea) (PEEU, K=4.7) and an aromatic

polythiourea (ArPTU, K=4.4), are combined. The resulting blend exhibits a very high

dielectric constant(K=7.5) while maintaining low dielectric loss (< 1%). The experimental

and simulation results demonstrate that blending these dissimilar dipolar polymers causes

a slight increase in the interchain spacing of the blend in its glassy state. This reduces the

barriers for the reorientation of dipoles in the polymer chains and generates a much higher

dielectric response than the neat polymers.

In addition to designing new dielectric materials with excellent dielectric

properties, it is crucial that we continue to improve the electrical properties of the state-of-

the-art materials. This allows us to utilize the existential large-scale manufacturing

facilities of these polymers. Polyetherimide (PEI), a high glass transition amorphous

polymer, is seen as the material of choice for high temperature capacitors. But it possesses

a moderate dielectric constant of 3.2, which limits its energy density. We present a

nanocomposite approach, where addition of small amounts of inorganic nanoparticles in

PEI can improve the dielectric constant by 60% while maintaining the breakdown strength,

thereby increasing the discharged energy density by 50%. This is a very promising

approach and a breakthrough experimental discovery for engineering nanostructures by

introducing low volume content of nanofillers with dielectric constant similar to that of the

matrix, to achieve markedly enhanced dielectric response. The results are extremely

intriguing and eliminate many undesirable features, primarily being low breakdown

strength of traditional dielectric nanocomposites containing high dielectric constant fillers,

which have been a focal point of study for the past 20 years in this area of research.

v

For practical applications, it is critical that the dielectric material possesses low

loss, especially the conduction loss, which could become significant at high temperatures

and high electric fields. In this pursuit, we developed a strongly dipolar polymer, poly

(ether methyl ether urea) (PEMEU) that exhibits a dielectric constant of 4 and is thermally

stable up to 150°C. The experimental results show that the ether units are effective in

softening the rigid polymer and making it thermally processable, while the high dipole

moment of urea units and glass structure of the polymer leads to a low dielectric loss and

low conduction loss. As a result, PEMEU high quality thin films exhibit exceptionally high

breakdown field of >1.5 GV/m, and a low conduction loss at fields leading up to the

breakdown. Consequently, the PEMEU films exhibit a high charge–discharge efficiency

of 90% and a high discharged energy density of 36 J/cm3.

Another key aspect is mitigating losses in available dielectric materials that show

promise for scalability and are attractive for high energy density capacitors. The conduction

at high fields and high temperatures of a semi-crystalline poly(tetrafluoroethylene-

hexafluoropropylene-vinylidene fluoride) terpolymer was investigated. Experimental

results show that the insulating nanofillers are very effective in reducing the conduction

current, i.e., more than two orders of magnitude reduction in conduction can be achieved

with less than 1 wt.% (<0.5 vol.%) of Al2O3 nanofillers. Experimental measurements are

compared with multiscale simulations, which provide insights into the dominant

conduction mechanism, i.e., the carrier hopping in the polymer. The conduction is

markedly reduced owing to a large decrease in the mobile carrier concentrations and

increased trap depth, caused by the nanofillers.

vi

In summary, this dissertation focusses on the development of next generation

capacitors by innovation in materials which possess high dielectric constant, low loss, high

breakdown strength, and high temperature thermal stability. We believe that the insightful

results and approaches shown by the introduction of localized free volume and low volume

content of nanoparticles may unravel new directions for future research in advanced

dielectrics.

vii

TABLE OF CONTENTS

List of Figures .............................................................................................................. ix

List of Tables ............................................................................................................... xiv

Acknowledgements ...................................................................................................... xv

Chapter 1 Introduction ............................................................................................................. 1

1.1 Fundamentals of capacitors ........................................................................................ 3 1.1.1 Static electric field ........................................................................................... 5 1.1.2 Time-varying electric field .............................................................................. 6 1.1.3 Polarization mechanisms ................................................................................. 8 1.1.4 Energy storage ................................................................................................. 11

1.2 Dielectric polymers .................................................................................................... 13 1.3 Statement of goals, objectives and dissertation organization ..................................... 17

Chapter 2 Introduction of free volume to achieve high dielectric constant in dipolar

polymers ........................................................................................................................... 19

2.1 Introduction ................................................................................................................ 19 2.2 Review of free volume theory .................................................................................... 23 2.3 Experimental section .................................................................................................. 27

2.3.1 Synthesis and film fabrication of ordered and disordered structures .............. 27 2.3.2 Measurement of dielectric properties of polymer powders ............................. 29 2.3.3 Details of characterization equipment ............................................................. 30

2.4 Results and discussion ............................................................................................... 30 2.5 Conclusion ................................................................................................................. 39

Chapter 3 Blending of dipolar polymers to enhance the free volume effect............................ 40

3.1 Introduction ................................................................................................................ 40 3.2 Experimental section .................................................................................................. 40

3.2.1 Synthesis and film fabrication of blends ......................................................... 40 3.2.2 Details of characterization equipment ............................................................. 42

3.3 Discussion of dielectric data ...................................................................................... 42 3.4 Structural analysis ...................................................................................................... 48 3.5 Conclusion ................................................................................................................. 54

Chapter 4 Enhanced dielectric response in dipolar polymers with inorganic nanodopants ..... 55

4.1 Introduction ................................................................................................................ 55 4.2 Composite theory ....................................................................................................... 55

4.2.1 Models for predicting effective permittivity ................................................... 58 4.3 Experimental section .................................................................................................. 59

4.3.1 Nanocomposites of polyetherimide ................................................................. 59 4.3.2 Nanocomposites of polystyrene ...................................................................... 60 4.3.3 Details of characterization equipment ............................................................. 61

viii

4.4 Results and discussion ............................................................................................... 62 4.4.1 Effect of alumina nanoparticles and the particle size ...................................... 62 4.4.2 Effect of nanoparticle type on dielectric constant ........................................... 70 4.4.3 Effect of high dielectric constant nanoparticles .............................................. 72 4.4.4 Importance of dipoles ...................................................................................... 74 4.4.5 Structural analysis ........................................................................................... 78 4.4.6 Multilayer core model for interfacial effect of nanocomposites ..................... 80

4.5 Conclusion ................................................................................................................. 82

Chapter 5 Dipolar polymers: high field behavior and study of conduction loss ...................... 83

5.1 Introduction ................................................................................................................ 83 5.2 Review of breakdown mechanisms ............................................................................ 84

5.2.1 Electronic breakdown ...................................................................................... 85 5.2.2 Thermal breakdown ......................................................................................... 86 5.2.3 Electromechanical breakdown ........................................................................ 87 5.2.4 Frohlich amorphous solid model ..................................................................... 90 5.2.5 Dependence of breakdown strength on film thickness .................................... 91

5.3 Experimental section .................................................................................................. 93 5.3.1 Synthesis and film fabrication of PEMEU ...................................................... 93 5.3.2 Details of characterization equipment ............................................................. 95

5.4 Results and discussion of PEMEU ............................................................................. 95 5.5 Introduction of nanoparticle dopants to reduce the conduction loss .......................... 102 5.6 Review of conduction in polymers ............................................................................ 103

5.6.1 Electrode limited conduction........................................................................... 105 5.6.2 Bulk limited conduction .................................................................................. 108

5.7 Film preparation and characterization of THV nanocomposites ................................ 113 5.8 Results and discussion of THV nanocomposites ....................................................... 115 5.9 Conclusion ................................................................................................................. 129

Chapter 6 Conclusion and recommendations for future work ................................................. 131

6.1 Summary .................................................................................................................... 131 6.2 Suggestions for future work ....................................................................................... 135

Appendix A Chapter 4 Supporting Information ..................................................................... 140

Appendix B Chapter 5 Supporting Information ...................................................................... 144

Bibliography ............................................................................................................................ 145

ix

LIST OF FIGURES

Figure 1-1 Ragone plot comparing various energy storage devices. [1] .................................. 2

Figure 1-2 Various generations of capacitor: (a) Leyden jar is a glass vessel coated inside

and out by conducting electrodes; (b) cylindrical capacitor is a rolled-up parallel

plate capacitor; (c) the multilayer capacitor with its staggered electrodes. [3] ................ 4

Figure 1-3 Schematic of parallel plate configuration. .............................................................. 6

Figure 1-4 Equivalent circuit diagram of a capacitor under AC field. ..................................... 8

Figure 1-5 The frequency dependence of the real and imaginary parts of the dielectric

constant in the presence of various polarization mechanism. [10]................................... 11

Figure 1-6 Polarization-electric field responses of: (a) linear; (b) relaxor ferroelectric; (c)

ferroelectric; (d) anti-ferroelectric [13]. ........................................................................... 12

Figure 1-7 Schematic of polarization-electric field response for a dielectric material at

high fields. ........................................................................................................................ 13

Figure 1-8 Organization flow of the objectives of this dissertation. ........................................ 18

Figure 2-1 Dipolar structure of urea and thiourea units. [47] .................................................. 21

Figure 2-2 Schematics of (a) aromatic polyurea (ArPU), (b) aromatic polythiourea

(ArPTU), (c) meta-phenylene polyurea (m-PhPU/meta-PU), and (d) methylene

polythiourea (MePTU). [50] ............................................................................................ 22

Figure 2-3 Schematic diagram illustrating free volume as calculated by Simha and

Boyer.[56] ........................................................................................................................ 25

Figure 2-4 Schematic of synthesis of meta-PU. ....................................................................... 28

Figure 2-5 Schematic of synthesis of PEEU. ........................................................................... 28

Figure 2-6 X-ray data for (a) ordered meta-PU structures; (b) films of disordered meta-

PU structure...................................................................................................................... 31

Figure 2-7 (a) DSC data and (b) TGA data of meta-PU. ......................................................... 32

Figure 2-8 (a) Dielectric constant vs. frequency, measured at room temperature; (b)

dielectric constant vs temperature, measured at 1 kHz; (c) P-E loop of meta-PU film. .. 34

Figure 2-9 Dielectric constant vs frequency for the mixture of meta-PU/castor oil (open

squares) and castor oil (open circles). .............................................................................. 35

Figure 2-10 (a) Dielectric constant and (b) dielectric loss as a function of frequency for

ordered and disordered structure of meta-PU. ................................................................. 36

x

Figure 2-11. X-ray data for (a) ordered PEEU structures, powder; (b) films of disordered

PEEU structure. ................................................................................................................ 37

Figure 2-12 Dielectric constant as a function of frequency for (a) disordered and (b)

ordered structure of PEEU. .............................................................................................. 38

Figure 3-1 Schematic of synthesis of aromatic polythiourea. .................................................. 41

Figure 3-2 Blend of two polymers: ArPTU and PEEU. .......................................................... 41

Figure 3-3 Dielectric data of the 1:1 blend of PEEU and ArPTU (a) as a function of

frequency at room temperature, including the inset which shows the dielectric data

of PEEU and ArPTU; (b) as a function of temperature at different frequencies. (c)

Dielectric constant vs. blend composition (weight ratio of PEEU:ArPTU) at room

temperature and 1 kHz. Data points are shown and the dashed line is drawn to guide

the eye. ............................................................................................................................. 44

Figure 3-4 Dielectric data of (a) ArPTU, (b) PEEU as functions of temperature measured

over a range of frequencies. ............................................................................................. 46

Figure 3-5 (a) Computational results of dielectric constant vs. specific volume for PEEU,

ArPTU, and blends for various supercells with different PEEU:ArPTU weight ratios.

(b) Comparison of simulation and experimental data of dielectric constant vs. blend

composition (PEEU:ArPTU weight ratio) at room temperature. Data points are

shown and the dashed line is drawn to guide the eye. [65] .............................................. 47

Figure 3-6 (a) X-ray diffraction data of ArPTU and PEEU, and their 1:1 blend.

Background subtracted data of (b) PEEU with peak at 18.6°, (c) ArPTU with peak at

18.6° and (d) blend data with peak at 17°. Wavelength of X-ray used was 1.54

angstroms. ........................................................................................................................ 49

Figure 3-7 AFM images: (a) amplitude and (b) phase for the PEEU:ArPTU 1:1 blend. ......... 50

Figure 3-8 DSC data of 1:1 PEEU:ArPTU blend. ................................................................... 50

Figure 3-9 TGA data of 1:1 PEEU:ArPTU blend. ................................................................... 51

Figure 3-10 PALS results of (a) positron lifetime, (b) spherical specific hole volume with

change in PEEU composition........................................................................................... 53

Figure 4-1 Molecular structure of the polyetherimide. [94] .................................................... 60

Figure 4-2 Molecular structure of the polystyrene. .................................................................. 61

Figure 4-3 (a) Room temperature dielectric properties of PEI/alumina (20 nm particle

size) nanocomposites at different alumina nanoparticle loading (in vol. %) vs

frequency. (b) Dielectric constant of nanocomposite films of PEI/alumina (20 nm

particle size) vs. nanofiller volume content and comparison with several widely used

dielectric models of diphasic dielectric composites (lines with no data points): curve

xi

(1) Parallel model, (2) Maxwell model, (3) Lichtenecker model, and (4) series

model. Inset shows an expanded view of the dielectric constants of the composite

films vs. alumina loading. Experimental data points are shown and lines are drawn

to guide the eye. (c) Dielectric properties vs. temperature of the PEI/alumina (20 nm

size) nanocomposite with 0.32 vol.% alumina loading at different frequencies. ............. 64

Figure 4-4 Dielectric properties at different frequencies of neat PEI as a function of

temperature....................................................................................................................... 65

Figure 4-5 (a) Charge-discharge cycles of PEI/alumina (20 nm) nanocomposites with

0.32 vol.% alumina under different electrical fields at 10 Hz and room temperature.

Inset: discharged energy density deduced from the charge-discharge cycle data. (b),

(c), (d) Discharged energy density of PEI/alumina (20 nm) nanocomposites with

0.32 vol.% loading at different temperatures (room temperature, 100oC, and 150oC),

and their comparison with that of neat PEI and BOPP (room temperature), measured

under 350 MV/m at 10 Hz. In (b), (c), and (d), Data points are shown and curves are

drawn to guide the eye. Data for BOPP were taken from Ref. [44] ................................. 66

Figure 4-6 Weibull plot showing failure distribution for PEI-0.32% Al2O3 nanocomposite

film. .................................................................................................................................. 67

Figure 4-7 Effect of nanofiller size on the dielectric response (at 1 kHz) of PEI/alumina

composite films vs filler volume content. ........................................................................ 68

Figure 4-8 Effect of nanofiller size on the dielectric response (at 1 kHz) of PEI/alumina

composite films vs filler volume content. Experimental data points are shown and

curves are drawn to guide the eye. ................................................................................... 71

Figure 4-9 Dielectric constant of PEI/BaTiO3 (50 nm size) nanocomposites vs. BaTiO3

volume content (experimental data points are shown and for > 3 vol.%

nanocomposites (orange squares) the data are from Ref. [95]). Experimental data are

compared with several commonly used composite models (Refs. [91]–[93]): (1)

Parallel model (black), (2) Maxwell model (red), (3) Lichtenecker model (green),

and (4) series model (blue), assuming the dielectric constant of BaTiO3 is 100X of

that of PEI. Inset is an expanded view of the enhanced dielectric response of

nanocomposites at very low volume content (< 1 vol.%) due to nanoparticle

interfacial effects, experimental data points are shown and solid curve is drawn to

guide the eye. ................................................................................................................... 73

Figure 4-10 Summary of dielectric constants of PEI nanocomposites with different

nanofillers (20 nm MgO; 20 nm SiO2; 20 nm alumina; 50 nm BaTiO3; 70 nm BN).

Experimental data points are shown and lines are drawn to guide the eye. ..................... 73

Figure 4-11 Dielectric constant measured at 1 kHz and room temperature vs. the

nanofiller content for PS nanocomposites. Data points are shown and solid curves

are drawn to guide the eye. .............................................................................................. 75

xii

Figure 4-12 (a) Dielectric data at different frequencies of PEI+0.32 vol.% Al2O3 20 nm

as a function of temperature, (b) Dielectric data of PEI and (c) PEI+0.32 vol.%

Al2O3 20 nm nanoparticle as a function of frequency at room temperature..................... 77

Figure 4-13 A representative TEM image of the PEI nanocomposite with 0.32 vol.%

alumina (20 nm particle size). Due to low volume content of nanoparticle in the

composite, only one nanoparticle is seen in the image area, as indicated. ....................... 79

Figure 4-14 (a) DSC and (b) X-ray diffraction data of PEI and the PEI nanocomposite

with 0.32 vol.% of alumina. ............................................................................................. 79

Figure 4-15 Tanaka’s multi-core model for interfaces between inorganic nanoparticles

and polymer matrix. [67], [97] ......................................................................................... 81

Figure 5-1 Schematic of synthesis and chemical structure of poly(ether methyl ether

urea), PEMEU. ................................................................................................................. 94

Figure 5-2 1H-NMR spectrum for PEMEU in DMSO-d6. ...................................................... 94

Figure 5-3 (a) Wide angle X-ray diffraction data at room temperature and (b) DSC data

of PEMEU film measured during heating. ....................................................................... 97

Figure 5-4 Dielectric constant and loss as functions of (a) frequency measured at room

temperature, and (b) temperature at frequencies from 1 kHz to 1 MHz of PEMEU

films. The error bars are attributed to the variation in thickness of film and the

electrode area. .................................................................................................................. 98

Figure 5-5 Electric breakdown field vs. film thickness for the PEMEU films measured at

room temperature. Dots represent y-axis error bars and symbols represent x-axis

error bars. ......................................................................................................................... 100

Figure 5-6 (a) AFM image, (b) Charging/discharging curves under different unipolar

fields, (c) Schematic showing calculation of discharged energy density and loss

under high field from the charging/discharging curves, (d) Discharged energy

density as a function of field of PEMEU thin films of 1.32 μm thick, measured at

room temperature. Dots represent y-axis error bars and symbols represent x-axis

error bar. ........................................................................................................................... 101

Figure 5-7 Schematic of conduction process in polymers. [12]............................................... 104

Figure 5-8 Schematic showing Schottky contact between metal and n-type polymer (a)

before contact, (b) after contact, (c) barrier lowering by image force and (d) barrier

lowering by external voltage. [126], [127] ....................................................................... 107

Figure 5-9 Schematic graph showing current density versus voltage for an ideal case of

space-charge limited current.[12] ..................................................................................... 110

Figure 5-10 Schematic showing random resistor network percolation. ................................... 113

xiii

Figure 5-11 Dielectric data of neat THV as a function of temperature.................................... 115

Figure 5-12 Dielectric data as a function of frequency for THV and THV nanocomposite

films, (b) DMA of THV and THV+0.5 wt.% Al2O3 films. .............................................. 117

Figure 5-13 (a) Current density, (b) conductivity as a function of field at different

temperatures, (c) conductivity as a function of temperature at different fields for neat

THV film. ......................................................................................................................... 118

Figure 5-14 (a) Current density, (b) conductivity as a function of field at different

temperatures, (c) conductivity as a function of temperature at different fields for

THV+0.5 wt.% Al2O3 film. .............................................................................................. 120

Figure 5-15 (a) Current density, (b) conductivity as a function of field at different

temperatures, (c) conductivity as a function of temperature at different fields for

THV+1 wt.% Al2O3 film. ................................................................................................. 122

Figure 5-16 (a) Comparison of conduction current of neat THV and different filler

loadings at 125°C, (b) scatter of conductivity at 60 MV/m as a function of alumina

nanofiller content. Dashed lines are drawn to guide the eye. ........................................... 123

Figure 5-17 Comparison of leakage conductivity from simulation and measurement at

85C and 125C for: (a) THV+0.5 wt.%, and (b) THV+1.0 wt.%, (c) carrier

concentration and trap depth as a function of filler content. ............................................ 125

Figure 5-18 X-ray diffraction data of neat THV polymer and THV+1 wt.% Al2O3

nanocomposite films. ....................................................................................................... 127

Figure 5-19 Two-dimensional X-ray diffraction data of neat THV polymer and THV+1

wt.% Al2O3 nanocomposite films. .................................................................................... 128

Figure 6-1 Current density as a (a) function of electric field over a range of temperatures,

(b) function of temperatures at 95.1 MV/m. .................................................................... 138

Figure 6-2 High frequency characterization of PEI-1wt.%Al2O3 nanocomposite films. ......... 139

xiv

LIST OF TABLES

Table 1-1 Summary of dielectric materials studied for capacitor applications. [20], [40] ...... 15

Table 2-1 Dipole moments of common dipolar units present in dielectric polymers. [15] ..... 20

Table 2-2 Summary of dielectric properties of ArPU, ArPTU, m-PhPU and MePTU. [50] ... 23

Table 2-3 Summary of experimental dielectric data of ordered and disordered structure of

meta-PU and PEEU. ......................................................................................................... 37

Table 3-1 Summary of the dielectric properties of the neat polymers and blends at 25 ˚C ..... 45

Table 4-1 Summary of dielectric data of polyetherimide (PEI) with alumina

nanoparticles. ................................................................................................................... 69

Table 4-2 Summary of dielectric data of polyetherimide (PEI) with different type of

nanoparticles. ................................................................................................................... 71

Table 4-3 Summary of dielectric data of polyetherimide (PEI) with barium titanate

(BaTiO3) nanoparticles. .................................................................................................... 74

Table 4-4 Summary of dielectric data of non-polar polystyrene (PS) nanocomposite films. .. 76

Table 5-1 Summary of dielectric theories of solids. [106]....................................................... 89

Table 5-2 Summary of fitting parameters of hopping conduction equation for the neat

THV and nanocomposites. ............................................................................................... 126

xv

ACKNOWLEDGEMENTS

First and foremost, I am deeply grateful to my adviser Prof. Qiming Zhang for his

constant guidance throughout this endeavor. He has tapped into my potential and brought

the best out in me. His work ethics are exemplary and have been a source of constant

motivation for me. Under his tutelage, I have become more disciplined and driven.

I sincerely appreciate my committee members: Prof. James Runt, Prof. Michael

Lanagan, Prof. Jerzy Ruzyllo and Prof. Noel Geibink. I am grateful to Prof. James Runt for

collaborating on my projects and I am indebted for his guidance and help with broadband

spectroscopy.

I would like to convey special thanks to Prof. Jerzy Ruzyllo, who was my master’s

adviser and has nurtured me through the early years of graduate studies. He gave me time

and guidance to prepare for the rigors of doctoral studies. A special mention to Prof.

Thomas Jackson who encouraged me to pursue doctoral degree and laid out strong

fundamentals of device physics. I appreciate his patience in solving my doubts and

strengthening my basics. Prof. Ashok has been a mentor and he has encouraged me

throughout my graduate studies.

I am thankful to my lab mates: Tian Zhang, Dr. Minren Lin, Dr. Shan Wu, Dr.

Xiaoshi Qian, Lu Yang. I sincerely acknowledge Dr. Minren Lin, who helped me in my

first year with polymer synthesis and film fabrication; Tian Zhang for nanocomposite film

preparation and Ciprian Iacob for performing broadband dielectric spectroscopy on my

samples. Special thanks to Jeff Long and Steve Perrini for their help and guidance in the

electrical characterization lab. I am grateful to Nicole Wonderling and Gino for their help

xvi

with XRD measurements, and rest of the MCL staff for their useful suggestions and help

with material characterization.

I would like to thank my collaborators: Prof. Jerry Bernholc at N.C. State and his

students Rui Dong and Bing Zhang, who carried out the simulation part of our free volume

study; Dr. Meng H. Lean (CTO of QEDone LLC) for his bipolar charge transport study;

Prof. Long-Qing Chen and his student Tiannan Yang for carrying out simulation study on

interfacial effect of nanoparticles; Prof. Qing Wang and his students Feihua Liu and Guang

Yang for their help with TGA and NMR study, and finally Prof. David Gidley from

University of Michigan for collaborating on PALS study. I sincerely acknowledge the

support of Office of Naval Research who supported this study.

A special mention to all my friends at Penn State especially Jared, Alyssa, Shruti,

Tanushree, Ganesh, Rahul Pandey, Rahul Simham Nitesh, Shantanab for their constant

support and company.

Last but not the least, I want to thank my beloved parents – Namita Singh and

Balwant Singh, grandmother Indra Shukla, and my extended family members – Dr. Amita

Dave, Mr. Atul Dave, Dr. Vasu Misra and Dr. Neeraj Tripathi for their constant support

and encouragement. I would like to dedicate this thesis to my loving parents.

1

Chapter 1

Introduction

The development of efficient and high-performance devices for electrical energy storage

is essential to meet the ever-increasing demands for electrical energy. For past century,

researchers have sought better ways to store energy and the continued research in this area

has led to the development of various energy storage devices: batteries, fuel cells,

supercapacitors and capacitors. [1] These energy storage devices are shown in Figure 1-1.

[1] For microsecond to fractional-second electrical energy storage, discharge, filtering and

power conditioning, capacitor technology is unparalleled in flexibility and adaptability to

meet the broad range of requirements of the present and the future. [2] The development

of advanced dielectrics, which enable capacitors to store more charge and withstand high

voltages can fulfill the need for compact, light-weight and reliable electrical power systems

to meet commercial, consumer and military requirements.

In ancient times, the Greeks were the first to recognize that they could separate

charge by rubbing certain dissimilar materials. Even today, the charge storage amounts to

separation of ions and electrons. This was demonstrated over two centuries ago by the

Leyden jar and Volta’s pile, the predecessor of the modern capacitors and battery,

respectively. Capacitors were originally known as condensers with reference to the ability

to store a higher density of electric charge than a normal isolated conductor. In the late

1950s, it was decided to harmonize the nomenclature of most electrical components, and

2

the term “capacitor” was coined to fall in line with “resistor” and “inductor”. The

progression of capacitors has been illustrated in Figure 1-2. [3]. A capacitor generally

consists of metallic conducting plates or foils separated by a dielectric material. These

metallic plates accumulate charge when the voltage is applied, resulting in electrical energy

stored in the dielectric material.

Figure 1-1 Ragone plot comparing various energy storage devices. [1]

Capacitor is a fundamental element of both digital and analog electronic circuits.

With the advent of modern electronics, the applications of capacitor have expanded, and

their utility seems limitless. One of the most important applications of capacitors is for the

energy storage component in a pulsed power supply (PPS). [4], [5] Pulse power technology

has been mainly developed for military applications that require extremely high peak

power, such as electromagnetic rail guns, lasers. [4], [6] In addition, these pulse power

supplies find application in external medical defibrillators, X-ray systems and various other

3

monitoring systems in medical industry. The capacitor-based system offers advantages of

high power density, little magnetic flux leakage, and graceful degradation.

Another important application is DC-bus capacitors used in power electronic

devices, especially inverters, where they can pave the way for high performance hybrid

electric vehicles (HEV’s). [7], [8] The capacitors have become an integral part of energy

saving systems in auto sector – such as auto ignition, regenerative braking etc. A modern

car may use as many as 1700 capacitors for various functions and accessories. [9] Next

generation power inverters with lower cost, high-efficiency, light-weight, better

performance and lifetime are the biggest challenges. Moreover, the high-power density

demand has led to significant challenges on thermal management since the power loss has

also proportionally increased. The present polypropylene capacitors can only be used at

105°C, if they are installed with a secondary cooling loop. [8] Thus, innovation in

capacitors, a crucial component, can bring a revolution in the electric car industry.

1.1 Fundamentals of capacitors

Capacitor is one of the three basic passive circuit components of any electrical

circuit. They provide electrical energy to be stored over a relatively long charging time and

then released over short (microseconds-milliseconds) periods.

The capacitor can be modeled as two conducting plates with area A separated by a

dielectric as shown in Figure 1-3. When a voltage (V) is applied across the plates, a charge

+q accumulates on one plate and a charge -q on the other. The capacitance, a characteristic

of the charge storage capability is expressed as:

4

𝐶 =𝑄

𝑉= 𝜖𝑟𝜖0

𝐴

𝑑 (1.1)

where ϵr is the relative permittivity or also known as the dielectric constant, and ϵ0 is the

permittivity of free space (ϵ0 = 8.854 x 10-12 F/m).

Figure 1-2 Various generations of capacitor: (a) Leyden jar is a glass vessel coated inside

and out by conducting electrodes; (b) cylindrical capacitor is a rolled-up parallel plate

capacitor; (c) the multilayer capacitor with its staggered electrodes. [3]

(a) (b)

(c)

5

1.1.1 Static electric field

The relationship between the static electric field E and the electric displacement D

can be derived from Maxwell equations:

𝐷 = 𝜖E (1.2)

where ϵ is the dielectric permittivity. In isotropic media, the relative permittivity

(ϵr) is given by ratio of dielectric permittivity (ϵ) and permittivity of free space (ϵ0).

𝜖𝑟 =𝜖

𝜖0 (1.3)

Instead of free space, let us consider a capacitor with a dielectric material inserted

between the two plates. The material will respond to the applied field by redistributing its

charge components, which will induce polarization charges P at the surface of the material.

In this case, a part of the charge density qs (qs = Q/A) is free charge as in the case of free

space. Another significant part of the charge density qs is bound at the boundaries of P for

charge compensation on the surfaces of the material in contact with the metal plates. This

bound surface qb is opposite in polarity and equal in magnitude to P. This can be

mathematically stated as:

𝑞𝑠 = (𝑞𝑠 − 𝑞𝑏) + 𝑞𝑏 = 𝐷 = 𝜖0𝐸 + 𝑃 = 𝜖𝑟𝜖0𝐸

= 𝜖0𝐸 + 𝜖0 (𝜖𝑟 − 1)𝐸 (1.4)

Polarization can be expressed as:

𝑃 = 𝜖0 (𝜖𝑟 − 1)𝐸 = 𝜖0𝜒𝑒𝐸

=𝑏𝑜𝑢𝑛𝑑 𝑐ℎ𝑎𝑟𝑔𝑒

𝑠𝑢𝑟𝑓𝑎𝑐𝑒 𝑎𝑟𝑒𝑎 (1.5)

6

= 𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑖𝑛𝑑𝑢𝑐𝑒𝑑 𝑑𝑖𝑝𝑜𝑙𝑒 𝑚𝑜𝑚𝑒𝑛𝑡

𝑣𝑜𝑙𝑢𝑚𝑒= 𝑁⟨�⃗� ⟩

where ⟨�⃗� ⟩ is the average dipole moment, 𝜒𝑒 is defined as the electric susceptibility.

Figure 1-3 Schematic of parallel plate configuration.

1.1.2 Time-varying electric field

When a time-varying voltage is applied to the dielectrics, the conduction current can be

expressed as:

𝐽𝑇 = 𝐽 + 𝑑𝐷

𝑑𝑡 (1.6)

where J is the conduction current. Consider a time-varying electric field is monochromatic

and a sinusoidal function with angular frequency ω as shown below:

𝐸 = 𝐸𝑚 exp(𝑗𝑤𝑡) (1.7)

7

Complex permittivity needs to be considered in this case and can be expressed as:

𝜖∗ = 𝜖′ − 𝑗𝜖′′ = (𝜀𝑟′ − 𝑗𝜀𝑟

′′)𝜖0 (1.8)

where 𝜀𝑟′ is the real part of the complex permittivity, also referred to as the dielectric

constant, and 𝜀𝑟′′ is the imaginary part. From here, the relative magnitude of losses can be

estimated by the dissipation factor (DF) or the loss tangent (tan δ), defined as:

tan δ = 𝜀𝑟′′

𝜀𝑟′ (1.9)

In a physical sense, the dielectric loss (tan δ) is due to the movement or rotation of

atoms or molecules in an alternating electric field. These losses depend on the temperature

as well as the frequency of the applied voltage. The dipoles or molecules cannot keep up

with change in the electric field when frequency is increased. The rotation of dipoles or

their ease of movement is a temperature dependent process, which will increase with rise

in temperature. Consequently, dielectric loss is directly proportional to both frequency and

temperature. There’s another loss in dielectrics known as the conduction loss, which

represents the flow of actual charge through the dielectric and occurs mostly at high field

and high temperature.

The capacitor under an AC field can be expressed by an equivalent diagram as

shown in Figure 1-4, where resistive component represents the loss. The quality factor (Q)

of a capacitor represents the efficiency of a given capacitor in terms of energy losses. It is

defined as:

𝑄 =𝑋𝑐

𝑅𝑐=

1

𝜔𝐶𝑅𝐶 (1.10)

where Xc is the reactance of the capacitor, C is the capacitance, Rc is the equivalent series

resistance and ω is the frequency at which the measurement is taken.

8

The dissipation is related to the quality factor by:

tan δ = 1

𝑄 (1.11)

The total current under electric field with a frequency of ω may be written as:

𝐽𝑇 = 𝐽 + 𝜖∗ 𝑑𝐸

𝑑𝑡= 𝜎𝐸 + 𝑗𝜔(𝜖′ − 𝑗𝜖′′)𝐸 (1.12)

Figure 1-4 Equivalent circuit diagram of a capacitor under AC field.

1.1.3 Polarization mechanisms

The frequency dependence of the real and imaginary parts of the permittivity are shown in

Figure 1-5. Each dielectric mechanism has a characteristic relaxation frequency. There are

four main types of polarization mechanisms:1) Electronic; 2) Atomic; 3) Orientational; 4)

Space-charge. [10]

9

Electronic polarization is associated with elastic displacement of electron cloud with

respect to nucleus under the influence of applied electric field. Atomic polarization is due

to displacement of ions/atoms. Both electronic and atomic polarization occur at frequencies

in the optical (>1015 Hz) and infrared (1012-1014 Hz) range, respectively. Thus, they are

classified in the resonance regime. Also, both the mechanisms are temperature

independent, as the phenomenon is intramolecular in nature. These mechanisms occur at

high frequencies and are instantaneous, which gives an indication that the dielectric

contribution in this regime is relatively smaller, irrespective of the polymer being polar or

non-polar. [11]

Orientation or dipolar polarization exists in polar materials consisting of dipoles,

for example, water has a permanent dipole in the structure. When electric field is applied,

the dipoles align along the direction of field, resulting in a net finite dipole moment per

molecule. The net polarization contributes to the effective dielectric constant of the

polymer. These dipoles need energy to overcome the resistance offered by the surrounding

molecules, which can be provided by thermal energy. Thus, this phenomenon is strongly

temperature dependent. When the field is removed, these dipoles take time to relax back to

the equilibrium. Hence, this type of polarization falls in the relaxation regime, and usually

this relaxation happens in audio frequencies.

Space charge polarization, also known as Maxwell-Wagner-Sillars interfacial

polarization occurs whenever there is an accumulation of charge at an interface between

two materials or between two regions within a material. The simplest example is interfacial

polarization due to the accumulation of charges in the dielectric near one of the electrodes.

10

It can also be due to the presence of impurities, non-homogeneity or incomplete contact of

the film with the electrode, which leads to regions of accumulated trap charges in the

dielectric medium. This mechanism is observed at low frequencies (<103 Hz). [10], [11]

In the presence of electronic, ionic, and dipolar polarization mechanisms, the

average induced dipole moment per molecule will be the sum of all the contributions in

terms of the local field.

𝑝𝑎𝑣 = 𝛼𝑒𝐸𝑙𝑜𝑐 + 𝛼𝑖𝐸𝑙𝑜𝑐 + 𝛼𝑑𝐸𝑙𝑜𝑐 (1.13)

where αe is the electronic polarizability, αi is the ionic polarizability and αd is the

dipolar polarizability. Each effect adds linearly to the net dipole moment per molecule.

Interfacial polarization cannot be simply added to the above equation because it occurs at

the interface and cannot be put into an average polarization per molecule in the bulk. [10]

11

Figure 1-5 The frequency dependence of the real and imaginary parts of the dielectric

constant in the presence of various polarization mechanism. [10]

1.1.4 Energy storage

During the charging and discharging of a capacitor, the stored (charged) and released

(discharged) energy density (Ue) can be calculated by the equation [12]

𝑈𝑒 = ∫𝐸𝑑𝐷 (1.13)

In linear dielectrics, the ratio between the polarization and the electric field is constant and

has typical P-E response as shown in Figure 1-6. For linear dielectrics, the equation can be

simplified to:

𝑈𝑒 = 1

2𝜀0𝜀𝑟𝐸

2 (1.14)

12

Other class of dielectrics are called non-linear dielectrics. They can be further divided into

three classes: relaxor ferroelectric, ferroelectric, and anti-ferroelectric. The P-E curves of

these materials are summarized in Figure 1-6.

Figure 1-6 Polarization-electric field responses of: (a) linear; (b) relaxor ferroelectric; (c)

ferroelectric; (d) anti-ferroelectric [13].

During the high field charging and discharging, most dielectrics exhibit a non-linear

increase of losses from the polarization hysteresis and high field conduction, especially at

fields greater than 100MV/m. This non-linear relationship between energy loss and the

electric field makes it very difficult to predict the high field loss from that at low electric

field. To make a quantitative comparison, Wu et. al. introduced the energy storage

efficiency (ƞ) for the charge-discharge cycle at high fields in Figure 1-7. [14]

𝜂 =𝑈𝑟

𝑈𝑠= 1 −

𝑈𝑙

𝑈𝑠 (1.15)

where the Us, Ur, and Ul are the charged, discharged and electrical loss energy densities in

the charging-discharging cycle respectively. The high field loss can be defined as 1- ƞ.

13

Figure 1-7 Schematic of polarization-electric field response for a dielectric material at

high fields.

1.2 Dielectric polymers

Dielectric materials store energy electrostatically through various polarization mechanisms

and release it by depolarization. Dielectric capacitors are unparalleled in flexibility,

adaptability, and efficiency for electrical energy storage, filtering, and power conditioning.

[2], [15]–[21] They are highly desirable for applications in the area of capacitive energy

storage, transistors, photovoltaic devices and electrical insulation.[2], [16], [22]–[24],

[24]–[28] The demand for capacitive energy storage has increased due to continuing

electrification of land and sea transportation, as well as military and civilian systems. [6],

[29]–[31] These applications require capacitors with high energy density, low loss, high

efficiency, and high operating temperature. Compared to ceramics and electrolytic

capacitors, polymer-based capacitors are attractive because they feature low manufacturing

14

cost and low dielectric loss, can be used under high voltage due to high breakdown strength,

and fail gracefully with an open circuit.[2], [15], [16] In many of these devices and systems,

capacitors constitute a substantial fraction of volume and weight (>30% volume and

weight).[7], [8], [32] To meet the demand of continued miniaturization of modern

electrical and electronic systems, the energy density of dielectric polymers must be

improved. In general, the energy stored in a capacitor is proportional to the dielectric

constant and the square of the electric field. Therefore, the materials of interest should

display high dielectric constant and high breakdown strength.

The present state-of-the-art high energy density film capacitors use biaxially

oriented polypropylene (BOPP). It is attractive for energy storage and regulation

applications, such as capacitors in HEVs and power grids due to its high dielectric

breakdown strength and low dielectric loss (< 0.018% when measured at low electric field).

However, the low dielectric constant (K ≈ 2.2) of BOPP limits its energy density. [20], [32]

Also, for many widely used linear dielectrics, including BOPP, it is found that the

conduction loss becomes more significant at higher applied fields. [12], [33] Normally,

these losses increase exponentially with the electric field, and cause Ohmic heating of the

capacitors. [12], [33], [34] This results in the need to have a cooling system to avoid

overheating of the BOPP film capacitors. For example, in hybrid electric vehicles, an extra

cooling loop has to be introduced in the BOPP capacitor banks in order to prevent a

runaway temperature increase caused by the conduction loss heating. [7], [8] Extensive

materials development efforts have led to several alternative dielectric polymers, including

polycarbonate (PC), poly(ethylene terephthalate) (PET), and poly(phenylene sulfide)

15

(PPS) with high operating temperatures (>125°C). However, the dielectric constant of

these polymers is still below 3.3. [20], [35] Table 1-1 summarizes the dielectric properties

of polymers used in industry and some recently developed dielectric polymers in research

labs.

On the other hand, the strong coupling among dipoles have led to high dielectric

constants of polyvinylidene fluoride(PVDF)-based ferroelectric polymers (K > 10). [15]–

[17], [36] By proper defect modifications of PVDF-based polymers, it has been shown that

these polymers can achieve either a high dielectric constant at room temperature (K > 50)

or a very high energy density (> 25 J/cm3). [37]–[39] However, the strong dipolar coupling

in these ferroelectric polymers causes high polarization hysteresis loss. The operating

temperature is still limited to below 100°C due to low Tm (< 140°C). Nevertheless, the

results demonstrate the potential of tailoring nano and meso-structures of dielectric

polymers to achieve high dielectric performance.

Table 1-1 Summary of dielectric materials studied for capacitor applications. [20], [40]

16

Material

Dielectric

Constant

(25°C)

Dielectric

Loss at

1 kHz

(10-3)

Working

Temperature

(°C)

Dielectric

Strength

(MV/m)

Energy

Density at

Breakdown

Strength

(J/cm3)

BOPP 2.2 0.2 90 820 6.2

PTFE 2.1 0.5 260 300 2.1

Solvent-cast PC 3.1 1.3 125 <820 9.2

Cyano-PC 3.2 3 180 710 7.1

PhONDI 3.2 7.8 150 350 1.7

PPS 3.1 0.5 150 470 3

PEEK 3.2 4 150 320 1.4

PEI 3.2 2 200 460 3

FPE 3.3 2.6 275 - -

PI 3.3 2 300 300 1.3

PEN/Si3N4 3.7 7 125 578 5.4

Polyurea 4.2 5 180 800 12

Nanolayer

PCPVDF-

HFP

4.6 10 125 750 14

Cyano-PEI 4.7 3 220 745 11.5

Modified polyurea 5.2 6.8 150 >700 >12

Modified PI 4–7 5 10 250 - -

Siloxane 8.6 60 150 - -

P(TFE-co-VDF) 10.2 30 270 >225 >2

PVDF-CTFE 11 50 125 750 27

Alkali-free barium

boroaluminosilicate

glass

6 5 180 >1000 >30

17

1.3 Statement of goals, objectives and dissertation organization

The objective of this work is to design nanostructured dielectric polymers based on

fundamental polymer physics for developing high energy storage capacitors. Considering

the rich polymer chemistry available for modifying and tuning the nanostructures, the new

directions developed by this dissertation will work towards the goal to generate polymeric

materials with high dielectric constant, low loss, high breakdown strength and high

operating temperature.

Chapter 2 and Chapter 3 discuss an unconventional approach to improve the

dielectric properties by introducing free volume in the dipolar polymers. The two

approaches: introducing disorder and blending of polymers, have been extensively

discussed. The introduction of free volume improves the dielectric constant while

maintaining low loss.

Chapter 4 discusses a nanocomposite approach, where addition of small amounts

of inorganic nanoparticles in PEI can improve the dielectric constant. This is a very

promising approach for engineering nanostructures by introducing nanofillers with

dielectric constant similar to that of the matrix, to achieve markedly enhanced dielectric

response.

In Chapter 5, the high field conduction and breakdown properties of dipolar

polymers is studied. The high quality thin films of the polyurea show colossal breakdown

strengths, demonstrating that introduction of polar units can be effective in scattering of

mobile charges, and defect free thin films can achieve high breakdown strength. In

addition, an approach to reduce the conduction loss in a semi-crystalline polymer, which

18

has been shown to be attractive for high energy storage capacitors has been discussed.

Experimental results show that introduction of tiny volume content of these nanoparticles

can dramatically reduce the conduction loss, thus making it practical for high temperature

operation.

Figure 1-8 Organization flow of the objectives of this dissertation.

Rational design of polymers

Approaches to improve the dielectric constant

Free volume approach

Introducing disorder in dipolar polymers

Blending of two dipolar polymers

Nanocomposite approach

Enhanced dielectric response using low volume content of inorganic dopants

Study of high field behaviour and

conduction loss

Ultra-thin films with high dipole moment

units

Doping of semi-crystalline polymer with low volume content of fillers

19

Chapter 2

Introduction of free volume to achieve high dielectric constant in

dipolar polymers

2.1 Introduction

In Chapter 1, we discussed the fundamentals of capacitors, and dielectric properties

of state-of-the-art polymers. The dielectric constant of these high energy density polymers

is still below 3.3. [15], [16], [20], [41] There has been many studies in the past where efforts

have been made to improve the dielectric constant. [11], [16], [20], [36], [42]. Still, for

next generation of materials, rational design of materials holds the key to fundamental

advances in energy storage, and is a smarter approach – given the exhaustive set of

materials available for selection. In dielectric polymers, a necessary condition for achieving

high dielectric constant is that they contain dipoles in the polymer chains. Strong coupling

among the dipoles can lead to high dielectric constant, as have been observed in semi-

crystalline polyvinylidene fluoride (PVDF)-based ferroelectric polymers (K > 10). [16],

[36] However, the strong coupling among dipoles causes large hysteresis loss, not desirable

for most polymer capacitor applications. Strongly-dipolar materials, in which dipole

moment is larger than 3 Debye with weak dipolar coupling have the potential to reach

relatively high dielectric constant than the widely used non-polar polymers including the

state-of-the-art dielectric polymer BOPP, and exhibit a lower loss. The dipole moments of

some functional groups of dielectric polymers are summarized in Table 2-1.

20

To reduce or even eliminate the polarization hysteresis loss, Zhang et. al. have

developed a class of amorphous polymers, containing high density dipoles of high dipole

moment, i.e., urea and thiourea (see Figure 2-1). [43]–[46] It has been shown that by

increasing the dipole moment and the dipole density, the dielectric constant in this series

of polymers increases from 4.1 to 5.7. The high dipole moments in these amorphous

polymers provide strong polar-scattering centers and traps, which significantly reduces the

conduction loss at high electric fields. As a result, these polymers exhibit an improved

electrical energy density than BOPP. [43] The high glass transition temperature also leads

to a higher operating temperature. In strongly dipolar polymer materials, such as the

polyurea and polythiourea, the orientation polarization is the dominant polarization

mechanism compared with the electronic, atomic or ionic polarization.

Table 2-1 Dipole moments of common dipolar units present in dielectric polymers. [15]

Dipole units Dipole moment (Debye)

Urea 4.56

Diphenyl urea 4.6

Thiourea 4.89

Diphenyl thiourea 4.9

PVDF 2.1

21

Urea Thiourea

Figure 2-1 Dipolar structure of urea and thiourea units. [47]

Compared with other models, the Frohlich model takes both the short-range

interaction between molecules, and the deformation polarizations into consideration. This

model has been used widely to describe the dielectric response in polymers. [48] In this

model, the dielectric constant, or the relative permittivity is proportional to the dipole

moment, volumetric dipole density, and correlation factor between the dipoles. It is given

as:

kT

Ngp

rrs

rrsrrs

0

2

2 9)2(

)2)((

(2.1)

where ϵrs, ϵr∞ are the dielectric constants at low frequency and optical frequency

respectively; N is the volumetric dipole density; g is the correlation factor; p is the dipole

moment; k is the Boltzmann constant; T is the temperature.

22

Four dielectric polymers based on polyurea and polythiourea have been developed, which

include aromatic polyurea (ArPU), aromatic polythiourea (ArPTU), meta-phenylene

polyurea (m-PhPU/meta-PU) and methylene polythiourea (MePTU), to study the influence

of dipole moment and dipole density on the dielectric properties. Figure 2-2 shows the

chemical structure of these polymers and Table 1.3 summarizes their dielectric properties.

Figure 2-2 Schematics of (a) aromatic polyurea (ArPU), (b) aromatic polythiourea

(ArPTU), (c) meta-phenylene polyurea (m-PhPU/meta-PU), and (d) methylene

polythiourea (MePTU). [50]

The enhanced dielectric constant observed in meta-PU is hard to explain by using

the Frohlich model as its difficult to experimentally predict the value of correlation factor

(g) shown in equation 2.1. The localized free volume can be used to explain the increased

dielectric constant of meta-PU and other dipolar polymers. In this chapter, free volume

theory is reviewed, followed by discussion of local free volume introduced by free volume.

23

Table 2-2 Summary of dielectric properties of ArPU, ArPTU, m-PhPU and MePTU. [50]

Polymer Dielectric constant

(1 kHz)

Loss

tangent

(1 kHz)

Breakdown

strength

Eb (MV/m)

Energy density at Eb

(J/cm3)

ArPU 4.1 0.87% 800 13.5

ArPTU 4.4 0.64% > 1000 20.1

meta-PU 5.7 1.71% 670 13

MePTU 5.7 1.55% 500 7.5

2.2 Review of free volume theory

Free volume is a semi-quantitative concept which has been employed in statistical

thermodynamic theories of the liquid state. (Lennard-Jones and Devonshire, 1939 [51],

[52]; Glasstone, Laidler and Eyring, 1941 [53]; Frenkel, 1946 [54]; Fowler and

Guggenheim, 1956 [55]. The earliest definition, according to Glasstone, Laidler and Eyring

(1941), of free volume is that it may be regarded as the volume in which each molecule of

a liquid moves in an average potential field due to its neighbors. However, theoretical

estimates of free volume depend on postulates regarding the compressibility of the

molecules and the nature of their packing in the liquid state. [55]

In terms of solids, the molecular motion in the bulk state depends on the presence

of vacancies, or voids. A similar model can be constructed for the motion of polymer

24

chains, the main difference being presence of multiple voids may be required to be in the

same locality, as cooperative motions are required. Therefore, for a polymeric segment to

move from its present position to an adjacent site, a critical void volume known as free

volume must first exist before the chain segments can move. [56] In molecular substances,

the transition from liquid to glass results in marked changes in viscosity, specific heat, and

thermal expansion coefficient within a narrow temperature interval centering about a glass

transition temperature (Tg). Thus, above Tg the marked changes in specific volume reduces

the constraints on the movement of polymer chain segments.

The definition of free volume often used in polymer studies is given by Doolittle

[57]–[59], which is given below:

if = v-vo (2.2)

where vf is the free volume per gram, v is the measured specific volume of the

polymer at temperature T and vo is termed as the occupied volume. In Doolittle’s studies,

vo is taken as the value of v extrapolated to 0°K and is therefore regarded as a constant

independent of temperature. This definition assumes that vf must tend to zero as the

temperature tends to absolute zero, and that the increase of v with temperature, due to

thermal expansion, is associated entirely with an increase in vf.

The basic idea underlying the free volume approach to relaxation phenomena is that

the molecular mobility at any temperature is dependent on the available free volume at that

temperature. As temperature increases, the free volume increases and molecular motions

become more rapid. A few molecular theories based on this free volume concept have been

proposed, with the ultimate aim of relating dynamic quantities such as the diffusion

coefficient, viscosity or relaxation time to free volume. [55], [56] These theories are

25

applicable to the liquid-like state and can therefore be applied to amorphous polymers at

temperature of the order of and above Tg. Figure 2-3 summarizes the conventional free

volume definition.

Figure 2-3 Schematic diagram illustrating free volume as calculated by Simha and

Boyer.[56]

It is important to note that the concept of “free volume” discussed in this work is focused

mainly on localized free volume pertaining to local motions in the polymer chain. Thus, it

is dependent more on the local nanostructures created than the glass transition of the

material. Fundamental dielectric theory suggests that strongly dipolar polymers have the

potential to realize a high dielectric constant. [55] In order to achieve high thermal stability,

these polymers should also possess a high glass transition temperature Tg. It has been

26

observed that in many dipolar polymers, the dielectric constant decreases markedly at

temperatures below Tg due to constraints of the glassy structure on the dipoles. In contrast,

at temperatures above Tg, the reduced constraints on the dipoles due to increased free

volume, lead to a large increase in dielectric constant. For example, polyvinyl chloride

(PVC), a simple polymer glass, exhibits a large increase in dielectric constant after

undergoing its glass transition, from K~3 below Tg to K>9 above Tg. [55], [60] The penalty

is that the dielectric loss also becomes high at temperatures above Tg (loss > 5%) due to

cooperative segmental motions in the rubbery state, which have long relaxation times. The

challenge is to introduce this excess free volume in strongly dipolar polymers at

temperatures far below Tg, thereby a relatively high dielectric constant may be achieved

without the penalty of high dielectric loss.

In this chapter, it has been shown that disorder in strongly dipolar polymers creates

a significantly larger free volume at temperatures far below Tg, enabling easier

reorientation of dipoles in response to an electric field in aromatic urea and thiourea

polymers. The net result is a substantial enhancement in the dielectric constant while

preserving low dielectric loss and high breakdown strength.

27

2.3 Experimental section

2.3.1 Synthesis and film fabrication of ordered and disordered structures

All chemicals used in this study were purchased from Sigma-Aldrich. Traditionally,

the aromatic polyurea were synthesized via polycondensation of aromatic diamine and

aromatic diisocyanate, which was synthesized from phosgene. In this study, a green

synthetic route which is isocyanate free, solvent free, and catalyst free is used for synthesis

of meta-aromatic polyurea (meta-PU). [49] Here, the urea units connect to the meta

position of aromatic rings, and thus it’s been named as meta-PU. The polymer was

synthesized by polycondensation of meta-phenylenediamine and diphenyl carbonate as

shown in Figure 2-4. After purification, meta-PU was isolated as pinkish powder. To

prepare the films, meta-PU powders were dissolved in dimethylformamide (DMF) to make

1.0 - 2.0 weight % solution. The thin films were prepared by casting the solution onto 1 cm

x 1 cm silicon substrates pre-coated with 40 nm of platinum. After casting, the films were

dried in a vacuum oven for 4 hours at room temperature, cured overnight at 110 °C, and

annealed at 140 °C for 12 hours under vacuum.

The meta-PU powders were crystalline in nature (ordered structures) and prepared

films turned out to be amorphous in nature (disordered structures), as confirmed from X-

ray measurements.

28

Figure 2-4 Schematic of synthesis of meta-PU.

Poly(arylene ether urea) (PEEU) was prepared from (m-phenylenedioxy) dianiline

and diphenyle carbonate by thermal poly condensation as shown in Figure 2-5. [46] The

mixture of the two monomers was stirred at 150 ℃ in vacuum for 4 hours, and PEEU

powder was obtained through purification with ethanol for 5-6 times. PEEU films were

prepared by dissolving the powders in DMF at elevated temperature, and solution cast at

80 ℃ for 2 days, followed by annealing at 150 ℃ for 2 days. After casting, the films were

dried in a vacuum oven for 4 hours at room temperature, cured overnight at 70 °C, and

annealed at 110 °C for 12 hours and then at 140 °C for 24 hours under vacuum.

Similar to meta-PU, the PEEU powders were crystalline in nature (ordered

structures) and prepared films turned out to be amorphous in nature (disordered structures),

as confirmed from X-ray measurements.

Figure 2-5 Schematic of synthesis of PEEU.

29

2.3.2 Measurement of dielectric properties of polymer powders

It is not easy to directly measure the dielectric constant of polymer powders. Here

a composite approach, i.e., mixing the powder with a fluid that does not affect the powder,

was employed for the dielectric characterization. Gold-sputtered glass slides were used as

the electrodes to form a parallel-plate capacitor. The electrodes were separated by Kapton

tape spacers with thickness of 64 µm. The area of the electrode was 1 cm x 1 cm (and

repeated with 0.5 cm x 0.5 cm electrode area). In order to measure the dielectric constant

of the meta-PU powder, we mixed the powder with dielectric fluids of different dielectric

constants and used the Lorenz-Lorentz equation (Equation (2.3)) [61] to deduce the

dielectric constant of the meta-PU powder from the dielectric constant of the mixture Ke.

It is noted that these dielectric fluids do not dissolve the meta-PU powder (as verified by

the X-ray diffraction data of the mixture). In the case of a two-phase, three-dimensional

medium, the Lorenz-Lorentz theory yields the following expression for eK in terms of the

dielectric constant1K (meta-PU powder),

2K (the fluid) and the powder volume fraction p1

of the mixture.

)2(

)(

2

)(

21

211

2

2

KK

KKp

KK

KK

e

e

(2.3)

High precision data can be obtained if the dielectric constant of the fluid is close to

that of the powder. The rationale behind this approach was that if the capacitance of a cell

filled with the reference fluid is lowered by adding powder, the dielectric constant of the

powder is lower than that of the fluid.

30

2.3.3 Details of characterization equipment

The dielectric data was obtained by using a HP 4294A Precision Impedance

Analyzer. The dielectric properties at variable temperature were measured using an HP

4284 impedance analyzer, which was connected to an environmental test chamber (Delta

9023). The Polarization-Electric field (P-E) response was measured with a modified

Sawyer-Tower circuit. The charged (stored) energy density, discharged (released) energy

density and efficiency were calculated from the P-E loop. The X-ray diffraction data were

collected using a Panalytical Xpert Pro MPD diffractometer.

2.4 Results and discussion

One of the ways to generate free volume in dipolar polymers is by introducing

disorder. In this work, through a combined theoretical and experimental investigation, it

has been shown that a disordered polymer with high Tg can be realized in several recently

developed strongly-dipolar polymers based on aromatic urea and thiourea units, with

dipole moments of 4.5 Debye and 4.89 Debye respectively,[50] which are much higher

than for VDF in PVDF based polymers, with the dipole moment of 2 Debye.[62] The

theoretical results are discussed briefly here and details of simulation work can be found at

Ref. [45], [63]. meta-PU was chosen because of its high dipole moment (4.5 Debye). The

theoretical results show that the specific volume is ~12% larger in the disordered structures

compared to the ordered structure of meta-PU. The larger volume gives urea units more

free space to reorient, thus large dipolar motion can lead to a larger permittivity. This was

confirmed by the experimental study of ordered and disordered structures. To compare

31

with the theoretical results, meta-PU with ordered (crystalline) and disordered (amorphous)

phases were prepared. It is interesting to note that the as synthesized meta-PU powder

shows relatively sharp X-ray diffraction peaks, visible in Figure 2-6 (a), indicating the

presence of a crystalline phase. In contrast, the films made from solution casting display a

broad X-ray peak centered at 2θ = 9.5°, shown in Figure 2-6 (b), indicating an amorphous

(disordered) structure.

Figure 2-7 shows DSC and TGA data. The DSC curves shows there is no glass

transition step up to 200°C. Also, TGA shows no sign of weight loss and transition below

200°C, which is a desirable feature for high temperature operation.

Figure 2-6 X-ray data for (a) ordered meta-PU structures; (b) films of disordered meta-

PU structure.

32

Figure 2-7 (a) DSC data and (b) TGA data of meta-PU.

The disordered structures (films) are discussed first. The dielectric properties of

meta-PU films are presented in Figure 2-8 (a) for the room temperature dielectric properties

as function of frequency and Figure 2-8 (b) for dielectric properties vs. temperature

33

measured at 1 kHz. The results show a high dielectric constant (K > 5.6) and low loss (loss

tangent ~ 1.5%) over a broad temperature range, due to the high glass transition

temperature. Even more importantly, the meta-PU films exhibit a linear dielectric response

and very low loss even at very high electric field, see Figure 2-8 (c) for the charge/discharge

curve at electric field close to 700 MV/m, measured at room temperature at 10 Hz. It is

the free volume effect (FVE) at temperatures below Tg (> 200°C) that leads to a high

dielectric constant (K > 5.6) in meta-phenylene polyurea (meta-PU).

The dielectric constant of meta-PU powder was measured by making slurries with

dielectric fluids of dielectric constant similar to the powder. Fluid (castor oil) with

dielectric constant of 4.7 was chosen to make slurry with meta-PU powder. The dielectric

constant of the mixture turns out as 4.6 at 1 kHz from the dielectric measurement. As shown

in Figure 2-9, the dielectric constant of the mixture of meta-PU powder (p1 = 25 vol %)

and castor oil (K2 = 4.75) is 4.5 (at 1 kHz), lower than that of the dielectric fluid, indicating

that the meta-PU powder has a dielectric constant smaller than 4.5. Using equation 2.3, the

calculated dielectric constant of the meta-PU powder is 3.8. The dielectric constant and

loss of disordered and ordered structure is summarized in the Figure 2-10.

34

Figure 2-8 (a) Dielectric constant vs. frequency, measured at room temperature; (b)

dielectric constant vs temperature, measured at 1 kHz; (c) P-E loop of meta-PU film.

35

Figure 2-9 Dielectric constant vs frequency for the mixture of meta-PU/castor oil (open

squares) and castor oil (open circles).

To further confirm that the results obtained here are not an isolated case, poly (arylene

ether urea) (PEEU) polymer (Figure 2-5) was also studied. Similar to meta-PU, the PEEU

powder shows sharp peaks, thus suggesting a crystalline phase (see Figure 2-11(a)). In

contrast, the solution-cast films display a broad X-ray peak (Figure 2-11(b)) indicating an

amorphous structure. The dielectric constant of PEEU films is 4.7, while the dielectric

constant of PEEU powder deduced from equation 2.3 using PEEU powder/fluid mixture is

3.65, which is smaller than that of the disordered structure (see Figure 2-12). The

experimental dielectric data of ordered and disordered structure is summarized in Table 2-

3.

36

Figure 2-10 (a) Dielectric constant and (b) dielectric loss as a function of frequency for

ordered and disordered structure of meta-PU.

(a)

(b)

37

Figure 2-11. X-ray data for (a) ordered PEEU structures, powder; (b) films of disordered PEEU

structure.

Table 2-3 Summary of experimental dielectric data of ordered and disordered structure of

meta-PU and PEEU.

Polymer Ordered structure (Powder) Disordered structure (Film)

meta-PU 3.8 5.7

PEEU 3.7 4.7

(a)

(b)

38

Figure 2-12 Dielectric constant as a function of frequency for (a) disordered and (b)

ordered structure of PEEU.

(a)

(b)

39

2.5 Conclusion

In conclusion, through combined theoretical and experimental studies, we show

that in meta-PU and PEEU, disordered phases exhibit higher dielectric constants compared

with those of ordered phases, due to the built-in free volumes in the disordered structures.

Localized free volumes in disordered phases of these strongly dipolar polymers at

temperatures far below the glass transition enable easier reorientation of dipoles in

response to an electric field, leading to high dielectric constants while preserving low

dielectric loss. At the same time, disorder enables longer wavelength vibrations, which also

increases permittivity. These concepts and the experimental results demonstrate a new and

promising approach for developing dielectric polymers with high dielectric constant, low

loss, and high operating temperature.

Acknowledgement

The part of this chapter has been published in Nanoenergy and IEEE Dielectrics and

Electrical Insulation Conference Proceedings. Reproduced here by the permission of

Elsevier.

[1] Y. Thakur et al., “Optimizing nanostructure to achieve high dielectric response with

low loss in strongly dipolar polymers,” Nano Energy, vol. 16, pp. 227–234, 2015

[2] Y. Thakur, M. Lin, S. Wu, and Q. Zhang, “Introducing free volume in strongly dipolar

polymers to achieve high dielectric constant,” in Electrical Insulation and Dielectric

Phenomena (CEIDP), 2015 IEEE Conference on, 2015, pp. 636–639

40

Chapter 3

Blending of dipolar polymers to enhance the free volume effect

3.1 Introduction

The understanding of localized free volume has been extended in this chapter

by blending two strongly dipolar polymers. Blending two polymers may create partial

mismatches between two dissimilar polymer chains, resulting in additional free volume.

Thus, reducing the constraints for dipole reorientation under applied field in the glassy

state, and raising the dielectric constant without causing high dielectric losses. Based on

these considerations, we investigate a class of nanostructured dipolar disordered polymers,

i.e., polymer blends. The introduction of blending in dipolar polymers increases the

dielectric constant significantly compared to the best state-of-the-art dielectric in current

use (biaxially oriented polypropylene, BOPP, K~2.2), while also tolerating much greater

operating temperature [1].

3.2 Experimental section

3.2.1 Synthesis and film fabrication of blends

All chemicals for synthesizing PEEU and ArPTU were purchased from Sigma-Aldrich.

ArPTU was synthesized via microwave-assisted polycondensation of diphenylmethane-diamino

(MDA) with thiourea in N-methyl-2-pyrrolidone (NMP) with p-toluenesulfonic acid (p-TSOH) as

41

a catalyst as shown in Figure 3-1. After purification, ArPTU was isolated as yellow powder, which

was used for film processing.

Figure 3-1 Schematic of synthesis of aromatic polythiourea.

The PEEU synthesis has been discussed in the previous chapter. The blend solution

was prepared by dissolving 1 wt.% of ArPTU and PEEU in DMF. The thin films were

prepared by casting the solution onto 1 cm x 1 cm silicon substrates pre-coated with 40 nm

of platinum. The films were kept in a drying oven under vacuum at room temperature for

4 hours, followed by heating to 70 °C overnight, 110 °C for 12 hours and then annealing

at 180 °C for 1 day.

Figure 3-2 Blend of two polymers: ArPTU and PEEU.

Blend

42

3.2.2 Details of characterization equipment

The dielectric data was obtained by using a HP 4294A Precision Impedance

Analyzer and a Novocontrol GmbH Concept 40 broadband dielectric spectrometer. The

grazing incidence X-ray scattering data were collected using a Panalytical X’Pert PRO

MPD diffractometer. The wavelength of X-ray was 1.54 angstroms. Background scattering

was subtracted using JADE analysis software and then the peak position was calculated for

the individual polymers and blends. Atomic force microscopy (AFM) was performed using

a Bruker Dimension AFM in tapping mode. Thermal gravimetric analysis (TGA) was

carried out in N2 at a heating rate of 10°C/min using a 2050 TGA from TA Instruments.

Differential scanning calorimetry (DSC) was carried out using a TA instruments Q2000 to

probe the thermal behavior. The data were taken at a scan rate of 10°C/min during heating.

The negative heat flow represents endothermic (heat absorbed) direction.

3.3 Discussion of dielectric data

An aromatic polythiourea (ArPTU) and a poly(ether urea) (PEEU), see Figure 3-2

for the chemical structures, were chosen as the blend components. ArPTU and PEEU have

dipole moments of 4.5 Debye and 4.89 Debye, respectively, leading to relatively high

dielectric constants, K = 4.4 and K = 4.7 for the two polymers in the glassy state. The Tg

of both polymers is above 200°C. As presented in Figure 3-3 (a), the 1:1 blend (by weight)

of the two polymers exhibits remarkably high and reproducible dielectric constant (K =

7.5) while maintaining low loss (<1%). The inset in Figure 3-3 (a) shows the dielectric

43

constants of the individual polymers. This is the first report on a polymer with such a high

dielectric constant of 7.5 and sufficiently low loss (below 1%) with minimum frequency

dispersion to be used for capacitor applications. Figure 3-3 (b) displays the dielectric

properties vs. temperature (at 10 kHz), which shows that the blend exhibits a high dielectric

constant and low loss up to 120°C.

Blends with different PEEU/ArPTU ratios were also prepared and their

dielectric properties at room temperature and 1 kHz are summarized in Figure 3-3 (c) and

Table 3.1. A very large increase in the dielectric constant was also observed in blends with

different ratios of the two polymers. The dielectric loss of these blends (above and below

a 1:1 ratio) shows a slight increase (> 1%) compared to those of the constituent polymers

and the 1:1 PEEU:ArPTU blend. The increase in dielectric constant may be attributed to

the increased free volume. However, that increase in free volume can also result in larger

scale polymer chain motions, which increase the dielectric loss (>1%).

In order to examine the possible effect of any sub-Tg transitions (β and ɣ

relaxations) on the dielectric constant of the neat polymers and 1:1 blend [64], a broad band

dielectric spectroscopy study was carried out. As presented in Figure 3-3 (b) and Figure 3-

4, there are no sub-Tg transitions down to -150°C over a broad range of frequencies.

44

Figure 3-3 Dielectric data of the 1:1 blend of PEEU and ArPTU (a) as a function of

frequency at room temperature, including the inset which shows the dielectric data of

PEEU and ArPTU; (b) as a function of temperature at different frequencies. (c) Dielectric

constant vs. blend composition (weight ratio of PEEU:ArPTU) at room temperature and 1

kHz. Data points are shown and the dashed line is drawn to guide the eye.

45

Table 3-1 Summary of the dielectric properties of the neat polymers and blends at 25 ˚C

Polymer Dielectric Constant (1 kHz) Loss (1 kHz)

PEEU 4.7 1.1%

ArPTU 4.4 0.64%

Blend (1:1) 7.5 0.77%

Blend (1:2) 7.9 1.62%

Blend (1:3) 8.6 1.84%

Blend (2:1) 8.3 1.45%

Blend (3:1) 7.4 1.35%

To provide further insight into molecular and nanoscale mechanisms responsible

for the observed enhancement of the dielectric response in the blends, simulations of

ArPTU, PEEU, and blends of ArPTU:PEEU in various blend compositions were carried

by Prof. Jerry Bernholc’s group at North Carolina State University.[65] In the calculations

of the dielectric properties, they combine the results from both molecular dynamics (MD)

and density functional theory (DFT) simulations to obtain the permittivity. The blend

simulations reveal a significantly larger specific volume (see Figure 3-5 (a)) due to an

increase in interchain spacing, again in accordance with the experimental results.

Comparison of simulation and experimental data of dielectric constant vs. blend

composition (PEEU:ArPTU ratio by wt.) at room temperature is presented in Figure 3-5

(b). The calculated enhancements are smaller, probably due to the relatively small size of

the simulation cell, which does not fully capture the effects of nanoscale morphology,

inaccuracies in interatomic potentials, and relatively short simulation time.

46

Figure 3-4 Dielectric data of (a) ArPTU, (b) PEEU as functions of temperature measured

over a range of frequencies.

47

Figure 3-5 (a) Computational results of dielectric constant vs. specific volume for PEEU,

ArPTU, and blends for various supercells with different PEEU:ArPTU weight ratios. (b)

Comparison of simulation and experimental data of dielectric constant vs. blend

composition (PEEU:ArPTU weight ratio) at room temperature. Data points are shown

and the dashed line is drawn to guide the eye. [65]

(a)

(b)

48

3.4 Structural analysis

Grazing incidence X-ray scattering of the ArPTU, PEEU, and blend with 1:1

PEEU:ArPTU ratio was carried out to probe structural changes and the data are presented

in Figure 3-6. As shown in many earlier studies, the broad X-ray peak around 2θ = 18o in

Figure 3-6 for the neat polymers arises from interchain segment scattering in the

amorphous state. The scattering data reveal: (i) there is only one broad X-ray diffraction

peak for the 1:1 blend; (ii) the broad X-ray peak for the blend is at ca. 2θ = 17o, indicating

that interchain spacing in the blend is more than 5 % larger than those of the individual

polymers. The expanded interchain spacing in the blend enables easier dipole reorientation

to the applied field and leads to a higher dielectric constant compared with those of the neat

polymers while maintaining low dielectric loss. These results indicate that the reduced

constraints achieved by molecular engineering of the dipolar polymers in the glassy phase

can significantly increase the dielectric constant without compromising the dielectric loss

[17].

The AFM image of the blend with 1:1 PEEU:ArPTU ratio is presented in Figure 3-

7, showing uniform mixing of the two polymers in the blend at the nanoscale. The DSC

data of the blend (Figure 3-8) does not show two glass transition steps till 250°C,

suggesting single phase behavior, which is consistent with the AFM data.

49

Figure 3-6 (a) X-ray diffraction data of ArPTU and PEEU, and their 1:1 blend.

Background subtracted data of (b) PEEU with peak at 18.6°, (c) ArPTU with peak at 18.6°

and (d) blend data with peak at 17°. Wavelength of X-ray used was 1.54 angstroms.

50

Figure 3-7 AFM images: (a) amplitude and (b) phase for the PEEU:ArPTU 1:1 blend.

Figure 3-8 DSC data of 1:1 PEEU:ArPTU blend.

The TGA data in Figure 3-9 shows no weight loss below 250°C, thus confirming the

thermal stability up to 250°C, which is a very desirable feature for high temperature

operation.

51

Figure 3-9 TGA data of 1:1 PEEU:ArPTU blend.

Both structure analysis and computer simulation results indicate that the nano

(molecular) scale mixing of the two polymers causes a slight increase of the interchain

spacing in the glassy blend, thus reducing the barriers for dipole reorientation along the

applied electric field, and generating a high dielectric response without compromising the

dielectric loss.

In addition, positron annihilation lifetime spectroscopy (PALS) study is being

carried out by Prof. David Gidley’s group at University of Michigan on the blend samples

for direct measurement of free volume. The details of the PALS spectrometer can be found

from Ref. [66]. The preliminary results look promising and show the presence of increased

void size in the blend structure. The Figure 3-10 below shows the positron lifetime and

specific hole volume determined from the fitted positron lifetime as a function of blend

ratio. If there is a simple mixing of the two components (separate microscopic regions of

52

pure ArPTU and PEEU) then we would expect the fitted single Ps lifetime to simply be a

weighted average of the two individual lifetimes and vary linearly along this grey line.

However, as confirmed by the DSC results in Figure 3.8 that the blends show single phase

behavior, so as expected, the lifetime does not vary linearly. Also, the specific hole volume

increases with change in blend ratio except for PEEU:ArPTU blend ratio of 2:1 and 3:1.

Thus, suggesting that there is agreement with our hypothesis of localized free volume,

where these voids occupying few angstroms of volume can reduce the constraint on dipoles

in the glassy state of the polymer.

53

0 20 40 60 80 100

1.70

1.75

1.80

1.85

1.90

1.95

2.00

2.05

Fixed 0.8 ns

Ps L

ife

tim

e (

ns)

% PEEU (%)

0 20 40 60 80 100

70

75

80

85

90

95

100

Sp

he

rica

l S

pe

cific

Ho

le V

olu

me

(

A3)

% PEEU (%)

Figure 3-10 PALS results of (a) positron lifetime, (b) spherical specific hole volume with

change in PEEU composition.

ArPTU

PEEU

ArPTU

PEEU

(a)

(b)

54

3.5 Conclusion

In conclusion, we demonstrate a low-cost approach to achieve dramatically

higher dielectric constants while preserving the low loss and high operating temperature in

strongly dipolar polymer systems. Specifically, we show that nanostructure engineering

through blending of two dissimilar strongly dipolar polymers creates sub-nanoscale

unoccupied volume, leading to polymers with a dielectric constant of 7.5 and a loss less

than 1%. The results demonstrated here, which can be applied to many existing dipolar

polymers, pave the way for a very low-cost approach for creating “new” dielectric

materials from existing ones, but with dramatically improved dielectric response. These

advantages can enable many more uses of high power density capacitors in portable and

automotive systems, aircraft control, and advanced weaponry.

Acknowledgement

The part of this chapter has been published in Nanoenergy. Reproduced here by the

permission of Elsevier.

Y. Thakur, B. Zhang, R. Dong, W. Lu, C. Iacob, J. Runt, J. Bernholc and Q.M. Zhang, “Generating

high dielectric constant blends from lower dielectric constant dipolar polymers using nanostructure

engineering” Nano Energy, vol. 32, pp. 73–79, 2017

55

Chapter 4

Enhanced dielectric response in dipolar polymers with inorganic nanodopants

4.1 Introduction

In the previous chapter, a free volume approach to improve the dielectric constant was

discussed. That approach entails designing “new” dielectric materials which utilize free

volume to enhance dielectric constant. Nevertheless, it is also important to explore ways to

improve the properties of state-of-the-art materials; so that we can employ the existential

large-scale manufacturing setups of these polymers. In contrast to the traditional polymer

nanocomposites which rely on the high dielectric constant fillers to raise the dielectric

constant of composites, an emerging theme for developing new dielectric polymers is to

explore nanostructure engineering, such as utilizing the large interfacial fraction in

nanocomposites and the associated effects on dielectric response. In this chapter, a

nanocomposite approach has been discussed where addition of small amounts of inorganic

nanoparticles (< 1 vol.%) can considerably improve the dielectric properties of dipolar

polymers.

4.2 Composite theory

The introduction of inorganic particles in the host polymer matrix is one of the

promising ways to improve the dielectric properties of present dielectric materials. The

premise of this approach resides in the combination of inorganic particles with high

permittivity and high dielectric strength of polymers to achieve high energy density in

56

polymers. In this pursuit, both micro-sized particle and nano-sized particle filled systems

have been extensively investigated. [67] The composites with microparticles showed

impairment in the dielectric properties due to the surface defects and stress cracking after

ageing. In addition, the thickness of the film, an important aspect of miniaturization of

modern capacitors, was limited by the size of the microparticles. Thus, the interest shifted

to nano-sized particle filled systems, which allows for relatively low filler loading and

nanometer sizes, without sacrificing some of the inherent polymeric properties, such as

density, flexibility, and ease of processing. As the particle shrinks in size from the

micrometer to nanometer scale, the percentage of atoms at the surface of the particle

becomes more significant, resulting to a dramatic change in the physical properties,

interfacial properties, and also the agglomeration behavior relative to bulk materials. In

addition, the inter-filler distances in the nanocomposites can be in the range of nanometers,

and the filler would interact chemically and physically with polymer matrices, resulting in

the emergence of intermediate or mesoscopic properties. These mesoscopic properties at

interfaces may bring unexpected but excellent macroscopic properties of nanocomposites.

[68] It has been demonstrated that, as the size of filler particles decreases to the nanometer

scale, large interfacial areas in the composite between the polymer and nanoparticle would

promote the exchange coupling effect through a dipolar interface layer, resulting in higher

polarization levels, dielectric response and breakdown strength. [67], [69]–[71] The

dielectric properties of inorganic particles are largely dependent on their size, for example

in BaTiO3, the dielectric permittivity decreases dramatically from 5000 to hundreds as the

particle size is reduced from 1μm to 30 nm. The decrease in dielectric permittivity is due

to the reduction of tetragonality with decrease in particle size and the transition to cubic

57

structure when the grain size drops down to 30 nm. Still the inorganic nanoparticles have

dielectric constant(K>100) way higher than the commercial dielectric polymers (K<3.5).

One of the promising ways to improve the dielectric constant is by using a composite

approach, in which nanoparticles with the dielectric constant much higher (e.g. 100X) than

that of the polymer are added to the polymer matrix. [11], [16], [42] The increase in the

dielectric constant of such composites is usually attributed to the high dielectric constant

of the inorganic fillers since the composite dielectric constant is a volume average of those

of the constituents. This avenue has been extensively explored with high dielectric constant

inorganic particles like TiO2 [72]–[75], BaTiO3 [11], [42], [70], [76]–[78], ZrO2[75], [79]–

[81], CaCu3Ti4O12[82]–[85], PZT [86]–[89]. However, this approach suffers from several

issues for energy storage applications. The primary challenge is that the inclusion of high

dielectric constant fillers results in highly inhomogeneous electric fields at the interface of

polymer and nanoparticle due to the large difference in permittivity of both phases. This

results in intensification of local electric fields in the polymer matrix near the filler

particles, leading to a large reduction of the electric breakdown strength. [38], [67], [90]

This has prevented the practical applications of dielectric organic/inorganic composites.

The second challenge is with the physical dispersion of oxide nanoparticles. The high

surface energy of nanoparticles usually results in agglomeration and phase separation from

the polymer matrix. This yields in poor quality films with weakened dielectric properties,

such as high dielectric loss and low field strength. Lastly, for the nanocomposite films, it

is important to rationally optimize the particle concentration to maximize the energy

storage density. Following the composite theory, it requires a high-volume content of high

dielectric constant nanoparticles to increment the dielectric constant of the polymer

58

composite films. On the other hand, the dielectric strength drops off dramatically with high

volume content of these fillers. Despite the increment in dielectric constant, the reduced

breakdown strength can only give slight improvements in energy density. Therefore, it is

required to optimize the volume content of nanoparticles to achieve maximum

improvement in energy density.

4.2.1 Models for predicting effective permittivity

The effective permittivity of a polymer nanocomposite depends on the individual

permittivity of the filler and polymer matrix along with different filler loadings and

interactions among them. To find the effective dielectric constant of the nanocomposites,

various models have been derived.

The following models have been commonly used for two-phase dielectric composites:

The Maxwell model is relatively easy for modeling due to its linearity for the dielectric

constant of composite Km for spherical inclusions in a continuous matrix. [91]

K𝑚 = K1𝐾2+2𝐾1−2𝑉2(𝐾1−𝐾2)

𝐾2+2𝐾1+𝑉2(𝐾1−𝐾2) (4.1)

where Km, K1, and K2 represent the dielectric constants of the composite, phase 1, and

phase 2, respectively, V1 and V2 are the volume fractions of phases 1 and 2 (V1 + V2 = 1).

Series and parallel mixing models represent the extreme cases. The series model

corresponds to alternating layers of each phase in the direction perpendicular to the applied

field (two capacitors in series). The parallel model corresponds to alternating layers of each

phase in the direction parallel to the applied field (two capacitors in parallel). [92]

(a) Series model:

1

𝐾𝑚=

𝑉1

𝐾1+

𝑉2

𝐾2 (4.2)

59

(b) Parallel model

𝐾𝑚 = 𝑉1𝐾1 + 𝑉2𝐾2 (4.3)

The Lichtenecker model represents a widely used empirical relationship without any

concern for the physical geometry of the composite system. It is a logarithmic mixture

formula and is most efficient in calculating the effective permittivity of the polymer

nanocomposites. [93]

𝑙𝑛𝐾𝑚 = 𝑉1𝑙𝑛𝐾1 + 𝑉2𝑙𝑛𝐾2 (4.4)

4.3 Experimental section

4.3.1 Nanocomposites of polyetherimide

Nanocomposites of polyetherimide (PEI) with various nanofillers were prepared by

a solution casting method. Ultem 1000 PEI polymer resin was purchased from General

Electric (GE), molecular structure shown in Figure 4-1. Alumina (Al2O3) nanoparticles

with mean particle diameters of 5 nm, 20 nm and 50 nm, magnesium oxide (MgO) and

silicon dioxide (SiO2) nanoparticles of 20 nm size, BaTiO3 (BTO) with 50 nm size, and

boron nitride (BN) with 70 nm size were purchased from US-Nano. To prepare

nanocomposite films, the nanoparticles with selected weight percent were added to PEI

powder and dissolved in dimethylformamide (DMF). Following this step, the mixed

solution was sonicated at room temperature using VWR Aquasonic 76T for 12 h to achieve

good nanoparticle dispersion. The solution was then cast onto a clean glass slide. The

60

solution cast films were kept in a drying oven at 70 °C for 12h to remove the solvent and

then heated to 100 °C and 150 °C for 1h, respectively, and 200 °C for 12h, followed by a

final drying step at 225 °C for 2h. Afterward, the films were kept in a vacuum oven at 200

°C for one day to further remove any residual solvent. Finally, films were peeled off from

the glass substrate and dried in a vacuum oven at 80°C for 4h.

Figure 4-1 Molecular structure of the polyetherimide. [94]

4.3.2 Nanocomposites of polystyrene

Nanocomposites of polystyrene (PS) with various nanofillers were prepared by a

solution casting method. PS resin was purchased from Sigma–Aldrich, see molecular

structure in Figure 4-2. The 2 wt.% solution of polystyrene in DMAc was stirred at 70℃

for 12 h. Alumina nanoparticles (20 nm and 50 nm) with selected weight percent were

dispersed in DMAc using a VWR Aquasonic 76T for 1h and then added to the PS solution.

Following this step, the mixed solution was sonicated using the same instrument for 12h to

get good dispersion and then cast on a clean glass slide. The solution was kept in a drying

oven at 90℃ for 24h and then vacuum dried at 100℃ for another day, followed by the final

61

drying step at 200℃ for 2h. Finally, the film was peeled from the substrate and dried in a

vacuum oven at 80°C for 4h.

Figure 4-2 Molecular structure of the polystyrene.

4.3.3 Details of characterization equipment

Nanocomposite film thickness was in the range of 8-12 microns measured by

Heidenhain ND287 digital micrometer. Sputtered gold electrodes of 2 mm and 6 mm

diameter were deposited on both surfaces of the composite films for electrical

characterization. Dielectric data were characterized using an HP 4294A Precision

Impedance Analyzer and a Novocontrol GmbH Concept 40 broadband dielectric

spectrometer. The charge-discharge response was measured with a modified Sawyer-

Tower circuit. The DC breakdown test was performed by applying a ramp rate of 100 V/s.

Differential scanning calorimetry (DSC) was carried out using a TA instruments Q2000 to

probe the thermal behavior. The data were taken at a scan rate of 10°C/min during heating.

The X-ray diffraction data were collected using a Panalytical X’Pert PRO MPD

diffractometer. An FEI Talos F200X (Scanning) Transmission Electron Microscope

(S/TEM) was used at 200 kV to probe the nanoparticles in the polymer matrix.

62

4.4 Results and discussion

4.4.1 Effect of alumina nanoparticles and the particle size

We chose polyetherimide (PEI), a high temperature polymer (glass transition

temperature Tg ~ 217oC), as the matrix. PEI contains dipolar units in the polymer chains

[94] (see Figure 4-1), exhibits a dielectric constant of 3.2, which for polymer dielectrics is

at the higher end of the dielectric constant spectrum, and low dielectric loss (0.29 %). Due

to its high operating temperature, modest K, and low loss, PEI has been considered for the

next generation of high performance capacitor dielectric material. [20], [35] Alumina

nanoparticles, whose dielectric constant (K = 9.5) is not far from that of the polymer matrix,

were used as the nanofillers, thus avoiding large local field concentration in polymer

interfacial regions near the nanofiller particles. As presented in Figures 4-3 (a) and 4-3 (b),

films with a very low volume content of alumina nanoparticles (0.32 vol.%) can achieve a

dielectric constant > 5, a more than 55% increase over that of PEI, while maintaining low

loss. The alumina nanofiller particle size in the composite is ca. 20 nm. Summarized in

Figure 4-3 (b) is the dielectric constant of the PEI/alumina (20 nm size) nanocomposites

vs. alumina volume content, which displays a sharp increase in the dielectric constant with

alumina nanofiller loading and reaches K > 5 at 0.32 vol.%. Moreover, this large increase

in the dielectric constant occurs in a small and narrow composition range. With additional

nanofiller, the dielectric constant decreases, and at 0.64 vol.% the dielectric constant K of

the film is 4. There have been several studies of PEI nanocomposites in the past. [41],[80],

[94] These earlier reports focused on high nanofiller content (> 2 vol. %). Our results for

PEI nanocomposites at > 0.62 vol.% nanofillers are consistent with the earlier work. It is

63

startling that the large enhancement in the dielectric response of PEI/alumina (20 nm size)

nanocomposites occurs at such a low filler volume content. Figure 4-3 (b) presents a

comparison of the experimental data with several widely-applied dielectric composite

models using the dielectric properties of PEI and alumina. None of these models can

describe the observed phenomenon, since they are based on the geometrical volume

average of the dielectric responses of the constituents.

The temperature dependence of the dielectric properties of the 0.32 vol.%

alumina (20 nm) nanocomposite film was also characterized (Figures 4-3 (c)). The data

show that the nanocomposite maintains low loss at temperatures in excess of 150°C, the

same as that of PEI (see Figure 4-4). When the temperature approaches the PEI Tg (~217

°C), the dielectric loss will increase in both the pure polymer and nanocomposites.

64

Figure 4-3 (a) Room temperature dielectric properties of PEI/alumina (20 nm particle size)

nanocomposites at different alumina nanoparticle loading (in vol. %) vs frequency. (b)

Dielectric constant of nanocomposite films of PEI/alumina (20 nm particle size) vs.

nanofiller volume content and comparison with several widely used dielectric models of

diphasic dielectric composites (lines with no data points): curve (1) Parallel model, (2)

Maxwell model, (3) Lichtenecker model, and (4) series model. Inset shows an expanded

view of the dielectric constants of the composite films vs. alumina loading. Experimental

data points are shown and lines are drawn to guide the eye. (c) Dielectric properties vs.

temperature of the PEI/alumina (20 nm size) nanocomposite with 0.32 vol.% alumina

loading at different frequencies.

65

Figure 4-4 Dielectric properties at different frequencies of neat PEI as a function of

temperature.

For the nanocomposites with low volume content of nanofillers, the breakdown

field is not compromised. Figure 4-5 (a) presents the charge-discharge curves of the 0.32

vol% alumina (20 nm) nanocomposite films at room temperature. These films exhibit a

high breakdown field of 525 MV/m, similar to that of pure PEI [35], achieving a discharged

energy density of 5.25 J/cm3 (see inset in Figure 4-5 (a). The increased discharged energy

density at different temperatures has also been measured. As shown in Figure 4-5 (b), the

room temperature discharged energy density is 2.9 J/cm3 under 350 MV/m, compared to

that of pure PEI (1.9 J/cm3) under the same field, due to the increased dielectric constant

of the nanocomposite film. Furthermore, Figure 4-5 (b) presents the energy density from

biaxially-oriented polypropylene (BOPP), which is widely used as the state-of-the-art

polymer capacitor. [44] With only 0.32 vol.% nanofiller, the energy density of the PEI

nanocomposite becomes more than 100 % higher than BOPP. At 100oC (Figure 4-5 (c))

the discharge energy density of the nanocomposite films is still more than 40% higher than

66

that of PEI. Due to its operating temperature limitation, BOPP was not measured at 100oC.

At 150oC (Figure 4-5 (d)), there is increased conduction loss in both PEI and

nanocomposite films at high field, causing a reduction in the discharged energy density.

Nevertheless, the discharge energy density of the nanocomposite is still 12% higher than

that of PEI, i.e., PEI films with only 0.32 vol% alumina nanofiller exhibit enhanced

discharge energy density and higher dielectric performance up to 150°C and above this

temperature.

Figure 4-5 (a) Charge-discharge cycles of PEI/alumina (20 nm) nanocomposites with 0.32

vol.% alumina under different electrical fields at 10 Hz and room temperature. Inset:

discharged energy density deduced from the charge-discharge cycle data. (b), (c), (d)

Discharged energy density of PEI/alumina (20 nm) nanocomposites with 0.32 vol.%

loading at different temperatures (room temperature, 100oC, and 150oC), and their

comparison with that of neat PEI and BOPP (room temperature), measured under 350

67

MV/m at 10 Hz. In (b), (c), and (d), Data points are shown and curves are drawn to guide

the eye. Data for BOPP were taken from Ref. [44]

The Weibull DC breakdown characteristics have been presented in Figure 4-6

showing characteristic breakdown strength of 406 MV/m and shape parameter of 5.5.

Figure 4-6 Weibull plot showing failure distribution for PEI-0.32% Al2O3

nanocomposite film.

To investigate the effect of nanoparticle size on the dielectric response of PEI

nanocomposites, films with alumina particles of 5 nm and 50 nm diameters were prepared

and characterized. As shown in Figure 4-7, the peak position of the dielectric enhancement

shifts to higher nanofiller volume content with nanoparticle size for these composites. For

68

PEI/alumina (5 nm) nanocomposites, the peak is at 0.24 vol.% with the dielectric constant

K=5, while for PEI/alumina (50 nm) nanocomposites the peak is at ca. 0.8 vol.% with the

K near 4.9. In addition, the composition range in which the dielectric enhancement occurs

is broader for larger-size nanoparticles. For nanoparticle interfacial effects, large size

nanofillers need higher volume content to reach a similar interfacial area surrounding the

nanofillers.

Figure 4-7 Effect of nanofiller size on the dielectric response (at 1 kHz) of PEI/alumina

composite films vs filler volume content.

69

Table 4-1 Summary of dielectric data of polyetherimide (PEI) with alumina

nanoparticles.

Neat Polymer Film Dielectric constant (1 kHz) Loss (1 kHz)

Neat PEI 3.17 0.29%

Nanocomposite Film Dielectric constant (1 kHz) Loss (1 kHz)

PEI with Al2O3 (5 nm)

PEI+0.08% Al2O3 (5 nm) (by vol.) 4.02 0.48%

PEI+0.16% Al2O3 (5 nm) (by vol.) 4.44 0.38%

PEI+0.24% Al2O3 (5 nm) (by vol.) 5.0 0.42%

PEI+0.32% Al2O3 (5 nm) (by vol.) 3.93 0.25%

PEI+0.48% Al2O3 (5 nm) (by vol.) 3.65 0.15%

PEI with Al2O3 (20 nm)

PEI+0.16% Al2O3 (20 nm) (by vol.) 4.56 0.46%

PEI+0.32% Al2O3 (20 nm) (by vol.) 5.0 0.46%

PEI+0.48% Al2O3 (20 nm) (by vol.) 4.74 0.5%

PEI+0.64% Al2O3 (20 nm) (by vol.) 4.05 0.27%

PEI+1.28% Al2O3 (20 nm) (by vol.) 4.01 0.68%

PEI+1.63% Al2O3 (20 nm) (by vol.) 3.93 0.61%

PEI with Al2O3 (50 nm)

PEI+0.27% Al2O3 (50 nm) (by vol.) 3.92 0.63%

PEI+0.32% Al2O3 (50 nm) (by vol.) 4.18 0.29%

PEI+0.64% Al2O3 (50 nm) (by vol.) 4.27 0.35%

PEI+0.83% Al2O3 (50 nm) (by vol.) 4.88 0.43%

PEI+1.1% Al2O3 (50 nm) (by vol.) 4.29 0.61%

PEI+1.66% Al2O3 (50 nm) (by vol.) 3.82 0.53%

70

4.4.2 Effect of nanoparticle type on dielectric constant

It was also investigated whether the observed enhancement in dielectric

response occurs with other nanofillers. PEI films with 20 nm size SiO2 (K = 3.9) and MgO

(K = 9.7) were prepared and characterized. Dielectric constants measured at room

temperature and 1 kHz are presented in Figure 4-8 and Table 4-2. For example, PEI/MgO

films show a dielectric constant of 4.95 (and maximum enhancement at 0.35 vol.% MgO

content). At 0.7 vol.%, the dielectric constant of the composite is reduced to 4.1, a trend

very similar to that observed in PEI/alumina nanocomposites, i.e., a significantly enhanced

dielectric response at very low nanofiller loading. It is interesting to note that the PEI/SiO2

nanocomposite has a dielectric constant K ~ 5, which is higher than both the pure polymer

matrix and the nanofiller. In addition, nanocomposites with boron nitride of 70 nm size

(hexagonal BN, dielectric constant ~ 5 - 7 and dielectric loss < 0.2%) were also prepared

and characterized. As shown in Figure 4-8 and Table 4-2, the PEI/BN films display an

enhanced dielectric constant, peaking at ca. 0.83 vol.% with K = 4.7 and displaying low

dielectric loss (Table 4-2). These results indicate that the enhanced dielectric constant in

the PEI films does not depend on the nanofiller type.

71

Figure 4-8 Effect of nanofiller size on the dielectric response (at 1 kHz) of PEI/alumina

composite films vs filler volume content. Experimental data points are shown and curves

are drawn to guide the eye.

Table 4-2 Summary of dielectric data of polyetherimide (PEI) with different type of

nanoparticles.

Nanocomposite Film Dielectric constant (1 kHz) Loss (1 kHz)

PEI with MgO (20 nm)

PEI+0.17% MgO (20 nm) (by vol.) 4.36 0.55%

PEI+0.35% MgO (20 nm) (by vol.) 4.95 0.23%

PEI+0.70% MgO (20 nm) (by vol.) 4.09 0.49%

PEI with SiO2 (20 nm)

PEI+0.26% SiO2 (20 nm) (by vol.) 4.88 0.24%

PEI+0.79% SiO2 (20 nm) (by vol.) 3.84 0.43%

PEI with BN (70 nm)

PEI+0.27% BN (70 nm) (by vol.) 3.73 0.25%

PEI+0.55% BN (70 nm) (by vol.) 3.97 0.58%

PEI+0.83% BN (70 nm) (by vol.) 4.71 0.89%

PEI+1.1% BN (70 nm) (by vol.) 4.55 0.19%

PEI+1.66% BN (70 nm) (by vol.) 4.21 0.18%

72

4.4.3 Effect of high dielectric constant nanoparticles

We note that PEI nanocomposites with high dielectric constant BaTiO3 (BTO)

nanofiller (50 nm particle size, K > 500) have been investigated earlier. [42], [95] To reach

a dielectric constant of K = 5, required more than 12 vol.% BTO nanofiller (See Figure 4-

9). The BTO volume content in these earlier studies was more than 3 vol.%. We thus

prepared PEI/BTO (50 nm) nanocomposites at very low volume filler content. As presented

in Figure 4-9 and Table 4-3, the dielectric response of PEI/BTO (50 nm) is nearly the same

as that of PEI/alumina (50 nm) in spite of the large difference in filler dielectric constant,

i.e., the peak enhancement (K=4.9) is at ca. 0.8 vol.%. Figure 4-10 summarizes all of the

nanocomposite experimental data in terms of percentage enhancement in the dielectric

constant of the nanocomposites, compared with that of the pure polymer matrix, showing

more than 50% enhancement in the nanocomposites studied.

73

Figure 4-9 Dielectric constant of PEI/BaTiO3 (50 nm size) nanocomposites vs. BaTiO3

volume content (experimental data points are shown and for > 3 vol.% nanocomposites

(orange squares) the data are from Ref. [95]). Experimental data are compared with several

commonly used composite models (Refs. [91]–[93]): (1) Parallel model (black), (2)

Maxwell model (red), (3) Lichtenecker model (green), and (4) series model (blue),

assuming the dielectric constant of BaTiO3 is 100X of that of PEI. Inset is an expanded

view of the enhanced dielectric response of nanocomposites at very low volume content (<

1 vol.%) due to nanoparticle interfacial effects, experimental data points are shown and

solid curve is drawn to guide the eye.

Figure 4-10 Summary of dielectric constants of PEI nanocomposites with different

nanofillers (20 nm MgO; 20 nm SiO2; 20 nm alumina; 50 nm BaTiO3; 70 nm BN).

Experimental data points are shown and lines are drawn to guide the eye.

74

Table 4-3 Summary of dielectric data of polyetherimide (PEI) with barium titanate

(BaTiO3) nanoparticles.

Nanocomposite Film Dielectric constant (1 kHz) Loss (1 kHz)

PEI with BaTiO3 (50 nm)

PEI+0.16% BaTiO3 (50 nm) (by vol.) 3.77 0.35%

PEI+0.32% BaTiO3 (50 nm) (by vol.) 3.94 0.43%

PEI+0.48% BaTiO3 (50 nm) (by vol.) 4.18 0.23%

PEI+0.64% BaTiO3 (50 nm) (by vol.) 4.36 0.24%

PEI+0.80% BaTiO3 (50 nm) (by vol.) 4.88 0.40%

PEI+0.96% BaTiO3 (50 nm) (by vol.) 4.44 0.45%

4.4.4 Importance of dipoles

A necessary condition for a polymer to exhibit a relatively high dielectric constant,

that is above 3, is that it should contain dipoles. This is the case for PEI in which the

phthalimide group possesses a high dipole moment, > 4 Debye. [94] In addition, to meet

the requirement of low dielectric loss (loss tangent < 0.01, or 1%), the dipoles in these

polymers should be weakly-coupled, which is also the case for PEI due to its amorphous

nature. [50], [55], [94] PEI has a high Tg (ca. 217 oC) and thus ensures high operating

temperature and low dielectric loss. However, the rigid structure of the polymer glass

imposes severe constraints on the responses of the dipoles to applied electric fields, limiting

the dielectric constant. [48], [55] On the other hand, if the constraints on the dipoles in the

glassy state can be significantly reduced, in this case by the interfacial effects of

75

nanocomposites, higher dielectric response can be achieved without the penalty of high

dielectric loss.

Analogous to PEI nanocomposites, the polar polyimide (PI) nanocomposites also

exhibit an enhanced dielectric constant occurring at very low filler volume content. [96] In

contrast, nanocomposites of a non-polar polymer, polystyrene, with alumina nanofillers of

20 nm and 50 nm, respectively, do not show dielectric enhancement, see Figure 4-11 and

Table 4-4. Thus, confirming that presence of polar polymer matrix is important for the

dielectric enhancement with low volume content of nanoparticles.

Figure 4-11 Dielectric constant measured at 1 kHz and room temperature vs. the nanofiller

content for PS nanocomposites. Data points are shown and solid curves are drawn to guide

the eye.

76

Table 4-4 Summary of dielectric data of non-polar polystyrene (PS) nanocomposite films.

Neat Polymer Film Dielectric constant (1 kHz) Loss (1 kHz)

Neat PS 2.77±0.04 0.30%

Nanocomposite Film Dielectric constant (1 kHz) Loss (1 kHz)

PS with Al2O3 (20 nm)

PS+0.16% Al2O3 (20 nm) (by vol.) 2.76±0.06 0.13%

PS+0.98% Al2O3 (20 nm) (by vol.) 2.79±0.03 0.14%

PS+1.63% Al2O3 (20 nm) (by vol.) 2.80±0.03 0.43%

PS with Al2O3 (50 nm)

PS+0.16% Al2O3 (50 nm) (by vol.) 2.83±0.02 0.33%

PS+0.48% Al2O3 (50 nm) (by vol.) 2.89±0.02 0.27%

PS+1.66% Al2O3 (50 nm) (by vol.) 2.87±0.03 0.24%

In order to check the effect of sub-Tg transitions (β and ɣ relaxations) on

dielectric constant of polymer nanocomposites [64], we also carried out broad band

dielectric spectroscopy study. The broadband dielectric data over a wide temperature range

indicates that there are no sub-Tg transitions down to -150oC, and no change in the dielectric

relaxation behavior between the PEI nanocomposites and the pure polymer. The only

difference is the increase in dielectric constant from 3.2 for pure PEI to 5 for the

nanocomposites, over the whole temperature range characterized (see Figures 4-3 (c), 4-4

and 4-12 (a) for the dielectric properties of PEI and nanocomposites of PEI/alumina (20

nm) with 0.32 vol.% nanofillers from – 150oC to 225oC at different frequencies, and

Figures 4-12 (b) and 4.12 (c) for the frequency spectra from 1 Hz to 1 MHz).

77

Figure 4-12 (a) Dielectric data at different frequencies of PEI+0.32 vol.% Al2O3 20 nm

as a function of temperature, (b) Dielectric data of PEI and (c) PEI+0.32 vol.% Al2O3 20

nm nanoparticle as a function of frequency at room temperature.

a

b

c

78

4.4.5 Structural analysis

A number of techniques were employed to characterize and analyze the changes

in PEI nanocomposites. Transmission electron microscopy images (see Figure 4-13 for a

representative image) show no evidence of appreciable nanofiller agglomeration in these

films. Differential scanning calorimetry (DSC) was performed on both pure PEI and the

PEI+0.32 vol.% alumina nanocomposite film. As shown in Figure 4-14 (a), the

nanocomposite film exhibits a reduced Tg (at ca. 210oC) compared with that of PEI (Tg

~217oC). In other words, average PEI segmental motion is somewhat faster in the presence

of a low volume fraction of well-dispersed alumina nanoparticles. This effect is presumably

dominated by changes in the dynamics of PEI segments near particle interfaces, leading to

reduced constraints for dipole reorientations in interfacial regions in the presence of applied

electric fields. Wide angle X-ray diffraction (XRD) was also carried out to probe possible

local structure changes due to the presence of nanofiller particles. As shown in Figure 4-

14 (b), the XRD pattern of the PEI - 0.32 vol.% alumina nanocomposite is nearly identical

to that of the neat PEI, displaying only the expected amorphous halo associated with PEI.

79

Figure 4-13 A representative TEM image of the PEI nanocomposite with 0.32 vol.%

alumina (20 nm particle size). Due to low volume content of nanoparticle in the composite,

only one nanoparticle is seen in the image area, as indicated.

Figure 4-14 (a) DSC and (b) X-ray diffraction data of PEI and the PEI nanocomposite with

0.32 vol.% of alumina.

20 nm

80

4.4.6 Multilayer core model for interfacial effect of nanocomposites

Tanaka’s multi-core model:

In order to explain how interfaces are formed chemically, physically and electrically,

Tanaka et. al. [97] proposed a hypothetical multi-layered core model to describe the

interfacial morphology between spherical inorganic particles and polymer matrix. The

multicore model involves a bonded layer, a bound layer and a loose layer as shown in the

Figure 4-15 below. The bonded layer corresponds to a transition layer with thickness of

about 1 nm, which is tightly bonded to both the spherical particle and the polymer. Such

bonding arises from either ionic, covalent or hydrogen bonds, or van der Waals force. The

bound layer is layer of polymer chains strongly bound and/or interacted to the bonded layer

and the surface of the particle. This layer plays a major role in altering the polymer chain

conformation, mobility and other stereographic structures. The thickness is in the range of

2-9 nm. The morphology and thickness of this layer is dependent on the interfacial

interaction strength in the bonded layer. The loose layer is a region loosely coupled to the

bound layer in which chain conformation, mobility, and even free volume and crystallinity

can differ from the polymer matrix. Comparatively, it has larger thickness, of several tens

of nanometers, than the inner layer. [67]

81

Figure 4-15 Tanaka’s multi-core model for interfaces between inorganic nanoparticles and

polymer matrix. [67], [97]

Based on Tanaka’s model Yang et al. [96] carried out the simulation study for our

experimental results. The interfacial region between the inorganic nanofiller and the

polymer matrix consists of multiple layers with different dielectric properties, where

dipoles in the inner layer may be restricted by chemical bonding or electric force, resulting

in a reduced local dielectric constant compared with the polymer matrix, whereas the outer

layer, which can be tens of nanometers thick, may contain more active dipoles with reduced

dipole rotation barriers, and thus show an increased local dielectric constant. The reduced

constraints for dipoles due to nanocomposites is consistent with the DSC data (Figure 4-

14 (a)). The analysis has been discussed in Appendix A.

82

4.5 Conclusion

We demonstrate a nanocomposite approach in which the dielectric response can be

enhanced markedly by addition of a very small amount of inorganic nanofiller, whose

dielectric constant can be similar to that of the polymer, thus eliminating many undesirable

features associated with the traditional nanocomposites using high dielectric constant

nanofillers to enhance the dielectric response. Although, the observed dielectric

phenomenon is beyond the present quantitative models of dielectric nanocomposites, still

these results will challenge and inspire fundamental research of interfacial effects in

nanocomposites. We expect that a successful development of theoretical understanding of

the observed phenomenon, plus the rich polymer chemistry available in tailoring the

nanostructures of dielectric polymers, will lead to a new generation of dipolar polymers

with much better performance than what is reported here. In addition to the higher dielectric

constant and breakdown field, nanocomposites with such a low volume content of

nanofillers allow for the use of melt extrusion method to produce dielectric films in large

quantities at low cost and high quality. Thus, opening the doors for commercialization of

this nanocomposite approach.

Acknowledgement

The part of this work has been submitted to Nanoscale published by RSC journal and has

been filed as a patent:

[1] Q. Zhang, Y. Thakur, T. Zhang and J. Runt, “Thin Film Capacitors” International

Patent Application PCT/US17/18307, February 17, 2017

83

Chapter 5

Dipolar polymers: high field behavior and study of conduction loss

5.1 Introduction

In the chapter 2, we discussed the rational design of dielectric polymers based on

weakly-coupled and strongly-dipolar polymers. Besides high dielectric constant, it is

important to maintain low conduction loss, especially at high electric fields. [44] In the

past few years, several aromatic polyurea and polythiourea polymers, possessing an

improved dielectric constant (K>4) have been developed. [50] The amorphous structure of

these polymers eliminates the polarization hysteresis loss, and high glass transition

temperature enables high thermal stability (>150°C operating temperature). However, most

of these aromatic polyurea and polythiourea polymers can only be processed into dielectric

films via solution-casting method, which causes higher manufacturing cost compared with

thermal (melt) process methods, and also solvents used may have deleterious impact on the

environment. Hence, one of the objective of this chapter is to investigate effective

approaches to chemically modify aromatic polyurea polymers, so that the resulting

polymer is thermally processable while maintaining high dielectric constant, low dielectric

loss, and exhibit high thermal stability.

One critical issue with many dielectric polymers which show low dielectric loss at

low electric field is the increased conduction loss at high electric fields. [31],[32]

Therefore, another objective is to examine the conduction loss of dipolar polymers. In the

84

first part of this chapter, the loss of PEMEU films at extremely high electric fields is

investigated.

The second part of this chapter is devoted to study of conduction loss at high field

and high temperature of another class of dipolar polymer – P(TFE-HFP-VDF) polymer,

which has been shown to be attractive for high temperature and high energy density

capacitors by Zhang et al. [98]. A nanocomposite approach has been proposed here, where

in order to suppress conduction at high temperature and high electric field, alumina (Al2O3)

nanofillers were added to the THV polymer matrix.

5.2 Review of breakdown mechanisms

In general, breakdown process in polymeric insulators is irreversible, and can be

destructive in many cases, resulting in a narrow conduction channel between the electrodes.

The term breakdown is used to describe processes in which a considerable current increase

results from a small voltage change. All the catastrophic breakdown in solids is electrically

power driven, and ultimately thermal in the sense that the discharge track involves at least

the melting and probably the carbonization or vaporization of the metalized dielectric. [99],

[100] An approach that can reduce the destructive effect of breakdown, and prolong the

lifetime is self-healing/self-clearing in metalized film capacitors. With a self-clearing

electrode, a fault in the dielectric will result in the thin metallized electrode in the

immediate vicinity being vaporized or turned from a metal conductor into a metal oxide

insulator. [2]

85

Deterministic models of breakdown are categorized to processes leading up to the

final stages of the breakdown. These are mainly subdivided into: electronic, thermal and

electromechanical breakdown. These processes are briefly discussed below:

5.2.1 Electronic breakdown

Electronic breakdown is initiated by local electric fields, in the range of 107-109 V/cm. At

high fields, the electrons accelerate with large mean free path by impact ionization,

resulting in catastrophic failure that can destroy the local lattice. The pioneering work in

this field was done by Von Hippel [101], who developed a single electron model to explain

the behavior in the low temperature range, followed by other models, such as the high-

energy criterion intrinsic breakdown by Frohlich, [102] and the avalanche breakdown by

Seitz [103].

The electronic breakdown in polymers can be considered of two types: intrinsic

breakdown and avalanche multiplication. The intrinsic breakdown occurs when the hot

electrons gain energy faster than the maximum rate at which they can lose energy by

electron-phonon scattering or inelastic collision to the lattice. Therefore, there must be a

critical field and corresponding electron energy above which the electrons indefinitely

acquire energy faster than they can lose it thereby leading to breakdown. This critical

electric field is referred as intrinsic breakdown field. [99]

The second is: avalanche multiplication or carrier impact ionization, where

electrons with a high energy, either as a result of acceleration in the field, or hot injection

from the electrode, or purely from chance fluctuations, collide with trapped or bound

86

electrons imparting sufficient energy for both electrons to be free after the collision. Given

a sufficiently high field both electrons rapidly gain enough energy to each cause a second

generation of collisions resulting in four free electrons. If this chain reaction continues, the

local concentration of high-energy electrons builds up to such an extent – that it is followed

by local destruction of the lattice. [99]

5.2.2 Thermal breakdown

Thermal breakdown occurs when electrical power dissipation causes heating of at least part

of the insulation to above a critical temperature, which results directly or indirectly in a

catastrophic failure. The conduction losses or losses due to polymer relaxation will result

in dissipation of power, and thereby cause Joule heating. The temperature will continue to

increase until the cooling of the insulator is equal to the electrical power dissipation, and

steady-state heat flow is set up. Breakdown occurs when this balance is disturbed by either

(1) physical change in the dielectric material, for example softening of the polymer; or (2)

the increase in conduction, as more carriers are available for conduction due to increase in

temperature. Alternatively, the increased segmental motion may increase the mobility for

intrinsic ionic conduction. Thus, electric power dissipation is increased causing a further

increase in temperature, and resulting in thermal runaway. [99], [104]

Mathematically, this mechanism can be expressed by the general equation (5.1) for the

thermal breakdown,

𝐶𝑣𝑑𝑇

𝑑𝑡− 𝑑𝑖𝑣(𝐾𝑡𝑔𝑟𝑎𝑑𝑇) = 𝜎𝐸2 (5.1)

87

where, Cv is the specific heat per unit volume; Kt is the thermal conductivity; T is the

temperature of the specimen; E is the applied field; σ is the electrical conductivity. The

first term on the left-hand side of equation represents heat absorbed by the material, and

the second term represents the heat lost to the surroundings. These are the two thermal

energy dissipation sources discussed above. The term on the right-hand side of the equation

represents the process of heat generation, which is dependent on the conductivity of the

material and applied field. The equation 5.1 clearly shows how the heat generation is

balanced by heat dissipation term. It is important to note that Cv, Kt and σ are functions of

both temperature and the applied field.

5.2.3 Electromechanical breakdown

Electromechanical breakdown takes place due to the electrostatic attraction of the

electrodes, which causes a mechanical compressive stress on the dielectric material

depending on its Young’s modulus. If the applied voltage is maintained, the field increases

due to the decrease in thickness, thereby increasing the attraction further. The breakdown

occurs when this mechanical compressive stress exceeds a critical value that cannot be

balanced by the dielectric’s elasticity. [99] Stark and Garton [105] gave the hypothesis for

this mechanism based on their observation of decreasing breakdown strength in

thermoplastics, when these materials start to soften with increase in temperature. This

breakdown mechanism can be mathematically evaluated by equating the two stresses for

the equilibrium situation before break down in a parallel-plate dielectric slab, i.e.,

electrostatic compressive stress = opposing elastic stress.

88

𝜖0𝜖𝑟

2 (

𝑉

𝑑)2

= 𝑌 𝑙𝑜𝑔𝑒 (𝑑0

𝑑) (5.2)

where Y is the Young’s modulus of elasticity, d0 is the initial dielectric thickness, and d is

the reduced thickness after application of voltage V. Rearranging the equation 5.2 in terms

of voltage,

𝑉 = 𝑑 (2𝑌

𝜖0𝜖𝑟𝑙𝑜𝑔𝑒 (

𝑑0

𝑑))

1

2

(5.3)

The above equation yields the critical voltage above which thickness goes to zero when

dV/d[d(V)]=0, which is expressed below:

𝑉𝑒𝑚 = 𝑑0 (𝑌

𝜖0𝜖𝑟exp (1))

1

2 (5.4)

Stark and Garton realized that their analysis pertaining to this model was rather

unrealistic as it assumed the dielectric material disappeared to an infinitesimal thickness at

V≥Vem, and it ignored plastic flow and dependence of Young’s modulus on time and stress.

The power law relation is used to characterize the polymers more accurately. [37]

𝜎 = 𝐾𝑒𝑁 = 𝐾 (𝑙𝑜𝑔𝑑0

𝑑)𝑁

(5.5)

1

2𝜖0𝜖𝑟 (

𝑉

𝑑)2

= 𝐾 (𝑙𝑜𝑔𝑑0

𝑑)𝑁

(5.6)

𝐸𝑐 =𝑉

𝑑𝑐=

2𝐾

𝜖0𝜖𝑟

1

2 (𝑁

2)

𝑁

2 (5.7)

89

where K is the Young’s modulus scales with the yield strength, N is the range of 0.1-0.6

for most polymers. For linear elasticity, K=Y and N=1; for the plasticity, N=0.

Table 5-1 Summary of dielectric theories of solids. [106]

I. Electronic breakdown process

Theories based on the single High energy criterion

electron approximation

Low energy criterion

Intrinsic breakdown

Collective critical field Single crystal

theories

Amorphous materials

Single avalanche model

Electron avalanche breakdown

Collective avalanche model

Field emission breakdown

Free volume breakdown

II. Thermal breakdown process

Steady state thermal breakdown process

Impulse thermal breakdown

III. Mechanical breakdown process

Electromechanical breakdown process

90

In addition to classification of various breakdown mechanism, the breakdown in

polymers depends on properties such as chemical structure, molecular motion, structural

irregularities, existence of additives etc. [99] Here, the focus will be on the electronic

breakdown mechanism and dependence of breakdown strength on the chemical structure,

specifically presence of polar groups in amorphous polymers. In the temperature

dependence of breakdown of non-polar polymers, the critical temperature (Tc) between low

and high temperature region clearly exists. However, in polar polymers a low temperature

region and Tc are not clearly defined. The introduction of a polar group in a polymer

structure increases the breakdown field in the low temperature region due to the scattering

of accelerated electrons by the dipoles from the polar group.

5.2.4 Frohlich amorphous solid model

Frohlich considered the effects of electron-electron interactions in an amorphous or

impure solid after giving his high-energy criterion theory. [102], [107] These types of

interactions are likely to dominate in amorphous insulators where traps are below the

conduction band. From his model, a collective breakdown field can be represented by the

following equation:

𝐸𝑐𝑜𝑙𝑙 = 𝐶 exp (𝛥𝐸

2𝑘𝐵 𝑇0) ,

where 𝐶 = (𝑚∗𝑛𝑣(𝑇0)𝐶2

exp (1)∆𝐸𝐶1)1/2 ℎ𝑣

𝑒𝜏𝑐 (5.8)

91

where C1 is an effective density of states and C2 = NtkBTe , Nt is the concentration of traps,

τc is the reciprocal of probability per unit time that an electron will make a transition in

energy.

In this mechanism, the breakdown strength drops rapidly with increasing

temperature, which has been observed in various amorphous materials. [107] The work

done by Austen and Pelzer [99] on polyethylene (PE) and polyvinylchloride (PVC)

confirmed the theory of Frohlich on amorphous solids. The long chains of PVC contain

strongly dipolar (H-C-Cl) side groups, which would act as scattering centers, thus the mean

free path for electrons in PVC should be much less than for PE. Frohlich theory predicts a

higher breakdown strength which was confirmed experimentally. [99] Moreover because

PVC is amorphous and PE is semi-crystalline, the transition to a negative temperature

coefficient of breakdown strength should be at a lower temperature; this was also found by

Austen and Pelzer. Apart from PVC, Wu et. al. showed that strongly dipolar and

amorphous polymers of thiourea can show high breakdown strength of (>1GV/m) at room

temperature due to presence of high dipole moment (>4Debye) polar groups, confirming

the Frohlich theory. [44], [108]

5.2.5 Dependence of breakdown strength on film thickness

Ideally, the dielectric breakdown strength of thin polymer films is a measure of the

intrinsic dielectric breakdown strength and should be independent of the thickness of film.

However, the situation is different for samples with thickness above 1 μm, which has been

supported by many studies in the past that show the dependence of breakdown strength on

92

the thickness. [109]–[113] It has been observed that the breakdown strength of the material

decreased with increasing the sample thickness. For instance, dc breakdown experiments

on very thin films deposited with Langmuir-Blodget technique have shown a power-law

dependence of the breakdown field on the film thickness. [109] In case of polypropylene,

it was shown that the electric strength of the polymer decreased with an increase in the

volume of an insulator, which was primarily due to the thickness dependence rather than

changes in area. [111]

In theory, the thickness dependence of breakdown strength is generally given by:

𝐸𝑏 = 𝑘𝑑−𝑛 (5.9)

where k is a constant, d is thickness of the sample, n is the exponent describing the

thickness dependence. For thermal breakdown mechanisms, this relationship is known for

steady-state and impulse thermal breakdown. Assumption of steady-state thermal

breakdown and field-independent conductivity leads to n=1 for thick slab approximation

(temperature distribution within the material) and n=0.5 for thin slab approximation

(constant temperature within material) .[99], [100] In case of field-dependent conductivity

and steady-state breakdown the exponents rely upon conduction mechanisms and show

independence or very weak dependence except space-charge limited conduction which

shows n=1.[99] The impulse thermal breakdown is independent or shows weak dependence

on the thickness.

In electronic breakdown, the intrinsic mechanism is independent of thickness. The

avalanche mechanism shows strong dependence on thickness for which Fowler-Nordheim

and Schottky emission result in n=0.5 and n=1. [114] While for electromechanical

breakdown mechanism, the breakdown strength is independent of thickness. The partial

93

discharge breakdown depends on the specimen dimension and void size; thickness

dependence shows n=0.39. [115]

5.3 Experimental section

5.3.1 Synthesis and film fabrication of PEMEU

Poly(ether methyl ether urea) (PEMEU) was synthesized from 2,2-Bis[4-(4-

aminophenoxy) phenyl] propane and diphenyl carbonate by polycondensation as shown in

Figure 5-1. The mixture of the two monomers was stirred at 150°C in vacuum for 4 h,

followed by washing with ethanol for 5-6 times to purify, and thus, PEMEU powder was

obtained. There is no solvent and no catalyst used in the synthesis. It is a green and low-

cost thermal polycondensation process. The 1H nuclear magnetic resonance (NMR) data

for PEMEU in DMSO-d6 is consistent with the structure of PEMEU (see Figure 5-2).

PEMEU peaks:1.48 (s, 6H, CH3-C-CH3); 7.07 (d, 4H, aromatic); 7.22 (d, 4H, aromatic);

7.48 (d, 4H, aromatic); 7.65 (d, 4H, aromatic); 8.75 (s, 2H, -NH-CO-NH-) and solvent

peaks: 2.45 (s, DMSO-d6); 3.35 (s, H2O in DMSO-d6). The free-standing films of PEMEU

were thermal processed (melt molding) under vacuum at 230°C, which is 70°C above its

Tg and 60°C below its decomposition temperature. The thickness of melt processed films

is in 70 μm to 100 μm range, that was used for dielectric characterization. To characterize

the high electric field response of the polymer films, solution-cast films of thickness down

to 1.32 μm were prepared. In this process, the PEMEU powder was dissolved in dimethyl

94

formamide (DMF) to make 1% solution by weight. The thin films were prepared by casting

the solution onto silicon substrates pre-coated with 40 nm of platinum at 70°C overnight,

and then annealed at 140°C under vacuum for 12 h. Circular gold electrodes of 2 mm to 6

mm in diameter were sputtered on the surfaces of the films for polarization loops

(charge/discharge) and dielectric characterization, respectively.

Figure 5-1 Schematic of synthesis and chemical structure of poly(ether methyl ether urea),

PEMEU.

Figure 5-2 1H-NMR spectrum for PEMEU in DMSO-d6.

95

5.3.2 Details of characterization equipment

The thickness of thin films was measured by Profilometer from KLA-Tencor. The

low-field dielectric constant and loss with varying temperature was measured using an HP

4294A precision impedance analyzer, which was connected to an environment test

chamber (Delta 9023). The Polarization-Electric field (P-E) response was measured with a

modified Sawyer-Tower circuit. The X-ray diffraction data were collected using a

Panalytical X’Pert PRO MPD diffractometer. Differential scanning calorimetry (DSC)

measurements were performed using the Q2000 DSC, TA instruments. The PEMEU

sample (film) was held in an aluminum pan and the scan was performed under nitrogen

ambient. The temperature was maintained at 40°C for 5 minutes, and then DSC scan was

performed as the sample was heated to 310°C at a heating rate of 10°C/min. The chemical

composition was characterized by 1H nuclear magnetic resonance (NMR) on a Bruker AM-

300 spectrometer. Atomic Force Microscopy (AFM) was performed using Bruker

Dimension AFM.

5.4 Results and discussion of PEMEU

In this work, a new type of aromatic polyurea, poly (ether methyl ether urea)

(PEMEU), was developed. Here, ether units are introduced to the polymer chain in order

to soften the rigid polymer backbone for developing thermally processable films. The

experimental results indicate that this approach is effective and free-standing films can be

fabricated using a laboratory melt molding machine. Concomitantly, the combination of

the high dipole moment of the urea units (dipole moment = 4.5 D), relatively high dipole

96

density in the polymer chains, and phenyl rings adjacent to the urea unit lead to a dielectric

constant of 4 with a low dielectric loss (~ 1%) over a broad temperature range (> 150°C).

The X-ray diffraction data of PEMEU film presented in Figure 5-3 (a) does not

show any sharp peak, just a broad halo at 2θ=17.5°, indicating that the polymer is

amorphous in nature. The amorphous nature of the PEMEU polymer can effectively

reduce the long-range polar coupling among the strongly dipolar urea units. Hence, in spite

of the very high dipole moment of the urea units in the polymer, the PEMEU films still

exhibit a low loss of 1 %. The DSC data in Figure 5-3 (b) shows a glass transition at 160°C

measured in the heating run. There is a heat absorption peak above 300°C due to the

decomposition of PEMEU. The high dipole moment and relatively high density of urea

units in the polymer chains impart a relatively high dielectric constant of 4 as presented in

Figure 5-4 (a), which is higher than most dielectric polymers reported in the literature with

dielectric constant below 3.3. [20], [32], [35] Moreover, the dielectric properties display

very little dispersion over a broad frequency range, a feature highly desirable for practical

dielectric devices. The temperature dependence of the polymer properties was also studied.

Figure 5-4 (b) shows that the dielectric properties of PEMEU are constant over a broad

temperature range and are stable up to 150°C.

97

Figure 5-3 (a) Wide angle X-ray diffraction data at room temperature and (b) DSC data

of PEMEU film measured during heating.

98

Figure 5-4 Dielectric constant and loss as functions of (a) frequency measured at room

temperature, and (b) temperature at frequencies from 1 kHz to 1 MHz of PEMEU films.

The error bars are attributed to the variation in thickness of film and the electrode area.

99

For dielectric films, the conduction loss at high electric field is a critical issue. [108],

[112], [113] Here, we examine the high field conduction loss of PEMEU films, especially

at very high electric fields. In dielectric films, it is well known that the dielectric breakdown

field increases with reduced film thickness. [112], [113] Earlier studies have shown that

the breakdown field of high-quality thin polymer films can reach > 1 GV/m. [44], [108]

Presented in Figure 5-5 is the electric breakdown field, measured from the polarization

loops at 10 Hz and room temperature, for the PEMEU films with different thickness. The

data shows that for PEMEU films of 1.32 µm thickness, the breakdown field can reach >

1.5 GM/m. The film quality (uniformity) was characterized by AFM and the result is

presented in Figure 5-6 (a). Indeed, across 2 µm x 2 µm surface area, the variation of the

film thickness is less than 0.5 nm for a 2.5 µm thick film. As observed in an earlier study,

avoidance of deep valleys in polymer films is essential for high dielectric strength, since

the sputtered electrode metal will penetrate into the defect areas and reduce the effective

dielectric thickness. [108] These thin spots in films will experience higher electric field,

increasing the conduction loss and causing breakdown.

The charge/discharge behavior was characterized in PEMEU thin films to

probe high field conduction and discharged energy density. The charging/discharging

behavior of the PEMEU films of 1.32 μm thick is presented in Figure 5-6 (b), measured at

10 Hz at room temperature. The data reveal that the 1.32 μm thick PEMEU films can reach

a breakdown field > 1.5 GV/m. It is interesting to note that the films even under 1.524

GV/m still show low conduction loss. The discharged energy density (UE, energy released)

and total stored energy density (US=UE + UL, where UL is the energy loss density) at high

fields were calculated from the charge-discharge curves as illustrated in Figure 5-6 (c),

100

where the charge-discharge efficiency ( ) is given by S

E

U

U .[14]

Figure 5-5 Electric breakdown field vs. film thickness for the PEMEU films measured at

room temperature. Dots represent y-axis error bars and symbols represent x-axis error

bars.

Due to the low conduction loss, the discharge energy efficiency at 1.524 GV/m is

90%, much higher than the non-polar polymers. [33] As a result, the polymer film delivers

a discharged energy density of 36 J/cm3, see Figure 5-6 (d). The results here are consistent

with the consideration that an effective approach to cut down the high field conduction loss

is to include high concentration of dipoles and deep traps in the polymer. The Coulombic

interaction between dipoles and charge carriers causes strong scattering compared with the

phonon-electron scattering, which reduces the conduction current and prevents dielectric

breakdown.[99], [116] Moreover, when the dipole moment exceeds a certain critical value,

as defined by, 64.0oaq

p

, the polar groups can also act as traps.[117]–[119] Here

q=1.6x10-19 C is the elementary charge, a0=5.29x10-11 m is the Bohr radius, and ϵ is the

101

effective dielectric constant that represents the dielectric screening effect due to the

medium. Considering the ultrashort electron transit time between scattering events under

very high electric field, the optical-frequency dielectric constant ϵ~2.48 is used for

estimation, which has been calculated from the refractive index of 1.577 at 632.8 nm for

polyurea [15]. This leads to a critical dipole moment of 4.04 D and dipole moment of urea

group (4.5 D) is much higher than this critical dipole moment.

Figure 5-6 (a) AFM image, (b) Charging/discharging curves under different unipolar

fields, (c) Schematic showing calculation of discharged energy density and loss under high

field from the charging/discharging curves, (d) Discharged energy density as a function of

field of PEMEU thin films of 1.32 μm thick, measured at room temperature. Dots represent

y-axis error bars and symbols represent x-axis error bar.

102

5.5 Introduction of nanoparticle dopants to reduce the conduction loss

In the first part of this chapter, an approach to reduce conduction loss was discussed

by incorporating high dipole moment units which act as scattering centers, thus reducing

the conduction loss. Still, it is important to explore other approaches for dielectric materials

that show promise for scalability and have been attractive for high energy density

capacitors.

Poly(vinylidene fluoride) (PVDF) based polymers have been considered for high

energy density capacitors as they exhibit high dielectric constant and high breakdown

strength. [16], [17], [36], [120] Dielectric constant from 10 to 50 can be achieved in these

semi-crystalline polymers. However, the strongly coupled dipoles in these polymers

exhibit pronounced polarization hysteresis at high fields, resulting in high loss. By

introducing tetrafluoroethylene (TFE) to VDF to provide low loss and high temperature

stability, and hexafluoropropylene (HFP) for ease of thermal processibility, Zhang et al.

[98] have shown that the P(TFE-HFP-VDF) polymer is attractive for high temperature and

high energy density capacitors. However, for high temperature applications, its conduction

loss at high temperatures and high fields needs to be reduced.

Nanocomposites belong to a new class of materials that are engineered for

improved material performance. [11], [81], [121] Ceramic nanoparticles can be blended

with dielectric polymer matrix to form polymer nanocomposites in which the large

interfacial areas and associated phenomena may be utilized for improving the dielectric

performance and electric insulation. For example, in previous studies, it has been shown

that the conduction can be reduced by an order of magnitude in a semi-crystalline polymer,

103

i.e., linear density polyethylene (LDPE), due to increased trap density in the nanocomposite

compared with the neat polymer.[122], [123] In this second part of the chapter, the

conduction at high temperatures and at fields up to 100 MV/m has been investigated in a

semi-crystalline poly(tetrafluoroethylene-hexafluoropropylene-vinylidene fluoride)

(THV) terpolymer, which has been shown to be attractive for high energy density

capacitors.

5.6 Review of conduction in polymers

Electrical conduction is governed by the manner of generating charge carriers and their

transport in a material. In general, the measured leakage current of a polymer film consists

of two parts:

𝐽 = 𝐽𝑐 + 𝐽𝑑 (5.10)

where Jc is the conduction current and Jd is the displacement current. The

displacement current is due to the change in electrical displacement with time, which is:

𝐽𝑑 =𝑑𝐷

𝑑𝑡 (5.11)

Displacement current is closely related to dielectric relaxation in the polymer as

shown by the equation (1.4) in Chapter 1. While, the conduction current is generated by

transport of charge carriers. The charge carriers that contribute to conduction current can

be either intrinsic carriers that are provided by the material itself or external carriers that

are injected from electrodes. [12]

104

For a metal-insulator-metal (MIM) structure, the conduction process can be

described in three steps: charge injection from electrode into the polymer, charge transport

in the bulk, and charge escaping from polymer to electrode. Each step contributes to the

total conduction current and the step with lowest charge transition rate will limit the overall

conduction current. Based on the different limitations, the conduction in polymers can be

classified as: electrode (injection) limited conduction and bulk limited conduction. Figure

5-7 illustrates the conduction mechanism in polymers. [12] HOMO (Highest Occupied

Molecular Orbital) and LUMO (Lowest Unoccupied Molecular Orbital) are used in

polymers or organics instead of valence band and conduction band.

Figure 5-7 Schematic of conduction process in polymers. [12]

105

5.6.1 Electrode limited conduction

In the electrode-limited case, the bulk material can take more carriers than the

electrode supplies, and correspondingly the current is limited by charge injection from the

electrodes into the insulator through an energy barrier. Since, this energy barrier prevents

the injection of the charge carriers from the electrodes, the charge carriers must overcome

these energy barriers for the current to flow.

The contact between the metal electrodes and polymers can be either Ohmic or

Schottky type depending on the relative position of the Fermi levels of metal electrodes

and polymers.

1. Schottky emission

In the Schottky theory, or also called thermionic emission, the current density is

generated by the electrons emitted from the electrode to the conduction band (LUMO) of

the insulator governed by the height of the Schottky barrier. [124], [125] With the

assistance of image force, and external field lowering barrier height, the charge carriers can

be injected by acquiring sufficient thermal energy to surpass the barrier. Figure 5-8 depicts

the Schottky contact between metal and n-type polymer. For Schottky emission, if the

electrons emitted from the cathode are not influenced by either space charges or traps, and

collected at the anode, then the thermionic emission is given by:

𝐽 = 𝐴∗𝑇2𝑒𝑥𝑝 (−𝜙𝑆

𝑘𝐵𝑇) 𝑒𝑥 𝑝 (

𝛽𝑆√𝐸

𝑘𝐵𝑇) (5.12)

where T is temperature, 𝜙𝑆 is the barrier height, 𝑘𝐵 is Boltzmann’s constant,

𝐴∗ is the Richardson constant, which is related to effective mass:

106

𝐴∗ =4𝜋𝑞𝑘𝐵𝑚∗

ℎ3 , (5.13)

and 𝛽𝑆 is expressed as:

𝛽𝑆 = √𝑞3

4𝜋𝜀0𝜀, (5.14)

Practically, the injected charge carriers can accumulate near the interface and form

space charges at the interfaces. In addition, the defects at the interface act as traps in

polymer and influence the electrode conduction. The disordered nature of most polymers,

narrow energy band from the weak-coupling between the polymer chains, and the presence

of localized states result in smaller mean free path of charge carriers in dielectric polymers

compared to crystalline materials. Therefore, the equation (5.12) needs to be modified to

account for these effects. A modified formula for Schottky emission, which takes into

account the finite bulk transportation rate in solids has been suggested by Simmons:

𝐽 = 2𝑞𝜇𝐸 (2𝜋𝑚∗𝑘𝐵𝑇

ℎ2 )

3

2𝑒𝑥𝑝 (−

𝜙𝑆−𝛽𝑆√𝐸

𝑘𝐵𝑇) (5.15)

107

Figure 5-8 Schematic showing Schottky contact between metal and n-type polymer (a)

before contact, (b) after contact, (c) barrier lowering by image force and (d) barrier

lowering by external voltage. [126], [127]

2. Field emission

The field emission is caused by the quantum mechanical tunneling of electrons

across the potential barrier at high electric field. [128] Usually, field emission dominates

at low temperature, as electrons can be emitted from the Fermi level of the metal to the

insulator. At high temperature, the electrons tunnel at energy levels higher than the Fermi

level. In this case, the emission will be thermally assisted field emission. As temperature

continues to rise, the thermionic emission will dominate. The field emission current in

defect free insulator is given by:

108

𝐽 = 𝐴∗𝑇2𝑒𝑥𝑝(−2𝛼𝜙𝐵

3/2/3𝑞𝐸)

(𝛼𝜋𝐵1/2

𝑘𝑇/𝑞𝐸)sin(𝜋𝛼𝜋𝐵1/2

𝑘𝑇/𝑞𝐸) (5.16)

At low temperatures, the equation can be simplified to:

𝐽 = 𝐴∗𝑇2

𝜋𝐵(

𝑞𝐸

𝛼𝑘𝑇)2

𝑒𝑥𝑝 (−2𝛼𝜋𝐵

3/2

3𝑞𝐸) (5.17)

In case the defects are considered, the field emission current will be largely affected

as the electrons will be partially captured by the traps in the insulator; resulting in space

charge near the interface of the metal-insulator. The Fowler-Nordheim tunneling current

density can be rewritten as:

𝐽(𝑡) = 𝑞2𝑚0[𝐹𝑐(𝑡)]

2

16𝜋2ℎ𝑚∗[𝑞𝜋𝐵(𝑡)] ×𝑒𝑥𝑝

4(2𝑚∗)1/2[𝑞𝜋𝐵(𝑡)]3/2

3ℎ𝑞𝐹𝑐(𝑡) (5.18)

5.6.2 Bulk limited conduction

When the injected charge carriers from electrode transit through the polymer film,

they may experience several resistances such as phonon scattering, defect trapping,

recombination etc. When the rate of bulk transportation is much slower than the supply of

carrier from the electrode, the total conduction will behave as a bulk limited current.

1. Space charge limited conduction

Space charge is generally referred to as the space filled with a net positive or

negative charge, and it is commonly observed in semiconductors and insulators. Such a

space charge can impose the number of charge carriers passing from one electrode to the

other, thus conduction is referred to as space charge limited. For example, if the cathode

109

emits more electrons per second than the polymer can transport, the remainder will

accumulate and form a negative space charge, which creates a field to reduce the rate of

electron emission from the cathode. Therefore, the current is controlled not by the electron-

injecting electrode but by the bulk of insulator or semiconductor. Without considering

traps, the ideal space charge limited current density can be expressed as:

J= 9/8ϵ0 ϵμV2/d3 (5.19)

where μ is the mobility in the polymer, V is the applied voltage and d is the

thickness of the sample. In space charge limited conduction, the current is proportional to

the square of the applied voltage. In the presence of traps in the polymer, part of space

charge may be trapped and cannot contribute to conduction. The current-voltage

characteristics of space-charge limited conduction has been summarized in Figure 5-9. The

mathematical derivation and details of space charge limited conduction with traps can be

accessed from Kao (2003). [12], [129], [130]

110

Figure 5-9 Schematic graph showing current density versus voltage for an ideal case of

space-charge limited current.[12]

2. Poole-Frenkel Conduction

The Poole-Frenkel conduction bears similarity with Schottky emission and it is also

referred as internal Schottky effect, since the mechanism of this effect is associated with

field-enhanced thermal excitation (or detrapping) of trapped electrons or holes. Both

effects are due to columbic interaction between the escaping electron and a positive charge,

but they differ in that the positive charge is fixed for the Poole-Frenkel trapping barrier,

while the positive charge is a mobile image charge for the Schottky barrier. In this case,

the effects of applied field would be strong enough to distort the potential well and lower

the energy barrier of the trap, effectively decreasing the depth of the trap. [12]

111

Polymers have intrinsic defects due to structure disorders and extrinsic defects due

to chain end groups and impurities left from processing. These defects act as traps ending

to capture charge carriers either from the electrode or from band conduction. With some

thermal excitation, the trapped electrons or holes can escape from these traps due to the

barrier lowering by applied high field. The current density due to Poole-Frenkel effect is

given by:

𝐽 = 𝜎0𝐸𝑒𝑥𝑝 (−𝐸𝑡

𝑘𝐵𝑇)×𝑒𝑥𝑝 (

𝛽𝑃𝐹√𝐸

𝑘𝐵𝑇) (5.20)

where 0 is the zero-field conductivity, Et is the trap energy barrier, βPF is Poole-

Frenkel constant given by:

𝛽𝑃𝐹 = √𝑞3

𝜋𝜖0𝜖 (5.21)

3. Hopping Conduction

A localized electron can drift in a bulk material by hopping from an occupied state

to an unoccupied state of a neighboring molecule if it acquires the energy necessary to

overcome the potential barrier. The carriers gain enough energy through random thermal

fluctuations and phonon interaction to escape their localized state and travel in an extended

state for a small amount of time before being recaptured by another localized state. [12],

[130], [131]

The probability of a hopping transition may be determined by both the distance

between the two sites and the potential barrier that must be overcome. If the potential

barrier width (i.e., the distance between the two sites) is larger than 10 Å, electron hop

112

from one molecule to the neighboring molecule. The probability of a hopping transition

may be determined by both the distance (l) between two sites and the barrier height Ea to

overcome:

𝐽 = 𝐽0exp (−𝐸𝑎

𝑘𝐵𝑇)×𝑒𝑥𝑝 (

𝑞𝑙𝐸

2𝑘𝐵𝑇) (5.22)

In disordered materials, like polymers, the individual carriers in presence of high

density of traps can exhibit percolation type transport. [131]–[133] When percolation is

applied to polymers, it typically takes the form of a spatially random resistor network with

each link of the network corresponding to the probability of a carrier hop between localized

states. [134], [135] The Figure 5-10 shows the current path through a hopping system

corresponding to a random resistor solution. The important feature of any percolation

model is the sudden appearance of long-range connectivity at a critical value, typically a

critical temperature (Tc). This transition point can be often linked to a physical transition

point, such as glass transition temperature. Below a critical temperature Tc, carriers are

restricted to interchain movement; above Tc the carriers can gain enough energy through

phonon interaction for long-range, interchain movement. In theory, this transition should

be quite sudden, however in practice, the wide variety of chain lengths and variability in

interconnectivity between crystalline and amorphous regions produces a continuous

transition, that can be difficult to observe sometimes. [99], [132], [133]

113

Figure 5-10 Schematic showing random resistor network percolation.

5.7 Film preparation and characterization of THV nanocomposites

The alumina nanoparticles of size 30-50 nm dispersed in IPA were purchased from Sigma

Aldrich. The THV polymer with composition – TFE/HFP/VDF-76/13/11 by weight, in form

of pellets was provided by PolyK Technologies. The nanoparticles were mixed with the

pellets in ethanol by wt.%. The solution was kept in vacuum for 24 hours. The resultant

THV polymer and nanoparticle mix was thoroughly mixed in a twin-screw micro

compounder. The resultant polymer nanocomposite was melt molded at 300°C to obtain

films of 30-40 µm thickness. It can be noted that the wt.% can be converted to vol.% by

using density of PVDF (ρ=1.8 g/cm3) and alumina nanoparticle (ρ=3.9 g/cm3); thereby

1wt.% equals to 0.46vol.%. The film thickness was evaluated using a digital micrometer.

114

The wide-angle X-ray scattering (WAXS) data were collected using a Panalytical X’Pert

PRO MPD diffractometer. The curve fitting and crystallinity was calculated using JADE

analysis software. WAXS two-dimensional images were obtained using a Rigaku DMAX-

Rapid II micro-diffractometer. The dielectric data was characterized by using HP 4294A

Precision Impedance Analyzer. The conduction current was characterized using a HP

4140B pA meter, which was connected to Trek High Voltage Amplifier (Model 609 D-6)

and environment test chamber (Delta 9023). A wait time of 20 seconds was given after every

voltage step for the current measurement. A longer wait time of 100 seconds was also used

at several temperatures and fields. There is no significant difference between the trends of

two sets of data. Same conditions were emulated when simulation work was performed.

The empirical data above and below critical temperature (Tc) are used to identify

direct tunneling and hopping conduction as the dominant charge injection and bulk

transport mechanisms, respectively, through nanocomposite film consisting of Al2O3 fillers

dispersed in the THV binder. Extracted parameters viz. barrier height, effective electron

mass, equivalent oxide thickness, mean hopping distance, trap depth, and density of state

in the conduction band are used together with measured I-V data and derived hopping

mobility for multiscale simulations with original and self-consistent continuum and particle

models to predict leakage conduction behavior at high temperature and high field.[136]–

[141] The multiscale models incorporate measured and extracted data from injection by

direct tunneling and hopping conduction transport together with charge

attachment/detachment, and recombination. The continuum model uses an effective

permittivity derived from the Lichtenecker logarithmic rule while the particle model caters

115

to an electrical double layer representation for the nanofiller. The details of simulation

models can be found in the supplemental section of Ref. [142].

5.8 Results and discussion of THV nanocomposites

The THV (TFE/HFP/VDF-76/13/11 mol%) polymer shows a dielectric constant of 4.3 and

low loss of <1% at 1 kHz as presented in Figure 5-11 and Figure 5-12 (a). The polymer

maintains its low loss at high temperatures (>100°C), see Figure 5-12, which is a desirable

feature for practical applications. The glass transition temperature (Tg) of the polymer is at

55°C as obtained from dynamic mechanical analysis (see Figure 5-12(b)). The data in Figure

5-12 (a) reveal that introducing various filler loadings of alumina nanoparticles into THV

polymer does not affect the dielectric properties and Tg.

Figure 5-11 Dielectric data of neat THV as a function of temperature.

116

As shown in Figure 5-13 (a), there is a large increase in the conduction current (J)

for neat THV at high temperature and high field, i.e., a 10X increase in the conduction

current as the polymer transitions from 45°C to 55°C and then another 10X increase from

55°C to 65°C, as a result of going through the Tg transition. Above Tg, the segmental

motions of the polymer chains will facilitate charge hopping, resulting in a large increase in

conduction current. [12], [34] The conductivity, defined here as the ratio of J(E)/E where E

is the applied field and J(E) the conduction current at E, as a function of field and

temperature is presented in Figure 5-13 (b) and 5-13 (c). The conductivity at 125 °C is

almost three orders of magnitude higher than that at room temperature.

The conduction current of THV nanocomposites with 0.5 wt%, 1 wt%, and 2 wt% of

alumina nanofillers (30 to 50 nm size, 1 wt % ~ 0.46 volume %) was characterized and is

presented in Figures 5-14 to 5-16. Figure 5-14 reveals that the high field conduction (> 40

MV/m) at high temperature (> 80 oC) is reduced by more than two orders of magnitude by

adding 0.5wt.% alumina nanoparticles to the THV polymer. While the nanofillers do not

affect the glass transition of the polymer matrix as observed in Figure 5-11(b), the

nanocomposites with 0.5 wt% nanofillers do not exhibit much changes in the conductivity

from room temperature to ca. 90oC, which is in sharp contrast to the large increase of

conductivity as the neat THV polymer passing through Tg. The result indicates that the

nanofillers effectively suppress the Tg effect on the leakage current of the neat THV

polymer. The nanofillers have a large surface area that leads to large amount of interface

regions between the nanoparticles and the polymer matrix. The local structure changes in

the interface area due to the polymer-nanoparticle interaction may increase the density and

117

depth of the charge carrier trapping sites, thus, reducing the conduction current. [71], [74],

[97], [143]

Figure 5-12 Dielectric data as a function of frequency for THV and THV nanocomposite

films, (b) DMA of THV and THV+0.5 wt.% Al2O3 films.

118

Figure 5-13 (a) Current density, (b) conductivity as a function of field at different

temperatures, (c) conductivity as a function of temperature at different fields for neat

THV film.

119

It is noted that the nanocomposites show an increase in the conductivity after 95°C

(Figure 5-14). The increase in the conduction current can be explained by percolation

theory. [131] In polymers, percolation takes the form of a spatially random resistor network

with each line of the network corresponding to the probability of carriers hopping between

localized states. The important feature of percolation is the appearance of long-range

connectivity at a critical value, such as at a critical temperature (Tc). The mobile charge

concentration n0 increases with temperature which reaches a critical concentration at Tc for

percolation. Below the temperature Tc, carriers are restricted to intra-chain movement while

above Tc, the carriers can gain enough energy through phonon interaction for long-range,

inter-chain movement. Even with such a percolation transition, n0 of the nanocomposites at

125oC is still 10X smaller and trap depth is more than 10% higher than those of the neat

THV polymer. Consequently, the high temperature (125oC) and high field (> 40 MV/m)

conductivity of nanocomposites with 0.5 wt.% of alumina nanofillers is more than two

orders of magnitude lower than that of the neat THV polymer.

120

Figure 5-14 (a) Current density, (b) conductivity as a function of field at different

temperatures, (c) conductivity as a function of temperature at different fields for

THV+0.5 wt.% Al2O3 film.

121

Similar reduction in the conduction current and conductivity is seen in the case of THV with

1 wt. percent alumina (Figure 5-15). The drastic reduction in conduction current by alumina

nanofillers at high temperatures is summarized in Figure 5-16. At 125°C, the conduction is

reduced by more than two orders of magnitude by addition of merely 0.5 wt.% alumina

nanocomposite film. The effect is confirmed in 1 wt.% alumina composition as well. A

threshold-like behavior is observed at 2 wt.% loading, i.e., further addition of nanoparticles

causes an increase in the dc conductivity. At high filler concentration, increased shallow

traps start behaving as conducting paths (forming percolation path), thus resulting in high

conductivity. [75], [123] Figure 5-16 (b) shows the conductivity at several temperatures as

functions of alumina nanofiller content. It’s clearly visible that there is an optimum

composition of filler loading (<2 wt.%), where dramatic reduction in conductivity can be

obtained.

122

Figure 5-15 (a) Current density, (b) conductivity as a function of field at different

temperatures, (c) conductivity as a function of temperature at different fields for THV+1

wt.% Al2O3 film.

123

Figure 5-16 (a) Comparison of conduction current of neat THV and different filler loadings

at 125°C, (b) scatter of conductivity at 60 MV/m as a function of alumina nanofiller

content. Dashed lines are drawn to guide the eye.

124

To provide insights into the experimental observations, we carried out simulation study on

the charge transport in these nanocomposite films. The multiscale models incorporate

measured and extracted data from injection by direct tunneling and hopping conduction

transport together with charge attachment/detachment, and recombination. [142], [144]

Figures S4 show the conductivity derived from measured I-V data and computed leakage

conductivities from multiscale simulations for T=85°C and 125°C as functions of electric

field for cases with THV+0.5wt.% Al2O3 and THV+1.0wt.% Al2O3, respectively. Both

continuum and particle models are in good agreement with each other and exhibit excellent

agreement with those derived from measured I-V data, thus, confirming the hopping

conduction as the dominant charge transport mechanism.

Fitting the experimental data of the neat THV polymer and nanocomposites with 0.5 wt%

nanofillers to the hopping current equation,

𝐽 = 𝑞𝑎𝑛0 𝑒[𝑞𝑎𝐸

𝑘𝑇−

𝐸𝑎𝑘𝑇

] (3.14)

yields the carrier concentration n0 as well as other parameters. Indeed, as can be seen, n0 of

2.85x1016/m3 for nanocomposites at 85 oC with 0.5 wt% nanofillers is about 20X lower than

that of the neat THV at 55oC. In addition, the trap depth is also increased in the

nanocomposites. Table 5-2 summarizes the fitting parameters of Eq. (1), hopping

conduction, for the neat THV and THV nanocomposites. Figure 5.17 plots the mobile

carrier concentration and trap depth vs. the filler content at 125oC. The mobile carrier

concentration decreases with the nanofillers in the composites, from 2.5x1018/m3 of neat

THV to 3.38x1015/m3 of the nanocomposites with 2 wt% alumina nanofillers. On the other

hand, the trap depth for the carrier hopping is the highest for the nanocomposites with 0.5

125

wt% alumina while the trap depth of 2 wt% nanocomposites becomes shallower (0.46 eV)

than that of the neat THV (0.52 eV). Combining the two effects leads to the lowest

conductivity for nanocomposites with < 2 wt% nanofillers.

Figure 5-17 Comparison of leakage conductivity from simulation and measurement at

85C and 125C for: (a) THV+0.5 wt.%, and (b) THV+1.0 wt.%, (c) carrier

concentration and trap depth as a function of filler content.

(

b)

126

Table 5-2 Summary of fitting parameters of hopping conduction equation for the neat

THV and nanocomposites.

Temperature

(°C)

Hopping

Distance (nm)

Carrier

Concentration

(#/m3)

Trap Depth

(eV)

Neat THV Polymer

35 0.7962 1.3219E+17 0.4879

55 0.7851 4.9620E+17 0.4849

125 0.6478 2.4968E+18 0.5204

THV+0.5wt% Al2O3

85 1.8510 2.8504E+16 0.6214

125 0.6288 2.2991E+17 0.6022

THV+1wt% Al2O3

85 0.9492 1.6181E+16 0.5525

125 1.0553 1.3082E+16 0.5550

THV+2wt% Al2O3

85 1.3662 3.4373E+15 0.4604

125 1.2494 3.3773E+15 0.4610

The wide-angle X-ray scattering (WAXS) results in Figure 5-18 show sharp crystalline

peak around 18° for both neat THV polymer and THV+1 wt.% Al2O3 nanocomposite film.

The crystallinity percentage calculated using Jade software for neat THV came out to be

47.61% with least-squares fitting residual (R)=2.9%. The high crystallinity can be

attributed to the presence of large wt.% of TFE as a monomer in THV. Interestingly, the

crystallinity decreases to 39.24% with R=3.3% in THV+1 wt.% Al2O3 films. To check if

there is a change in orientation of the polymer chains by addition of nanoparticles, we

performed 2-D WAXS (see Figure 5-19). Both neat polymer and nanocomposite films

show ring-like, isotropic scattering patterns suggesting no change in the orientation.

127

Figure 5-18 X-ray diffraction data of neat THV polymer and THV+1 wt.% Al2O3

nanocomposite films.

128

Figure 5-19 Two-dimensional X-ray diffraction data of neat THV polymer and THV+1

wt.% Al2O3 nanocomposite films.

(a) neat THV

(b) THV+1wt.%Al2O3

129

5.9 Conclusion

In summary, a thermally processable polymer, PEMEU, which shows a dielectric

constant of 4 with a low loss and thermal stability up to 150°C, was developed. In the high-

quality thin polymer films of 1.32 μm thick, an exceptionally high breakdown strength

(>1.5 GV/m) has been obtained. Because of the high dipole moment of urea units and

amorphous glass structure, the polymer films exhibit a low conduction loss even under 1.5

GV/m field. Consequently, the films generate a discharged energy density of 36 J/cm3.

In addition, it is shown that addition of small fraction (< 0.5 vol%) of alumina

nanofillers significantly reduces the conduction current in the THV polymer, which is

effective in introducing deep traps in the polymer matrix to suppress the mobile carrier

concentration. The results further reveal that there might exists a threshold nanofiller

concentration beyond which the suppression of conduction current becomes less effective.

Simulation study is in agreement with experimental findings. Moreover, the very low

volume content as demonstrated in the THV nanocomposites makes it possible to fabricate

nanocomposite films using melt extrusion technology, which is a very low cost and large-

scale fabrication technology for creating high quality polymer dielectric films. Thus, this

study paves the path for development of practical high temperature polymer-based

dielectrics using nanocomposites which exhibit low conduction loss at high fields and high

temperature.

130

Acknowledgement

The part of this chapter has been published in Journal of Electronic Materials, Applied

Physics Letters and International Conference of Dielectrics Proceedings. Reproduced here

by the permission of Springer and AIP Publishing LLC.

[1] Y. Thakur, M. Lin, S. Wu, and Q. M. Zhang, “Aromatic Polyurea Possessing High

Electrical Energy Density and Low Loss,” J. Electron. Mater., vol. 45, no. 10, pp. 4721–

4725, 2016

[2] Y. Thakur, T. Zhang, M. Lin, Q. Zhang, and M. H. Lean, “Mitigation of conduction

loss in a semi-crystalline polymer with high dielectric constant and high charge-discharge

efficiency,” in Dielectrics (ICD), 2016 IEEE International Conference on, 2016, vol. 1,

pp. 59–63

[3] Y. Thakur, M. H. Lean, and Q. Zhang, “Reducing conduction losses in high energy

density polymer using nanocomposites,” Appl. Phys. Lett., vol. 110, no. 12, p. 122905,

2017.

131

Chapter 6

Conclusion and recommendations for future work

6.1 Summary

The rational design of dielectric materials for high energy storage film capacitors

is one of the most active academic research areas in advanced functional materials. The

development of high energy storage capacitors will open ways for portable power supply

units for advanced weaponry; weight reduction for hybrid electric vehicles and medical

devices. This dissertation focused on designing nanostructured dielectric materials based

on fundamental concepts of applied polymer physics for application in high energy storage

capacitors. The key components for selection of dielectric materials are: dielectric constant,

loss, breakdown strength and thermal stability. In this dissertation, all of these components

have been discussed and addressed using rational design of materials.

Free volume approach

It is well known that strongly dipolar polymers that are weakly coupled exhibit low

dielectric loss, but also low dielectric constant at temperatures far below the glass transition

temperature Tg, due to constraints of the glassy structure on the dipoles. In contrast, at

temperatures above Tg, the reduced constraints on the dipoles, due to increased free

volume, lead to a large increase in dielectric constant. However, the loss also increases

markedly. The challenge is to introduce the free volume at temperatures below Tg, so that

132

large segment motion can be avoided and low loss is maintained. In this dissertation, we

showed through combined theoretical and experimental studies that by optimizing the

nanostructures in a family of weakly-coupled strongly dipolar polymers, such as aromatic

urea and thiourea polymers, we can overcome this challenge and develop a high energy

density polymer with low loss and high operating temperature. Two approaches to

introduce free volume below Tg was discussed: (1) by introducing disorder; (2) by blending

of two dissimilar dipolar polymers.

By introducing disorder in meta-phenylene polyurea (meta-PU), it is the free volume effect

(FVE) at temperatures below Tg (> 200°C) that leads to a high dielectric constant (K > 5.6).

It also possesses very low loss (high charge/discharge efficiency) even at very high electric

fields (> 600 MV/m). These results uncover that a disordered structure with a significantly

larger free volume enables easier reorientation of dipoles in response to an electric field

even at temperatures far below Tg, leading to a high dielectric constant while preserving a

low dielectric loss.

In the second approach, we showed that by blending two polymers from a family

of weakly-coupled strongly dipolar polymers, such as aromatic urea and thiourea polymers,

a high-energy density polymer with low loss and high operating temperature can be

realized. The resulting polymer blend of two strongly dipolar polymers, e.g., poly(arylene

ether urea) (PEEU, K = 4.7) and aromatic polythiourea (ArPTU, K = 4.4) exhibits an

exceptionally high dielectric constant, K = 7.5, while maintaining a low dielectric loss (<

1%). Both structure analysis and computer simulation results indicate that the nano

(molecular)-scale mixing of the two polymers causes a slight increase of the interchain

spacing in the glassy blend, thus reducing the barriers for dipole reorientation along the

133

applied electric field and generating a high dielectric response without compromising the

dielectric loss. Furthermore, we are able to show that the increase in the dielectric constant

of the mixture is due to increase in its specific volume, which enables easier reorientation

of the dipoles with the applied field.

Nanocomposite approach in dipolar polymers for high dielectric constant

In order to raise the dielectric constants of polymer-based dielectrics, composite

approaches, in which inorganic fillers with much higher dielectric constants are added to

the polymer matrix, have been investigated. However, the high dielectric constant fillers

cause high local electric fields in the polymer, resulting in a large reduction of the electric

breakdown strength. We show that a significant increase in the dielectric constant can be

achieved in polyetherimide nanocomposites with nanofillers whose dielectric constant can

be similar to that of the matrix. The presence of nanofillers reduces the constraints on the

dipole response to the applied electric field, thus enhancing the dielectric constant. Our

results demonstrate that through nanostructure engineering, the dielectric constant of

nanocomposites can be enhanced markedly without using high dielectric constant

nanofillers.

The realization of this proposed interfacial effect and successful demonstration of

the improvement in dielectric properties of dipolar polymers opens up new pathways in the

future research of nanocomposites. Most importantly, this approach can be applied to many

existing high temperature dielectric polymers, thus these nanocomposite polymers can be

easily scaled up by utilizing the existing manufacturing process and be readily developed

into capacitive devices.

134

Thermally processable polymer with high breakdown strength

Based on design of high dipole moment polyureas and polythioureas, a thermally

processable polyurea polymer, poly (ether methyl ether urea) (PEMEU), was designed by

introducing ether units in the structure to soften the rigid polymer. The polymer possesses

a dielectric constant of 4 and is thermally stable up to 150°C. The high dipole moment of

urea units and glass structure of the polymer leads to a low dielectric loss and low

conduction loss. As a result, PEMEU high quality thin films can be fabricated which

exhibit exceptionally high breakdown field of >1.5 GV/m, and a low conduction loss at

fields up to the breakdown. Consequently, the PEMEU films exhibit a high charge-

discharge efficiency of 90% and a high discharged energy density of 36 J/cm3. This is the

highest breakdown strength reported among class of polyureas and polythioureas

developed.

Nanocomposite approach for reducing conduction current

Apart from improving the energy density, it is also important to address the loss at

high fields and high temperature, known as the conduction loss. Normally, these losses

increase exponentially with the electric field, and cause heating of the capacitors. This

results in the need to have a cooling system to avoid overheating of the film capacitors. To

address this issue, a small fraction of alumina nanoparticles was added to a semi-crystalline

polymer, poly(tetrafluoroethylene-hexafluoropropylene-vinylidene fluoride) (THV)

terpolymer; resulting in the reduction of conduction loss by more than two orders of

magnitude at high temperatures (125°C), a desirable feature for practical applications. To

135

develop the theoretical understanding of the charge transport in these nanocomposite films,

we collaborated with Dr. Meng H. Lean, (CTO of QE Done LLC), who has developed

dynamic charge mapping models capable of simulating bipolar charge transport in both

layered and nanocomposite films. The combined theoretical and experimental study show

that the nanoparticles reduce the conduction losses in semi-crystalline polymers by

introducing deep traps in the polymer matrix, and thus trapping the charge carriers which

cause the conduction loss. The results further reveal that there might exists a threshold

nanofiller concentration (2 wt.%) beyond which the suppression of conduction current

becomes less effective.

6.2 Suggestions for future work

Utilization of Interfacial effect

The surface of nanofillers can be modified by chemical or physical means. There

has been a growing research interest in investigation of surface modification of the filler

on the dielectric properties. Zhou et al. [145] have shown that by modifying BaTiO3

(diameter=85-100 nm) nanoparticles with hydrogen peroxide (H2O2) improved the

morphology of PVDF-BaTiO3 nanocomposites due to hydrogen bonding between fluorine

and the hydroxyl group of PVDF and H2O2-modified BaTiO3. In addition, it made the

dielectric constant stable with frequency as well as temperature while reducing the

dielectric loss by restricting the movement of side groups and polymer chains near the

interface. Similarly, BaTiO3 has been modified using dopamine by Lin et al. [146],

resulting in increased dielectric constant by improved interface. On similar lines, there have

136

been many studies on surface modification of the fillers. [11], [67] However, all this work

has been on high volume content and mostly high dielectric constant fillers. There has been

hardly any work done on low volume content and low dielectric constant fillers. As it has

been presented in dissertation work that the low volume content nanocomposite films can

achieve high dielectric constant, it will be important to explore the interfacial phenomena

by surface modification of these nanofillers. There are few ideas along this line: chemical

synthesis of ceramic nanoparticles; i.e. utilizing the ability to choose ceramic particles

which can provide more interfacial area, thus providing more interfacial effect. Point

defects can also be introduced in these ceramic particles to introduce free volume. Another

nanoparticle system to look at would be conductive nanoparticles, which include metal

nanoparticles, e.g. bronze and copper nanoparticles; graphene nanoparticles. In addition,

effect on breakdown strength can be studied with surface modification of nanoparticles.

Thus, tailoring the nanoparticle surfaces as well as tuning the interfacial polymer shell

structures are recognized as crucial challenges along this direction of research.

Direct measurement of free volume

One of the ways to characterize free volume is by PALS (Positron Annihilation

Lifetime Spectroscopy). Till now, free volume measured in class of polyureas and

polythioureas has been using grazing incidence X-ray diffraction and simulation studies. It

is important to directly measure these sample using PALS. [66] Prof. David Gidley’s group

at University of Michigan studies nanoscale defects and open volumes in condensed matter

using positron annihilation spectroscopy. The initial PALS study of blends showed some

promising results of presence of large void sizes in blends compared to the individual

137

polymers. The blends of strongly dipolar polymers designed in future and surface modified

nanocomposite films can be studied using PALS technique.

Another method to probe changes in specific volume is by applying hydrostatic

pressures and measuring the dielectric properties. [147], [148] Changes in volume as a

function of pressure can give an insight into localized free volume effect.

Conduction mechanism of PEI nanocomposites

For charge transport study in Chapter 3, we collaborated with Dr. Meng H. Lean

(CTO of QE Done LLC), who has developed dynamic charge mapping models capable of

simulating bipolar charge transport in both layered and nanocomposite films. Till now, we

have studied THV, a semi-crystalline polymer, and seen reduction in charge conduction by

addition of nanocomposites. It is important to study an amorphous polymer as well,

preferably with high glass transition temperature like polyetherimide (PEI) or polyimide

(PI). [149] It’s a simpler system as it is a single polymer system, no copolymers plus they

are of commercial interest as well. The Figure 6-1 shows the conduction characteristics of

the neat PEI discussed in Chapter 4. The conduction is a function of both temperature and

field, and as seen in Figure 6-1 (b), the conduction increases by four orders of magnitude

when the neat polymer is heated from room temperature to 225°C. As demonstrated in

THV polymer system in Chapter 5, nanoparticles can be effective in reducing the

conduction current. Similar ideas can be applied to the PEI system as well. The

nanocomposites can be made by solution casting, so in that case it is easy to tune the size

of nanocomposites as well and study the size effect on charge conduction. Similarly, a non-

polar polymer like BOPP can also be studied. From the preliminary bipolar charge

138

transport simulations, it has been seen that the leakage current shows sensitivity to shape,

size and orientation of the nanoparticles. In addition, thermally stimulated discharging

current (TSDC) measurements can be conducted to study the trap levels, activation energy

and relaxation process of these nanocomposite films.

Figure 6-1 Current density as a (a) function of electric field over a range of temperatures,

(b) function of temperatures at 95.1 MV/m.

(a)

(b)

139

High Frequency Characterization

The high frequency characterization is important to understand the relaxation behavior of

these nanocomposite films. From initial discreet resonant experiments by Prof. Vid

Bobnar’s group at Jozef Stefan Institute in Slovenia, it is seen that these PEI nanocomposite

films start showing relaxation in GHz region as dielectric constant drops from 5 to 3.4. A

sweep of continuous high frequency scan in future will give more information on the

relaxation behavior but still discrete frequency results show signs of relaxation in these

nanocomposite films.

Figure 6-2 High frequency characterization of PEI-1wt.%Al2O3 nanocomposite films.

140

Appendix A

Chapter 4 Supporting Information

The following simulation work was performed by Tiannan Yang in Prof. Long-Qing

Chen’s group.

Based on Tanaka’s model presented in Chapter 4, a theoretical model is proposed

with a spatially varying local dielectric constant around the interfacial region within the

polymer matrix as:

2

interface 2 1 0,g

MK r K K g K e g r r

where r is the distance away from the surface of the nanofiller, KM is the dielectric

constant of the polymer matrix, K1 is the reduction of the dielectric constant in the inner

layer (r=0) compared to the polymer matrix, K2 determines the increase of the dielectric

constant in the outer layer, and r0 is a characteristic width of the interfacial region. Such

K(r) function features a quadratic growth in the inner layer (r/r0<<1), and an exponential

decay to KM far away from the interface (r/r0>>1). In calculating the polarization response

of the PEI/Al2O3 nanocomposites, the parameters are chosen as r0=50 nm, K1 = 0.7, and K2

= 8.5, as fitted from the experimentally measured effective dielectric constants of

PEI/Al2O3 nanocomposites with 20 nm filler size; KM = 3.2 is the dielectric constant of

PEI. See the distribution of dielectric constant in the interfacial region as a function of r in

Figure S1. These parameters are then fixed to reproduce the dielectric response of PEI

nanocomposites with different nanofiller sizes and volume contents.

The polarization response of the nanocomposite to an external field is simulated

through numerically solving the electrostatic equilibrium equation in a system with

periodically aligned 3-dimensional array of fillers in the polymer matrix described by the

phase-field method [150], [151], whereby the effective dielectric constant of the

nanocomposite can be calculated. Here, with the dielectric constant distributions around a

filler particle fitted to a well-established multilayer core model and a phase-field

description of a composite nanostructure, Yang et al. [96] successfully reproduced the

experimentally observed rapid increase in dielectric constant with filler volume fraction

141

and the appearance of a peak at very small volume fractions, as well as the shift of the peak

to higher volume fraction as the filler size increases. Figure S2 presents the spatial

distribution of the polarization in the PEI nanocomposites on applying an electric field of

1MV/m. As seen, at the dielectric peak of around 0.3 vol% nanofiller in PEI/Al2O3

nanocomposites with 20 nm filler size, a large region with greatly enhanced polarization

emerges in the middle of two nearest fillers on applying an electric field, due to its high

local dielectric constant resulting from an optimal distance to the surface of the fillers.

Changing the filler content to 0.1 vol% or 0.9 vol% leads to an increased or decreased

distance between nearest fillers, both of which will result in a reduced overall polarization

response. The dielectric peak shifts to a larger nanofiller volume content with an increased

filler size of 50 nm, due to a reduced surface-area-to-volume ratio. The distances between

adjacent nanofillers are listed in Table S1. As shown in Figure S3, with the parameters in

the multilayer core model fitted to specific filler materials, the computational results can

successfully reproduce the observed large increase of dielectric constants and the

appearance of a dielectric peak at low nanofiller volume content, as well as the shift of the

dielectric peak to higher nanofiller volume content with nanofiller size.

Figure S1. Distribution of the dielectric constant of polymer at the interfacial region as a

function of the distance r from the surface of nanofiller.

142

Figure S2. Polarization distribution in nanocomposites with (a) nanofillers of 20 nm at filler

content of 0.1 vol%, 0.3 vol%, and 0.9 vol%; and (b) nanofillers of 50 nm at filler content of 0.3

vol% and 0.9 vol%, on applying an electric field E3=1MV/m, within the cross section passing

through the centers of two nearest nanofillers.

Figure S3. Modeling results of the dielectric constant of PEI nanocomposites with 5 nm,

20 nm, and 50 nm nanofillers vs. the volume fraction of nanofillers.

143

Table S1. Effect of particle size and volume fraction on the distance between the

neighboring nanoparticles.

Particle diameter

(nm) Volume fraction (%)

Distance between neighboring

particles (nm)

20 0.1 141

20 0.3 92

20 0.9 57

50 0.3 230

50 0.9 144

144

Appendix B

Chapter 5 Supporting Information

The following simulation was performed by Dr. Meng H. Lean from QE Done LLC.

Figure S4. Comparison of leakage conductivity from simulation and measurement at

85C and 125C for: (a) THV+0.5 wt.%, and (b) THV+1.0 wt.%.

(a)

(b)

145

Bibliography

[1] M. Winter and R. J. Brodd, What are batteries, fuel cells, and supercapacitors? ACS

Publications, 2004.

[2] W. J. Sarjeant, J. Zirnheld, and F. W. MacDougall, “Capacitors,” Plasma Sci. IEEE Trans.

On, vol. 26, no. 5, pp. 1368–1392, 1998.

[3] S. Ducharme, “An inside-out approach to storing electrostatic energy,” ACS Nano, vol. 3, no.

9, pp. 2447–2450, 2009.

[4] H. Wisken and T. H. Weise, “Capacitive pulsed power supply systems for ETC guns,” IEEE

Trans. Magn., vol. 39, no. 1, pp. 501–504, 2003.

[5] H. Akiyama, T. Sakugawa, T. Namihira, K. Takaki, Y. Minamitani, and N. Shimomura,

“Industrial applications of pulsed power technology,” IEEE Trans. Dielectr. Electr. Insul.,

vol. 14, no. 5, 2007.

[6] F. MacDougall et al., “High energy density capacitors for pulsed power applications,” in

Pulsed Power Conference, 2009. PPC’09. IEEE, 2009, pp. 774–778.

[7] T. A. Burress et al., “Evaluation of the 2007 Toyota Camry hybrid synergy drive system,”

Oak Ridge National Laboratory (ORNL), Oak Ridge, TN, 2008.

[8] T. A. Burress et al., “Evaluation of the 2008 Lexus LS 600H hybrid synergy drive system,”

Oak Ridge National Laboratory (ORNL), Oak Ridge, TN, 2009.

[9] R. Deshpande, Capacitors: technology and trends. Tata McGraw-Hill Education, 2012.

[10] S. O. Kasap, Principles of electronic materials and devices. McGraw-Hill, 2006.

[11] V. K. Thakur and R. K. Gupta, “Recent progress on ferroelectric polymer-based

nanocomposites for high energy density capacitors: synthesis, dielectric properties, and

future aspects,” Chem. Rev., vol. 116, no. 7, pp. 4260–4317, 2016.

[12] K. C. Kao, Dielectric phenomena in solids. Academic press, 2004.

[13] I. Burn and D. Smyth, “Energy storage in ceramic dielectrics,” J. Mater. Sci., vol. 7, no. 3,

pp. 339–343, 1972.

[14] S. Wu, M. Lin, S. Lu, L. Zhu, and Q. Zhang, “Polar-fluoropolymer blends with tailored

nanostructures for high energy density low loss capacitor applications,” Appl. Phys. Lett.,

vol. 99, no. 13, p. 132901, 2011.

[15] H. S. Nalwa, Handbook of Low and High Dielectric Constant Materials and Their

Applications, Two-Volume Set. Academic Press, 1999.

[16] Q. Chen, Y. Shen, S. Zhang, and Q. Zhang, “Polymer-Based Dielectrics with High Energy

Storage Density,” Annu. Rev. Mater. Res., vol. 45, no. 1, 2015.

[17] B. Chu et al., “A dielectric polymer with high electric energy density and fast discharge

speed,” Science, vol. 313, no. 5785, pp. 334–336, 2006.

[18] V. Sharma et al., “Rational design of all organic polymer dielectrics,” Nat. Commun., vol. 5,

2014.

[19] Q. Li et al., “Flexible high-temperature dielectric materials from polymer nanocomposites,”

Nature, vol. 523, no. 7562, pp. 576–579, 2015.

[20] D. Tan, L. Zhang, Q. Chen, and P. Irwin, “High-Temperature Capacitor Polymer Films,” J.

Electron. Mater., vol. 43, no. 12, pp. 4569–4575, 2014.

[21] Y. Cao, P. C. Irwin, and K. Younsi, “The future of nanodielectrics in the electrical power

industry,” IEEE Trans. Dielectr. Electr. Insul., vol. 11, no. 5, pp. 797–807, 2004.

[22] M.-H. Yoon, H. Yan, A. Facchetti, and T. J. Marks, “Low-voltage organic field-effect

transistors and inverters enabled by ultrathin cross-linked polymers as gate dielectrics,” J.

Am. Chem. Soc., vol. 127, no. 29, pp. 10388–10395, 2005.

146

[23] S. Bertolazzi, J. Wünsche, F. Cicoira, and C. Santato, “Tetracene thin film transistors with

polymer gate dielectrics,” Appl. Phys. Lett., vol. 99, no. 1, p. 13301, 2011.

[24] M. E. Roberts, N. Queraltó, S. C. Mannsfeld, B. N. Reinecke, W. Knoll, and Z. Bao, “Cross-

linked polymer gate dielectric films for low-voltage organic transistors,” Chem. Mater., vol.

21, no. 11, pp. 2292–2299, 2009.

[25] A. Facchetti, “Π-conjugated polymers for organic electronics and photovoltaic cell

applications†,” Chem. Mater., vol. 23, no. 3, pp. 733–758, 2010.

[26] M. T. Dang, L. Hirsch, and G. Wantz, “P3HT: PCBM, best seller in polymer photovoltaic

research,” Adv. Mater., vol. 23, no. 31, pp. 3597–3602, 2011.

[27] L. Huo, S. Zhang, X. Guo, F. Xu, Y. Li, and J. Hou, “Replacing alkoxy groups with

alkylthienyl groups: a feasible approach to improve the properties of photovoltaic polymers,”

Angew. Chem., vol. 123, no. 41, pp. 9871–9876, 2011.

[28] T. L. Hanley, R. P. Burford, R. J. Fleming, and K. W. Barber, “A general review of

polymeric insulation for use in HVDC cables,” Electr. Insul. Mag. IEEE, vol. 19, no. 1, pp.

13–24, 2003.

[29] J. Ennis et al., “Recent advances in high voltage, high energy capacitor technology,” in

Pulsed Power Conference, 2007 16th IEEE International, 2007, vol. 1, pp. 282–285.

[30] H. Bluhm, “Power Systems,” 2006.

[31] T. Osaka and M. Datta, Energy storage systems in electronics. CRC Press, 2000.

[32] M. Rabuffi and G. Picci, “Status quo and future prospects for metallized polypropylene

energy storage capacitors,” Plasma Sci. IEEE Trans. On, vol. 30, no. 5, pp. 1939–1942,

2002.

[33] Q. Chen, Y. Wang, X. Zhou, Q. Zhang, and S. Zhang, “High field tunneling as a limiting

factor of maximum energy density in dielectric energy storage capacitors,” Appl. Phys. Lett.,

vol. 92, no. 14, p. 142909, 2008.

[34] C. Xu and S. Boggs, “Measurement of resistive and absorption currents in capacitor films up

to breakdown,” in Electrical Insulation, 2006. Conference Record of the 2006 IEEE

International Symposium on, 2006, pp. 249–252.

[35] J. Ho and R. Jow, “Characterization of high temperature polymer thin films for power

conditioning capacitors,” DTIC Document, 2009.

[36] L. Zhu and Q. Wang, “Novel ferroelectric polymers for high energy density and low loss

dielectrics,” Macromolecules, vol. 45, no. 7, pp. 2937–2954, 2012.

[37] X. Zhou et al., “Electrical breakdown and ultrahigh electrical energy density in poly

(vinylidene fluoride-hexafluoropropylene) copolymer,” Appl. Phys. Lett., vol. 94, no. 16, p.

162901, 2009.

[38] X. Zhou, B. Chu, B. Neese, M. Lin, and Q. Zhang, “Electrical energy density and discharge

characteristics of a poly (vinylidene fluoride-chlorotrifluoroethylene) copolymer,” Dielectr.

Electr. Insul. IEEE Trans. On, vol. 14, no. 5, pp. 1133–1138, 2007.

[39] Q. Zhang, V. Bharti, and X. Zhao, “Giant electrostriction and relaxor ferroelectric behavior

in electron-irradiated poly (vinylidene fluoride-trifluoroethylene) copolymer,” Science, vol.

280, no. 5372, pp. 2101–2104, 1998.

[40] N. J. Smith, B. Rangarajan, M. T. Lanagan, and C. G. Pantano, “Alkali-free glass as a high

energy density dielectric material,” Mater. Lett., vol. 63, no. 15, pp. 1245–1248, 2009.

[41] J. Ho and T. Jow, “Characterization of High Temperature Polymer Thin Films for Power

Conditioning Capacitors; Army Research Laboratory: Adelphi, MD, 2009,” There No

Corresp. Rec. This Ref.

[42] D. Tan, Y. Cao, E. Tuncer, and P. Irwin, “Nanofiller dispersion in polymer dielectrics,”

2013.

147

[43] Y. Wang, X. Zhou, M. Lin, and Q. Zhang, “High-energy density in aromatic polyurea thin

films,” Appl. Phys. Lett., vol. 94, no. 20, p. 202905, 2009.

[44] S. Wu et al., “Aromatic polythiourea dielectrics with ultrahigh breakdown field strength, low

dielectric loss, and high electric energy density,” Adv. Mater., vol. 25, no. 12, pp. 1734–1738,

2013.

[45] Y. Thakur et al., “Optimizing nanostructure to achieve high dielectric response with low loss

in strongly dipolar polymers,” Nano Energy, vol. 16, pp. 227–234, 2015.

[46] Z. Cheng et al., “Aromatic poly (arylene ether urea) with high dipole moment for high

thermal stability and high energy density capacitors,” Appl. Phys. Lett., vol. 106, no. 20, p.

202902, 2015.

[47] W. Kumler and G. M. Fohlen, “The Dipole Moment and Structure of Urea and Thiourea1,” J.

Am. Chem. Soc., vol. 64, no. 8, pp. 1944–1948, 1942.

[48] H. FROHLICH, Theory of Dielectrics: Dielectric Constant and Dielectric Loss. Oxford

University Press, 1987.

[49] S. Wu, M. Lin, Q. Burlingame, and Q. Zhang, “Meta-aromatic polyurea with high dipole

moment and dipole density for energy storage capacitors,” Appl. Phys. Lett., vol. 104, no. 7,

p. 72903, 2014.

[50] Y. Thakur, M. Lin, S. Wu, Z. Cheng, D.-Y. Jeong, and Q. Zhang, “Tailoring the dipole

properties in dielectric polymers to realize high energy density with high breakdown strength

and low dielectric loss,” J. Appl. Phys., vol. 117, no. 11, p. 114104, 2015.

[51] J. Lennard-Jones and A. Devonshire, “Critical and co-operative phenomena. III. A theory of

melting and the structure of liquids,” Proc. R. Soc. Lond. Ser. Math. Phys. Sci., pp. 317–338,

1939.

[52] J. Lennard-Jones and A. Devonshire, “Critical and co-operative phenomena. IV. A theory of

disorder in solids and liquids and the process of melting,” Proc. R. Soc. Lond. Ser. Math.

Phys. Sci., pp. 464–484, 1939.

[53] S. Glasstone, H. Eyring, and K. J. Laidler, The theory of rate processes. McGraw-Hill, 1941.

[54] J. Frenkel, “Viscous flow of crystalline bodies,” Zhurnal Eksp. Teor. Fiz., vol. 16, no. 1, pp.

29–38, 1946.

[55] N. G. McCrum, B. E. Read, and G. Williams, “Anelastic and dielectric effects in polymeric

solids,” 1967.

[56] L. H. Sperling, Introduction to physical polymer science. John Wiley & Sons, 2005.

[57] A. K. Doolittle, “Studies in Newtonian flow. I. The dependence of the viscosity of liquids on

temperature,” J. Appl. Phys., vol. 22, no. 8, pp. 1031–1035, 1951.

[58] A. K. Doolittle, “Studies in Newtonian flow. II. The dependence of the viscosity of liquids on

free-space,” J. Appl. Phys., vol. 22, no. 12, pp. 1471–1475, 1951.

[59] A. K. Doolittle, “Studies in Newtonian flow. III. The dependence of the viscosity of liquids

on molecular weight and free space (in homologous series),” J. Appl. Phys., vol. 23, no. 2,

pp. 236–239, 1952.

[60] H. Djidjelli, M. Kaci, T. Sadoun, D. Benachour, J. Farenc, and J.-J. Martinez-Vega, “Effect

of hydrothermic ageing on dielectric and mechanical properties of rigid poly (vinyl

chloride),” Macromol. Mater. Eng., vol. 286, no. 7, pp. 407–411, 2001.

[61] D. J. Bergman, “The dielectric constant of a composite material—a problem in classical

physics,” Phys. Rep., vol. 43, no. 9, pp. 377–407, 1978.

[62] K. Tashiro, Ferroelectric Polymers, ed. HS Nalwa. Dekker, New York, 1995.

[63] R. Dong, V. Ranjan, M. B. Nardelli, and J. Bernholc, “Atomistic simulations of aromatic

polyurea and polyamide for capacitive energy storage,” Phys. Rev. B, vol. 92, no. 2, p. 24203,

2015.

148

[64] J. Wei et al., “Achieving High Dielectric Constant and Low Loss Property in a Dipolar Glass

Polymer Containing Strongly Dipolar and Small-Sized Sulfone Groups,” ACS Appl. Mater.

Interfaces, vol. 7, no. 9, pp. 5248–5257, 2015.

[65] Y. Thakur et al., “Generating High Dielectric Constant Blends from Lower Dielectric

Constant Dipolar Polymers using Nanostructure Engineering,” Nano Energy, 2016.

[66] H. Hristov, B. Bolan, A. Yee, L. Xie, and D. Gidley, “Measurement of hole volume in

amorphous polymers using positron spectroscopy,” Macromolecules, vol. 29, no. 26, pp.

8507–8516, 1996.

[67] Z.-M. Dang, J.-K. Yuan, S.-H. Yao, and R.-J. Liao, “Flexible nanodielectric materials with

high permittivity for power energy storage,” Adv. Mater., vol. 25, no. 44, pp. 6334–6365,

2013.

[68] T. Tanaka, G. Montanari, and R. Mulhaupt, “Polymer nanocomposites as dielectrics and

electrical insulation-perspectives for processing technologies, material characterization and

future applications,” IEEE Trans. Dielectr. Electr. Insul., vol. 11, no. 5, pp. 763–784, 2004.

[69] Q. Wang and L. Zhu, “Polymer nanocomposites for electrical energy storage,” J. Polym. Sci.

Part B Polym. Phys., vol. 49, no. 20, pp. 1421–1429, 2011.

[70] P. Kim et al., “High energy density nanocomposites based on surface-modified BaTiO3 and

a ferroelectric polymer,” ACS Nano, vol. 3, no. 9, pp. 2581–2592, 2009.

[71] T. Lewis, “Interfaces are the dominant feature of dielectrics at the nanometric level,” IEEE

Trans. Dielectr. Electr. Insul., vol. 11, no. 5, pp. 739–753, 2004.

[72] J. Li, S. I. Seok, B. Chu, F. Dogan, Q. Zhang, and Q. Wang, “Nanocomposites of

ferroelectric polymers with TiO2 nanoparticles exhibiting significantly enhanced electrical

energy density,” Adv. Mater., vol. 21, no. 2, pp. 217–221, 2009.

[73] B. Balasubramanian, K. L. Kraemer, N. A. Reding, R. Skomski, S. Ducharme, and D. J.

Sellmyer, “Synthesis of Monodisperse TiO2- Paraffin Core- Shell Nanoparticles for

Improved Dielectric Properties,” ACS Nano, vol. 4, no. 4, pp. 1893–1900, 2010.

[74] J. K. Nelson and J. C. Fothergill, “Internal charge behaviour of nanocomposites,”

Nanotechnology, vol. 15, no. 5, p. 586, 2004.

[75] J. K. Nelson and others, Dielectric polymer nanocomposites. Springer, 2010.

[76] J. Li, J. Claude, L. E. Norena-Franco, S. I. Seok, and Q. Wang, “Electrical energy storage in

ferroelectric polymer nanocomposites containing surface-functionalized BaTiO3

nanoparticles,” Chem. Mater., vol. 20, no. 20, pp. 6304–6306, 2008.

[77] Z. Li et al., “In situ catalytic encapsulation of core-shell nanoparticles having variable shell

thickness: dielectric and energy storage properties of high-permittivity metal oxide

nanocomposites,” Chem. Mater., vol. 22, no. 18, pp. 5154–5164, 2010.

[78] Q. Li et al., “Sandwich-structured polymer nanocomposites with high energy density and

great charge–discharge efficiency at elevated temperatures,” Proc. Natl. Acad. Sci., vol. 113,

no. 36, pp. 9995–10000, 2016.

[79] N. Guo, S. A. DiBenedetto, P. Tewari, M. T. Lanagan, M. A. Ratner, and T. J. Marks,

“Nanoparticle, size, shape, and interfacial effects on leakage current density, permittivity,

and breakdown strength of metal oxide- polyolefin nanocomposites: experiment and theory,”

Chem. Mater., vol. 22, no. 4, pp. 1567–1578, 2010.

[80] C. Zou, D. Kushner, and S. Zhang, “Wide temperature polyimide/ZrO 2 nanodielectric

capacitor film with excellent electrical performance,” Appl. Phys. Lett., vol. 98, no. 8, p.

82905, 2011.

[81] X. Huang and C. Zhi, Polymer Nanocomposites: Electrical and Thermal Properties.

Springer, 2016.

[82] Z.-M. Dang et al., “Advanced calcium copper titanate/polyimide functional hybrid films with

high dielectric permittivity,” Adv. Mater., vol. 21, no. 20, pp. 2077–2082, 2009.

149

[83] W. Yang, S. Yu, R. Sun, and R. Du, “Nano-and microsize effect of CCTO fillers on the

dielectric behavior of CCTO/PVDF composites,” Acta Mater., vol. 59, no. 14, pp. 5593–

5602, 2011.

[84] B. S. Prakash and K. Varma, “Dielectric behavior of CCTO/epoxy and Al-CCTO/epoxy

composites,” Compos. Sci. Technol., vol. 67, no. 11, pp. 2363–2368, 2007.

[85] F. Amaral, C. Rubinger, F. Henry, L. Costa, M. Valente, and A. Barros-Timmons, “Dielectric

properties of polystyrene–CCTO composite,” J. Non-Cryst. Solids, vol. 354, no. 47, pp.

5321–5322, 2008.

[86] H. Tang, Y. Lin, C. Andrews, and H. A. Sodano, “Nanocomposites with increased energy

density through high aspect ratio PZT nanowires,” Nanotechnology, vol. 22, no. 1, p. 15702,

2010.

[87] J. Yao et al., “Enhancement of dielectric constant and piezoelectric coefficient of ceramic–

polymer composites by interface chelation,” J. Mater. Chem., vol. 19, no. 18, pp. 2817–2821,

2009.

[88] S. Banerjee and K. Cook-Chennault, “Influence of Al particle size and lead zirconate titanate

(PZT) volume fraction on the dielectric properties of PZT-epoxy-aluminum composites,” J.

Eng. Mater. Technol., vol. 133, no. 4, p. 41016, 2011.

[89] H. Tang, Y. Lin, and H. A. Sodano, “Enhanced energy storage in nanocomposite capacitors

through aligned PZT nanowires by uniaxial strain assembly,” Adv. Energy Mater., vol. 2, no.

4, pp. 469–476, 2012.

[90] L. An, S. A. Boggs, and J. P. Calame, “Energy storage in polymer films with high dielectric

constant fillers [Feature Article],” IEEE Electr. Insul. Mag., vol. 24, no. 3, pp. 5–10, 2008.

[91] R. C. Buchanan, Ceramic materials for electronics, vol. 68. CRC press, 2004.

[92] D. Barrow, T. Petroff, R. Tandon, and M. Sayer, “Characterization of thick lead zirconate

titanate films fabricated using a new sol gel based process,” J. Appl. Phys., vol. 81, no. 2, pp.

876–881, 1997.

[93] K. Lichtenecker, “Dielectric constant of artificial and natural mixtures,” Phys Z, vol. 27, pp.

115–158, 1926.

[94] J. Bendler and T. Takekoshi, “Molecular modeling of polymers for high energy storage

capacitor applications,” in Power Sources Symposium, 1992., IEEE 35th International, 1992,

pp. 373–376.

[95] J. Lomax et al., “Electrical properties of BaTiO3 nanoparticles in poly (ether imide),” Smart

Mater. Struct., vol. 21, no. 8, p. 85017, 2012.

[96] Y. Thakur et al., “Enhancement of the Dielectric Response in Polymer Nanocomposites with

Low dielectric Constant Fillers,” Nanoscale.

[97] T. Tanaka, M. Kozako, N. Fuse, and Y. Ohki, “Proposal of a multi-core model for polymer

nanocomposite dielectrics,” Dielectr. Electr. Insul. IEEE Trans. On, vol. 12, no. 4, pp. 669–

681, 2005.

[98] S. Zhang et al., “Semicrystalline polymers with high dielectric constant, melting temperature,

and charge-discharge efficiency,” Dielectr. Electr. Insul. IEEE Trans. On, vol. 19, no. 4, pp.

1158–1166, 2012.

[99] L. A. Dissado and J. C. Fothergill, Electrical degradation and breakdown in polymers. IET,

1992.

[100] J. J. O’Dwyer, The theory of dielectric breakdown of solids. Clarendon Press, 1964.

[101] A. Von Hippel and G. Lee, “Scattering, trapping, and release of electrons in NaCl and in

mixed crystals of NaCl and AgCl,” Phys. Rev., vol. 59, no. 10, p. 824, 1941.

[102] H. Fröhlich, “Electric breakdown of ionic crystals,” Phys. Rev., vol. 61, no. 3–4, p. 200,

1942.

150

[103] F. Seitz, “On the theory of electron multiplication in crystals,” Phys. Rev., vol. 76, no. 9, p.

1376, 1949.

[104] K. W. Wagner, “The physical nature of the electrical breakdown of solid dielectrics,” J.

Am. Inst. Electr. Eng., vol. 41, no. 12, pp. 1034–1044, 1922.

[105] K. Stark and C. Garton, “Electric strength of irradiated polythene,” Nature, vol. 176, no.

4495, pp. 1225–1226, 1955.

[106] M. Ieda, “Dielectric breakdown process of polymers,” IEEE Trans. Electr. Insul., no. 3, pp.

206–224, 1980.

[107] H. Frohlich, “On the theory of dielectric breakdown in solids,” in Proceedings of the Royal

Society of London A: Mathematical, Physical and Engineering Sciences, 1947, vol. 188, pp.

521–532.

[108] Q. Burlingame, S. Wu, M. Lin, and Q. Zhang, “Conduction Mechanisms and Structure–

Property Relationships in High Energy Density Aromatic Polythiourea Dielectric Films,”

Adv. Energy Mater., vol. 3, no. 8, pp. 1051–1055, 2013.

[109] V. Agarwal and V. Srivastava, “Thickness dependence of breakdown field in thin films,”

Thin Solid Films, vol. 8, no. 5, pp. 377–381, 1971.

[110] C. Lhymn, P. Kosel, and R. Vaughan, “Thickness dependence of dielectric breakdown

voltage,” Thin Solid Films, vol. 145, no. 1, pp. 69–74, 1986.

[111] S. Cygan and J. Laghari, “Dependence of the electric strength on thickness area and volume

of polypropylene,” IEEE Trans. Electr. Insul., no. 6, pp. 835–837, 1987.

[112] G. A. Schneider, “A Griffith type energy release rate model for dielectric breakdown under

space charge limited conductivity,” J. Mech. Phys. Solids, vol. 61, no. 1, pp. 78–90, 2013.

[113] B. Helgee and P. Bjellheim, “Electric breakdown strength of aromatic polymers:

dependence on film thickness and chemical structure,” Electr. Insul. IEEE Trans. On, vol.

26, no. 6, pp. 1147–1152, 1991.

[114] F. Forlani and N. Minnaja, “Electrical breakdown in thin dielectric films,” J. Vac. Sci.

Technol., vol. 6, no. 4, pp. 518–526, 1969.

[115] R. Gerson and T. C. Marshall, “Dielectric breakdown of porous ceramics,” J. Appl. Phys.,

vol. 30, no. 11, pp. 1650–1653, 1959.

[116] M. Mittleman and R. Von Holdt, “Theory of low-energy-electron scattering by polar

molecules,” Phys. Rev., vol. 140, no. 3A, p. A726, 1965.

[117] J.-M. Lévy-Leblond, “Electron capture by polar molecules,” Phys. Rev., vol. 153, no. 1, p.

1, 1967.

[118] J. Ugalde and C. Sarasola, “Bound electronic states in a statically screened electric-dipole

potential,” Phys. Rev. A, vol. 54, no. 4, p. 2868, 1996.

[119] C. Lü, Z. Cui, C. Guan, J. Guan, B. Yang, and J. Shen, “Research on preparation, structure

and properties of TiO2/polythiourethane hybrid optical films with high refractive index,”

Macromol. Mater. Eng., vol. 288, no. 9, pp. 717–723, 2003.

[120] Q. Zhang, V. Bharti, and G. Kavarnos, “Poly (vinylidene fluoride)(PVDF) and its

copolymers,” Encycl. Smart Mater., 2002.

[121] J. K. Nelson and others, Dielectric polymer nanocomposites. Springer, 2010.

[122] F. Tian, Q. Lei, X. Wang, and Y. Wang, “Investigation of electrical properties of

LDPE/ZnO nanocomposite dielectrics,” Dielectr. Electr. Insul. IEEE Trans. On, vol. 19, no.

3, pp. 763–769, 2012.

[123] A. T. Hoang, L. Pallon, D. Liu, Y. V. Serdyuk, S. M. Gubanski, and U. W. Gedde, “Charge

transport in LDPE nanocomposites Part I—Experimental approach,” Polymers, vol. 8, no. 3,

p. 87, 2016.

[124] E. H. Rhoderick and R. Williams, Metal-semiconductor contacts. Clarendon Press Oxford,

1978.

151

[125] P. Emtage and W. Tantraporn, “Schottky emission through thin insulating films,” Phys.

Rev. Lett., vol. 8, no. 7, p. 267, 1962.

[126] S. M. Sze and K. K. Ng, Physics of semiconductor devices. John wiley & sons, 2006.

[127] Y. Wang, “The application and characterization of dielectric and piezoelectric polymer

devices,” The Pennsylvania State University, 2009.

[128] M. Lenzlinger and E. Snow, “Fowler-Nordheim tunneling into thermally grown SiO2,” J.

Appl. Phys., vol. 40, no. 1, pp. 278–283, 1969.

[129] M. A. Lampert, “Simplified theory of space-charge-limited currents in an insulator with

traps,” Phys. Rev., vol. 103, no. 6, p. 1648, 1956.

[130] M. A. Lampert and P. Mark, “Current injection in solids,” 1970.

[131] R. Zallen, The physics of amorphous solids. John Wiley & Sons, 2008.

[132] J. Brunson, “Hopping conductivity and charge transport in low density polyethylene,”

2010.

[133] G. Deutscher, R. Zallen, and J. Adler, “Percolation structures and processes,” Princet. Ser.

Phys., 1983.

[134] D. Das-Gupta, “Conduction mechanisms and high-field effects in synthetic insulating

polymers,” IEEE Trans. Dielectr. Electr. Insul., vol. 4, no. 2, pp. 149–156, 1997.

[135] A. Hunt, “New developments in the theory of the hopping conductivity of spatially random

systems,” J. Phys. Condens. Matter, vol. 6, no. 6, p. 1239, 1994.

[136] M. H. Lean and W.-P. L. Chu, “Dynamic charge mapping in layered polymer films,”

Dielectr. Electr. Insul. IEEE Trans. On, vol. 21, no. 3, pp. 1319–1329, 2014.

[137] M. H. Lean and W.-P. L. Chu, “Simulation of charge packet formation in layered polymer

film,” COMPEL Int. J. Comput. Math. Electr. Electron. Eng., vol. 33, no. 4, pp. 1396–1415,

2014.

[138] M. H. Lean and W.-P. L. Chu, “Effect of gaseous void on bipolar charge transport in

layered polymer film,” J. Phys. Appl. Phys., vol. 47, no. 7, p. 75303, 2014.

[139] M. H. Lean and W.-P. L. Chu, “Effective permittivity of nanocomposites from 3D charge

transport simulations,” J. Appl. Polym. Sci., vol. 133, no. 15, 2016.

[140] M. H. Lean and W.-P. L. Chu, “Modeling effect of ferroelectric nanofiller size on bipolar

charge transport in nanocomposite film,” J. Polym. Sci. Part B Polym. Phys., vol. 53, no. 19,

pp. 1380–1390, 2015.

[141] M. H. Lean and W.-P. L. Chu, “Effect of degree of crystallinity on charge transport in

ferroelectric semicrystalline polymer film,” IEEE Trans. Dielectr. Electr. Insul., vol. 22, no.

3, pp. 1444–1454, 2015.

[142] Y. Thakur, M. H. Lean, and Q. Zhang, “Reducing conduction losses in high energy density

polymer using nanocomposites,” Appl. Phys. Lett., vol. 110, no. 12, p. 122905, 2017.

[143] T. Tanaka, “Dielectric nanocomposites with insulating properties,” Dielectr. Electr. Insul.

IEEE Trans. On, vol. 12, no. 5, pp. 914–928, 2005.

[144] Y. Thakur, T. Zhang, M. Lin, Q. Zhang, and M. H. Lean, “Mitigation of conduction loss in

a semi-crystalline polymer with high dielectric constant and high charge-discharge

efficiency,” in Dielectrics (ICD), 2016 IEEE International Conference on, 2016, vol. 1, pp.

59–63.

[145] T. Zhou, J.-W. Zha, R.-Y. Cui, B.-H. Fan, J.-K. Yuan, and Z.-M. Dang, “Improving

dielectric properties of BaTiO3/ferroelectric polymer composites by employing surface

hydroxylated BaTiO3 nanoparticles,” ACS Appl. Mater. Interfaces, vol. 3, no. 7, pp. 2184–

2188, 2011.

[146] V. KumaráThakur, E. JináTan, P. SeeáLee, and others, “Surface functionalization of BaTiO

3 nanoparticles and improved electrical properties of BaTiO 3/polyvinylidene fluoride

composite,” RSC Adv., vol. 1, no. 4, pp. 576–578, 2011.

152

[147] C. Roland and R. Casalini, “Effect of hydrostatic pressure on the viscoelastic response of

polyurea,” Polymer, vol. 48, no. 19, pp. 5747–5752, 2007.

[148] G. Williams, “Effect of pressure on the dielectric β-relaxation in polyethylene

terephthalate,” Trans. Faraday Soc., vol. 62, pp. 1321–1328, 1966.

[149] S. Diaham, F. Saysouk, M. Locatelli, and T. Lebey, “Huge nanodielectric effects in

polyimide/boron nitride nanocomposites revealed by the nanofiller size,” J. Phys. Appl.

Phys., vol. 48, no. 38, p. 385301, 2015.

[150] L.-Q. Chen, “Phase-field models for microstructure evolution,” Annu. Rev. Mater. Res., vol.

32, no. 1, pp. 113–140, 2002.

[151] T. Yang, J.-M. Hu, C. Nan, and L. Chen, “Predicting effective magnetoelectric response in

magnetic-ferroelectric composites via phase-field modeling,” Appl. Phys. Lett., vol. 104, no.

5, p. 52904, 2014.

153

VITA

YASH THAKUR

Yash Thakur was born in Lucknow on January 1991 and was raised in Slapper,

Himachal Pradesh, India. In 2012, he received his bachelor of engineering degree from

University Institute of Engineering and Technology – Panjab University, Chandigarh,

India. He joined the Electrical Engineering at The Pennsylvania State University for his

graduate studies in Fall 2012, and worked on nanocrystalline quantum dots for photovoltaic

applications in Prof. Jerzy Ruzyllo’s group for his master’s research. He joined Prof.

Qiming Zhang’s group in Summer 2014 with research focus on dielectric polymers for

high energy density capacitors. He has won IEEE DEIS Fellowship in the academic year

2015-16, a prestigious award given annually to three students in the world in the area of

dielectrics and electrical insulation. He has also been awarded the Melvin P. Bloom

memorial outstanding doctoral research award in Electrical Engineering department at

Penn State. He has authored 12 journal papers and has a provisional US patent. He is an

active member of IEEE, DEIS and MRS.