oxidative and initiated chemical vapor deposition of

138
i Oxidative and Initiated Chemical Vapor Deposition of Polymer Electronic Materials for Applications in Energy Conversion and Storage A Thesis Submitted to the Faculty of Drexel University By Siamak Nejati In Partial Fulfilment of the Requirements for the Degree Of Doctor of Philosophy April 2013

Upload: others

Post on 11-Feb-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

i

Oxidative and Initiated Chemical Vapor Deposition of Polymer Electronic Materials

for Applications in Energy Conversion and Storage

A Thesis

Submitted to the Faculty

of

Drexel University

By

Siamak Nejati

In Partial Fulfilment of the

Requirements for the

Degree

Of

Doctor of Philosophy

April 2013

ii

© Copyright 2013

Siamak Nejati. All Rights Reserved.

iii

DEDICATION

To Mona for her true love and inspiration

iv

ACKNOWLEDGEMENTS

My first and foremost acknowledgement should and must go to my advisor, Dr. Kenneth Lau, a

patient man who walked me through my path in the past few years. His endless source of support

and wisdom inculcated in me a thirst for exploring innovative ideas. I was so lucky to work with

him and he will be a memorable figure in my life. I owe any measure of success I have so far

achieved to him. His kindness was above all. Not only I learned the path to be a good scientist

from him but also he taught me how to educate and teach. I am also grateful for the support and

guidance of my thesis committee, Dr. Jason Baxter, professor Masoud Soroush, Dr. Caroline

Schauer and Dr. Jonathan Spanier, who have all made measurable impacts on my development

during my time at Drexel. I also wish to express my sincere gratitude to professor Palmese, who

has generously offered help whenever I asked for and supported me financially and emotionally

when I find myself struggling with illness.

I would like to thank my other colleague at Drexel University who shared their knowledge with

me as well as hands-on training in lab. My past and present group mates, Dr. Ranjita Bose,

Stefani Susilo, Zakiya Carter, Arpit Patel, Danielle Martine, Laura Wu, Jahnavi Deshmukh,

Thomas Minford, Gregoty Wallowitch, Devin Cody, Sajeewa Ranasinghe, Noah Watson, Yuriy

Smolin and Chia-Yun Hsieh. I was also lucky enough to know many people from other

departments and groups. I would like to thank, Professor Anthony Addison, Dr. Volker Presser,

Dr. Edward Basgall, Timothy Wade, Dr. Stephen Dicker, Jared Cole and Elizabeth Plowman for

helping me in the lab with hands-on training. I am heartily grateful to many former and current

graduate students at Drexel University for their continuous help and support: Dr. Harsh Sharma,

v

Dr. Ali Emileh, Dr. Micheal Walter, Dr. Yuesheng Ye, Arianna Waters, James Throckmorton,

Taha Mohseni, Glenn Guglietta, Majid Sharifi, Min Heon, Eric Laird and Steven Spurgeon. To

all of my friends and colleagues your kindness will never be forgotten.

I would like to acknowledge the past and present staff at Chemical and Biological Engineering

department for their continuous help and support during my time at Drexel. Dan Luu, Dorothy

Gould, Jennifer Bing, Katie Brumbelow and Tracy McClure , thank you for all your help and

guidance. I am also grateful to the help and support provided by the Centralized Research Facility

at Drexel University and its core staff, Ms. Sahar Javedani and Craig Johnosn.

Finally, my path has led me thus far only with the support of my caring parents Mostafa Nejati,

Behrouz Bavarian, Roohangiz Beigi and Fatemeh Baradaran, my brothers Babak Nejati, Sina

Nejati and Mohamad Bavarian, my sisters Solmaz Alipour, Maryam Bavarian and Sara Bavarian

and my wife, Mona Bavarian. Mona was not only a great source of encouragement for me but

also she was a wise, knowledgeable colleague and collaborator who contributed to the solar cell

project with her sophisticated and smart approach in mathematical modelling of the system. It is

with you that I want to share this achievement.

vi

Table of Contents

List of Tables ................................................................................................................................... x

List of Figures ................................................................................................................................ xi

Abstract ...................................................................................................................................... xvii

1 Chapter 1: Introduction and Background ........................................................................... 1

1.1 Polymer synthesis using chemical vapor deposition ...................................................... 1

1.2 Initiated chemical vapor deposition ............................................................................... 2

1.3 iCVD polymer electrolytes for solar cell application..................................................... 4

1.4 Oxidative chemical vapor deposition ............................................................................. 5

1.5 Redox polymers for supercapacitor application ............................................................. 7

1.6 References ...................................................................................................................... 8

2 Chapter 2: Overall Objective and Specific Aims .............................................................. 15

Specific Aim 1: Synthesis of polymer electronic materials ...................................................... 15

Specific Aim 2: Integration of polymer electrolytes in dye sensitized solar cells .................... 15

Specific Aim 3: Integration of conducting polymers in supercapacitors .................................. 16

3 Chapter 3: Synthesis of Polymer Electronic Materials .................................................... 17

3.1 Introduction .................................................................................................................. 17

3.2 Poly(2-hydroxyethylmethacrylate) as a polymer electrolyte ....................................... 17

3.3 Experimental ................................................................................................................ 18

3.3.1 Polymer synthesis ................................................................................................. 18

3.3.2 Polymer characterization ...................................................................................... 18

3.4 Results and discussion ................................................................................................. 19

3.4.1 Fourier transform infrared spectroscopy .............................................................. 19

vii

3.4.2 X-ray photoelectron spectroscopy ........................................................................ 20

3.4.3 Gel permeation chromatography .......................................................................... 22

3.4.4 Electrochemical impedance spectroscopy ............................................................ 23

3.5 Unsubstituted polythiophene as a redox-active polymer ............................................. 25

3.6 Experimental ................................................................................................................ 27

3.6.1 Polymer deposition ............................................................................................... 27

3.6.2 Characterization ................................................................................................... 28

3.7 Results and discussion ................................................................................................. 29

3.7.1 Vibrational spectroscopy ...................................................................................... 30

3.7.2 X-ray photoelectron spectroscopy ........................................................................ 38

3.7.3 UV-vis spectroscopy ............................................................................................ 43

3.7.4 Cyclic voltammetry .............................................................................................. 45

3.7.5 Effect of oCVD processing conditions on polymer properties............................. 47

3.8 Conclusions .................................................................................................................. 51

3.9 References .................................................................................................................... 52

4 Chapter 4: Integration of Polymer Electrolytes in Dye Sensitized Solar Cells .............. 56

4.1 Introduction .................................................................................................................. 56

4.1.1 DSSC working principle ...................................................................................... 57

4.1.2 Replacing the liquid electrolyte ............................................................................ 59

4.1.3 Pore filling methodology ...................................................................................... 59

4.2 Experimental ................................................................................................................ 61

4.2.1 Electrode preparation ........................................................................................... 61

4.2.2 Pore filling ............................................................................................................ 61

4.2.3 Pore filing estimation ........................................................................................... 62

4.2.4 Dye sensitized solar cell fabrication ..................................................................... 63

viii

4.2.5 Solar cell characterization .................................................................................... 63

4.3 Results and discussion ................................................................................................. 64

4.3.1 TiO2 electrode porosity and surface area ............................................................. 64

4.3.2 Effect of process parameters on pore filling ......................................................... 65

4.3.3 Understanding pore filling dynamics ................................................................... 68

4.3.4 Pore filling quality ................................................................................................ 72

4.3.5 Polymer integrated DSSC performance ............................................................... 74

4.3.6 Electrochemical impedance spectroscopy ............................................................ 81

4.4 Conclusions .................................................................................................................. 83

4.5 References .................................................................................................................... 83

5 Integration of conducting polymers in supercapacitors ................................................... 87

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

5.2 Supercapacitors ............................................................................................................ 88

5.3 Experimental ................................................................................................................ 89

5.3.1 Polymer synthesis ................................................................................................. 89

5.3.2 Polymer characterization ...................................................................................... 90

5.3.3 TiO2 electrode fabrication .................................................................................... 92

5.3.4 Activated carbon electrode fabrication ................................................................. 92

5.3.5 Current collectors ................................................................................................. 93

5.3.6 oCVD polythiophene integration ......................................................................... 93

5.3.7 Polymer weight measurement .............................................................................. 94

5.4 Results and discussion ................................................................................................. 94

5.4.1 Vibrational spectroscopy ...................................................................................... 95

5.4.2 UV-vis spectroscopy ............................................................................................ 98

5.4.3 Cyclic voltammetry of the deposited film on FTO .............................................. 99

ix

5.4.4 Polymer integration within nanostructures ........................................................... 99

5.4.5 Effect of nanostructure on charge storage .......................................................... 102

5.4.6 Polymer integrated supercapacitors .................................................................... 108

5.5 Conclusions ................................................................................................................ 111

5.6 References .................................................................................................................. 111

6 Chapter 6: Conclusions and Future Directions .............................................................. 115

6.1 Specific aim 1............................................................................................................. 115

6.2 Specific aim 2............................................................................................................. 116

6.3 Specific aim 3............................................................................................................. 117

6.4 List of publications..................................................................................................... 118

x

List of Tables

Table ‎3.1. XPS elemental analysis of oCVD polythiophene and standard oligothiophenes. ....... 41

Table ‎3.2. Effect of oCVD synthesis conditions on polythiophene properties.Sample run.......... 49

xi

List of Figures

Figure ‎1.1. Reaction scheme and mechanism proposed for iCVD polymerization. ..................................... 3

Figure ‎1.2. iCVD reactor chamber. .............................................................................................................. 3

Figure ‎1.3. oCVD reactor showing the oxidant (blue) and monomer (red) vapor flows. ............................. 6

Figure ‎3.1. FTIR spectra of HEMA monomer, standard PHEMA and iCVD PHEMA. ............................ 20

Figure ‎3.2. XPS survey of 100 nm PHEMA deposited on Si (001) wafer. The C1s and O1s peak

intensities were used to evaluate the ratio of C:O. The peak around 1000 eV is the Auger peak of

oxygen. .......................................................................................................................................... 21

Figure ‎3.3. High resolution XPS signals of (a) oxygen 1s, and (b) carbon 1s............................................ 21

Figure ‎3.4. Refractive index detector intensity as a function of elution time in gel permeation

chromatography for (a) PEO calibration standards, and (b) iCVD PHEMA (A and B are deposited

using 0.2 and 0.4 sccm flow of initiator, respectively. .................................................................. 22

Figure ‎3.5. The Nyquist plot for a polymer electrolyte sandwiched between two Pt-coated FTO glass. The

data was fitted to a Randles equivalent circuit that included a Warburg impedance to derive the

conductivity and diffusion coefficient. .......................................................................................... 24

Figure ‎3.6. Conductivity of iCVD PHEMA incorporating different dielectric additives containing 0.05 M

I2 and 0.5 M LiI as a redox couple. ............................................................................................... 25

Figure ‎3.7. FTIR spectra of as-deposited and washed polythiophene. The doping induced vibrational

bands in the 1100-1500 cm-1

region for the as-deposited film disappear after washing as a result

of dedoping. ................................................................................................................................... 31

Figure ‎3.8. The (a) calculated and (b) experimental FTIR spectra of thiophene. ....................................... 33

Figure ‎3.9. The (a) calculated and (b) experimental FTIR spectra of bithiophene. .................................... 33

Figure ‎3.10. The (a) calculated and (b) experimental FTIR spectra of terthiophene. ................................ 34

Figure ‎3.11. The (a) calculated and (b) experimental FTIR spectra of quarterthiophene. ......................... 34

xii

Figure ‎3.12. Raman spectra of as-deposited and washed polythiophene. The spectra can be divided into

three main regions in the range of 1350-1500, 1000-1250 and 650-740 cm-1

. Insets magnify the

latter two regions of lower intensity. Washing results in narrowing of the peaks in the highest

intensity region as a result of a loss of the quinoid vibration. ....................................................... 36

Figure ‎3.13. Raman spectral fitting in the 1300-1500 cm-1

region of (a) as-deposited and (b) washed. .... 37

Figure ‎3.14. Raman spectral fitting in the (a) 1000-1250 and (b) 620-740 cm-1

regions of as-deposited

PTh. ............................................................................................................................................... 38

Figure ‎3.15. High resolution S2p XPS spectra of (a) washed and (b) as-deposited PTh. .......................... 39

Figure ‎3.16. High resolution Cl2p XPS spectra of (a) washed and (b) as-deposited PTh. ......................... 40

Figure ‎3.17. XPS survey spectra of washed polythiophene with C60 sputter depth profiling.

Compositional analysis yields a C:S ratio of 4.3:1, which is close to the theoretical value of 4:1. A

small amount of oxygen present (2.1 at%) is attributed to adsorbed oxygen species. Sputter depth

profiling using Ar+

shows the dopant has been effectively removed with the washing. The

maximum depth of acquisition is 10 nm and the spectra are separated equally by 2nm. .............. 41

Figure ‎3.18. High resolution (a) C1s and (b) S2p XPS spectra of washed polythiophene. Peak fits reveal

the expected 1:1 ratio of α and‎β‎carbons‎as‎well‎as‎a‎single‎sulfur‎species‎of‎the‎thiophene‎ring.

....................................................................................................................................................... 42

Figure ‎3.19. UV-vis spectra of as-deposited and washed polythiophene. The as-deposited film shows the

characteristic broad polaron absorption of the doped form while the dedoped film after washing

gives a single absorption peak at λmax = 470 nm of the‎π-π*‎transition. ........................................ 44

Figure ‎3.20. UV-vis spectrum of the dissolved portion of PTh. ................................................................. 44

Figure ‎3.21. Correlation of λmax with the number of thiophene chain units, comparing literature data and

experimental data of undoped standard oligothiophenes with washed polythiophene and its

dissolved soluble fraction after washing. The correlation indicates a longer conjugation length

xiii

with a higher degree of polymerization. The correlation also indicates that the solid portion of the

washed film is much longer than 6-7 thiophene units while the dissolved fraction consists mainly

of 5 repeat units. ............................................................................................................................ 45

Figure ‎3.22. Cyclic voltammogram of washed polythiophene in a supporting electrolyte solution of

acetonitrile and tetraethylammonium hexafluorophosphate. The anodic and cathodic peaks are

observed at 1.1 and 0.6 V vs SCE, respectively. The corresponding film color in the oxidized and

reduced forms are also shown. ...................................................................................................... 46

Figure ‎3.23. Cyclic voltammetry sweep of PTh. ........................................................................................ 47

Figure 4.1. Dye sensitized solar cell layout. .............................................................................................. 57

Figure ‎4.2. Schematic of the pore filling process during initiated chemical vapor deposition (iCVD).

Monomer (M) and initiator (I) molecules are delivered into the reaction chamber in the form of

gaseous vapors. The initiator is selectively activated by a series of heated filaments. The activated

initiator (R) and monomer (M) adsorb onto the TiO2 surface within the nanostructured

mesoporous electrode that is kept cooled to enhance the adsorption-limited process. Addition

polymerization of the monomer at activated initiator sites results in the formation of a growing

polymer inside the pores. ............................................................................................................... 60

Figure ‎4.3. Nitrogen adsorption/desorption isotherm of the mesoporous electrode. .................................. 64

Figure ‎4.4. Effect of total pressure on pore filling, showing cross sectional SEM of the 4 µm thick TiO2

electrodes that are (a) uncoated, and after iCVD treatment at a total pressure of (b) 60, (c) 125,

and (d) 200 mtorr (scale bar=100 nm). Complete pore filling is observed at intermediate

pressures. ....................................................................................................................................... 66

Figure ‎4.5. The individual effect of the initiator and monomer on pore filling of TiO2 electrodes, showing

lack of complete pore filling at initiator flow rates of (a) 0.2, and (b) 0.3 sccm, and at monomer

flow rates of (c) 0.2, and (d) 0.6 sccm (scale bar=200 nm). Complete pore filling is observed at

initiator flow rates of 0.4 sccm and higher, and at a monomer flow rate of 0.4 sccm. .................. 67

xiv

Figure ‎4.6. Estimated time constants for monomer diffusion and reaction as a function of the monomer

relative‎pressure‎(solid‎line‎is‎τD and‎dashed‎line‎is‎τR)................................................................. 71

Figure ‎4.7. Pore filling of TiO2 electrodes of different thicknesses of (a) 4 µm (scale bar=500 nm), and (b)

12 µm (scale bar=1 µm), the inset shows the bottom of the sample (scale bar=500 nm). By

carefully controlling mass transport (gas and surface diffusion)................................................... 72

Figure ‎4.8. Thermogravimetric‎analysis‎of‎the‎12.6‎μm‎thick‎pore‎filled‎titanium‎dioxide.‎Cross-sectional

SEM showed 97±10 nm thick polymer overlayer. ........................................................................ 73

Figure ‎4.9. The filled mesoporous of thickness L with δ polymer overlayer. ............................................ 74

Figure ‎4.10. Current-voltage characteristics of DSSCs fabricated with the polymer electrolyte containing

50:50‎vol%‎propylene‎carbonate‎and‎γ-butyrolactone‎(),‎and‎with‎the‎standard acetonitrile liquid

electrolyte‎()‎incorporating‎ruthenium‎dyes‎(a)‎535‎or‎N3,‎and‎(b)‎505. .................................... 75

Figure ‎4.11. Effect of redox solvent on the performance of DSSCs fabricated with the polymer electrolyte

()‎and‎the‎corresponding‎ liquid‎electrolyte‎()‎containing‎(a)‎50:50‎vol%‎propylene‎carbonate‎

and γ-butyrolactone, and (b) pure propylene carbonate. In each case, Voc is enhanced while Jsc

remains relatively unchanged in the polymer electrolyte cell. ...................................................... 77

Figure ‎4.12. Current-voltage characteristics of DSSCs containing 50:50 vol% propylene carbonate and γ-

butyrolactone‎pore‎ filled‎ polymer‎ electrolyte‎ (),‎ ‎ partially‎ pore‎ filled‎ polymer‎ electrolyte‎ (),‎

and‎liquid‎electrolyte‎(). .............................................................................................................. 77

Figure ‎4.13. Comparison between the performance of DSSCs fabricated with a quasi-solid state PHEMA

polymer electrolyte containing 50:50 vol% propylene carbonate and γ-butyrolactone‎(),‎and‎with‎

a‎standard‎liquid‎electrolyte‎containing‎acetonitrile‎()‎for‎different‎TiO2 thicknesses, showing (a)

open circuit voltage Voc, (b) short circuit current Jsc, (c) fill factor FF, and (d) power conversion

efficiency η (error bar=1 SD). Polymer electrolyte DSSCs provide similar efficiency as the liquid

electrolyte cells with TiO2 electrode thicknesses that are nearly three times thinner. ................... 79

xv

Figure ‎4.14. The effect of illuminated light intensity on short circuit current density of DSSCs utilizing

(a) liquid electrolyte, and (b) polymer electrolyte, each containing propylene carbonate with

different redox concentrations (in I2(M):LiI(M)) of 0.01:0.1 (∎), 0.08:0.8 (), and 0.2:2.0 ()

iodine and lithium iodide, respectively. ........................................................................................ 80

Figure ‎4.15. Comparison between the Bode diagrams at their respective Voc of DSSCs fabricated with the

polymer electrolyte‎ ()‎ and‎ the‎ corresponding‎ liquid‎ electrolyte‎ ()‎ containing‎ propylene‎

carbonate. The shift of the mid-frequency peak to lower frequency of the polymer electrolyte cell

indicates a decrease in charge recombination at the electrolyte-electrode interface. .................... 82

Figure ‎5.1. FTIR spectrum of an as-deposited doped film on silicon. ....................................................... 95

Figure ‎5.2. Raman spectra of PTh deposited on silicon, excited using (a) 488, and (b) 633 nm laser beams.

....................................................................................................................................................... 97

Figure ‎5.3. Raman spectra of polythiophene deposited at different oCVD condition. The shift in quinod

peak to a lower wavenumber is concomitant with an increase in film electrical conductivity. .... 97

Figure ‎5.4. UV-vis spectrum of an as-deposited film after dedoping the film to its undoped form (by

exposure to 2.0 M methylamine in methanol for 2 min followed by washing with neat methanol).

Inset shows the control over absorption spectrum by changing oCVD parameters, the increase in

conjugation length obtained by reducing the oxidant concentration. ............................................ 98

Figure ‎5.5. Cyclic voltammogram of an as-deposited film on FTO glass. ................................................. 99

Figure ‎5.6. Conformal coating of polythiophene within porous nanostructures using oCVD. (a) Uncoated

anodized aluminum oxide (AAO) membrane, 57 µm thick and 200 nm pore diameter. (b)

Polythiophene coated AAO in the as-deposited doped state of the polymer. (c) Polythiophene

coated AAO in the undoped state of the polymer after dedoping. (d) Cross-sectional SEM of an

AAO membrane showing the porous channels (darker shade). (e) Cross-sectional SEM of a

polythiophene coated AAO membrane showing conformal and uniform. The conditions used for

xvi

filling AAO without mass transfer limitations up to 40 nm thick film on the inner wall was as

follow: 30 °C stage temperature, and 0.1, 2 and 2 sccm of initiator, monomer and N2, and 800

mtorr total pressure. ..................................................................................................................... 101

Figure ‎5.7. Polythiophene integrated TiO2 electrodes. (a) Undoped PTh (by exposure to 2.0 M

methylamine in methanol for 2 min followed by washing with neat methanol). (b) Cross-

sectional SEM of the mesoporous electrode with conformal PTh coating (scale bar is 200 nm).

..................................................................................................................................................... 102

Figure ‎5.8. Cyclic voltammograms at different scan rates of PTh deposited within mesoporous TiO2

electrodes recorded in a three electrode set up versus an Ag/AgCl reference electrode. ............ 103

Figure ‎5.9. Effect of polymer thickness and 3D nanostructure on the specific capacitance of oCVD

polythiophene. (a) SEM images (scale bar is 200 nm) and specific capacitance values of thin (250

nm) and thick (800 nm) films on planar FTO electrodes, and of thin (4 nm) and thick (6 nm) films

inside mesoporous electrodes of TiO2; specific capacitance is based on per mass of polymer and

reported with 2 standard deviations. The current density of the anodic peak (doping peak) as a

function of scan rate for different thickness of films on (b) planar substrates and (c) within

mesoporous nanostructures. ........................................................................................................ 107

Figure ‎5.10. Electrochemical behavior of oCVD polythiophene coated activated carbon. SEM images of

activated carbon electrodes (a) without coating, and coated with polythiophene at polymer-to-

activated carbon mass ratios of (b) 1:1, (c) 1.5:1, and (d) 2.7:1 (insets show a magnified region of

the surface morphology of each sample; scale bar is 1 µm and 50 nm for the inset). (e) Specific

capacitance values based on per total mass of electrode for different polymer mass loadings.

Measurements were made at 100 mV/s for five different samples at each loading with the error

bars representing two standard deviations. (f) Cyclic voltammograms of the 1.5:1 mass ratio

pesudocapacitor recorded at different scan rates between –2 and 2 V. (g) Capacitance of the same

sample measured up to 5000 cycles (at 100 mV/s ). ................................................................... 110

xvii

Abstract

Oxidative and Initiated Chemical Vapor Deposition of Polymer Electronic Materials

for Applications in Energy Conversion and Storage

Siamak Nejati

Kenneth K. S. Lau Supervisor, Ph.D.

Initiated and oxidative chemical vapor deposition (iCVD/oCVD) are novel surface

polymerization techniques for the formation of polymer thin films. Taking advantage of a vacuum

environment, polymer CVD enables the synthesis of stoichiometric polymers that can be applied

on different surface geometries down to nanometer scale. In iCVD and oCVD, the surface

adsorption of reactive species is typically the limiting step, which means the polymer growth and

properties can be tuned by adjusting the surface availability of the reactants. In this work, iCVD

and oCVD were utilized for the synthesis and integration of electronic polymer materials in the

nanostructured electrodes of energy conversion and storage devices. iCVD was used to synthesize

poly(2-hydroxyethyl methacrylate) (PHEMA) as a potential polymer electrolyte while oCVD was

used to make unsubstituted polythiophene (PTh) as a potential electrically conducting polymer.

PHEMA was then integrated within the mesoporous electrode of dye sensitized solar cells and

iCVD conditions that enabled successful pore filling were identified. The resulting devices

fabricated with iCVD PHEMA polymer electrolyte showed superior performance when compared

to their liquid counterparts. PTh was integrated in the porous electrode of activated carbon

supercapacitors and oCVD conditions that enabled conformal ultrathin coating were found. The

resulting devices with oCVD PTh was found to have significantly higher charge storage capacity

as a result of the synergistic effect of the redox polymer and the nanostructured topology. Overall,

this work demonstrated the viability of iCVD and oCVD pathways for the design, synthesis, and

xviii

processing of polymers in nanostructured architectures for energy applications. Success

ultimately depended on a clear understanding of the fundamental kinetic and transport

mechanisms for enabling polymer integration.

1

1 Chapter 1: Introduction and Background

1.1 Polymer synthesis using chemical vapor deposition

Polymers, the building blocks of many things around us, are usually synthesized through

different chemistries in a liquid media. Whether the polymer is synthesized through step or chain

polymerization, the processing of the polymer requires additional processing steps. Except for

in-situ polymerization which is not always an easy process to undertake in the liquid phase, to

use the resulting polymers as an active component of devices spray coating, spin coating and

solvent casting are among the most frequently used methods.1 All these methods utilize polymer

solutions which necessitates the use of solvents. In many cases solvent entrapment in the final

product or device that benefit from the polymeric material is detrimental to product performance.

Also for many applications integrating polymers within high aspect ratio geometries is needed.

In this case, using conventional liquid processing methods might be difficult due to surface

tension forces, lack of wettability, and solution viscosity that tend to lead to partial blockage of

the structure and non-conformal coating through the structure.

The ability to integrate polymers within confined geometries and nanoscale domains,2-5

such

as in the electrodes6 of solar cells and organic light emitting diodes (OLEDs), is expected to

enhance interfacial properties that could lead to improved device performance. Chemical vapor

deposition (CVD) is a viable approach for the synthesis and integration of polymers that

combines polymerization and coating in a single step at an interface. This way the polymer is

being directly applied to the substrate of interest. Using CVD the main challenges associated

with polymer integration can be addressed as the process is solvent free and the polymerization

is done in-situ and under vacuum conditions. There are two novel CVD processes that have

demonstrated suitability for polymer synthesis and coating which have been studied in this

2

thesis, namely initiated chemical vapour deposition (iCVD) and oxidative chemical vapour

deposition (oCVD).

1.2 Initiated chemical vapor deposition

In iCVD, as shown in Figure 1.1, a gaseous mixture of initiator and monomer vapors is

introduced into a vacuum chamber. The initiator is activated in the gas phase by a resistively

heated filament suspended 1-3 cm above a cooled substrate. The filament temperature is mild

enough that the monomer is not thermally degraded. The monomer and the activated initiator

(typically a primary free radical species) then adsorb onto the substrate. The monomer adds to

the activated initiator sites to initiate and propagate polymer chain radicals. Finally, these

polymer radical chains terminate leading to the formation of a polymer thin film.7 Figure 1.2

shows an iCVD reactor in our facility. iCVD is suitable technique for depositing polymer films

on heat sensitive substrates as the typical substrate temperature is around room temperature. In

typical iCVD synthesis conditions, the process is adsorption limited as evidenced by the faster

polymer growth rate with lower substrate temperatures. When using free radical initiation, the

mechanism of iCVD follows a radical polymerization scheme that has been confirmed by kinetic

modeling.8 In iCVD reactant vapors are metered through needle valves or mass flow controllers

via a heated delivery manifold upstream of the reactor. A vacuum pump downstream creates the

vacuum and a downstream throttle valve along with a Baratron capacitance pressure gauge are

used together with a pressure controller to maintain a desired pressure. To control resulting

polymer properties, iCVD parameters including the gas phase concentration of the reactants,

which can be adjusted by the total pressure and flow rate, as well as the substrate temperature,

which determines surface adsorption, are among the major iCVD parameters to consider

3

Figure ‎1.1. Reaction scheme and mechanism proposed for iCVD polymerization.8

Figure ‎1.2. iCVD reactor chamber.

By adjusting these iCVD parameters carefully, iCVD can produce exceptionally clean

polymers with stoichiometric composition and tunable molecular weight with no residual

solvents or impurities, and importantly iCVD polymers can be integrated into complex

4

topological geometries. Since the birth of iCVD, there has been tremendous development in the

polymers that can be synthesized through iCVD, including polytetrafluoroethylene,9-13

acrylates,7,8

methacrylates,14-19

and styrenes.20,21

iCVD has also shown to be a promising method

for producing copolymers22-29

and crosslinked polymers26,30-38

by utilizing more than one

monomer or a crosslinker. Due to the wide range of polymer chemistries and functionalities that

can be made use of, iCVD has been demonstrated in many different applications such as

biomedical implants,39-42

controlled chemical release,25,30,42,43

sensors,29,44-46

antimicrobial,47-48

low surface energy,49-52

dielectric coatings,53,54

and polymer electrolytes and polyelectrolytes.55-57

As mentioned earlier, in iCVD due to the absence of liquids conformal coating at the micro and

nanoscale can be achieved easily. Attempts to coat complex 3D structures using iCVD has

shown great promise30,42,43,58

and there seems to be a new wave of polymer applications relying

on the engineering of polymer coatings within nanostructures.59-63

1.3 iCVD polymer electrolytes for solar cell application

Polymer electrolytes are attractive from the point of practical applications as they have decent

ion conductivity, are low cost, have good stability, and are easy to process. Polymer electrolytes

can be used as a dry solid or they can be gelled by trapping an organic liquid such as ethylene

carbonate (EC), propylene carbonate (PC) or acetonitrile (AcN) into the polymer matrix to

improve ion conductivity. The application of polymer electrolytes in lithium batteries,64

electrochromic devices,65-67

and especially in dyes sensitized solar cells (DSSCs)68-70

have been

of great interest and numerous studies have been devoted to integrating these polymers within

active electrode materials.

Here, iCVD is used for depositing poly(2-hydroxyethyl methacrylate) (PHEMA) as a

potential polymer gel electrolyte for DSSCs to replace the current liquid electrolyte used in the

5

cell. Previous work with poly(methyl methacrylate) (PMMA) has shown improved performance

for gel electrolyte base devices.64,65,70

The importance of this study is to evaluate the potential of

iCVD in the synthesis and integration of PHEMA polymer electrolyte into DSSCs and to assess

the resulting device performance.

1.4 Oxidative chemical vapor deposition

In oCVD, as shown in Figure 1.3, a gaseous mixture of monomer and oxidant is introduced into

a vacuum chamber.71,72

Unlike iCVD, there is no need to thermally activate the oxidant initiator

molecule. If chosen properly, the oxidant will spontaneously enable the oxidative polymerization

of a redox active monomer such as thiophene or pyrrole. The proposed mechanism is believed to

be the formation of radical cations similar to a Friedel-Crafts chemistry using oxidants that show

Lewis acid character like FeCl3.72

The radical cation act as a propagation unit and can go through

step polymerization by attaching to another monomer unit and transferring the cation to the new

unit or by reacting with another radical cation.73,74

Oxidative chemistry has been widely

practiced through chemical and electrochemical pathways in the liquid phase for synthesizing

conductive polymers.75

The wealth of synthesis knowledge and the current growing interest in conductive polymers

for different applications such as solar cells,76,77

light emitting diodes,78-80

supercapacitors,81-83

antistatic surfaces,84

and electrochromic devices65,85

present a great opportunity for oCVD

particularly with its unique capability to form thin conformal wrinkle-free coatings. The initial

attempt with CVD polymerization of conjugated polymers date back to 1986 when Mohamadi

and co-workers exposed a cooled substrate which was covered by an oxidant (FeCl3) to a vapor

of monomers (thiophene and pyrrole).86-88

Later Winther-Jensen and other researchers adapted a

6

similar method named vapor phase polymerization (VPP) and synthesized exceptionally high

conductivity polymer.89,90

In comparison with VPP, oCVD is a continuous method which relies on the delivery of the

reactants (both oxidant and monomer) as vapors instead of relying on an oxidant-coated

substrate. In a similar fashion to iCVD, reactant vapors are metered through needle valves or

mass flow controllers via a heated delivery manifold upstream of the reactor again a vacuum

pump downstream creates the vacuum and downstream throttle valve along with a Baratron

capacitance pressure gauge are used together with a pressure controller to maintain a constant

desired pressure. To tune polymer properties again oCVD synthesis conditions, including the

surface and gas phase concentrations of the reactants can be adjusted by changing vapor flow

rates of the components and the substrate temperature. By adjusting oCVD parameters carefully,

conjugated polymer with desirable properties can be integrated into different geometries and

since oCVD is typically operated at low pressure again using proper condition high aspect ratio

structures can be filled or coated as desired.

Figure ‎1.3. oCVD reactor showing the oxidant (blue) and monomer (red) vapor flows.

7

1.5 Redox polymers for supercapacitor application

Conjugated polymers have been widely investigated as energy storage materials, due to their

large pseudocapacitance that emanate from their ability to be p or n doped. The fact that upon

doping the material becomes conductive means that they can store charge in their whole volume,

which when compared to a physical electrochemical double layer capacitance should result in

higher charge storage capacity.

In the past 20 years a great deal of attention has been devoted to the design and fabrication

of supercapacitors that makes use of conductive polymers. Among different classes of

conjugated polymers, polythiophenes due to their environmental stability have been the subject

of many studies. Polythiophene, poly(3,4-ethylenedioxythiophene), polypyrrole, polyaniline and

composites of these materials have been evaluated in supercapacitor devices,81-83, 91-98

and very

recently the importance of nanostructuring the electrodes incorporating these polymers have

resulted in enhanced charge storage and stability.94-97

oCVD have shown to be a very promising method in the synthesis and coating of conjugated

polymers with control over coating down to the nanoscale.99,100

Here oCVD is used to synthesis

polythiophene using a Lewis acid oxidant to enable oxidative polymerization. By adjusting

oCVD conditions, polythiophene can be integrated within porous activated carbon electrodes to

create pseudocapacitors. The importance of this study is to understand the role of the

nanostructure on charge storage in redox-active polythiophene.

The thesis therefore continues with the following: Chapter 2 describes the overall objective

and specific aims of this thesis. Chapters 3, 4 and 5 describe the work in each of the three

specific aims in detail. Chapter 3 discusses the iCVD synthesis of PHEMA and the oCVD

synthesis of unsubstituted polythiophene as potential polymer electronic materials. Chapter 4

discusses the integration of iCVD polymer electrolytes within the nanostructured electrodes dye

8

sensitized solar cells and the impact of the polymer on device behavior. Chapter 5 discusses the

integration of unsubstituted polythiophene within the nanostructured electrodes of

supercapacitors and the impact of the polymer on device behavior, Finally, Chapter 6 gives the

overall conclusions and describes the unique contributions of this thesis as well as offers some

possible future directions stemming from this work.

1.6 References

1. L. Ferrante, Handbook of Applied Polymer Processing Technology, CRC Press, Harpers

Ferry, USA 1996.

2. S.-J. Moon, Y. Itzhaik, J. H. Yum, S. M. Zakeeruddin, G. Hodes and M. Grätzel,

Journal of Physical Chemistry Letters, 2010, 1, 1524-1527.

3. G. K. Mor, S. Kim, M. Paulose, O. K. Varghese, K. Shankar, J. Basham and C. A.

Grimes, Nano Letters, 2009, 9, 4250-4257.

4. D. C. Olson, Y. J. Lee, M. S. White, N. Kopidakis, S. E. Shaheen, D. S. Ginley, J. A.

Voigt and J. W. P. Hsu, Journal of Physical Chemistry C, 2007, 111, 16640-16645.

5. K. Shankar, G. K. Mor, H. E. Prakasam, O. K. Varghese and C. A. Grimes, Langmuir,

2007, 23, 12445-12449.

6. J. S. Miller, Advanced Materials, 1993, 5, 671-676.

7. K. K. S. Lau and K. K. Gleason, Macromolecules, 2006, 39, 3688-3694.

8. K. K. S. Lau and K. K. Gleason, Macromolecules, 2006, 39, 3695-3703.

9. B. A. Cruden, K. K. Gleason and H. H. Sawin, Journal of Vacuum Science &

Technology B, 2002, 20, 690-695.

10. E. D. Laird, R. K. Bose, W. Wang, K. K. S. Lau and C. Y. Li, Macromolecular Rapid

Communications, 2013, 34, 251-256.

9

11. K. K. S. Lau, J. Bico, K. B. K. Teo, M. Chhowalla, G. A. J. Amaratunga, W. I. Milne,

G. H. McKinley and K. K. Gleason, Nano Letters, 2003, 3, 1701-1705.

12. S. Limb, K. S. Lau, D. Edell, E. Gleason and K. Gleason, Plasmas and Polymers, 1999,

4, 21-32.

13. H. G. Pryce Lewis, J. A. Caulfield and K. K. Gleason, Langmuir, 2001, 17, 7652-7655.

14. R. Bakker, V. Verlaan, C. H. M. van der Werf, J. K. Rath, K. K. Gleason and R. E. I.

Schropp, Surface and Coatings Technology, 2007, 201, 9422-9425.

15. K. K. S. Lau and K. K. Gleason, Advanced Materials, 2006, 18, 1972-1977.

16. M. Ma, Y. Mao, M. Gupta, K. K. Gleason and G. C. Rutledge, Macromolecules, 2005,

38, 9742-9748.

17. Y. Mao, N. M. Felix, P. T. Nguyen, C. K. Ober and K. K. Gleason, Journal of Vacuum

Science & Technology B, 2004, 22, 2473-2478.

18. Y. Mao and K. K. Gleason, Langmuir, 2004, 20, 2484-2488.

19. Y. Mao and K. K. Gleason, Macromolecules, 2006, 39, 3895-3900.

20. W. E. Tenhaeff and K. K. Gleason, Surface and Coatings Technology, 2007, 201, 9417-

9421.

21. W. Jo, K. Freedman, D. K. Yi, R. K. Bose, K. K. S. Lau, S. D. Solomon, M. J. Kim,

Biofabrication, 2011, 3, 015002.

22. A. K. H. Achyuta, A. J. White, L. H. G. Pryce and S. K. Murthy, Macromolecules, 2009,

42, 1970-1978.

23. S. G. Im, K. W. Bong, C.H. Lee, P. S. Doyle and K. K. Gleason, Lab on a Chip, 2009, 9,

411-416.

24. S. G. Im, B.-S. Kim, W. E. Tenhaeff, P. T. Hammond and K. K. Gleason, Thin Solid

Films, 2009, 517, 3606-3611.

25. S. J. P. McInnes, E. J. Szili, S. A. Al Bataineh, J. Xu, M. E. Alf, K. K. Gleason, R. D.

Short and N. H. Voelcker, ACS Applied Materials & Interfaces, 2012, 4, 3566-3574.

10

26. V. Raghunathan, J. L. Yague, J. Xu, J. Michel, K. K. Gleason and L. C. Kimerling,

Optics Express, 2012, 20, 20808-20813.

27. J. Xu and K. K. Gleason, Chemistry of Materials, 2010, 22, 1732-1738.

28. R. Yang and K. K. Gleason, Langmuir, 2012, 28, 12266-12274.

29. Y. Ye, Y. Mao, H. Wang and Z. Ren, Journal of Materials Chemistry, 2012, 22, 2449-

2455.

30. R. K. Bose, A. M. Heming and K. K. S. Lau, Macromolecular Rapid Communications,

2012, 33, 1375-1380.

31. R. K. Bose and K. K. S. Lau, Biomacromolecules, 2010, 11, 2116-2122.

32. K. Chan and K. K. Gleason, Langmuir, 2005, 21, 8930-8939.

33. K. Chan, L. E. Kostun, W. E. Tenhaeff and K. K. Gleason, Polymer, 2006, 47, 6941-

6947.

34. A. M. Coclite, G. Ozaydin-Ince, R. d'Agostino and K. K. Gleason, Macromolecules,

2009, 42, 8138-8145.

35. Lau, K. K. S.; Mao, Y.; Lewis, H. G. P.; Murthy, S. K.; Olsen, B. D.; Loo, L. S.;

Gleason, K. K., Thin Solid Films 2006, 501, (1-2), 211-215.

36. N. Mari-Buye, S. O'Shaughnessy, C. Colominas, C. E. Semino, K. K. Gleason and S.

Borros, Advanced Functional Materials, 2009, 19, 1276-1286.

37. I. G. Ozaydin, G. Demirel, K. K. Gleason and M. C. Demirel, Soft Matter, 2010, 6,

1635-1639.

38. G. Ozaydin-Ince and K. K. Gleason, Chemical Vapor Deposition, 2010, 16, 100-105.

39. A. K. H. Achyuta, V. S. Polikov, A. J. White, H. G. P. Lewis and S. K. Murthy,

Macromolecular Bioscience, 2010, 10, 872-880.

40. W. S. O'Shaughnessy, D. J. Edell and K. K. Gleason, Thin Solid Films, 2009, 517, 3612-

3614.

11

41. W. S. O'Shaughnessy, S. K. Murthy, D. J. Edell and K. K. Gleason, Biomacromolecules,

2007, 8, 2564-2570.

42. K. K. S. Lau and K. K. Gleason, Surface and Coatings Technology, 2007, 201, 9189-

9194.

43. K. K. S. Lau and K. K. Gleason, Macromolecular Bioscience, 2007, 7, 429-434.

44. L. Montero, G. Gabriel, A. Guimera, R. Villa, K. K. Gleason and S. Borros, Vacuum,

2012, 86, 2102-2104.

45. C. D. Petruczok, S. Y. Yang, A. Asatekin, K. K. Gleason and G. Barbastathis,

Microelectromechanical Systems, 2013, 22,54-61

46. W. E. Tenhaeff, L. D. McIntosh and K. K. Gleason, Advanced Functional Materials,

2010, 20, 1144-1151.

47. T. P. Martin, K. Chan and K. K. Gleason, Thin Solid Films, 2008, 516, 681-683.

48. T. P. Martin, S. E. Kooi, S. H. Chang, K. L. Sedransk and K. K. Gleason, Biomaterials,

2006, 28, 909-915.

49. M. Gupta and K. K. Gleason, Langmuir, 2006, 22, 10047-10052.

50. M. Ma, Y. Mao, M. Gupta, K. K. Gleason and G. C. Rutledge, Macromolecules, 2005,

38, 9742-9748.

51. Y. Mao and K. K. Gleason, Macromolecules, 2006, 39, 3895-3900.

52. C. T. Riche,B. C. Marin, N. Malmstadt and M. Gupta, Lab on a Chip, 2011,11, 3049-

3052

53. R. Bakker, P. Weijers, D. A. Spee, S. M. J. van, d. W. C. H. M. van, J. K. Rath and R. E.

I. Schropp, Thin Solid Films, 2011, 519, 4418-4420.

54. N. J. Trujillo, Q. Wu and K. K. Gleason, Advanced Functional Materials, 2010, 20, 607-

616.

55. R. K. Bose, S. Nejati and K. K. S. Lau, ECS Transanctions, 2009, 25, 1229-1235.

56. A. M. Coclite, P. Lund, R. Di Mundo and F. Palumbo, Polymer, 2013, 54, 24-30.

12

57. S. Nejati and K. K. S. Lau, Thin Solid Films, 2011, 519, 4551-4554.

58. K. K. Lau, J. Bico, K. B. Teo, M. Chhowalla, G. A. Amaratunga, W. I. Milne, G. H.

McKinley and K. K. Gleason, Nano Letters, 2003, 3, 1701-1705.

59. A. Asatekin and K. K. Gleason, Nano Letters, 2010, 11, 677-686.

60. A. Asatekin and K. K. Gleason, ACS Symposium Series, 2011, 1078, 39-50.

61. M. Gupta, V. Kapur, N. M. Pinkerton and K. K. Gleason, Chemistry of Materials, 2008,

20, 1646-1651.

62. P. D. Haller, C. A. Flowers and M. Gupta, Soft Matter, 2011, 7, 2428-2432.

63. W. E. Tenhaeff and K. K. Gleason, Chemistry of Materials, 2009, 21, 4323-4331.

64. Y. H. Liao, D. Y. Zhou, M. M. Rao and W. S. Li, Dianchi Gongye, 2008, 13, 428-432.

65. A. A. Argun, A. Cirpan and J. R. Reynolds, Advanced Materials, 2003, 15, 1338-1341.

66. S. Y. Lin, C. M. Wang, P. T. Hsieh, Y. C. Chen, C. C. Liu and S. C. Shih, Progress in

Colloid and Polymer Science, 2009, 287, 1355-1358.

67. R. Sydam, M. Deepa and A. K. Srivastava, RSC Advances, 2012, 2, 9011-9021.

68. F. Bella, D. Pugliese, J. R. Nair, A. Sacco, S. Bianco, C. Gerbaldi, C. Barolo and R.

Bongiovanni, Physical Chemistry Chemical Physics, 2013, 15, 3706-3711.

69. B. Muthuraaman, G. Will, H. Wang, P. Moonie and J. Bell, Electrochimica Acta, 2013,

87, 526-531.

70. H. Yang, M. Huang, J. Wu, Z. Lan, S. Hao and J. Lin, Materials Chemistry and Physics,

2008, 110, 38-42.

71. S. G. Im and K. K. Gleason, Macromolecules, 2007, 40, 6552-6556.

72. J. P. Lock, S. G. Im and K. K. Gleason, Macromolecules, 2006, 39, 5326-5329.

73. J. Heinze, B. A. Frontana-Uribe and S. Ludwigs, Chemical Reviews, 2010, 110, 4724-

4771.

13

74. P. Kovacic and M. B. Jones, Chemical Reviews, 1987, 87, 357-379.

75. H. Higashimura and S. Kobayashi, in Encyclopedia of Polymer Science and

Technology, John Wiley & Sons, 2002.

76. S. Gunes, H. Neugebauer and N. S. Sariciftci, Chemical Reviews, 2007, 107, 1324-1338.

77. Y. Kim, S. Cook, S. M. Tuladhar, S. A. Choulis, J. Nelson, J. R. Durrant, D. D. Bradley,

M. Giles, I. McCulloch and C.S. Ha, Nature Materials, 2006, 5, 197-203.

78. Y. Cao, G. Yu, C. Zhang, R. Menon and A. Heeger, Synthetic Metals, 1997, 87, 171-

174.

79. D. Kabra, L. P. Lu, M. H. Song, H. J. Snaith and R. H. Friend, Advanced Materials,

2010, 22, 3194-3198.

80. Y. H. Kim, J. Lee, S. Hofmann, M. C. Gather, L. Müller-Meskamp and K. Leo,

Advanced Functional Materials, 2013.

81. L. Z. Fan, Y. S. Hu, J. Maier, P. Adelhelm, B. Smarsly and M. Antonietti, Advanced

Functional Materials, 2007, 17, 3083-3087.

82. A. Laforgue, P. Simon, C. Sarrazin and J. F. Fauvarque, Journal of Power Sources,

1999, 80, 142-148.

83. V. Khomenko, E. Frackowiak and F. Béguin, Electrochimica Acta, 2005, 50, 2499-

2506.

84. D. C. SUN and D. S. SUN, Polymer Materials Science and Engineering, 2009, 7, 030.

85. A. A. Argun, P.-H. Aubert, B. C. Thompson, I. Schwendeman, C. L. Gaupp, J. Hwang,

N. J. Pinto, D. B. Tanner, A. G. MacDiarmid and J. R. Reynolds, Chemistry of

Materials, 2004, 16, 4401-4412.

86. A. Mohammadi, M. A. Hasan, B. Liedberg, I. Lundström and W. Salaneck, Synthetic

Metals, 1986, 14, 189-197.

87. A. Mohammadi, I. Lundström, W. Salaneck and O. Inganäs, Synthetic Metals, 1987, 21,

169-173.

88. W. R. Salaneck, C. R. Wu, J. O. Nilsson and J. L. Brédas, Synthetic Metals, 1987, 21,

57-61.

14

89. M. Mueller, M. Fabretto, D. Evans, P. Hojati-Talemi, C. Gruber and P. Murphy,

Polymer, 2012, 53, 2146-2151.

90. B. Winther-Jensen, J. Chen, K. West and G. Wallace, Macromolecules, 2004, 37, 5930-

5935.

91. D. Y. Liu and J. R. Reynolds, ACS Applied Materials & Interfaces, 2010, 2, 3586-3593.

92. K. Lota, V. Khomenko and E. Frackowiak, Journal of Physics and Chemistry of Solids,

2004, 65, 295-301.

93. J. H. Park, J. M. Ko, O. O. Park and D.-W. Kim, Journal of Power Sources, 2002, 105,

20-25.

94. H. Wang, Q. Hao, X. Yang, L. Lu and X. Wang, Nanoscale, 2010, 2, 2164-2170.

95. J. Xu, K. Wang, S.-Z. Zu, B.-H. Han and Z. Wei, ACS Nano, 2010, 4, 5019-5026.

96. K. Zhang, L. L. Zhang, X. S. Zhao and J. Wu, Chemistry of Materials, 2010, 22, 1392-

1401.

97. M. Ertas, R. M. Walczak, R. K. Das, A. G. Rinzler and J. R. Reynolds, Chemistry of

Materials, 2012, 24, 433-443.

98. J. Zhang, L. B. Kong, B. Wang, Y. C. Luo and L. Kang, Synthetic Metals, 2009, 159,

260-266.

99. S. Vaddiraju, H. l. Cebeci, K. K. Gleason and B. L. Wardle, ACS Applied Materials &

Interfaces, 2009, 1, 2565-2572.

100. S. G. Im, D. Kusters, W. Choi, S. H. Baxamusa, M. C. M. van de Sanden and K. K.

Gleason, ACS Nano, 2008, 2, 1959-1967.

15

2 Chapter 2: Overall Objective and Specific Aims

The overall objective of this thesis is to study novel chemical vapor deposition pathways to

synthesize polymers and integrate them within nanostructured electrodes of energy conversion

and storage devices. To achieve this objective, the work has been carried out with the following

three specific aims.

Specific Aim 1: Synthesis of polymer electronic materials

The fundamental question here is whether iCVD and oCVD can be used to synthesize polymers

with properties suitable for optoelectronic applications. Our hypothesis is that a fundamental

understanding of each polymer CVD process enables us to produce polymer electronic materials

with the appropriate properties and function for the desired device application. To prove our

hypothesis, we synthesized PHEMA through iCVD and investigated the composition, structure

and ionic conductivity of the material. Likewise, we synthesized unsubstituted polythiophene by

oCVD and investigated the composition, structure, redox behavior and electrical conductivity of

the material.

Specific Aim 2: Integration of polymer electrolytes in dye sensitized solar cells

The fundamental question here is whether iCVD is a viable route for the integration of PHEMA

polymer electrolytes within nanostructured porous electrode of a dye sensitized solar cell. Our

hypothesis is that a fundamental understanding of the dynamic processes that govern iCVD

polymerization will enable us to identify the parameters that lead to the successful integration of

the polymer within these high aspect ratio structures and importantly enhance device

performance. To prove our hypothesis, we deposited PHEMA from different iCVD conditions

16

within the porous electrodes and then investigated the quality of pore filling and the performance

of the solar cell fabricated using the polymer integrated electrodes.

Specific Aim 3: Integration of conducting polymers in supercapacitors

The fundamental questions here are whether the unsubstituted polythiophene can be integrated

effectively within the nanostructured electrodes of supercapacitors to create suitable

pseudocapacitors. Our hypothesis is that a fundamental understanding of the oxidative

polymerization mechanism and CVD process behavior will enable us to choose the proper

oxidant and deposit unsubstituted polythiophene and more importantly the polymer will act as an

effective pseudocapacitor for charge storage. To prove our hypothesis, we first optimized the

polythiophene deposition chemistry and then integrated the active material within the porous

structure and evaluated the charge storage capacity of the polymer.

17

3 Chapter 3: Synthesis of Polymer Electronic Materials

3.1 Introduction

Initiated chemical vapor deposition (iCVD) has been shown to be a clean, solvent-free and

practical way to deposit polymer thin films. iCVD relies on the delivery of reactant vapors into a

vacuum chamber where the concentration of each reactant can be adjusted to influence the

resulting polymer structure and properties. In particular, the iCVD process uses a polymerization

initiator to start the polymerization reactions. Given its general approach, iCVD is a versatile

process which can be applied for the synthesis of different polymers for various applications.

Here, iCVD is studied for the synthesis of methacrylate-type polymer electrolytes.

Oxidative chemical vapor deposition (oCVD) also relies on the delivery of a gaseous

mixture of reactants. However, unlike iCVD, an oxidant is used to enable oxidative

polymerization of a monomer on the surface.1 oCVD is very similar to conventional vapor phase

polymerization2 with the exception that the oxidant instead of being deposited on the surface is

being delivered continuously as a vapor so the process will not be limited by the availability of

the oxidant initially coated on the surface but can be operated continuously. Here, oCVD is

explored in the synthesis of redox-active polythiophenes.

3.2 Poly(2-hydroxyethylmethacrylate) as a polymer electrolyte

Poly(2-hydroxyethyl methacrylate) (PHEMA) is one of the most studied synthetic hydrogel in

biological applications due to its ability to uptake water and non-cytotoxicity. The solvent uptake

of a polymer is an essential property for making a good polymer electrolyte as well. Poly(methyl

methacrylate) (PMMA) has shown promise as a gelled polymer electrolyte when used in

different devices.3-5

Likewise, PHEMA is able to uptake polar solvents and could make a viable

ionic conductor.

18

3.3 Experimental

3.3.1 Polymer synthesis

For the synthesis of PHEMA, the monomer 2-hydroxyethyl methacrylate (HEMA; 97% Aldrich)

and the initiator di-tert-amyl peroxide (TAPO; 97% Aldrich were used without further

purification. Depositions were carried out in a stainless steel, custom-built vacuum reactor,

similar to those described previously.6 The monomer HEMA was heated to 70 °C to achieve

sufficient vapor pressure while TAPO was kept at room temperature. Vapors of HEMA and

TAPO were metered into the reactor using needle valves. The backside cooled stage was

maintained at 25 °C and the filament wire array was heated to 300 °C. Depositions were carried

out under a reactor pressure of 100-150 mTorr controlled using a downstream throttle valve and

pressure controller (MKS Instruments) together with a dry vacuum pump (iH80, Edwards

Vacuum).

3.3.2 Polymer characterization

PHEMA was deposited on different substrates including Si(100), FTO glass and regular glass

slides to facilitate various polymer characterization. Film thickness was determined using an M-

2000U variable angle spectroscopic ellipsometer (J. A. Woollam). Fourier transform infrared

(FTIR) spectra were acquired in normal transmission mode using an MCT/A detector over the

range of 650-4000 cm-1

at a resolution of 4 cm-1

and averaged over 256 scans. X-ray

photoelectron spectroscopy (XPS) was performed on a Physical Electronics PHI 5000

VersaProbe with a scanning monochromatic source from an Al anode and with dual beam charge

neutralization. Survey XPS spectra were acquired at 100 W with pass energy of 117 eV over the

range of 0-1100 eV with 1 eV resolution and 100 ms dwell time, and averaged over 5 scans.

High resolution XPS spectra of C1s and O1s core electrons were acquired in high power mode

19

of 100 W with a pass energy of 23.5 eV. The polymer electrolyte conductivity was measured by

performing electrochemical impedance spectroscopy (EIS) (Gamry Reference 600) on redox-

incorporated PHEMA sandwiched between two platinized FTO electrodes. Platinized electrode

was prepared by spin coating of 50 µl of 5 mM solution of chloroplatinic acid hydrate (Aldrich,

99.9%) in 2-propanol (Aldrich, 99%) on FTO glass (Hartford‎ Glass,‎ 15‎ Ω/) followed by

annealing at 400 °C for 40 min. By modeling the EIS data with an appropriate equivalent circuit7

the ionic conductivity was evaluated. Gel permeation chromatography (GPC) was performed on

a Waters GPC system containing Styragel DMF columns HR 3 and HR 4 (Waters) with a

molecular weight range of 500-600,000 g/mol and at 40 °C. DMF (Aldrich, >99.9%) with 0.05

M LiBr (Aldrich, >99.9%) was used as the mobile phase eluting at a flow rate of 1.0 mL/min.

PEG/PEO standards were used for calibration.

3.4 Results and discussion

3.4.1 Fourier transform infrared spectroscopy

The iCVD process for PHEMA has been reported to be successful and the application of the

iCVD deposited PHEMA has been studied for biological purposes.6 The deposited film in our

reactor has shown identical features to that of standard PHEMA available commercially, as seen

in Figure ‎3.1. The C–O stretch (1200-1300 cm-1

), C–H bend (1350-1500 cm-1

), C=O stretch

(1700-1750 cm-1

) and C–H stretch (2700-3050 cm-1

) are evident in both the standard and iCVD

films. Their spectra are in good agreement with each other. Our iCVD deposited PHEMA

spectrum is in good agreement with that of iCVD PHEMA reported elsewhere.8 The broad peak

cantered at ~3450 cm-1

shows the hydroxyl group and the carbonyl group is evident with the

presence of the strong peak cantered at 1727 cm-1

. Comparing the FTIR spectra of HEMA and

PHEMA, the sharp band around 1670 cm-1

assigned to C=C stretch of the vinyl bond that is

20

observed in the monomer disappears in the polymer, confirming polymerization through the

double bond.

Figure ‎3.1. FTIR spectra of HEMA monomer, standard PHEMA and iCVD PHEMA.

3.4.2 X-ray photoelectron spectroscopy

The XPS survey spectrum of iCVD PHEMA is shown in Figure 3.2. The carbon-to-oxygen ratio

of 2:1 derived from the spectrum confirms that the polymer has the proper stoichiometry of

linear homopolymer of PHEMA. To further investigate the quality and chemical environment of

the deposited polymer film high resolution O1s and C1s spectra were acquired As shown in

Figure 3.3, the oxygen signal was fitted to three different bond environments with equal

intensities that are located at 532.5, 533.1 and 533.9 eV, which are in good agreement with the

reported values for O–C=O, OH and O–C=O oxygens for stoichiometric PHEMA.9

The carbon

signal was resolved into four different peaks at 285, 285.8, 286.6 and 288.9 eV corresponding to

21

–CH, –CCO, CO and –COO carbons with the expected intensity ratio for homopolymer

PHEMA.9

Figure ‎3.2. XPS survey of 100 nm PHEMA deposited on Si (001) wafer. The C1s and O1s peak

intensities were used to evaluate the ratio of C:O. The peak around 1000 eV is the Auger peak of

oxygen.

Figure ‎3.3. High resolution XPS signals of (a) oxygen 1s, and (b) carbon 1s.

22

This demonstrates iCVD's ability to produce clean, well-defined polymers with full preservation

of chemical functionality and absence of any impurity in the deposited film. This makes it an

extremely simple one-step method for synthesizing and forming polymer films without the

multi-step procedure typical of liquid synthesis and processing.

3.4.3 Gel permeation chromatography

Having a low molecular weight polymer is favorable for the purpose of ion conduction and in

our case the polymer molecular weight will be a function of iCVD growth kinetics.10

It has been

previously shown that polymerization rate is directly related to polymer kinetic chain length in

free radical polymerization i.e., slower deposition rate results in lower molecular weight

polymer. Polymer molecular weight was estimated for PHEMA deposited at two different

initiator flow rates with all else unchanged. Figure 3.4 shows the elution trace for the

poly(ethylene oxide) (PEO) standards and the two PHEMA samples. By increasing the flow rate

of initiator from 0.2 to 0.4 sccm while keeping the other parameters constant (monomer flow

rate FM = 0.4 and FN2 = 0.4, 125 mtorr total pressure) the average polymer molecular weight

increased from 2470 to 3590 g/mol.

Figure ‎3.4. Refractive index detector intensity as a function of elution time in gel permeation

chromatography for (a) PEO calibration standards, and (b) iCVD PHEMA (A and B are

deposited using 0.2 and 0.4 sccm flow of initiator, respectively.

23

3.4.4 Electrochemical impedance spectroscopy

To evaluate polymer ionic conductivity, EIS was performed on the Pt|electrolyte|Pt system to

derive the Nyquist plot and the Randles equivalent cell that included a finite Warburg impedance

was chosen for the purpose of fitting.11

As shown in Figure 3.5, the response at high frequencies

in the Nyquist plot can be attributed to the charge transfer at the electrode|electrolyte interface,

while the response at low frequencies can be associated with diffusion processes in the

electrolyte. The apparent diffusion coefficient of ionic species and the ionic conductivity of the

polymer were estimated from the Warburg impedance knowing the distance between the two

electrodes, which was measured using an optical microscope. The equivalent circuit shown in

Figure 3.5 consists of a pure resistor, a constant phase element (CPE) and a diffusion impedance

(Warburg impedance). The pure resistance in this circuit includes both the electrode resistance

(FTO) and the charge transfer resistance at the Pt|electrolyte interface, and the CPE is a non-

ideal capacitor12

that accounts for the double layer capacitance at electrode|electrolyte interface.

Here, due to the non-homogeneity of the surface a CPE element was used instead of a pure

capacitor. The last element which is a Warburg impedance was chosen for the diffusion

impedance however since the electrode distance was very short (~15 µm) a finite Warburg12

was

used instead of an infinite Warburg impedance. The following equation for the finite Warburg

impedance was integrated into the EIS software, and the output parameter conductivity (σ

[S/cm]) and diffusion coefficient (D [cm2/s]) were back calculated.

) (3-1)

In Equation 3-1, Zw is the finite Warburg impedance, σ is the ionic conductivity of the polymer

electrolyte, δ diffusion length, and ω is the frequency at which Zw is recorded. As can be seen in

Figure 3.5, the series resistance can be read off as 35.3 Ω (as the first arc at high frequency starts

from this value which shows the pure resistance of the electrodes.) to reduce the number of

24

degrees‎of‎ freedom.‎For‎ the‎ initial‎guess‎ for‎ the‎CPE‎element‎ the‎Z”‎value‎ for‎ the‎ first‎ semi-

circle was used and knowing the corresponding frequency of this point the equivalent ideal

capacitor can be back‎calculated‎using‎Z”‎=‎1/(ωC)

Figure ‎3.5. The Nyquist plot for a polymer electrolyte sandwiched between two Pt-coated FTO

glass. The data was fitted to a Randles equivalent circuit that included a Warburg impedance to

derive the conductivity and diffusion coefficient.

Ion diffusion inside iCVD PHEMA was evaluated using propylene carbonate and γ-

butyrolactone as gel plasticizers and dielectric additives for improving ionic conductivity inside

the polymer electrolyte.13

Here, we investigated the impact of the solvent composition on

polymer electrolyte conductivity. Among different solvent for the redox species of interest in

this study (LiI and I2) propylene carbonate seems to be a good candidate as it has been

frequently used as gel plasticizer and have a very high dielectric constant (89.6 at 40 °C).

However, since propylene carbonate has a high viscosity at room temperature, γ-butyrolactone

25

was added to adjust viscosity so ion diffusion will not be hindered due to the high viscosity

knowing that the diffusion coefficient is inversely proportional to the viscosity of the media. We

have tested a mixture of propylene carbonate and γ-butyrolactone with different volume ratios

containing 0.5 M LiI and 0.05 I2M as redox species, to find the conductivity of the polymer

electrolyte that was prepared using these solutions by fitting the EIS signal of the cell in a

symmetrical assembly. As can be seen in Figure 3.6, a 50:50 volume ratio of γ-butyrolactone to

propylene carbonate showed the highest ionic conduction.

Figure ‎3.6. Conductivity of iCVD PHEMA incorporating different dielectric additives

containing 0.05 M I2 and 0.5 M LiI as a redox couple.

3.5 Unsubstituted polythiophene as a redox-active polymer

Polythiophenes are an important representative class of conjugated polymers that form some of

the most environmentally and thermally stable materials that can be used as electrical

26

conductors, non-linear optical devices, polymer LEDs, transistors, electrochromic or smart

windows, photoresists, antistatic coatings, sensors, batteries, electromagnetic shielding materials,

artificial noses and muscles, solar cells, electrodes, microwave absorbing materials, new types of

memory devices, batteries, nanoswitches, optical modulators and valves, imaging materials,

polymer electronic interconnects, nanoelectronic and optical devices.14

Unsubstituted polythiophene (PTh) was first reportedly synthesized from polycondensation

of the 2,5-bromine substituted thiophene monomer in solution in the presence of a Ni(II)

catalyst.15,16

This method is an extension of the Kumada coupling of Grignard reagents to aryl

halides.17

Since PTh, even at low molecular weights, is insoluble in THF, the precipitation of the

polymer under the above reaction conditions limits the formation of higher molecular weights.

Other 2,5-dihalothiophenes have also been used, and finally in 1980 Wudl and co-workers

produced a high conductivity polymer of 10 S/cm when it was doped with AsF618

and later

Yammamoto et al reported19

on a polymer with 50 S/cm. Although the above methods that relied

on the 2,5-dihalothiophenes and Ni catalyst have been generally used to prepare high quality

PTh, other methods have been reported.20

An early report by Sugimoto described the synthesis

of PTh by treating thiophene with FeC13.20,21

Nowadays, the FeC13 method is a well-established

method for polymerizing polythiophenes22,2320

and continues to be the most widely used and

straightforward method to prepare polythiophene and its derivatives. In general, processing of

PTh due to its insolubility is challenging and that is one of the reason that its application has

been limited. Here to come over the challenges associated with the polymer insolubility and

difficulty in processing, we developed a method similar to the FeCl3 route based on a new

oxidant that is vaporizable and has enough oxidation potential to initiate the oxidative

polymerization of the PTh. We tuned the polymer film properties by changing the parameters for

synthesis of the polymer using our oCVD method.

27

3.6 Experimental

3.6.1 Polymer deposition

To enable oxidative chemical vapor deposition, we have used an oCVD reactor system described

in detail elsewere.24

Briefly, the reactor chamber was evacuated to base pressure (ca. 5 mtorr)

using a dry vacuum pump (Edwards Vacuum). Monomer, thiophene (97%, Sigma Aldrich), and

oxidant initiator, vanadium oxytrichloride (99%, Strem Chemicals), were used as received and

metered independently from glass source vessels into the chamber using precision metering

valves (Swagelok). The initiator was heated up to 45 °C to achieve sufficient vapor pressure and

its temperature was kept constant using a temperature controller (Omega Engineering). The

monomer had sufficient vapor pressure at room temperature and was not heated. The chamber

pressure was measured with a pressure transducer (MKS Instruments) and automatically

maintained by using a downstream throttle valve connected to a pressure controller (MKS

Instruments). The substrate temperature was kept constant through backside cooling of the

reactor stage by using a recirculating chiller (Thermo Scientific Neslab).

In order to tune the oCVD polymerization reaction and synthesis chemistry at the surface,

the ratio of the reactant (monomer, initiator) partial pressure to its saturated vapor pressure at the

temperature of the substrate (i.e., Pr/Pr,sat) was carefully adjusted and controlled (see Discussion

3.7.5) Thus, pressures ranging from 12-22 torr, and monomer and initiator flow rates of 0.5-7.0

and 0.1-1.0 sccm (standard cm3/min), respectively, were studied. The substrate temperature was

set at 5 °C. Polythiophene films were deposited on various substrates, including fluorine-doped

tin oxide glass (15/,‎ Hartford‎ Glass),‎ silicon‎ wafers‎ (WRS‎ Materials),‎ microscope‎ glass‎

slides (Fisher Scientific), and quartz glass (Chemglass). The silicon wafers were used as

received while all the other substrates were sonicated in dilute detergent solution (Citranox) and

28

thoroughly rinsed in deionized water prior to use. After each oCVD deposition, the substrate

temperature was raised to 80 °C for 4 h before the samples were extracted for analysis.

3.6.2 Characterization

Fourier transform infrared spectra (FTIR) were acquired on a Thermo Nicolet 6700 spectrometer

in normal transmission mode using an MCT/A detector at a resolution of 4 cm-1

averaged over

64 scans. UV-vis spectra of deposited films on quartz glass were acquired between 280-800 nm

with 1 nm resolution using a Shimadzu UV-1700 spectrophotometer. X-ray photoelectron

spectroscopy (XPS) was performed on a Physical Electronics PHI 5000 VersaProbe with a

scanning monochromatic source from an Al anode and with dual beam charge neutralization.

Survey XPS spectra were acquired at 100 W with pass energy of 117 eV over the range of 0-

1100 eV with 0.5 eV resolution and 50 ms dwell time, and averaged over 5 scans. High

resolution XPS spectra of C1s, O1s, V2p, Cl2p and S2p core electrons were acquired in high

power mode of 100 W with a pass energy of 23.5 eV using different acquisition times chosen

based on the observed intensity of the elements from the surveys. For depth profiling, C60 was

used as the ion source with a sputtering time of 30 s in between each depth acquisition and a

sputtering rate of ~4 nm/min. Raman spectra were collected on a Renishaw RM1000

microspectrometer using an Ar ion laser 488 nm with ~1 µm lateral spot size and 11 mW total

power. Cyclic voltammograms were recorded in a three electrode setup under a nitrogen blanket

with a Gamry Reference 600 potentiostat. The polythiophene samples deposited on FTO glass

served as the working electrode while a 2.5x2.5 cm platinum gauze (Princeton Applied

Research) was used as the counter electrode. The silver reference electrode (Princeton Applied

Research) was filled with 0.1 M silver nitrate (99.9999%, Sigma Aldrich) and 0.1 M

tetraethylammonium perchlorate (electrochemical grade, Sigma Aldrich) in acetonitrile (ACS

29

grade, Sigma Aldrich). A supporting electrolyte of 0.1 M tetrabutylammonium

hexafluorophosphate (electrochemical grade, Fluka) in acetonitrile was bubbled for 1 h with

nitrogen prior to use. The potential was swept between –0.4 and 1.2 V vs. Ag/AgNO3 with a

sweep rate of 80 mV/s. Polythiophene film conductivity was estimated through measuring sheet

resistivity using an Alessi four point probe connected to a Keithley 2400 source meter.

All characterizations were performed on as-deposited polythiophene films. In addition,

characterizations were done on films washed by first soaking in 0.1 M hydrochloric acid (Sigma

Aldrich) for 1 h to remove any vanadium compounds, then neutralized with 0.1 M sodium

hydroxide (BDH Chemicals), and finally washed thoroughly with distilled water and dried in air.

Besides iCVD polythiophenes, analysis was also performed for comparison on oligothiophene

standards,‎ including‎2,2’-bithiophene (97%, Acros Organics), terthiophene (99%, Alfa Aesar),

-quarterthiophene (TCI America) and -sexithiophene (Sigma Aldrich).

3.7 Results and discussion

In the oCVD of PTh, thiophene monomer and vanadium oxytrichloride oxidant/initiator were

continuously introduced as vapors into the chamber, where adsorption of the reactants on the

cooled substrate resulted in polymer film formation at all the different oCVD conditions of

pressure, flow rates, and substrate temperature explored. It was found that film deposition rate is

a strong function of reactant surface concentration, particularly of the initiator, with increasing

concentration leading to faster growth kinetics. For example, at a fixed substrate temperature (5

°C) and constant flow rates of monomer and initiator (2.0 and 0.5 sccm, respectively), increasing

the total pressure resulted in a higher deposition rate. Similarly, with all other conditions

remaining unchanged, an increase in substrate temperature by 2 °C significantly reduced the

deposition rate. Importantly, these observations point to the fact that oCVD is an adsorption-

30

limited process. Thus, a substrate temperature of 5 °C was chosen as a balance between the

decrease in reactant reactivity at lower temperatures and reduced adsorption at higher

temperatures. With the flow rates of monomer and initiator set at 2.5 and 0.6 sccm, respectively,

and the total reactor pressure of the reactor at 18 torr, the as-deposited film in its doped state

showed an electrical conductivity as high as 20 mS/cm. Dedoping the film by washing resulted

in the loss of electrical conductivity and a color change of the film from brown to orange.

3.7.1 Vibrational spectroscopy

Figure 3.7 shows the FTIR spectrum of the as-deposited polythiophene film. It has strong

distinct features in between 1100 and 1500 cm-1

which can be assigned to the doping induced

vibrations of polythiophene.19,25

The peaks observable at 790 and 690 cm-1

are characteristic of

C–H out-of-plane bending of the thiophene ring.26

The C–H stretches located above 3000 cm-1

also indicate that the aromatic thiophene ring is preserved. The peak located at 1490 cm-1

on the

shoulder of the doping induced vibration is attributed to the antisymmetric C=C stretch. Also

shown in Figure ‎3.7 is the FTIR spectrum of the washed polythiophene film. After washing with

acid, the C–H out-of-plane bending modes have become the most prominent features while the

doping induced peaks seem to have disappeared indicating that the acid wash removes the

dopant from the as-deposited film. There are small broad features at 1140 and 1260 cm-1

that

remain which are possibly from residual doping. Also noticeable besides the peak at 1491 cm-1

of the antisymmetric C=C stretch is the appearance of a peak located at 1439 cm-1

due to its

symmetric vibration that was previously obscured in the as-deposited film by the doping induced

bands. Additional peaks at 1042 cm-1

and 1069 cm-1

are assigned to C–C stretches, and ring

deformation can be observed at 740 cm-1

. The small broad peak at 950 cm-1

in only the as-

31

deposited polythiophene is possibly due to the presence of vanadium compounds formed during

deposition that is removed with acid washing.

Figure ‎3.7. FTIR spectra of as-deposited and washed polythiophene. The doping induced

vibrational bands in the 1100-1500 cm-1

region for the as-deposited film disappear after washing

as a result of dedoping.

The above assignments have been confirmed by literature19,25

and comparing with

experimental spectra of undoped standard oligothiophenes as well as computational modeling

results of neutral -oligothiophene molecules from density functional theory (DFT) calculations

in GAUSSIAN 03.27

To assign the vibrational frequencies observed in the FTIR spectra of

unsubstituted polythiophene (PTh) synthesized using oCVD,‎ a‎ series‎ of‎ undoped‎ α-

32

oligothiophenes were solvent-cast on silicon wafers using chloroform (ACS grade, Sigma

Aldrich). The samples were then dried in a vacuum oven for 3 h before the cast films were

analyzed by FTIR and the spectra are shown in Figure 3.8-3.11. Gaussian 03 was used to

perform geometry optimization and frequency calculations using a 6-31G*‎basis‎set‎and‎Becke’s‎

three-parameter exchange functional in combination with the Lee-Yang-Parr correlation

functional (B3LYP) in the gas phase.27

The calculated vibrational frequencies were then scaled

by a global scaling factor of 0.97 that was obtained by minimizing the root mean square (RMS)

error RMS of key vibrational bands between the experimental and simulated FTIR spectra of

thiophene. This was done as DFT calculations of similar aromatic ring structures pyrrole and

furan at the B3LYP/6-31G* level have shown the calculated frequencies to overestimate that of

corresponding experimental data.28

The same scaling factor used for thiophene was also used to scale the calculated spectra of

bithiophene, terthiophene and quarterthiophene, and comparisons with experimental spectra are

shown in Figure 3.9, 3.10 and 3.11, respectively. Here the strong bands at around 700 and 800

cm-1

correspond to C–H in-plane and out-of-plane bending, respectively, and are consistent with

experimental results.29

As the number of thiophene repeat units increases, the band near 700 cm-1

reduces in intensity while the one at 800 cm-1

is enhanced. This is mainly due to an increase in

the rigidity of the chain that favors in-plane C–H bending as well as a decrease in C–H groups at

the α and α’‎ (2‎ and 5) positions relative to those at the β and β’ (3 and 4) positions with

increasing chain length. A small feature between 700 and 800 cm-1

, which is observed to grow

with an increase in chain length, corresponds to C–S–C ring deformation. This frequency is

distinctively observable in washed PTh and suggests that the C–S–C vibration in the PTh film is

preserved. Together with the C–S–C vibration, the symmetric and antisymmetric C=C stretches

33

in the region of 1400-1600 cm-1

observed in PTh confirm that the thiophene ring structure is

Figure ‎3.8. The (a) calculated and (b) experimental FTIR spectra of thiophene.

Figure ‎3.9. The (a) calculated and (b) experimental FTIR spectra of bithiophene.

34

Figure ‎3.10. The (a) calculated and (b) experimental FTIR spectra of terthiophene.

Figure ‎3.11. The (a) calculated and (b) experimental FTIR spectra of quarterthiophene.

35

intact. A splitting of the band at 800 cm-1

becomes visible as the number of thiophene backbone

units increases. By examining the associated geometry and vibration of this mode, the splitting

can be attributed to the presence of different C–H in-plane vibrations induced by torsion of the

chain backbone in the neutral state. An increase in the number of thiophene repeat units also

leads to a general decrease in the peaks above 3000 cm-1

assigned to aromatic C–H ring

stretches. Examining more closely, this region contains contributions from two main C–H

stretching modes: one at ~3100 cm-1

from positions 2 and 5, and one at ~3050 cm-1

from

positions 3 and 4 of the thiophene ring. An increase in chain length results in a decrease in the α

and α’‎ vibration‎ relative to the β and β’‎ one, which is also observed for washed PTh and

indicates that oCVD successfully leads to the formation of unsubstituted polythiophene of

reasonable chain length.

As Figure 3.12 shows, the Raman spectra of the as-deposited and washed polythiophene

share similar features. The features can be divided into three main regions between the range of

1300-1510, 1000-1250, and 620-740 cm-1

. The strongest peaks can be observed in the 1300-

1500 cm-1

region that can be decomposed into four peaks located at 1365, 1417, 1457 and 1504

cm-1

.30,31

The vibration bands at around 1500 (1) and 1460 cm-1

(2) are assigned to asymmetric

and symmetric C=C ring stretching modes of polythiophene, respectively, and are of significant

importance in characterizing the polymer chain length and effective conjugation length.26,32

A

shift of the asymmetric C=C band from 1500 to 1502 and 1504 cm-1

has been reported when

polythiophene was electrochemically synthesized using bithiophene and terthiophene,

respectively, as monomer instead of thiophene.33

In our case, this peak is higher at 1504 cm-1

.

These shifts are small compared to the shift observed in oligomers of conjugated polymer

system, nevertheless they suggests a lower conjugation length for the doped film.32,34

The weak

36

shoulder at 1365 cm-1

is assigned to the C– C’ vibration in the thiophene ring and the peak

located at 1417 cm-1

can be attributed to the quinoid vibration in polythiophene (Figure ‎3.13).

Figure ‎3.12. Raman spectra of as-deposited and washed polythiophene. The spectra can be

divided into three main regions in the range of 1350-1500, 1000-1250 and 650-740 cm-1

. Insets

magnify the latter two regions of lower intensity. Washing results in narrowing of the peaks in

the highest intensity region as a result of a loss of the quinoid vibration.

37

Figure ‎3.13. Raman spectral fitting in the 1300-1500 cm-1

region of (a) as-deposited and (b)

washed.

Lastly, we observe in this region that although the as-deposited and washed films are

comparable, the latter shows narrower peaks that can be attributed to the loss of the quinoid

vibration of the doped state which consequently explains the loss of film conductivity as a result

of dopant removal during washing. The second region in the 1000-1250 cm-1

range can also be

decomposed into four distinct peaks (Figure ‎3.14a). The strongest peak located at 1045 cm-1

is

attributed to the C–H bending mode and the band at 1222 cm-1

is assigned to the vibration of the

Cα–Cα’ linkage between adjacent thiophene rings. The remaining two bands in this region can be

assigned to inter-ring C-C stretches of the distorted part of the molecule.35

The third region

within 620-740 cm-1

showing the weakest intensity of the three regions can also be resolved into

four distinct peaks (Figure ‎3.14b). The peak at 697 cm-1

describes the C–S–C coplanar vibration

of the thiophene ring, and distortional vibrations can be seen at 680 and 652 cm-1

. The band at

740 cm-1

can be also assigned to ring deformation. The intensity ratio of the sharp peak at 697

cm-1

in this region to the distortion peak at 680 cm-1

has been used to correlate with electrical

conductivity and defects in polythiophene.26

In our case, by deconvoluting this region into its

38

constituents peaks, the intensity ratio of 680 (D) to 697 (7) is calculated to be 0.67, which is

higher compared with the typical values (<0.6) reported for highly conductive polythiophene

prepared through electrochemical polymerization.29

As the film washed this ratio was also

decreased which is also an expected phenomena when the oxidation state of PTh samples

changes.36

Lastly, the washed film did not show any measurable conductivity which might be

due to the dopant loss.

Figure ‎3.14. Raman spectral fitting in the (a) 1000-1250 and (b) 620-740 cm-1

regions of as-

deposited PTh.

3.7.2 X-ray photoelectron spectroscopy

XPS was used for elemental analysis and understanding the chemical environments within the

PTh films. As shown in Figure ‎3.15, the as-deposited film shows a small presence of doped PTh

at a higher binding energy in the S2p spectra compared to the neutral state that disappears with

washing.37,38

Given that most of the as-deposited polymer is actually in an undoped state appears

to agree with the lower electrical conductivity observed. Likewise, as shown in Figure ‎3.16, the

presence of chlorine in as-deposited PTh that is essentially removed with washing evidences the

39

presence of dopant as well. The Cl2p peak resolved into its 2p3/2 and 2p1/2 doublet split shows a

single chlorine environment at ~198 eV that is characteristic of metal-bound chlorine. Since

vanadium is also observed in as-deposited PTh, this suggests that a vanadium compound

associated with chlorine might be the active species for doping PTh in a similar way as in

polypyrrole doped with perchlorate.39

The lack of covalently bound chlorine to organic carbon

excludes the possibility that the thiophene ring is chlorinated. In addition, washing of the sample

results in the near complete loss of vanadium and chlorine, which further supports them as

related to the dopant and not structurally attached to the polymer.

Figure ‎3.15. High resolution S2p XPS spectra of (a) washed and (b) as-deposited PTh.

40

Figure ‎3.16. High resolution Cl2p XPS spectra of (a) washed and (b) as-deposited PTh.

To obtain more quantitative compositional data, we performed XPS measurements on acid

washed polythiophene. The acid treatment removed any dopant that would interfere with

analysis of the polymer alone. As shown in Figure ‎3.17, the XPS survey reveals that only

carbon, sulfur and slightly adsorbed oxygen species is present in the film. The survey remains

unchanged with C60 sputter depth profiling, which indicates that the dopant was effectively

removed with washing. The survey also shows there is about 2.1 at% oxygen, < 0.1 at%

41

vanadium and 0.4 at% chlorine in the film. Interestingly, all the undoped oligothiophene

standards tested also show more than 1 at% oxygen (see Figure 3.17) so we believe the oxygen

could be due to adsorbed moisture on the samples.40,41

Figure ‎3.17. XPS survey spectra of washed polythiophene with C60 sputter depth profiling.

Compositional analysis yields a C:S ratio of 4.3:1, which is close to the theoretical value of 4:1.

A small amount of oxygen present (2.1 at%) is attributed to adsorbed oxygen species. Sputter

depth profiling using Ar+

shows the dopant has been effectively removed with the washing. The

maximum depth of acquisition is 10 nm and the spectra are separated equally by 2nm.

Table ‎3.1. XPS elemental analysis of oCVD polythiophene and standard oligothiophenes.

C (%) S (%) O (%) C:S ratio

terthiophene 3T 78.2 18.2 3.6 4.29

quarterthiophene 4T 81.0 17.6 1.4 4.60

sexithiophene 6T 78.7 17.6 3.7 4.47

iCVD polythiophene 78.1 18.2 2.1 4.35

42

Figure ‎3.18. High resolution (a) C1s and (b) S2p XPS spectra of washed polythiophene. Peak

fits reveal the expected 1:1 ratio of α and‎ β‎ carbons‎ as‎well‎ as‎ a‎ single‎ sulfur‎ species‎ of‎ the‎

thiophene ring.

As shown in Figure ‎3.18, high resolution C1s XPS spectra were also taken to elucidate specific

chemical bonding environments. The C1s region was fitted with two peaks located at 284.6 and

285.4 eV that can be attributed to β and carbons of the thiophene ring, respectively.39,42

The

43

relative intensity of these two peaks is 0.95 (theoretical value = 1.0), which lends support that

the thiophene ring structure is preserved during oCVD and stoichiometric polythiophene is

obtained. The S2p peak was decomposed into its constituent 2p3/2 and 2p1/2 peaks with a known

band split of 1.19 eV and a relative intensity of 2:1. The single 2p3/2 sulfur peak at 163.5 eV

indicates that the sulfur environment in the washed sample is intact in the thiophene ring.38

3.7.3 UV-vis spectroscopy

The UV-vis spectra of as-deposited and washed polythiophene are given in Figure ‎3.19. The as-

deposited sample shows two broad absorptions at 450 and 690 nm that are typical of doped

polythiophene.43,44

With washing, the sample was dedoped and the spectrum now shows a single

prominent peak with a maximum at 480 nm. Thismax is comparable to the reported value for

polythiophene‎ synthesized‎ electrochemically‎ and‎ has‎ been‎ assigned‎ to‎ the‎ π-π* transition.45-48

Based‎on‎the‎onset‎of‎the‎π-π*‎transition,‎the‎optical‎band‎gap‎is‎estimated‎to‎be‎1.9‎eV,‎which‎is‎

similar to that reported for electrochemically synthesized polythiophene.45

It should be noted that

the washing in this case was done using chloroform instead of acid as we wanted to dissolve and

separate out any soluble fraction for UV-vis analysis as well. Interestingly, themax of the

soluble portion is observed at 418 nm, which is lower than the peak at 480 nm for the washed

solid fraction (Figure ‎3.20).

This shift can be related directly to a difference in conjugation length, as shown in Figure

‎3.21 where max for various undoped standard oligothiophenes is plotted against the number of

thiophene repeat units, and compared with the soluble and undissolved portions of the washed

polythiophene film. We clearly see with the oligothiophene standards that a higher peak

wavelength value, which indicates a greater conjugation length, correlates naturally to a longer

44

Figure ‎3.19. UV-vis spectra of as-deposited and washed polythiophene. The as-deposited film

shows the characteristic broad polaron absorption of the doped form while the dedoped film after

washing gives a single absorption peak at λmax =‎470‎nm‎of‎the‎π-π*‎transition.

Figure ‎3.20. UV-vis spectrum of the dissolved portion of PTh.

45

thiophene chain. Based on this trend, the soluble fraction is estimated to consist mainly of chain

oligomers of five thiophene units long while the washed polythiophene film is expected to be

much longer than six thiophene units given that its peak wavelength is much higher than all the

oligothiophene standards tested. In fact, it is higher than that reported for polybithiophene and

comparable to that of polythiophene synthesized electrochemically.48

Figure ‎3.21. Correlation of λmax with the number of thiophene chain units, comparing literature

data49

and experimental data of undoped standard oligothiophenes with washed polythiophene

and its dissolved soluble fraction after washing. The correlation indicates a longer conjugation

length with a higher degree of polymerization. The correlation also indicates that the solid

portion of the washed film is much longer than 6-7 thiophene units while the dissolved fraction

consists mainly of 5 repeat units.

3.7.4 Cyclic voltammetry

A cyclic voltammetry sweep of the acid washed polythiophene film between –0.2 and 1.2 V vs.

the saturated calomel electrode (SCE) is shown in Figure 3.22. The anodic and cathodic peaks

observed at 1.1 and 0.6 V, respectively, are in good agreement with reported data for

46

polythiophene synthesized by electrochemical methods.44,45

As the potential was swept in the

anodic direction, the film color changed from orange-red in its initial dedoped state to blue

representing the redoped state. Sweeping in the reverse direction caused the film to revert back

to its original orange-red color.

Figure ‎3.22. Cyclic voltammogram of washed polythiophene in a supporting electrolyte solution

of acetonitrile and tetraethylammonium hexafluorophosphate. The anodic and cathodic peaks are

observed at 1.1 and 0.6 V vs SCE, respectively. The corresponding film color in the oxidized

and reduced forms are also shown.

Further potential cycling for 10 cycles led to a gradual decrease in the anodic peak by 10% and a

slightly peak shift to a higher potential (Figure 3.23). This can be explained by an irreversible

degradation of the film due to the nucleophilic attack of the supporting electrolyte anion on the

thiophene rings leading to polythiophene overoxidation and a loss of chemical integrity.47,48

47

Figure ‎3.23. Cyclic voltammetry sweep of PTh.

3.7.5 Effect of oCVD processing conditions on polymer properties

To investigate the effect of oCVD synthesis and process parameters on resulting film properties,

we relied on the relative pressure Pr/Pr,sat (where Pr is the partial pressure of the reactant species

and Pr,sat is its vapor pressure at the substrate temperature) as a key adsorption parameter for

controlling the surface concentration of monomer (M) and initiator (I), and in turn the

polymerization growth kinetics during deposition.50

For the studies here, the substrate

concentration was kept constant at 5 °C, which yielded a monomer and initiator vapor pressure

of 27 and 7 torr, respectively, based on literature vapor-liquid equilibrium data.51

Then by

changing the flow rate of the reactants, we were able to design a series of synthesis experiments

to explore the oCVD space for influencing polymer film characteristics. Table 3.2 shows

polythiophene synthesized by oCVD at various (PM/PM,sat, PI/PI,sat) coordinates i.e., at different

surface monomer and initiator concentrations, and the corresponding electrical conductivity of

48

the as-deposited films. Although the values lie in the lower range reported for polythiophene,29

more importantly, conductivity can be controlled by tuning iCVD synthesis. In addition, Table

3.2 gives the corresponding Raman intensity ratio 1 / 2 of the as-deposited polythiophene as

well as the UV-vis absorption peak max of the washed polythiophene.

Qualitatively, we can see from sample runs 1-3 in which monomer concentration was kept

roughly constant while initiator concentration was increased that the UV-vis absorption peak

shifted to lower wavelengths. From our discussion of Figure 3.21, this shift indicates a decrease

in polythiophene chain length and since the UV-vis‎peak‎corresponds‎to‎the‎π-π*‎transition,‎it‎is‎

reasonable to expect that this also means a shorter conjugation length. Likewise, comparing

sample runs 3 and 5, for a similar initiator concentration, a decrease in monomer concentration

also resulted in a decrease in peak wavelength and effective conjugation length. These

observations suggest that a greater presence of monomer relative to the initiator at the growth

surface leads to a greater conjugation length of the deposited polythiophene. However, this

generalization does not fully explain the results. If we compare runs 2 and 4 in which the relative

amount of monomer to initiator is the same, we do not get the same absorption maximum. In

addition, the max values do not fully correlate with the observed conductivities even though the

peak‎represents‎a‎π-π*‎transition.‎This‎can‎be‎rationalized‎by‎realizing‎ that‎the‎ max data only

represents washed polythiophene in which the dopant has been removed. In addition, the

washing also removed the soluble oligothiophene portion of the polymer. Thus, we should not

expect the max data to fully follow the conductivity data of the as-deposited polythiophene.

49

Table ‎3.2. Effect of oCVD synthesis conditions on polythiophene properties.Sample run.

PM/PM,sat PI/PI,sat λmaxa (nm) I(ν1)/I(ν2) Conductivity

(mS/cm)

1 0.55 0.10 490 0.28 6.2

2 0.53 0.53 480 0.26 4.0

3 0.50 0.73 472 0.32 20.1

4 0.25 0.25 460 0.21 0.71

5 0.15 0.78 445 0.20 NAb

a λmax obtained from the washed film

b No stable reading.

To fully rationalize the observed data, we should recognize that the oxidant utilized in

oCVD synthesis acts not only as an initiator to enable the polymerization of polythiophene but

also simultaneously dopes the formed polymer. As such, the absolute amount of initiator is

equally important in determining electrical conductivity. Thus, looking at runs 2 and 4 again,

with more initiator present a higher electrical conductivity is observed even though the relative

amounts of monomer and initiator are equivalent in the two runs. Quantitatively, we have found

that the Raman intensity ratio 1 / 2 correlates well with film conductivity, noting that both sets

of measurements were taken from as-deposited polythiophene. The larger the 1 / 2 intensity

ratio, the higher the film conductivity. This phenomena observed by others for VVP and

attributed to the lower chain length which suggest less conformational defect in the chain and

suggest higher conjugation length in the doped film.32

So a combination of sufficient chain

length, effective conjugation length and effective doping in the as deposited film resulted in the

observed behavior. It is worth mentioning that our results along with other reports on the values

50

of the π-π*‎suggests that the saturation limit for wavelength of this transition is not necessary

471 nm and can exceed this value.45

This observation pose that effective conjugation length in

the polythiophene system is smaller than what is predicted (17 units) with the nonlinear

approach introduced by Meier et al.52

It is widely recognized that the synthesis of polythiophene in the liquid phase using an

oxidant initiator follows an oxidative polymerization mechanism.54

In a vapor deposition

environment, the oxidative polymerization mechanism has also been suggested for the synthesis

of PEDOT.1,54

Here we propose that the polymerization of polythiophene using vanadium

oxytrichloride as the oxidant initiator follows a similar scheme whereby a cation radical is

formed on the thiophene ring and stabilized by vanadium compound counteranions acting as the

dopant. This enables the addition of thiophene rings through - carbon coupling and allows the

formation of extended conjugation. The specificity of this coupling is supported by our test of

using 2,5-dibromothiophene with a comparable oxidation potential as the thiophene monomer

but which did not result in any film formation at least in the wide oCVD parameter space that

was explored here. However, the possibility of - and - linkages during polymerization

cannot be unequivocally ruled out at this point. It is possible that the presence of these defects

could account for the lower electrical conductivity of the polythiophene films particularly when

they were deposited at a high initiator concentration relative to the monomer.55

One hypothesis is

that the higher initiator concentration at low monomer surface coverage simply provides more

active initiator species that can attack the less reactive β sites, leading to a greater probability of

structural defect linkages being formed. Unfortunately, Raman spectroscopy was not sensitive

enough to detect these vibrational defect modes. However, by re-examining the FTIR spectra,

we have found that the peak intensity ratio of the C–H bending vibrations at 690 and 790 cm-1

,

which has been correlated with chain length,26

actually underestimated the chain length predicted

51

by UV-vis. This weakening in the C–H vibration modes could be an indication of defects, and in

fact has been observed in the electrochemical synthesis of polythiophene and introduced as a

reason for lower polythiophene conductivity.55

3.8 Conclusions

Polymer electrolyte with exceptionally clean and accurate stoichiometry were synthesized using

iCVD. The polymer ionic conductivity was measured and optimized for the gel electrolyte

application. The polymer was fully characterized and the molecular weight found to be in a

suitable range for gel electrolyte application. Despite having exceptional electroactive properties,

applications of unsubstituted polythiophene have been limited due to its insolubility. Here we

showed for the first time that the unsubstituted polythiophene can be deposited through oCVD

process using vanadium oxytrichloride as an effective oxidant. The deposited films were well

characterized and the control over polymer properties in the space parameter of the oCVD

obtained. Vibrational spectroscopy along with the conductivity results and UV-vis data suggest

that the deposited film is not fully doped polymer. The possibility of - and - linkages

during polymerization cannot be unequivocally ruled out at this point. It is possible that the

presence of these defects could account for the lower electrical conductivity of the

polythiophene films particularly when they were deposited at a high initiator concentration

relative to the monomer. However, the route of this defect could not be in chemical degradation

of the ring as XPS did not indicate any chemical change in the ring structure. Significantly,

polymer conjugation length and electrical conductivity can be tuned by controlling oCVD

process variables. Polymerization is found to be adsorption-limited so by providing sufficient

monomer and limiting the amount of initiator at the growth surface, polythiophene is believed to

be formed through - thiophene linkages.

52

3.9 References

1. J. P. Lock, S. G. Im and K. K. Gleason, Macromolecules, 2006, 39, 5326-5329.

2. B. Winther-Jensen, J. Chen, K. West and G. Wallace, Macromolecules, 2004, 37, 5930-

5935.

3. Y.H. Liao, D.Y. Zhou, M.-M. Rao and W.S. Li, Dianchi Gongye, 2008, 13, 428-432.

4. H. Yang, M. Huang, J. Wu, Z. Lan, S. Hao and J. Lin, Materials Chemistry and Physics,

2008, 110, 38-42.

5. A. A. Argun, A. Cirpan and J. R. Reynolds, Advanced Materials, 2003, 15, 1338-1341.

6. R. K. Bose and K. K. S. Lau, Biomacromolecules, 2010, 11, 2116-2122.

7. C. Longo, A. F. Nogueira, M.-A. De Paoli and H. Cachet, Journal of Physical Chemistry

B, 2002, 106, 5925-5930.

8. K. Chan and K. K. Gleason, Langmuir, 2005, 21, 8930-8939.

9. G. P. López, D. G. Castner, B. D. Ratner, Surface and Interface Analysis, 1991 17, ( 5),

267–272.

10. K. K. S. Lau and K. K. Gleason, Macromolecules, 2006, 39, 3695-3703.

11. Bard, A. J.; Faulkner, L. R., Electrochemical Methods: Fundamentals and Applications.

Wiley, Hoboken, USA 2000.

12. Barsoukov, E.; Macdonald, J. R., Impedance Spectroscopy: Theory, Experiment, and

Applications. Wiley, Hoboken, USA 2005.

13. R. Komiya, L. Y. Han, R. Yamanaka, A. Islam and T. Mitate, Journal of Photochemistry

and Photobiology, A, 2004, 164, 123-127.

14. C. Li, H. Bai and G. Shi, Chemical Society Reviews, 2009, 38, 2397-2409.

15. T. Yamamoto, K. Sanechika and A. Yamamoto, Journal of Polymer Science: Polymer

Letters, 1980, 18, 9-12.

53

16. J. W. P. Lin and L. P. Dudek, Journal of Polymer Science: Polymer Chemistry 2003, 18,

2869-2873.

17. W. T. Robinson and B. Belleau, Journal of the American Chemical Society, 1972, 94,

4376-4378.

18. Kobayashi, M.; Chen, J.; Chung, T. C.; Moraes, F.; Heeger, A. J.; Wudl, F. Synthetic

Metals 1984, 9, (1), 77-86.

19. T. Yamamoto, A. Morita, Y. Miyazaki, T. Maruyama, H. Wakayama, Z. H. Zhou, Y.

Nakamura, T. Kanbara, S. Sasaki and K. Kubota, Macromolecules, 1992, 25, 1214-

1223.

20. K. Yoshino, S. Hayashi and R.i. Sugimoto, Japanese Journal of Applied Physics, 1984,

23, 899-900.

21. R. Sugimoto, S. Takeda, H. B. Gu and K. Yoshino, Chemistry Express, 1986, 1, 635-

638.

22. O. Inganäs, W. R. Salaneck, J. E. Österholm and J. Laakso, Synthetic Metals, 1988, 22,

395-406.

23. J. R. Reynolds, J. P. Ruiz, A. D. Child, K. Nayak and D. S. Marynick, Macromolecules,

1991, 24, 678-687.

24. S. Nejati and K. K. S. Lau, Nano Letters, 2011, 11, 419-423.

25. S. Hotta, W. Shimotsuma and M. Taketani, Synthetic Metals, 1984, 10, 85-94.

26. Y. Furukawa, M. Akimoto and I. Harada, Synthetic Metals, 1987, 18, 151-156.

27. R. D. Gaussian 03, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A.

Robb, J. R. Cheeseman, J. A. Montgomery, Jr., T. Vreven, K. N. Kudin, J. C. Burant, J.

M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani,

N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J.

Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E.

Knox, H. P. Hratchian, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.

E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y.

Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S.

Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K.

Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J.

Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L.

Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M.

54

Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, and J. A.

Pople, Gaussian, Inc., Wallingford CT, 2004.

28. G. Rauhut and P. Pulay, Journal of Physical Chemistry, 1995, 99, 3093-3100.

29. M. Akimoto, Y. Furukawa, H. Takeuchi, I. Harada, Y. Soma and M. Soma, Synthetic

metals, 1986, 15, 353-358.

30. E. A. Bazzaoui, J. Aubard, N. Félidj, G. Laurent and G. Lévi, Journal of Raman

Spectroscopy, 2005, 36, 817-833.

31. G. Poussigue, C. Benoit, J. L. Sauvajol, J. P. Lere-Porte and C. Chorro, Journal of

Physics: Condensed Matter, 1991, 3, 8803-8824.

32. P. M. Bayley, B. Winther-Jensen, D. R. MacFarlane, N. M. Rocher and M. Forsyth,

Reactive Functional Polymers, 2008, 68, 1119-1126.

33. G. Compagnini, A. De Bonis, R. S. Cataliotti and G. Marletta, Physical Chemistry

Chemical Physics, 2000, 2, 5298-5301.

34. C. Castiglioni, M. Del Zoppo and G. Zerbi, Journal of Raman Spectroscopy, 2005, 24,

485-494.

35. E. Bazzaoui, G. Levi, S. Aeiyach, J. Aubard, J. Marsault and P. Lacaze, Journal of

Physical Chemistry, 1995, 99, 6628-6634.

36. E. A. Bazzaoui, S. Aeiyach, J. Aubard, N. Felidj, P. C. Lacaze, N. Sakmeche and G.

Lévi, Journal of Raman Spectroscopy, 1998, 29, 177-183.

37. G. Tourillon and Y. Jugnet, Journal of Chemical Physics, 1988, 89, 1905-1913.

38. E. T. Kang, K. G. Neoh and K. L. Tan, Physical Reviews B, 1991, 44, 10461-10467.

39. L. Atanasoska, K. Naoi and W. H. Smyrl, Chemistry of. Materials, 1992, 4, 988-994.

40. H. Hintz, H. Peisert, H. J. Egelhaaf and T. Chassé, Journal of Physical Chemistry C,

2011, 115, 13373-13376.

41. D. Oeter, C. Ziegler and W. Göpel, Synthetic Metals, 1993, 61, 147-150.

42. S. S. Jeon, S. J. Yang, K. J. Lee and S. S. Im, Polymer, 2010, 51, 4069-4076.

55

43. G. Zotti and G. Schiavon, Synthetic Metals, 1989, 31, 347-357.

44. D. Zhang, J. Qin and G. Xue, Synthetic Metals, 1999, 106, 161-164.

45. S. Alkan, C. A. Cutler and J. R. Reynolds, Advanced Functional Materials, 2003, 13,

331-336.

46. H. Harada, T. Fuchigami and T. Nonaka, Journal of Electroanalytical Chemistry, 1991,

303, 139-145.

47. J. Roncali, F. Garnier, M. Lemaire and R. Garreau, Synthetic Metals, 1986, 15, 323-331.

48. S. Wang, K. Tanaka and T. Yamabe, Synthetic Metals, 1989, 32, 141-150.

49. I. F. Perepichka and D. F. Perepichka, Handbook of Thiophene-Based Materials, Wiley,

Chichester, UK, 2009.

50. K. K. S. Lau and K. K. Gleason, Macromolecules, 2006, 39, 3688-3694.

51. C. L. Yaws, Yaw's Handbook of Antoine Coefficients for Vapor Pressure, Knovel,

2009.

52. H. Meier, U. Stalmach and H. Kolshorn, Acta Polymerica, 1997, 48, 379-384.

53. S. Sadki, P. Schottland, N. Brodie and G. Sabouraud, Chemical Society Reviews, 2000,

29, 283-293.

54. S. G. Im, P. J. Yoo, P. T. Hammond and K. K. Gleason, Advanced. Materials, 2007, 19,

2863.

55. K. Tanaka, T. Shichiri and T. Yamabe, Synthetic Metals, 1986, 16, 207-214.

56

4 Chapter 4: Integration of Polymer Electrolytes in Dye Sensitized Solar Cells

4.1 Introduction

The global demand for energy is increasing with population and for the major portion of this

demand we are currently relying on fossil fuel. The increase in green house gases associated

with burning fossil fuel and the dwindling fossil fuel supply have fueled attention to other

sources of energy that are cleaner.1

Photovoltaic (PV) devices which rely on abundant sun

energy are among the promising energy conversion devices for the future of the planet. However

due to the higher cost of solar cells for energy conversion and delivery this technology has not

been fully exploited. Having more economical PV cells with respectable efficiency that can last

for 20 years or more could be a way to bring this technology to reality. Consequently, the quest

for making current PV technologies cheaper has led to many inventions in the last 20 years

among which the dye sensitized solar cell (DSCC) has evolved quite rapidly to a point that it is

currently undergoing commercialization.2 The dye sensitized solar cell first introduced by

Michael‎ Grätzel‎ and‎ Brian‎ O’Regan‎ in‎ 1992‎ has gone through many refinements and

efficiencies have exceeded 11%.3 Lately, devices in the lab scale with efficiency as high as

12.3% have been fabricated and it appears that the efficiency of this sophisticated design can

reach upward of 15%.4,5

A higher efficiency DSSC certainly will help pave the road for more extensive

commercialization of this technology, there remains challenges left to be addressed for

enhancing the performance of the cell and modules. The first and most important problem

associated with the cell design is leakage of the liquid electrolyte at elevated temperature which

has detrimental effect on cell performance and eventually results in cell failure.6 Significant

attention has been devoted to this problem and many materials, including organic and inorganic

solid state hole transport materials,7-9

conjugated polymers,10,11

and solid and gel polymer

57

electrolytes,12-15

have been explored to replace the liquid electrolyte. Among these choices,

polymer electrolytes are a promising candidate as they can be designed towards enhancing ionic

conductivity.16

Nevertheless the effort to integrate polymer within the cell has been less than

effective and polymer infiltration is constrained by the high aspect ratio of the mesoporous

nanostructured electrode.17-20

To surmount this problem here we are relying on our unique

method to integrate polymer materials within the nanostructures.

.

4.1.1 DSSC working principle

As shown in Figure.‎4.1, a DSSC is composed of five main components: a transparent conductive

oxide (TCO) electrode, a wide band gap semiconductor, a chromophore sensitizer, a hole

conductor media, and a counter electrode to complete the circuit.

Figure.‎4.1. Dye sensitized solar cell layout.

58

To benefit from a higher surface area and more dye incorporation in the semiconductor

layer, the electrode is fabricated as a porous network of interconnected nanoparticles with

electrode roughness (the ratio of actual available surface area to geometric surface area) as high

as 1000. To form these types of electrode, a sol-gel process is being employed using

nanoparticles of a semiconductor with nominal particle size from 15-25 nm that results in 45-

75% porosity.18,21,22

Currently, TiO2 in the anatase form is being used for electrode fabrication

and the specific surface area ranging from 55-150 m2/g depending on the size of the particle used

is obtained. Upon illumination, the ground state of the dye is excited to its unrelaxed metal-to-

ligand-charge-transfer (MLCT) state. The molecule can relax to its original state within tens of

femto to a pico second or can inject an electron into the semiconductor from any of the

unthermalized states.23

The electrons injected into the semiconductor are transferred within the

network of TiO2 through a random walk assisted by transfer between localized sites which are

distributed in energy or tunneling over a longer range.24

These electrons find their way to the

conductive substrate on which the electrode assembly is fabricated. The dye in the oxidize form

on the other hand is being reduced by a redox couple in the electrolyte solution. In the typical

DSSC, iodide and triiodidie are the redox species in the electrolyte.25

The way the redox species

works in the system is that the I– will donate an electron to regenerate the dye and form I3

–. The

I3– diffuse to the platinum interface at the counter electrode where it receives back electrons from

the external circuit in a two electron transfer to reform I–. In this way, the photogenerated

electrons from the dye traveling through the external circuit can continuously be created by the

regeneration of the dye.

However, the electrons injected into the semiconductor diffuse within TiO2 structure can

recombine with the ionic species in the solution at the semiconductor-liquid electrolyte interface.

Different mechanisms have been proposed for this recombination processes and the detail of

59

possible mechanism can be found elsewhere.22,26,27

Recombination can also occur by the

recombination of the electron in the TiO2 network with the dye but this pathway is of lesser

importance and has been often being ignored in most studies as it is insignificant compared to

recombination with the redox ions in the electrolyte.28

4.1.2 Replacing the liquid electrolyte

As mentioned earlier, the medium for idodide/triiodide transport is a liquid electrolyte which in

this case is an aprotic solvent, and the most commonly used in DSSC is acetonitrile. It is known

that the acetonitrile can escape the cell and cause cell failure. This effect is exacerbated at higher

temperature and has led many researchers to find alternatives.6 To replace the liquid with a solid

material which potentially eliminates the problem associated with leaking, many attempts had

been made to infiltrate polymers and solid state hole transport materials.17,25

Although the

importance of pore filling and pore wetting is recognized as a necessary technological leap for

advancing DSSC performance, currently the effective electrode thickness of the device

fabricated with these materials has been limited by their penetration depth into the mesoporous

TiO2 layer to ~2 µm.16-18

The nanoscale pore diameter and narrow pore structure make it

especially difficult for liquid based techniques to transport these materials due to viscous and

steric effects. In addition, these processes introduce solvents into the system that make their

complete removal nontrivial, and the presence of any unwanted chemical will most likely

deteriorate cell performance.

4.1.3 Pore filling methodology

iCVD was used to polymerize poly(2-hydroxyethyl methacrylate) (PHEMA) as a potential

polymer electrolyte material inside the mesoporous TiO2 electrode of the DSSC. PHEMA has

60

been chosen as the presence of the ester and hydroxyl groups are expected to facilitate ion

transport and enable the formation of a stable gel electrolyte with propylene carbonate and γ-

butyrolactone typically used as dielectrics in DSSCs.30

Using conventional free radical

polymerization of 2-hydroxyethyl methacrylate (HEMA) and t-amyl peroxide (TAPO) as the

monomer and initiator, respectively, the effect of critical iCVD parameters in influencing the

rate of polymer formation inside the pores has been investigated in order to obtain a systematic

way for filling the mesoscopic TiO2 electrode effectively. Figure ‎4.2 shows the iCVD scheme to

integrate polymer electrolyte within the mesoporous electrode structure. To obtain pore filling of

the mesoporous TiO2 electrode with PHEMA, we performed iCVD on 3-4 µm thick electrode

layers to systematically investigate different iCVD parameters and observe the resulting

electrode SEM cross sections.

Figure ‎4.2. Schematic of the pore filling process during initiated chemical vapor deposition

(iCVD). Monomer (M) and initiator (I) molecules are delivered into the reaction chamber in the

form of gaseous vapors. The initiator is selectively activated by a series of heated filaments. The

activated initiator (R) and monomer (M) adsorb onto the TiO2 surface within the nanostructured

mesoporous electrode that is kept cooled to enhance the adsorption-limited process. Addition

polymerization of the monomer at activated initiator sites results in the formation of a growing

polymer inside the pores.

61

4.2 Experimental

4.2.1 Electrode preparation

The titanium dioxide colloidal paste consisting of TiO2 nanoparticles P25 (Envonik) was

prepared using a published procedure.31

The prepared paste was spin coated at 700-1200 rpm

onto‎fluorine‎doped‎tin‎oxide‎conductive‎glass‎(Hartford‎Glass,‎15‎Ω/) to obtain a film of the

paste between 2 and 8 µm in thickness. To create the mesoporous structure, the spin coated TiO2

paste was heated up to 500°C and annealed for 30 min. For pore filling experiments, samples

were cooled to room temperature and stored until further use. For dye sensitized solar cell

fabrication, samples were cooled down to 100 °C and immersed in a 3x10-4

M solution of

photosensitizer dye cis-dicyano-bis(2,2’-bipyridyl-4,4’-dicarboxylic acid) ruthenium(II)

(Ruthenizer 505, Solaronix) or cis-bis(isothiocyanato)bis(2,2'-bipyridyl-4,4'-dicarboxylato)-

ruthenium(II) (Ruthenizer 535, Solaronix) in pure ethanol, and left in the dark for 20 h.

4.2.2 Pore filling

Each prepared mesoporous TiO2 on glass was first heated up to 200 °C for 30 min to remove any

adsorbed water or contamination inside the pores and after cooling the sample was then placed

inside our iCVD chamber to enable polymerization of poly(2-hydroxyethyl methacrylate)

(PHEMA) inside the mesoporous TiO2. Briefly, the chamber was evacuated to a base pressure of

< 7 mtorr using a dry pump system (Edwards iH160). The monomer, 2-hydroxyethyl

methacrylate (Aldrich, 97%) and initiator and t-amyl peroxide (Aldrich, 97%) were used as

received, and metered independently at set flow rates through precision needle valves

(Swagelok) into the chamber. The chamber pressure was controlled by a downstream throttle

valve (MKS instruments 253B) and a pressure transducer (MKS instruments 626A) connected to

a pressure controller (MKS instruments 651C). The substrate temperature of the sample was kept

62

constant through backside cooling water controlled with a recirculating chiller (Neslab RTE-7).

A series of equally spaced Chromaloy filaments (Goodfellow) was suspended above the sample

and resistively heated using a DC power supply (Sorensen DLM 60-10) to thermally activate the

initiator. With the cooling water and filament temperatures at 27 and 350 °C, respectively, the

substrate temperature was measured to be 50 °C. To investigate the iCVD parameter space for

pore filling, the monomer and initiator flow rates were set at different values from 0.2 to 0.8

sccm (std cm3

per min), and the chamber was maintained at pressures in the range of 60 to 200

mtorr. After iCVD polymerization and deposition was completed, the chamber was pumped

down while a stream of argon was introduced to purge the reactor until base pressure was

restored. The chamber was then vented to atmosphere and the sample was removed for

characterization with cross sectional SEM (Zeiss Supra 50 VP) after a ~5 nm Pt/Pd coating to

evaluate the degree of pore filling.

4.2.3 Pore filing estimation

Each sample was cleaved into two parts, one part was used for cross sectional SEM and the

remainder was used for thermogravimetric analysis (TGA). Using cross sectional SEM, the

polymer overlayer coating and the mesoporous thickness were measured at 5 different locations

along the edge and the average values used for pore filling estimation. TGA was performed on a

Q50 (TA Instruments) with a 10°C/min ramp from room temperature to 800 °C under a nitrogen

purge.. The porosity of the starting unfilled electrode was estimated using N2 adsorption

isotherm at 77 K. The mesoporous TiO2 deposited on a silicon substrate was first placed into a

macro cell (Quantachrome) and degassed at 180 °C for 12 h. The cell was then connected to the

Autosorb-1 (Quantachrome) and adsorption/desorption isotherms were recorded in between 5

and 100% of P/P0 where P0 was set to 760 torr.

63

4.2.4 Dye sensitized solar cell fabrication

Each prepared TiO2 electrode was removed from the dye solution in an enclosed argon

environment and rinsed thoroughly with pure ethanol to remove any excess and unadsorbed dye

from the TiO2 surface. The sample was then transferred immediately to the iCVD reactor for

incorporating PHEMA with the procedure described above and using iCVD conditions that gave

complete pore filling. After deposition, the sample was removed from the chamber, and some of

the deposited polymer were removed from the glass edge with a cotton swab dipped in dimethyl

formamide (EMD Chemicals, >99.9%) to create the electrical contact. To incorporate the redox

couple in the PHEMA, the sample was immersed for 6 h in a redox solution composed of 0.5 M

lithium iodide (Aldrich, 99.9%) and 0.05 M iodine (Aldrich, 99%) in 50:50 or 0:100 vol% of γ-

butyrolactone (Aldrich, >99%) and propylene carbonate (Aldrich, >99%). A platinized counter

electrode was prepared by spin coating of 50 µl of 5 mM solution of chloroplatinic acid hydrate

(Aldrich, 99.9%) in 2-propanol‎(Aldrich,‎99%)‎on‎FTO‎glass‎(Hartford‎Glass,‎15‎Ω/) followed

by annealing at 400 °C for 40 min. The counter electrode and photoanode were pressed against

each other and clipped together. Comparisons were made with a standard liquid electrolyte cell

prepared in the same way except that the polymer was replaced by acetonitrile (Alfa Aesar,

99.7%), which was used as the redox solvent and electrolyte, and two L-shaped spacers (Surlyn,

Solaronix) were heat-sealed between the two electrodes to avoid a short in the cell. The distance

between the two electrodes is measured to be 25 µm.

4.2.5 Solar cell characterization

Photovoltaic measurements of each assembled DSSC was carried out in a custom solar simulator

equipped with a 300 W Xe lamp filtered to AM 1.5 spectral conditions. Data were taken using a

Gamry Reference 600 with a 0.26 cm2 mask on the samples under 100 mW/cm

2 irradiance and

64

an externally applied bias potential. For each electrode thickness, ten samples were

characterized. Each assembled DSSC was also characterized with electrochemical impedance

spectroscopy under 1 sun illumination (corresponding to global AM 1.5 and 100 mW/cm2

irradiance) with a frequency sweep between 800 kHz and 20 mHz, and an AC amplitude of 10

mV‎at‎the‎cell’s‎open‎circuit‎voltage.‎The‎platinized‎electrode‎was‎used‎as‎both‎the‎counter‎and‎

reference electrodes. When needed the light intensity was reduced with neutral density filters

(Newport). For each electrolyte tested, measurements were carried out on five samples prepared

identically.

4.3 Results and discussion

4.3.1 TiO2 electrode porosity and surface area

The results of nitrogen sorption shown in Figure ‎4.3 were used to evaluate the porosity and the

available surface area of the electrode. The specific surface area of the electrode was found by

using the Brunauer–Emmett–Teller (BET) model from a linear part of BET plot (P/P0 = 0.10–

Figure ‎4.3. Nitrogen adsorption/desorption isotherm of the mesoporous electrode.

65

0.30) to be 55 m2/g, and average pore diameter was calculated by the Barrett–Joyner–Halenda

(BJH) method from the desorption part of isotherm to be about 17 nm.32

The average pore size

evaluated from BJH analysis matches well with that estimated using SEM image and imageJ

software from top down images. The porosity, found to be around 42-45%. This value is close to

reported values elsewhere for mesoporous TiO2.33

4.3.2 Effect of process parameters on pore filling

To investigate pore filling in the nanostructured electrode, the influence of partial pressure of the

reactant and substrate temperature during iCVD were studied as these parameters are known to

determine reactant surface concentration through the relative pressure parameter, z = P/Psat,

where P is the reactant partial pressure and Psat is the reactant vapor pressure at the substrate

temperature.10

To see the effect of reactant surface concentration on the pore filling quality,

initially a set of experiments by changing the total reactor pressure in between from 60 to 200

mtorr was investigated. This was essentially done to see the effect of system pressure on the

amount of PHEMA polymer which has filled the pores during iCVD (all other conditions

remaining constant, with HEMA, TAPO and N2 flow rates each set at 0.4 sccm). Compared to a

clean cross section of mesoporous TiO2 before iCVD (Figure ‎4.4a), deposition at 60 mtorr yields

only a polymer coating forming on top of the electrode without any polymer being visible within

the porous network (Figure ‎4.4b). As total pressure increased above 80 mtorr, filling of the inner

pore volume starts taking place (not shown). At 125 mtorr, a completely pore filled cross section

together with a top overcoat is observed (Figure 4.4c). Further increase in total pressure to 200

mtorr results in the quality of fill deteriorating again, with evidence of partial filling within the

mesoporous layer and premature coating on top (Figure 4.4d).

66

Figure ‎4.4. Effect of total pressure on pore filling, showing cross sectional SEM of the 4 µm

thick TiO2 electrodes that are (a) uncoated, and after iCVD treatment at a total pressure of (b) 60,

(c) 125, and (d) 200 mtorr (scale bar=100 nm). Complete pore filling is observed at intermediate

pressures.

To separate the individual effects of monomer and initiator partial pressures on pore filling,

each reactant flow rate was varied while keeping the other constant and compensating with

nitrogen inert to maintain a constant total flow rate (1.2 sccm) at a fixed total pressure (125

mtorr). Considering the initiator, adjusting TAPO flow rate from 0.2 to 0.8 sccm reveals that

flows above 0.4 sccm give complete pore filling while lower flows yield partial filling with

greater filling fraction as flow increases (Figure ‎4.5c). This is to be expected since the initiator

67

contributes to the amount of initiated polymer chain radicals, and once sufficient initiator

concentration has been reached polymer growth becomes monomer limited. This trend has been

observed previously with iCVD polymerization of acrylate and methacrylate polymers.34-36

Likewise, considering the monomer, adjusting the flow rate between 0.2 and 0.6 sccm (while

keeping a fixed total flow of 1.2 sccm and total pressure of 125 mtorr) shows that the fill quality

initially improves with increasing monomer flow rate and after passing a fully filled state, pore

filling deteriorates again (Figure ‎4.5).

Figure ‎4.5. The individual effect of the initiator and monomer on pore filling of TiO2 electrodes,

showing lack of complete pore filling at initiator flow rates of (a) 0.2, and (b) 0.3 sccm, and at

monomer flow rates of (c) 0.2, and (d) 0.6 sccm (scale bar=200 nm). Complete pore filling is

observed at initiator flow rates of 0.4 sccm and higher, and at a monomer flow rate of 0.4 sccm.

68

This observation indicates that pore filling is much more sensitive to the monomer than the

initiator flow rate. As an important note, although the initiator and monomer flows and their gas

concentrations are comparable (which might suggest long chain polymer formation would not be

possible), the surface concentration based on surface adsorption of the initiator is actually much

lower than for the monomer due its higher vapor pressure (by a factor of 10). In addition, the

initiator is activated in the gas phase so we would expect only a fraction of the activated initiator

to actually reach the surface.

Thus, we attribute the observed complete pore filling at intermediate total pressures (Figure

‎4.4) to mainly the effect of the monomer. To explain this pore filling behavior, we consider the

process to be composed of three main steps: (1) diffusion of the reactants inside the mesoporous

structure; (2) adsorption and surface diffusion of the reactants; and (3) polymerization reaction at

the surface. Like all porous media, we expect mass transport and reaction kinetics to each play a

critical role in polymer growth inside the pores. Diffusion needs to take into account both gas

and surface diffusion. Under iCVD conditions in mesoporous TiO2, Knudsen diffusion

dominates gas phase transport (Knudsen number Kn=λ/dpore~104, where λ and dpore are estimated

to be 160 µm and 17 nm, respectively). Surface diffusion, on the other hand, requires an

understanding of the adsorption behavior within mesoporous materials, and experimental studies

have shown that surface diffusivity reaches a maximum at some intermediate surface coverage

of the adsorbate,33,34

a clue that this phenomenon might capture the observed trend in our work.

4.3.3 Understanding pore filling dynamics

To understand the observed behavior of the monomer on pore filling, the fate of the monomer

during iCVD must be considered. Monomer is transported from the gas phase within the

mesoporous TiO2 structure to the TiO2 surface, where it adsorbs and then reacts. Thus, monomer

69

diffusion both in the gas phase and on the surface as well as monomer reaction at the surface has

to be taken into account. Our hypothesis is that complete pore filling will require surface

reaction to be rate limiting i.e., diffusional transport resistance is small. To determine whether

this is the case, the time constants for diffusion (τD) and reaction (τR) have been determined as a

function of monomer relative pressure z, as shown in Figure 4.6. The time constants give a

relative measure of how fast each process is taking place at different monomer surface

concentrations. The diffusion time constant is estimated as follows:

τD 2

eff

(4-1)

where L is the mesoporous TiO2 layer thickness and Deff is the effective diffusivity of the

monomer that accounts for both gas phase and surface diffusion acting as parallel resistances to

transport:

eff g k s s (4-2)

where Dk and Ds are the monomer diffusion coefficients in the gas phase and on the surface,

respectively, and pg and ps are the fractions of the monomer occupying the gas phase and

surface, respectively. Interestingly, experimental data on diffusion into micro- and mesoporous

media have shown a maximum diffusivity at the intermediate partial pressures or surface

coverage. This phenomenon has been well reported in detailed reviews37

and attempts have been

made to model this behavior,38,39

which we have adopted here to calculate the transport

parameters:

k pore3

1‎ α

(4-3)

s 0 1‎ F

2

F

(4-4)

α ad

pore

‎= F

pore

(4-5)

70

F

1

(4-6)

g 1 αα

g s

‎ (4-7)

s 1 g‎ (4-8)

Here, Dk is the Knudsen diffusion coefficient since gas phase diffusion inside mesoporous

TiO2 is found to be dominated by Knudsen diffusion (Kn = λ/dpore ~ 104, where λ and dpore are

estimated as 160 µm and 17 nm respectively). The Knudsen diffusivity has been corrected by a

(1 − α) term to account for the decrease in pore size with adsorption, where α is the ratio of the

volume of monomer adsorbed relative to the entire pore volume available. To estimate α we

used α = ( FLVml)/Vpore, where Vpore is the total pore volume, Vml is the volume of a monolayer of

the adsorbed monomer, and FL is the surface coverage that is based on the generalized

Freundlich isotherm appropriate for less than a monolayer adsorption and thus applies to the low

z values of the iCVD conditions used here. Based on similar diffusion in mesoporous structures,

the isotherm parameters are taken as m = 0.2 and K = 3,40

and 0 is estimated to be 4.6x10

-5

cm2/s.

37 In addition, monomer (HEMA) properties used are: molecular weight MW = 130.14

g/mol, liquid density s = 1.071 g/cm3, and gas density g that is derived from the ideal gas law

and is a function of z. Here, T is the mesoporous layer temperature of 50 °C. At this temperature,

the reaction time constant is calculated as:

τR 1

p M (4-9)

where kp is the second order propagation reaction rate constant for HEMA polymerization

(reported as 2527 L/mol.s at 50 °C41

), and [M] = s FL/MW is the monomer surface concentration,

which is a function of z and calculated based on a similar reported approach.36

Now plotting τD

and τR as a function of z (Figure ‎4.6) reveals an intermediate z range in which monomer

71

diffusion becomes faster than reaction, and importantly this range corresponds to our

experimentally observed range over which complete pore filling is achieved.

At either lower or higher monomer concentrations, pore diffusion becomes limited, and this

leads to a lack of polymer growth within the pores and premature pore filling. With this

knowledge, we are able to control iCVD conditions to achieve complete pore filling successfully

for electrode thicknesses up to 12 µm (Figure ‎4.7 b), which is much thicker than is possible with

other reported procedures (~2 µm).16,17,35,36

Figure ‎4.6. Estimated time constants for monomer diffusion and reaction as a function of the

monomer‎relative‎pressure‎(solid‎line‎is‎τD and‎dashed‎line‎is‎τR).

72

Figure ‎4.7. Pore filling of TiO2 electrodes of different thicknesses of (a) 4 µm (scale bar=500

nm), and (b) 12 µm (scale bar=1 µm), the inset shows the bottom of the sample (scale bar=500

nm). By carefully controlling mass transport (gas and surface diffusion).

4.3.4 Pore filling quality

To provide a more quantitative measure of pore filling, cross sectional SEM was complemented

with TGA measurements to estimate the % pore filling, defined as the percent of the total pore

volume initially within the mesoporous TiO2 electrode (Vpore,L) which has been grown with

polymer by iCVD polymerization (Vpolymer,L), refer to Figure 4.8.

(4-10)

To‎estimate‎the‎%‎pore‎filling‎of‎the‎nominal‎12‎μm‎thick‎electrode, first the porosity (ε) of the

initial, unfilled TiO2 electrode layer is derived from the N2 adsorption and desorption isotherms.

From the total volume of the adsorbate (~136 cm3/g STP), ε is estimated to be 0.43-0.45 Then,

cross sectional SEM was performed on the polymer filled TiO2 electrode to obtain an average

value of the thickness of the TiO2 electrode layer (L) and the top overcoating (δ)‎as‎12.7±0.1‎μm‎

73

and 97 ± 5 nm, respectively. TGA was then used to obtain the total mass of polymer (mpolymer)

and TiO2 (mTiO2) since the polymer selectively decomposes at a much lower temperature (< 450

°C) while the TiO2 remains thermally stable (up to 800 °C),see Figure 4.8.

Figure ‎4.8. Thermogravimetric‎analysis‎of‎the‎12.6‎μm‎thick‎pore‎filled‎titanium‎dioxide. Cross-

sectional SEM showed 97±10 nm thick polymer overlayer.

By taking the density of the polymer (PHEMA) and TiO2 (P25) to be polymer = 1.09 g/cm3 and

TiO2 = 3.98 g/cm3, respectively, we can determine the total volume of the polymer (Vpolymer =

mpolymer/ polymer) and TiO2 (VTiO2= mTiO2/ TiO2). The actual polymer volume within the electrode

layer is then derived by subtracting the top overcoating volume from the total polymer volume,

while the TiO2 volume is used to derive the pore volume based on the calculated porosity. Thus,

the % pore filling can be written as:

(4-11)

74

Finally, by rewriting Equation 4-11 in terms of measured and literature values, the % pore filling

is calculated as:

(4-12)

Figure ‎4.9. The filled mesoporous of thickness L with δ polymer overlayer.

Thus, for the nominal 12 µm thick mesoporous TiO2 electrode pore filled with PHEMA polymer

electrolyte by iCVD, we estimate the % pore filling to be within the range of 92-100%. The

uncertainty is a result of the sensitivity range of the TGA and porosity measurements as well as

variability in the thickness and density values. Although the % pore filling can only be treated as

a rough estimate, the near 100% value indicates effective pore filling using the iCVD approach.

4.3.5 Polymer integrated DSSC performance

4.3.5.1 Effect of photosensitizer

Once proper conditions for pore filling were found, we applied iCVD and integrated PHEMA as

a polymer electrolyte in the dye sensitized TiO2 electrode. After incorporating the iodide-

triiodide redox couple in 50:50 propylene carbonate and γ-butyrolactone and assembling into a

75

complete quasi solid state DSSC, photocurrent-voltage measurements were made and compared

with a standard liquid electrolyte cell containing acetonitrile. As shown in Figure ‎4.10., although

the short circuit current density (Jsc) of the polymer electrolyte DSSC is reduced, the open circuit

voltage (Voc) is notably higher when compared with the acetonitrile liquid electrolyte cell. Now,

looking at cells of the same electrolyte across the two different dyes, the cells with dye 535 show

better performance than with dye 505. Dye 535 has a maximum light absorption that is red

shifted by about 50 nm when compared with dye 505.1 As a result, the incident photon to

electron conversion efficiency (IPCE) of the cell with dye 535 is superior to the one with dye

505. We therefore utilized dye 535 in the rest of this work.

Figure ‎4.10. Current-voltage characteristics of DSSCs fabricated with the polymer electrolyte

containing‎ 50:50‎ vol%‎ propylene‎ carbonate‎ and‎ γ-butyrolactone‎ (),‎ and with the standard

acetonitrile‎liquid‎electrolyte‎()‎incorporating‎ruthenium‎dyes‎(a)‎535‎or‎N3,‎and‎(b)‎505.

4.3.5.2 Effect of electrolyte

To further understand the resulting Voc enhancement in the polymer electrolyte DSSC, we

compared the current-voltage characteristics of this cell with liquid cells containing the same

76

redox solvent rather than acetonitrile. As shown in Figure ‎4.11, when either a 50:50 % of

propylene carbonate and γ-butyrolactone or pure propylene carbonate was used to incorporate

the iodide-triiodide redox couple, we see that Jsc does not change significantly between the liquid

and polymer electrolyte cells using the same solvent as it does between the acetonitrile and

polymer electrolyte cells. In contrast, Voc is again consistently higher in both cases for the

polymer cell when compared with their liquid counterparts. This implies that since Jsc is not

affected the increase in Voc cannot have come from a difference in ion conduction as a result of

changing from a liquid to a quasi solid state polymer electrolyte. We further compared polymer

cells with complete fill and partial fill showing premature blockage. As shown Figure ‎4.12 the

partially filled polymer cells again show little change in Jsc while the increase in Voc is now much

smaller than if the electrode were completely filled.

The Voc for the sample with complete pore filling is much higher when compared with the

partially pore filled, which is higher than that with the liquid electrolyte. The Jsc however

remains in the same range. This demonstrates the need for complete pore filling to derive the

maximum benefit of the polymer electrolyte in enhancing DSSC performance as a result of the

largest gain in Voc.

77

Figure ‎4.11. Effect of redox solvent on the performance of DSSCs fabricated with the polymer

electrolyte‎()‎and‎the‎corresponding‎liquid‎electrolyte‎()‎containing‎(a)‎50:50‎vol%‎propylene‎

carbonate and γ-butyrolactone, and (b) pure propylene carbonate. In each case, Voc is enhanced

while Jsc remains relatively unchanged in the polymer electrolyte cell.

Figure ‎4.12. Current-voltage characteristics of DSSCs containing 50:50 vol% propylene

carbonate and γ-butyrolactone‎pore‎filled‎polymer‎electrolyte‎(),‎‎partially‎pore‎filled‎polymer‎

electrolyte‎(),‎and‎liquid‎electrolyte‎().

78

4.3.5.3 Effect of photoanode thickness

To further elucidate the effect of PHEMA pore filling, we investigated the changes in DSSC

performance with increasing TiO2 thickness and again made comparisons between polymer

electrolyte and acetonitrile liquid electrolyte cells, as detailed in Figure ‎4.13 Based on current-

voltage measurements, power conversion efficiency η of each cell has been derived from Jsc and

Voc as well as its fill factor FF, defined as the ratio of the actual maximum power to the

theoretical derivable power (i.e., Jsc×Voc). Knowing the power density of the illuminated

light(Wo), cell efficiency has been calculated as η=(FF×Jsc×Voc)/Wo.

As expected, the standard liquid cell shows a considerably smaller Voc with thicker

electrodes, which is typically attributed to an increase in recombination sites for the redox

couple (triiodide) to recapture dye injected electrons from the TiO2 as more surface area is

available.42,43

Similarly, the increase in surface area with thicker electrodes leads to higher Jsc in

the liquid cell with the greater amount of photosensitization. In contrast, for the completely pore

filled polymer electrolyte cell, Voc and Jsc appear to be less sensitive to electrode thickness

especially as electrode thickness increases above 4 µm. The smaller effect on Voc could be due to

passivation of the TiO2 surface when in contact with the iCVD polymer as has been observed

with small molecules.43

However, it is clearly evident that Voc is consistently much higher for the

79

Figure ‎4.13. Comparison between the performance of DSSCs fabricated with a quasi-solid state

PHEMA polymer electrolyte containing 50:50 vol% propylene carbonate and γ-butyrolactone

(),‎ and‎ with‎ a‎ standard‎ liquid‎ electrolyte‎ containing‎ acetonitrile‎ ()‎ for different TiO2

thicknesses, showing (a) open circuit voltage Voc, (b) short circuit current Jsc, (c) fill factor FF,

and (d) power conversion efficiency η (error bar=1 SD). Polymer electrolyte DSSCs provide

similar efficiency as the liquid electrolyte cells with TiO2 electrode thicknesses that are nearly

three times thinner.

polymer cell compared to the liquid one at all electrode thicknesses we tested. There seems to be

a leveling off of Jsc which suggests some saturation effect, and could be a result of an increase in

diffusion resistance to ion conduction of the redox couple within the polymer matrix.

It is known that as the concentration of the iodide-triiodide redox species (using acetonitrile

as the redox solvent and liquid electrolyte) falls below some threshold, Jsc will plateau to a

limiting value with increasing light intensity.43

Likewise, in our case as shown in Figure ‎4.13, as

80

illuminated light intensity increases for the polymer electrolyte cell at constant electrode

thickness (~4 µm), Jsc becomes limiting at sufficiently low redox concentration. In contrast, for

the liquid electrolyte cell here, no plateau is observed for the range of redox concentrations

tested. The observed limiting behavior in the polymer cell can be attributed to the onset of

diffusion limited ion transport in the electrolyte due to a decrease in ion concentration gradient.43

Since the concentration gradient is influenced not only by the amount of redox species present

Figure ‎4.14. The effect of illuminated light intensity on short circuit current density of DSSCs

utilizing (a) liquid electrolyte, and (b) polymer electrolyte, each containing propylene carbonate

with different redox concentrations (in I2(M):LiI(M)) of 0.01:0.1 (∎), 0.08:0.8 (), and 0.2:2.0

() iodine and lithium iodide, respectively.

but also by the distance for ion transport, the diffusional limitation can be used to explain the

Jsc plateau observed in

Figure ‎4.14 for polymer cells when the TiO2 electrode thickness increases above 4 µm. Given

that the plateau is not observed for liquid electrolyte cells in all our cases, this suggests that ion

transport in these cells do not reach the threshold to become diffusion limited.

81

Since this behavior is only observed in the polymer electrolyte cell, the trends suggest

charge transport rather than charge generation becomes limiting, see Supporting Information for

more details. In Figure ‎4.13, although Jsc is lower, the significantly higher Voc together with a

relatively constant FF yield higher device efficiency for the polymer electrolyte cell compared

with the liquid acetonitrile cell at all the electrode thicknesses considered. In fact, the polymer

cell efficiencies remain fairly constant as a result of a lack of sensitivity of the device

parameters. As an important note, the difference in efficiencies observed between the polymer

and liquid electrolyte cells is solely due to the replacement of the liquid with the polymer, and

there has been no attempt to fully optimize cell performance. Significantly, the DSSC integrating

the iCVD PHEMA electrolyte can provide similar efficiency as the acetonitrile liquid electrolyte

DSSC using TiO2 electrodes that are a nearly factor of three thinner. Importantly, this implies

that there could be potential advantages when we consider a similar factor in reducing materials

usage, especially of the expensive dye, as well as the possibility of greater cell robustness,

particularly when thinking of applying the cell architecture on flexible substrates.

4.3.6 Electrochemical impedance spectroscopy

We therefore performed electrochemical impedance spectroscopy on the polymer and liquid

cells at their respective Voc to gain further insight. Figure ‎4.15 shows the Bode plots of the

polymer and liquid electrolyte DSSCs containing propylene carbonate, and illuminated under 1

sun corresponding to global AM 1.5 and 100 mW/cm2 irradiance. We observe a shift in the mid-

frequency peak towards lower frequency when the liquid electrolyte is replaced with the

polymer. This spectral peak frequency is a measure of the recombination time constant of the

electrons with the redox couple44

(triiodide) if we assume that charge recombination with the

oxidized dye molecules can be neglected due to fast reduction of the dye through the redox

82

couple.28

Thus, the shift to lower frequency suggests an increase in effective electron lifetime

from ~12 ms in the liquid cell to ~50 ms in the polymer electrolyte cell due to reduced

recombination with the electrolyte. We believe this decrease in charge recombination is the

cause for the substantially higher Voc. It is reasonable to expect that the contact of the quasi solid

polymer with the TiO2 electrode surface through complete pore filling can alter interfacial

properties. For example, charge transfer via surface trap states could be reduced by the polymer

Figure ‎4.15. Comparison between the Bode diagrams at their respective Voc of DSSCs fabricated

with‎ the‎ polymer‎ electrolyte‎ ()‎ and‎ the‎ corresponding‎ liquid‎ electrolyte‎ ()‎ containing‎

propylene carbonate. The shift of the mid-frequency peak to lower frequency of the polymer

electrolyte cell indicates a decrease in charge recombination at the electrolyte-electrode

interface.

blocking active sites on the TiO2 surface.45

Also, it is possible that Li+ ions (from the redox

species) could coordinate with the polymer matrix thereby limiting lithium adsorption on the

TiO2 surface, which would result in band edge movement and consequently an increase in Voc of

the cell.46

83

4.4 Conclusions

We have successfully utilized iCVD to integrate a polymer electrolyte into the mesoporous TiO2

electrode of the DSSC. By carefully controlling diffusion transport and surface kinetics during

iCVD, complete pore filling can be obtained. The model prediction showed a good match with

the observed phenomena. This results in significant enhancement in cell properties. The

efficiency of the quasi solid state PHEMA electrolyte DSSC is higher when compared with their

liquid electrolyte counterparts. We attribute the observed increase in open circuit cell voltage to

the suppression of electron recombination at the electrolyte-electrode interface as a result of the

exceptional pore filling achieved by iCVD. iCVD promises to be a valuable synthesis and

processing pathway which as demonstrated here has the potential to overcome the major

technological constraint of using a liquid electrolyte in the DSSC.

4.5 References

1. EIA, Annual Energy Outlook with Projection to 2035, Released 2010.

2. J. B. Baxter, Journal of Vacuum Science & Technology A, 2012, 30, 020801-020819.

3. Y. Chiba, A. Islam, Y. Watanabe, R. Komiya, N. Koide and L. Y. Han, Japanese

Journal of Applied Physics, 2006, 45, L638-L640.

4. A. Yella, H.-W. Lee, H. N. Tsao, C. Yi, A. K. Chandiran, M. K. Nazeeruddin, E. W.-G.

Diau, C.-Y. Yeh, S. M. Zakeeruddin and M. Grätzel, Science, 2011, 334, 629-634.

5. M. D. McGehee, Science, 2011, 334, 607-608.

6. M. Ikegami, J. Suzuki, K. Teshima, M. Kawaraya and T. Miyasaka, Solar Energy

Materials and Solar Cells, 2009, 93, 836-839.

84

7. U. Bach, D. Lupo, P. Comte, J. E. Moser, F. Weissortel,‎ .‎Salbeck,‎H.‎Spreitzer‎and‎M.‎

Grätzel, Nature, 1998, 395, 583-585.

8. .‎ ruger,‎R.‎Plass,‎M.‎Grätzel‎ and‎H. J. Matthieu, Applied Physics Letters 2002, 81,

367-369.

9. Q. B. Meng, K. Takahashi, X. T. Zhang, I. Sutanto, T. N. Rao, O. Sato, A. Fujishima, H.

Watanabe, T. Nakamori and M. Uragami, Langmuir, 2003, 19, 3572-3574.

10. D. Gebeyehu, C. J. Brabec and N. S. Sariciftci, Thin Solid Films, 2002, 403, 271.

11. Y. Saito, N. Fukuri, R. Senadeera, T. Kitamura, Y. Wada and S. Yanagida,

Electrochemical Communications, 2004, 6, 71-74.-274

12. F. Cao, G. Oskam and P. C. Searson, Journal of Physical Chemistry, 1995, 99, 17071-

17073.

13. R. Komiya, L. Y. Han, R. Yamanaka, A. Islam and T. Mitate, Journal of Photochemistry

and Photobiology A, 2004, 164, 123-127.

14. A. F. Nogueira, J. R. Durrant and M. A. De Paoli, Advanced Materials, 2001, 13, 826-

830.

15. J. H. Wu, Z. Lan, J. M. Lin, M. L. Huang, S. C. Hao, T. Sato and S. Yin, Advanced

Materials, 2007, 19, 4006-4011.

16. J. Wu, S. Hao, Z. Lan, J. Lin, M. Huang, Y. Huang, P. Li, S. Yin and T. Sato, Journal of

the American Chemical Society, 2008, 130, 11568-11569.

17. G. P. Bartholomew and A. J. Heeger, Advanced Functional Materials, 2005, 15, 677-

682.

18. B.‎O’Regan,‎F.‎ enzmann,‎R.‎Muis‎and‎ .‎Wienke,‎Chemistry of Materials, 2002, 14,

5023-5029.

19. .‎Schmidt-Mende‎and‎M.‎Grätzel, Thin Solid Films, 2006, 500, 296-301.

20. .‎H.‎ um,‎P.‎Chen,‎M.‎Grätzel‎and‎M. K. Nazeeruddin, ChemSusChem, 2008, 1, 699-

707.

21. S. Nejati and K. K. S. Lau, Nano Letters, 2010, 11, 419-423.

85

22. S. Pavasupree, S. Ngamsinlapasathian, Y. Suzuki and S. Yoshikawa, Material Research

Bulletin, 2008, 43, 149-157.

23. J. B. Asbury, N. A. Anderson, E. Hao, X. Ai and T. Lian, Journal of Physical Chemistry

B, 2003, 107, 7376-7386.

24. J. Nelson and R. E. Chandler, Coordination Chemistry Reviews, 2004, 248, 1181-1194.

25. B.‎O’Regan‎and‎M.‎Grätzel, Nature, 1991, 353, 737-740.

26. A. N. M. Green, R. E. Chandler, S. A. Haque, J. Nelson and J. R. Durrant, Journal of

Physical Chemistry B, 2004, 109, 142-150.

27. A. J. Frank, N. Kopidakis and J. v. d. Lagemaat, Coordination Chemistry Reviews, 2004,

248, 1165-1179.

28. A.‎Hagfeldt‎and‎M.‎Grätzel, Accounts of Chemical Research, 2000, 33, 269-277.

29. A. C. Fisher, L. M. Peter, E. A. Ponomarev, A. B. Walker and K. G. U. Wijayantha,

Journal of Physical Chemistry B, 2000, 104, 949-958.

30. H. Yang, M. Huang, J. Wu, Z. Lan, S. Hao and J. Lin, Materials Chemistry and Physics,

2008, 110, 38-42.

31. A. Kay, M. Nazeeruddin, I. Rodicio, R. Humpbry-Baker, E. Müller, P. Liska, N.

Vlachopoulos and M. Grätzel, Journal of the American Chemical Society, 1993, 115,

6382-6390.

32. S. S. Lowell, J.E.; Thomas M.A.; Thommes, M. , Characterization of Porous Solids and

Powders: Surface Area, Pore Size and Density, Springer, Dordrecht, Netherlands 2005.

33. G. L. Rumble, T.; Murakoshi, K., Electron transfer in nanomaterials, Proceedings of the

Electrochemical Society, 2004, 22, 126-136.

34. G. Ozaydin-Ince and K. K. Gleason, Chemical Vapor Deposition, 2010, 16, 100-105.

35. S. H. Baxamusa and K. K. Gleason, Chemical Vapor Deposition, 2008, 14, 313-318.

36. K. K. S. Lau and K. K. Gleason, Macromolecules, 2006, 39, 3695-3703.

86

37. J. G. Choi, D. D. Do and H. D. Do, Industrial & Engineering Chemistry Research, 2001,

40, 4005-4031.

38. R. Valiullin, P. Kortunov, J. Kärger and V. Timoshenko, Magnetic Resonance Imaging,

2005, 23, 209-213.

39. R. Valiullin, P. Kortunov, J. Karger and V. Timoshenko, Journal of Chemical Physics,

2004, 120, 11804-11814.

40. R. Valiullin, J. Karger and R. Glaser, Physical Chemistry Chemical Physics, 2009, 11,

2833-2853.

41. S. Beuermann and M. Buback, Progress in Polymer Science, 2002, 27, 191-254.

42. S. Ito, S. M. Zakeeruddin, R. Humphry-Baker, P. Liska, R. Charvet, P. Comte, M. K.

Nazeeruddin, P. Pechy,‎ M.‎ Takata,‎ H.‎ Miura,‎ S.‎ chida‎ and‎ M.‎ Grätzel, Advanced

Materials, 2006, 18, 1202-1205.

43. S.‎ .‎ Huang,‎ G.‎ Schlichth rl,‎ A.‎ .‎ Nozik,‎ M.‎ Grätzel‎ and‎ A. J. Frank, Journal of

Physical Chemistry B, 1997, 101, 2576-2582.

44. R. Kern, R. Sastrawan, J. Ferber, R. Stangl and J. Luther, Electrochimica Acta, 2002, 47,

4213-4225.

45. .‎Moser,‎S.‎Punchihewa,‎P.‎P.‎ nfelta‎and‎M.‎Grätzel, Langmuir, 1991, 7, 3012-3018.

46. G. Redmond and D. Fitzmaurice, Journal of Physical Chemistry, 1993, 97, 1426-1430

87

5 Integration of conducting polymers in supercapacitors

5.1 Introduction

With increased interest in more sustainable energy production and resources a challenging

question that remains to be answered is how to store the produced energy. Fossil fuels possess

very high energy densities that can be released by their burning. At the current moment there is

no alternative that matches their efficiency although their efficiency (Carnot efficiency) is not

phenomenal. Although the sun promises to be a clean energy resource, energy storage is

challenging given its intermittent availability. One possible answer to the energy storage

challenge is to store the energy in a similar fashion to fossil fuels, which is to utilize chemical

reactions and transformations of chemical bonds. This can be achieved through the use of redox-

active materials, which have the ability to be involved in electron transfer reactions.

Electrochemical storage is a way of storing generated power. Among different

electrochemical devices, batteries and capacitors are among the most studied systems and

currently for the huge portion of our needs for energy storage on a small scale we have relied on

batteries. Capacitors on the other hand when compared with batteries have much higher power

density but they suffer from significantly lower energy density. Comparing the electrochemical

capacitor and battery, it is worth pointing out the major differences that cause the observed

differences in their performance. In batteries, the electrochemical reaction of an ion with the

active electrode materials takes place within the bulk of the solid material and to reach the

charged state ions need to diffuse into the bulk solid. This is a slow process that makes its

charge-discharge far slower when compared with the electrochemical capacitor which store

charge only at the surface of the polarized electrodes in the form of an electrical double layer.

Nevertheless the charge being stored due to the screening of the electrical field of the electrode

within the solution is not significant in electrochemical capacitors as the double layer

88

capacitance dimension shrinks to the nm range.1 The fast charge-discharge of the

electrochemical capacitor when compared with a battery is the reason behind their higher power

density, however their total capacitance suffers from the larger dimension at which charge

storage must take place. Thus, there is a need to introduce new materials that have both high

power and energy densities. In the past twenty years, pesudocapacitors have been introduced that

incorporate battery-like redox reactions into electrochemical capacitors and there are hopes that

these material can act as an intermediate between batteries and capacitors or even outperform

both.1

An interesting aspect of conjugated polymers is their ability to be doped and therefore the

possibility to take part in redox reactions. The doping processes in conjugated polymers

correspond to oxidation in the case of p-doping or reduction in the case of n-doping. That means

the doping processes correspond to redox reactions in the polymer matrix. From an applications

point of view, it is important to know the phenomenological details of such redox reactions,

which potential range charging occurs and what is the maximum level of oxidation that can be

achieved before the material starts to degrade. Charge storage and transport in conducting

polymers were both of interest from the early stages of conducting polymer development.2 From

a practical point of view, a high charge storage capacity in these polymers is favorable for device

fabrication. Here, we explore the application of conducting polymers synthesized through oCVD

and integrated within nanostructured electrodes of supercapacitors.

5.2 Supercapacitors

Electrochemical capacitor is not a new field. It dates back to 1957 when Becker used a porous

carbon material to store charge in a double layer at the solid/electrolyte interface.3

Electrochemical capacitors however can rely on both an electrical double layer and

89

pseudocapacitance.1 Pseudocapacitance arises whenever there is a continuous charge that passes

faradaically with electrode voltage. In a sense pseudocapacitance is a phenomenon that take

place on the surface but has the nature of faradaic processes and is highly reversible. It passes

the charge through the double layer as in battery charging-discharging, but the capacitance arises

on account of special thermodynamic reasons between the extent of charge and the potential of

the electrode.1

Different materials have shown to possess high pseudocapacitance including some oxides

such as RuO2 , MnOx and conducting polymers.5,6

Among these, conducting polymers due to

their, low cost, flexibility, and ease of processing seems to be good a candidate. However

compared to oxides with cycle stability reported as high as 105 cycles,

7 conducting polymers still

need improvement. Another aspect of conducting polymers that has been reported lately is the

structural dependency of charge storage.8-11

It has been recently realized that the polymers when

spread into higher surface area by introducing nano- and mesostructured domains can actually

accumulate more charge.11

Here, we used oCVD as a powerful tool to synthesize high quality

polythiophene and integrated polymeric material within nanostructure electrode to investigate

the effect of nanostructure in polymer charge storage capacity.

5.3 Experimental

5.3.1 Polymer synthesis

To enable oxidative chemical vapor deposition, we have used an oCVD reactor system described

in detail elsewhere.12

Briefly, the reactor chamber was evacuated to base pressure (ca. 5 mtorr)

using a dry vacuum pump (Edwards Vacuum). Monomer, thiophene (97%, Sigma Aldrich), and

oxidant initiator, antimony pentachloride (99%,Across organic), were used as received and

metered independently from glass source vessels into the chamber using precision metering

90

valves (Swagelok). The initiator was heated up to 50 °C to achieve sufficient vapor pressure and

its temperature was kept constant using a temperature controller (Omega Engineering). The

monomer had sufficient vapor pressure at room temperature and was not heated. The chamber

pressure was measured with a pressure transducer (MKS Instruments) and automatically

maintained by using a downstream throttle valve connected to a pressure controller (MKS

Instruments). The substrate temperature was kept constant through backside cooling of the

reactor stage by using a recirculating chiller (Thermo Scientific Neslab).

In order to tune the oCVD polymerization reaction and synthesis chemistry at the surface,

the ratio of the reactant (monomer, initiator) partial pressure to its saturated vapor pressure at the

temperature of the substrate (i.e., Pr/Pr,sat) was carefully adjusted and controlled (see Discussion).

Thus, pressures ranging from 0.4-2 torr, and monomer and initiator flow rates of 0.5-5 and 0.1-

0.5 sccm (standard cm3/min), respectively, were studied. The substrate temperature was set at 30

°C. Polythiophene films were deposited on various substrates, including fluorine-doped tin oxide

glass (15/,‎Hartford‎Glass),‎silicon‎wafers‎(WRS‎Materials),‎microscope‎glass‎slides‎(Fisher‎

Scientific), and quartz glass (Chemglass) and within porous structure of TiO2 and anodized

aluminum oxide membrane (Anodisc, Whatman).

5.3.2 Polymer characterization

Fourier transform infrared spectra (FTIR) were acquired on a Thermo Nicolet 6700 spectrometer

in normal transmission mode using an MCT/A detector at a resolution of 4 cm-1

averaged over

64 scans. UV-vis spectra of deposited films on quartz glass were acquired between 280-800 nm

with 1 nm resolution using a Shimadzu UV-1700 spectrophotometer. X-ray photoelectron

spectroscopy (XPS) was performed on a Physical Electronics PHI 5000 VersaProbe with a

scanning monochromatic source from an Al anode and with dual beam charge neutralization.

91

Survey XPS spectra were acquired at 100 W with pass energy of 117 eV over the range of 0-

1100 eV with 1 eV resolution and 100 ms dwell time, and averaged over 2 scans. High

resolution XPS spectra of C1s, O1s, Sb3d, Cl2p and S2p core electrons were acquired in high

power mode of 100 W with a pass energy of 11.5 eV for C1s, Cl2p and S2p and 23.5 eV for O1s

and Sb3d using different acquisition times chosen based on the observed intensity of the

elements from the surveys. Raman spectra were collected on a Renishaw RM1000

microspectrometer using an Ar ion laser 488 nm with ~1 µm lateral spot size and 11 mW total

power and He-Ne laser 633nm with ~1 µm lateral spot size and 10 mW total power. Cyclic

voltammograms were recorded in a three electrode setup under a nitrogen blanket with a Gamry

Reference 600 potentiostat. The polythiophene samples deposited on FTO glass served as the

working electrode while a 2.5x2.5 cm platinum gauze (Princeton Applied Research) was used as

the counter electrode. The silver reference electrode (Princeton Applied Research) was filled

with 0.1 M silver nitrate (99.9999%, Sigma Aldrich) and 0.1 M tetraethylammonium perchlorate

(electrochemical grade, Sigma Aldrich) in acetonitrile (ACS grade, Sigma Aldrich). A

supporting electrolyte of 0.1 M tetrabutylammonium hexafluorophosphate (electrochemical

grade, Fluka) in acetonitrile was bubbled for 1 h with nitrogen prior to use. The potential was

swept between -0.4 and 1.2 V vs. Ag/AgNO3 with a sweep rate of 80 mV/s. The charge storage

capability of the samples deposited on planar FTO glass, and within mesoporous titanium

dioxide and activated carbon electrodes were tested by performing cyclic voltammetry in a three

electrode setup using organic salts e.g. tetraethylammonium tetrafluoroborate in acetonitrile as

the electrolyte, an Ag/AgCl reference electrode. The reference was calibrated using Fe+/Fe in

acetonitrile using a glassy carbon electrode as the working electrode and a platinum mesh as the

counter electrode. Polythiophene film conductivity was estimated through measuring sheet

resistivity using an Alessi four point probe connected to a Keithley 2400 source meter.

92

5.3.3 TiO2 electrode fabrication

The titanium dioxide colloidal paste consisting of TiO2 nanoparticles P25 (Envonik) was

prepared using a published procedure.13

The prepared paste was spin coated at 800 rpm onto

fluorine‎doped‎tin‎oxide‎conductive‎glass‎(Hartford‎Glass,‎15‎Ω/) to obtain a film of the paste

between 4 µm in thickness. To create the mesoporous structure, the spin coated TiO2 paste was

heated up to 500°C and annealed for 30 min.

5.3.4 Activated carbon electrode fabrication

Activated carbon electrodes were fabricated in two different ways and tested on two different

current collectors in two different configurations. Activated carbon particle YP-50 (Kurary inc)

with 1500-1700 cm2/g and average particle size in between 5-20 µm were used as received. A

suspension of activated carbon in ethanol % 0.5 w/w was made and mixed for 2 h at room

temperature. 50 µl of this solution were pipetted in between two piece tap placed parallel on the

FTO and treated stainless steel and allowed to dry in the oven for 24 h at 100 °C and were kept

in the oven at room temperature till they were used for electrochemical measurements.

The activated carbon mat was fabricated according to published protocol14

using YP-50 and

%5.0 binder (% 60 aqueous solution of PTFE) to obtain 100 µm thick electrodes. Briefly the

particles were mixed with the ethanol and then binder was added drop-wise while stirring the

mixture. The whole mixture was heated for 2 h at 60 °C and semi-dry pastes were spread on a

piece of glass and kneaded till a gum-quality paste was obtained. The mat was dried in the oven

at 120 °C for 24 h and stored under an inert atmosphere till they were further used.

93

5.3.5 Current collectors

As for FTO, the glass slides were sonicated in dilute detergent solution (Citranox) and

thoroughly rinsed in deionized water and sonicated again in ethanol for 10 min followed by

heating to 400 °C to remove any contamination from the surface caused by solvent adsorption.

They were then stored in an inert environment until they were further used.

For the stainless steel plate, 0.5 mm thick stainless steel plates (McMaster Carr) were cut to

size. The plates were mechanically roughened using 400 grit SiC paper and then washed and

rinsed in DI water and sonicated in ethanol for 10 min. The surface of the current collectors were

etched in a solution of 1 M NaOH (Aldrich) for 10 min at room temperature and then rinsed with

DI water and immediately dipped into a 1 M HCl ( Aldrich) at 80 ºC for 10 min. The current

collectors were rinsed with water and dried in vacuum oven at 80 ºC for 15 min and kept under

vacuum until further use.

For the Swagelok setup the electrode set up was fabricated as described elsewhere.15

Two

stainless steel rods (3/8”,‎McMaster Carr) were used as the electrodes and their surfaces were

roughened with 40 grit SiC and then rinsed thoroughly with water and sonicated in ethanol for

20 min. The cleaned electrodes were dried in vacuum oven for 20 min at 80 °C and then were

transferred to a sputter coater to be coated with 5 nm of Pt coating.

5.3.6 oCVD polythiophene integration

As we have shown, polymer integration within nanostructured electrodes can be obtained by

carefully tuning the CVD parameters to be under a reaction-limited regime.16

The effect of

various oCVD parameters was studied first using a well-defined pore geometry of the AAO

membrane to determine the conditions at which conformal polythiophene could be achieved on

the inner walls of the pores. Pressure was varied to obtain a range at which the inner wall is

94

coated without premature pore blockage. So at 30 °C stage temperature, and 0.5, 5 and 2 sccm of

initiator, monomer and N2, respectively, we have changed pressure from 200 mtorr to 2 torr and

obtained control over coating inside the structure as thick as 40 nm before the pores were

blocked at the top. We then adjusted each reactant flow rate at a time to find the reaction-

limiting conditions. Ar was also used instead of nitrogen to eliminate any possible reaction of

nitrogen during deposition. This approach was repeated for the conformal deposition on

mesoporous TiO2 and activated carbon. The proper oCVD conditions for conformal coating

were different depending on the nature of the pores in each substrate.

5.3.7 Polymer weight measurement

Weight measurements for the polymer on planar substrates were done on a precision lab scale.

For very thin films (< 100 nm), due to the lack of sensitivity of the scale, thickness was instead

measured by cross-sectional SEM and related to weight based on a thickness-weight calibration

made with thicker films. For the polymer deposited within the nanaostructure of TiO2,

thermogravimetric analysis (TGA) was performed to estimate the mass of polymer. For the

polymer within the activated carbon, the mass was estimated from added mass to the support

using the lab scale.

5.4 Results and discussion

The oCVD synthesized polythiophene were deposited on different substrates. The films adhered

well to the underlying substrates and none of the common solvents we tested, including

tetrahydrofuran, chloroform, methanol, and methyl formamide, were able to dissolve the films.

Electrical conductivity reached values as high as 50-70 S/cm.

95

5.4.1 Vibrational spectroscopy

Figure ‎5.1 shows the FTIR spectrum of the as-deposited polythiophene. The strong vibration

band in between 1100 and 1500 cm-1

can be assigned to the doping induced vibrations of

polythiophene.17,18

The peaks observable at 790 and 690 cm-1

are characteristic of C–H out-of-

plane bending of the thiophene ring.19

The C–H stretches located above 3000 cm-1

also indicate

that the aromatic thiophene ring is preserved. Comparing this FTIR spectrum with the one

obtained for PTh deposited using VOCl3 as the oxidant, the first striking difference is the

absence of any oxidant residue peak at 900 cm-1

and between 1600-1700 cm-1

. This could

suggest less defects in the film as there is no peak associated with the oxidant residue in the film.

Figure ‎5.1. FTIR spectrum of an as-deposited doped film on silicon.

96

Since undoped polythiophene films exhibits very strong visible absorption bands related to

the π-π* electronic transition, whose maximum is located between 470-520 nm, 488 nm

excitation allowed these peaks to be enhanced and therefore indicative of the native polymer. In

addition, Raman spectra were acquired using 633 nm laser excitation to reduce the resonance

enhancement of the native state so that the effect of doping can be more clearly observed, see

Figure ‎5.2. The strong bands in between 1000-1300 cm-1

are indicative of highly doped polymer

when compared with the Raman spectrum of a washed, dedoped polymer. These results are in

agreement with previous reports20

on the doped polythiophene excited with the 633 nm beam

and suggest that the oCVD deposition of polythiophene using SbCl5 oxidant generated a highly

doped state.

To investigate the effect of oCVD synthesis parameters on polymer quality, the oxidant

concentration was adjusted by changing its flow rate from 0.1 to 0.5 sccm while keeping other

conditions constant (total pressure (P) =800 mtorr, monomer flow rate FM = 2 sccm, stage

temperature (Ts) = 30 °C). The Raman spectra of the polythiophene deposited at three different

condition that resulted in different conductivity ranging from 10 to 50 S/cm are shown in Figure

‎5.3. As seen, the region of interest that include the intra C=C vibration corresponding to the

quinoid structure shifts from 1450 to 1420 cm-1

as the conductivity increased. The film deposited

at 0.3 sccm of oxidant seems to have the maximum conductivity and as the oxidant

concentration is increased further the conductivity is reduced as shown by the shift of the

quinoid band to higher wavenumbers.

97

Figure ‎5.2. Raman spectra of PTh deposited on silicon, excited using (a) 488, and (b) 633 nm

laser beams.

Figure ‎5.3. Raman spectra of polythiophene deposited at different oCVD condition. The shift in

quinod peak to a lower wavenumber is concomitant with an increase in film electrical

conductivity.

98

5.4.2 UV-vis spectroscopy

UV-vis spectroscopy also confirms that the polythiophene films deposited using the new oxidant

has better structural properties as evidenced by the longer conjugation length that is indicated by

the red shift in the maximum absorption peak (max) to 515 nm compared to 480 nm for the film

deposited using the VOCl3 oxidant. Once again, by changing oCVD synthesis conditions film

properties can be tuned. Figure ‎5.4 shows the UV-vis spectrum of the deposited polymer on

quartz glass and the shift in absorption maximum as a result of changing the oxidant

concentration in the gas feed

Figure ‎5.4. UV-vis spectrum of an as-deposited film after dedoping the film to its undoped form

(by exposure to 2.0 M methylamine in methanol for 2 min followed by washing with neat

methanol). Inset shows the control over absorption spectrum by changing oCVD parameters, the

increase in conjugation length obtained by reducing the oxidant concentration.

99

5.4.3 Cyclic voltammetry of the deposited film on FTO

Figure ‎5.5 shows the cyclic voltammogram of an as-deposited polythiophene film measured in

an electrolyte of 0.5 M tetraethylammonium tetrafluoroborate (TEABF4) in acetonitrile. The p-

doping (oxidation) and dedoping (reduction) peaks are clearly observed at 0.8 and 0.25 V vs.

Ag/AgCl, respectively, and are accompanied by a color change from red in the undoped state to

deep blue in the doped form. The locations of these peaks match reported values of p-doping and

dedoping peaks for unsubstituted polythiophene.21

Figure ‎5.5. Cyclic voltammogram of an as-deposited film on FTO glass.

5.4.4 Polymer integration within nanostructures

Chemical vapor deposition is well-known to produce conformal coatings on topographically

complex substrates. Even with porous structures where the aspect ratio (length-to-diameter) of

100

the pores can exceed 1000:1 and pore sizes are on the nanometer scale, CVD is able to achieve

uniform and conformal coating on the pore walls,22

and if so desired, the pore spaces can even be

entirely filled.16

Here, we demonstrate a unique way to integrate unsubstituted polythiophene

within porous nanostructures with exceptional control over film thickness and coating

conformality by simple adjusting of oCVD synthesis parameters. We have chosen to deposit

within anodized aluminum oxide (AAO) membranes having well-defined pore geometries in

order to investigate the oCVD parameter space that leads to effective polymer integration. By

reducing total pressure and oxidant concentration so as to be in the reaction-limited regime that

favors mass transport inside the pores, we obtained conformal growth of the polymer within

porous nanostructures with average deposition rates as low as 0.5 nm/min (deposition rate was

estimated by taking the ratio of SEM-measured cross sectional film thickness and film

deposition time).

Figure 5.6 shows 57 µm thick AAO membranes with a nominal pore diameter of 200 nm,

comparing the white uncoated membrane (Figure ‎5.6a) to the polythiophene coated one which in

the as-deposited doped form has a dark grayish blue color (Figure ‎5.6b), and the polythiophene

coated disc after dedoping to yield the orange-red undoped state (Figure ‎5.6c). Cross-sectional

scanning electron microscopy (SEM) images show that the vertically aligned pore channels of

the AAO membrane (Figure ‎5.6d) have been conformally coated with polythiophene, in this case

with a film thickness of ~30 nm (Figure ‎5.6e). This uniform coating is observed throughout the

entire thickness of the AAO membrane and demonstrates oCVD's ability to create conformal

films along pore walls inside porous nanostructures by manipulating the rate of polymerization

at the surface relative to mass transport dynamics. It should be noted that with faster reaction

rates at higher oxidant flow rate, preferential deposition on the outer membrane surface rather

than inside the pores led to non-conformal coating and eventually pore blockage (not shown).

101

Figure ‎5.6. Conformal coating of polythiophene within porous nanostructures using oCVD. (a)

Uncoated anodized aluminum oxide (AAO) membrane, 57 µm thick and 200 nm pore diameter.

(b) Polythiophene coated AAO in the as-deposited doped state of the polymer. (c)

Polythiophene coated AAO in the undoped state of the polymer after dedoping. (d) Cross-

sectional SEM of an AAO membrane showing the porous channels (darker shade). (e) Cross-

sectional SEM of a polythiophene coated AAO membrane showing conformal and uniform. The

conditions used for filling AAO without mass transfer limitations up to 40 nm thick film on the

inner wall was as follow: 30 °C stage temperature, and 0.1, 2 and 2 sccm of initiator, monomer

and N2, and 800 mtorr total pressure.

We have applied our knowledge of oCVD to enable in-situ polymerization and conformal

coating of PTh within the nanostructure to mesoporous TiO2 supported on TO glass. As seen in

Figure ‎5.7, again by adjusting oCVD parameters we were able to deposit polymer within the

nanostructure in a conformal fashion. Since the pore structure was more tortuous and the aspect

ratio was even higher when compared with the AAO disk, we have increased FM from 2 to 5

sccm while reducing the total pressure to 500 mtorr as this allowed the rate of reaction to be

controlled within the TiO2 pores more effectively.

102

Figure ‎5.7. Polythiophene integrated TiO2 electrodes. (a) Undoped PTh (by exposure to 2.0 M

methylamine in methanol for 2 min followed by washing with neat methanol). (b) Cross-

sectional SEM of the mesoporous electrode with conformal PTh coating (scale bar is 200 nm).

5.4.5 Effect of nanostructure on charge storage

With the ability to create conformal coatings inside porous nanostructures, we turned our focus

to study the effect of a porous nanostructure on the electrochemical behavior of polythiophene

films. Here, oCVD was used to deposit polythiophene films on planar FTO substrates as well as

inside 4 µm mesoporous TiO2 layers supported on FTO substrates. The TiO2 layers consisted of

a network of 25 nm TiO2 interconnected nanoparticles spin coated from a suspension onto FTO

glass and annealed at 450 °C to remove the solvent, resulting in a mesoporous network structure.

Conformal coating inside mesoporous layers as thick as 4 µm was obtained at a slow deposition

rate of 1 nm/min (based on estimating an average film thickness by knowing the mass of

deposited polymer, the total available substrate surface area, and assuming a polymer density16

of 1.35 g/cm3). Before performing any electrochemical tests, the samples were soaked in

tetrahydrofuran (THF) for 3 h and dried in a vacuum oven for 8 h to remove THF completely.

103

The electrochemical charge storage capacity of the polymer films deposited on planar electrodes

and within porous nanostructured electrodes was then investigated in a three electrode setup

using cyclic voltammetry. Voltage was swept over a potential window of 1.5 V (–0.5 to +1.0 V)

vs. Ag/AgCl in a liquid electrolyte of 0.5 M TEABF4 in acetonitrile and using an oversized

activated carbon as the counter electrode. The cyclic voltammogram was used to derive the

specific capacitance of the polymer using the following equation:

(5-1)

where α is the scan rate, m only considers the mass of the polymer, i and V are the current and

voltage on the voltammogram,‎ and‎ ΔV is the window over which cyclic voltammetry was

performed. The TiO2 porous layer itself was found to contribute minimal capacitance (< 1 F/g)

so the capacitance measurements are expected to only probe the effect of the polymer. The

cyclic voltammograms at different scan rates of polythiophene integrated into TiO2 are shown in

Figure ‎5.8.

Figure ‎5.8. Cyclic voltammograms at different scan rates of PTh deposited within mesoporous

TiO2 electrodes recorded in a three electrode set up versus an Ag/AgCl reference electrode.

104

Using the cyclic voltammograms we estimated the specific capacitance of the polymer knowing

the mass of the sample deposited on the electrodes. Figure ‎5.9a gives the specific capacitance

(per mass of the polymer) for polythiophene thin films deposited on planar substrates and inside

mesoporous TiO2 nanostructures. Clearly, there is a dependence of the specific capacitance on

polymer thickness both on the planar substrate (250 and 800 nm) and within the nanostructure (4

and 6 nm). Film thickness of polythiophene on the flat substrates was measured directly through

cross-sectional SEM while that inside the nanostructures was estimated by measuring the mass

gain due to the polymer after oCVD using thermogravimetric analysis (TGA) and relating it to

polythiophene density (1.2 g/cm3)

23 and the nanostructure specific surface area (55 m

2 /g,).

16 For

polythiophene deposited on planar substrates, the specific capacitance of thin films (down to 50

nm as the minimum film thickness practically possible for electrochemical measurements) over

the 1.5 V window was found to be 150 F/g and for thicker samples (> 450 nm) this value was

reduced to 75 F/g. To understand this difference, the anodic peak current taken from the cyclic

voltammogram was traced as a function of scan rate, see Figure 3b. For a planar thin film (250

nm), the linear response is characteristic of a reaction limited behavior in the redox process. In

contrast for a thick film (800 nm), the deviation from a linear response indicates that ion

diffusion and penetration inside the polymer becomes limiting. This is confirmed by the linear

relationship obtained by plotting the peak current with the square root of the scan rate (not

shown) predicted for the ion diffusion limited case, and which allows an estimate of the BF4– ion

diffusion coefficient of 8.9x10-9

m2/s to be obtained.

24 This value is in the range of diffusion

coefficient values reported for thiophene polymers in planar geometry.25

The difference in

specific capacitance between thin and thick polythiophene films is also seen for the case of the

polymer deposited within the mesoporous TiO2 nanostructures. As the polymer thickness

increased from 4 to 6 nm, the latter being close to the limit at which the pores become filled, the

105

specific capacitance was significantly reduced from 275 to 75 F/g, see Figure ‎5.9a. By plotting

the anodic peak current as a function of scan rate, as seen in Figure ‎5.9c, again the thicker

coating this time within the nanostructure shows that the redox process is highly diffusion

limited. Plotting the current with the square root of the scan rate yields a linear behavior for the

ion diffusion limited process, leading to a diffusion coefficient for BF4– ions to be 8.2x10

-8 m

2/s,

which is an order higher than that for the dense, thick planar polythiophene film but is

comparable to that reported for the diffusion of ions in a liquid electrolyte within nanoporous

electrodes.26

This suggests that diffusion in this case is limited by ion diffusion in the liquid and

not in the polymer, presumably due to the much narrower pore channels with the thicker

polymer coating inside the porous nanostructure.

What is more important with the results in Figure ‎5.9a is the significant enhancement in

specific capacitance when a thin polythiophene film is inside the mesoporous TiO2 layer

compared to a thin film on a planar substrate, where in both cases ion diffusion is not limiting. If

the thin polythiophene film inside the nanostructure were solely dictated by reaction controlled

redox processes, then the expected specific capacitance (per mass of polymer) should be similar

to that of the thin planar film. However, the measured capacitance for the film inside the porous

nanostructure is almost 1.6 times higher. The data suggests that the three-dimensional

nanostructured pore surface adds a significant component to the electrochemical behavior of

polythiophene that is not available on a two-dimensional planar surface. There are reports to

indicate that a nanoconfinement effect can significantly increase the electrical double layer

capacitance (EDLC) in nanoporous electrodes. This has been attributed to enhanced collision

frequency of ions with the pore walls as ions diffuse within a porous media (analogous to gas

phase Knudsen transport) as well as enhanced charge screening with the electric field

distribution inside the nanoporous structure.27

This increase in double layer capacitance due to

106

nanoconfinement has been shown to be a strong function of roughness (defined as the ratio of

actual to geometric surface area) and ion concentration in the bulk.28

However, for our case, the

enhanced charge capacity should come from Faradaic reactions of the polythiophene rather than

an electrical double layer since the bare TiO2 nanostructure had negligible measurable

capacitance. There is the possibility that the polythiophene might have a double layer

capacitance component that could lead to the nanoconfinement enhanced capacity as observed in

other double layer nanoporous structures since we see a flat rectangular portion in the cyclic

voltammograms (not shown) that suggest some double layer contribution (see below). It is also

possible that the faradaic processes in polythiophene could be augmented in a nanoconfined

polymer. The increase in ion collision frequency, the greater local ion concentration and the

changes in electric field distribution within the pores might help the ions to interact more

effectively with the polymer chains. The thinness and curvature of the polymer film might also

open additional sites for ion incorporation. These factors could lead to a greater amount of

charge being transferred during doping and dedoping. It is also possible that an increase in

faradaic current could arise from more stable oxidation of the polymer. In the bulk, further

oxidation can be achieved at higher electrode potentials,29

however in the case of unsubstituted

polythiophene this easily leads to over-oxidation in the presence of presence of nucleophillic in

the media that result in polymer degradation.30

However, with the polymer confined in the

nanopores, more oxidation could potentially be achieved without irreversible chemical

degradation since the pores might prevent these solvent nucleophiles from interacting with the

polymer.31,32,27

Although the reasons for enhanced capacity is not entirely understood, the effect

of a porous nanostructure on enhancing the specific capacitance of ultrathin polythiophene is

clearly observed.

107

Figure ‎5.9. Effect of polymer thickness and 3D nanostructure on the specific capacitance of

oCVD polythiophene. (a) SEM images (scale bar is 200 nm) and specific capacitance values of

thin (250 nm) and thick (800 nm) films on planar FTO electrodes, and of thin (4 nm) and thick

(6 nm) films inside mesoporous electrodes of TiO2; specific capacitance is based on per mass of

polymer and reported with 2 standard deviations. The current density of the anodic peak (doping

peak) as a function of scan rate for different thickness of films on (b) planar substrates and (c)

within mesoporous nanostructures.

108

5.4.6 Polymer integrated supercapacitors

Given the ability to significantly enhance charge storage with ultrathin oCVD polythiophene

within porous TiO2 nanostructures, we investigated further by using oCVD to integrate

polythiophene thin films within the nanostructure of activated carbon electrodes to create

supercapacitors. The activated carbon mat with 100 µm thick electrodes was cut into 1 cm2 size

for oCVD coating and electrochemical characterization was performed in a two electrode set up.

Unsubstituted polythiophene was then conformally coated within the porous activated carbon by

oCVD at various deposition times using a constant set of oCVD deposition conditions (1 Torr

pressure, 0.3 sccm oxidant flow rate, 2.5 sccm monomer flow rate, and 30 °C substrate

temperature) to obtain electrodes with different mass loadings of the polymer within the matrix.

Figure ‎5.10 show the top-down SEM images of the activated carbon particles in the fabricated

electrodes before and after oCVD with several mass loadings of polymer to activated carbon

(1:1, 1.5:1, and 2.7:1). As seen in Figure ‎5.10a, uncoated activated carbon particles have a very

rough surface consisting of microscale topologies and mesopores. With an increasing amount of

polythiophene as in Figure ‎5.10b and c, the surface morphology becomes noticeably smoother,

and eventually at high enough loading of the polymer as in Figure ‎5.10d the coating is sufficient

to cover the microscale features and fill in many of the pores.

The electrochemical performance of the fabricated electrodes after polythiophene dedoping

and THF washing and dried, were tested in a symmetric two electrode Swagelok cell15 using 1

M TEABF4 in acetonitrile as the electrolyte. A voltage range of –2 to 2 V was found to be an

appropriate window over which the fabricated electrodes were stable (this was based on separate

cyclic voltammetry measurements made beforehand in a three electrode setup where the

electrodes were used in a symmetric configuration and the voltage on the electrode was probed

independently vs. an Ag/AgCl electrode).33

Specific capacitance of the assembled Swagelok

109

cells was derived from cyclic voltammetry and using Equation 5-1, the only difference being

that the total mass of the electrode (including the polymer, activated carbon and binder) was

used in the calculation. As shown in Figure ‎5.10e, specific capacitance values of oCVD

polythiophene coated activated carbon pseudocapacitors are highly dependent on the mass

loading of polymer. In particular, at a polymer-to-activated carbon mass ratio of ~1.5, specific

capacitance reaches a maximum, a value ~50% higher than that for bare activated carbon (145

vs. 92 F/g). More significantly, this capacitance translates to over a 250% increase in volumetric

capacitance since the volume contribution of the ultrathin polymer coating is negligible (120 vs.

47 F/cm3). Figure ‎5.10f shows cyclic voltammetry sweeps of the polythiophene coated activated

carbon at the 1.5:1 mass ratio over a 4.0 V window at different scan rates. The dependency of

the peak current value upon the scan rate is not entirely linear but there is no sign of diffusion

limitations up to 100 mV/s. The pseudocapacitor is shown to be stable up to the 5000 cycles

tested with a small 5% capacitance drop during the first 100 cycles, see Figure ‎5.10g.

Looking back at each cyclic voltammogram in Figure ‎5.10f, we observe that the signal has

both characteristics of a pure electrical double layer capacitor (EDLC) and faradaic processes

associated with a pseudocapacitor. The flat portion of the curves around 0 V is characteristic of

the rectangular plot typically obtained with an EDLC such as activated carbon while the peaks

noticeable towards either edge of the voltage window relate to the redox reactions of a

pseudocapacitor like polythiophene. This suggests that the ultrathin polythiophene coating did

not completely block access to pores so that a considerable portion of the activated carbon

surface is still available for ions to dock at the surface and contribute to charge storage in

electrical double layers. This area could be mainly from the micropores that are inaccessible to

oCVD for conformal polymer coating and therefore activated carbon remains accessible to ions.

However, when the polymer coating becomes too thick, these pores no longer become accessible

110

and further; the surface features in activated carbon become more rounded and ill-defined. This

could lead to lower capacitance due to pore blockage, less surface area and even the onset of

diffusion limitations. This could explain the behavior observed in Figure ‎5.10e, in which the

maximum capacitance observed could be understood as an initial increase in the contribution

from pseudocapacitance with having more redox active polymer to an eventual decrease in

capacitance due to a drop in surface area and pore blocking with thicker polymer coatings.

Figure ‎5.10. Electrochemical behavior of oCVD polythiophene coated activated carbon. SEM

images of activated carbon electrodes (a) without coating, and coated with polythiophene at

polymer-to-activated carbon mass ratios of (b) 1:1, (c) 1.5:1, and (d) 2.7:1 (insets show a

magnified region of the surface morphology of each sample; scale bar is 1 µm and 50 nm for the

inset). (e) Specific capacitance values based on per total mass of electrode for different polymer

mass loadings. Measurements were made at 100 mV/s for five different samples at each loading

with the error bars representing two standard deviations. (f) Cyclic voltammograms of the 1.5:1

mass ratio pesudocapacitor recorded at different scan rates between –2 and 2 V. (g) Capacitance

of the same sample measured up to 5000 cycles (at 100 mV/s ).

111

5.5 Conclusions

We have demonstrated oCVD's unique ability to synthesize unsubstituted polythiophene based

on the direct vapor-to-solid oxidative polymerization of thiophene monomer using antimony

pentachloride oxidant. Ultrathin polythiophene films have been successfully integrated within

high aspect ratio porous nanostructures of AAO, TiO2 and activated carbon by operating oCVD

under slow kinetics that favor reactant mass transport. Significantly, ultrathin conformal

polythiophene coatings that preserve the high surface area of the porous nanostructures showed

enhanced capacitance 1.6 times over that of planar films that could be attributed to

nanoconfinement effects not available in open planar geometries. As a result, type I symmetric

pseudocapacitors consisting of polythiophene coated activated carbon electrodes displayed

significantly greater energy storage capacity compared to bare activated carbon, with an increase

of 50% and 250% in specific and volumetric capacitance, respectively, at an optimal polymer-to-

activated carbon mass ratio of 1.5:1. Capacitance was stable up to the 5000 cycles tested with a

small decrease of 5% within the first 100 cycles. The capacitance enhancement could be

attributed to the preservation of nanostructure surface area and accessible pore space that is

afforded by ultrathin films. oCVD promises to be a valuable synthesis and processing pathway

for making conductive polymers and creating enhanced energy storage solutions that takes

advantage of nanostructured device architectures.

5.6 References

1. B. E. Conway, Electrochemical Supercapacitors: Scientific Fundamentals and

Technological Applications, Plenum, New York, 1999.

2 M. Nowak, D. Spiegel, S. Hotta, A. Heeger and P. Pincus, Macromolecules, 1989, 22,

2917-2926.

3. H. I. Becker, United States patent, US 2800616, 1957.

112

4. D. Galizzioli, F. Tantardini and S. Trasatti, Journal of Applied Electrochemistry, 1974,

4, 57-67.

5. J. Zheng, P. Cygan and T. Jow, Journal of the Electrochemical Society, 1995, 142,

2699-2703.

6. A. Laforgue, P. Simon, C. Sarrazin and J.-F. Fauvarque, Journal of Power Sources,

1999, 80, 142-148.

7. T. Jow and J. Zheng, Journal of the Electrochemical Society, 1998, 145, 49-55.

8. H. Wang, Q. Hao, X. Yang, L. Lu and X. Wang, Nanoscale, 2010, 2, 2164-2170.

9. J. Xu, K. Wang, S.-Z. Zu, B.-H. Han and Z. Wei, ACS Nano, 2010, 4, 5019-5026.

10. K. Zhang, L. L. Zhang, X. S. Zhao and J. Wu, Chemistry of Materials, 2010, 22, 1392-

1401.

11. R. B. Ambade, S. B. Ambade, N. K. Shrestha, Y.-C. Nah, S.-H. Han, W. Lee and S.-H.

Lee, Chemical Communications, 2013, 49, 2308-2310.

12. S. Nejati and K. K. S. Lau, Langmuir, 2011, 27, 15223-15229.

13. A. Kay, M. Nazeeruddin, I. Rodicio, R. Humpbry-Baker, E. Müller, P. Liska, N.

Vlachopoulos and M. Grätzel, Journal of the American Chemical Society, 1993, 115,

6382-6390.

14. J. Chmiola, Pore-size ion-size correlations for carbon supercapacitors, Drexel

University, Philadelphia, USA 2009.

15. S. D. Beattie, D. M. Manolescu and S. L. Blair, Journal of Electrochemical Society,

2009, 156, A44-A47.

16. S. Nejati and K. K. S. Lau, Nano Letters, 2011, 11, 419-423.

17. S. Hotta, W. Shimotsuma and M. Taketani, Synthetic Metals, 1984, 10, 85-98.

113

18. T. Yamamoto, A. Morita, Y. Miyazaki, T. Maruyama, H. Wakayama, Z. H. Zhou, Y.

Nakamura, T. Kanbara, S. Sasaki and K. Kubota, Macromolecules, 1992, 25, 1214-

1223.

19. Y. Furukawa, M. Akimoto and I. Harada, Synthetic Metals, 1987, 18, 151-157.

20. M. Fu, G. Shi, F. Chen and X. Hong, Physical Chemistry Chemical Physics, 2002, 4,

2685-2690.

21. R. J. Waltman, J. Bargon and A. F. Diaz, Journal of Physical Chemistry, 1983, 87,

1459-1463.

22. A. Asatekin and K. K. Gleason, Nano Letters, 2010, 11, 677-686.

23. Z. Mo, K. B. Lee, Y. B. Moon, M. Kobayashi, A. J. Heeger and F. Wudl,

Macromolecules, 1985, 18, 1972-1977.

24. P. M. S. Monk, R. J. Mortimer and D. R. Rosseinsky, in Electrochromism, Wiley-VCH

Verlag GmbH, 2007, pp. 143-171.

25. Y. Kim, Y. Kim, S. Kim and E. Kim, ACS Nano, 2010, 4, 5277-5284.

26. M. Zuleta, P. Björnbom, A. Lundblad, G. Nurk, H. Kasuk and E. Lust, Journal of

Electroanalytical Chemistry, 2006, 586, 247-259.

27. S. Park, H. C. Kim and T. D. Chung, The Analyst, 2012, 137, 3891-3903.

28. H. Boo, S. Park, B. Ku, Y. Kim, J. H. Park, H. C. Kim and T. D. Chung, Journal of the

American Chemical Society, 2004, 126, 4524-4525.

29. M. Dietrich and J. Heinze, Synthetic Metals, 1991, 41, 503-506.

30. H. Zhou, H. Chen, S. Luo, G. Lu, W. Wei and Y. Kuang, Journal of Solid State

Electrochemistry, 2005, 9, 574-580.

31. C. Peng, D. Hu and G. Z. Chen, Chemical Communications, 2011, 47, 4105-4107.

32. D. Ofer, R. M. Crooks and M. S. Wrighton, Journal of the American Chemical Society,

1990, 112, 7869-7879.

114

33. T. W. Lewis, G. M. Spinks, G. G. Wallace, A. Mazzoldi and D. De Rossi, Synthetic

Metals, 2001, 122, 379-385

115

6 Chapter 6: Conclusions and Future Directions

Overall, this work demonstrates polymer CVD is a powerful method in the synthesis of polymer

electronic materials. We showed that iCVD and oCVD both can be used for polymer integration

within high aspect ratio porous electrodes which is not accessible to conventional liquid phase

processing. We also developed the necessary synthesis and processing knowledge required to

achieve suitable polymer properties and function.

6.1 Specific aim 1

We have successfully utilized iCVD and oCVD to design and synthesize new material for

optoelectronic applications. Polymer electrolytes with exceptionally clean and well-defined

stoichiometry were synthesized using iCVD. The polymer ionic conductivity and molecular

weight were found to be in a suitable range for gel electrolytes. For future work, different

polymer chemistries can be explored to further enhance ion conductivity as this will allow the

development of a better gel electrolyte. Poly(4-vinyl pyridine) will be a good starting point as

theas the pyridine group can coordinate effectively with counter ions and facilitate ion transport

in the polymer. Other methacrylate polymers including poly(glycidyl methacrylate) (PGMA)

and poly(methyl methacrylate) (PMMA) could be tested. Both As for the current application of

gel electrolyte in DSSC both charge transfer processes at the semiconductor interface and ion

diffusion in polymer gel are of importance, and for every new material tested these parameters

should be carefully considered.

We showed for the first time that unsubstituted polythiophene can be deposited through

oCVD using vanadium oxytrichloride as an effective oxidant. Control over polymer properties

can be achieved by changing oCVD processing parameters. Vibrational spectroscopy along with

conductivity measurements and UV-vis data suggest that the deposited film is not fully doped.

116

The possibility of - and - linkages during polymerization cannot be unequivocally ruled out

at this point. It is possible that the presence of these defects could account for the lower electrical

conductivity of the polythiophene films particularly when they were deposited at a high initiator

concentration relative to the monomer. However, the nature of this defect could not be from the

chemical degradation of the ring during oCVD as XPS did not indicate any chemical change in

the ring structure. Significantly, polymer conjugation length and electrical conductivity can be

tuned by controlling oCVD process variables. Polymerization is found to be adsorption-limited

so by providing sufficient monomer and limiting the amount of initiator at the growth surface,

polythiophene is believed to be formed through the normal - thiophene linkages. To verify

this mechanism, further work can be done to explore the polymerization chemistry. As an

example, 3,4-dibromothiophene would be good candidates as the bromine in the 3 and 4 ring

positions will remove the possibility of mislinkages and this will help with understanding the

oCVD synthesis on resulting polymer structure more accurately.

6.2 Specific aim 2

We showed how iCVD design parameters can play a crucial role in successful polymer

integration within porous nanostructures. By carefully controlling diffusion transport and surface

kinetics during iCVD, complete pore filling was obtained. To understand the pore filling

process, we developed a model to predict the time constants for transport and reaction, and the

model prediction was shown to be in good agreement with the observed deposition phenomena.

The impact of pore filling on DSSC performance was studied. The results suggested a significant

enhancement in cell properties. The efficiency of the quasi solid state PHEMA electrolyte DSSC

was higher when compared with their liquid electrolyte counterparts. We concluded that the

117

observed increase in open circuit cell voltage is related to the decrease in electron recombination

at the electrolyte-electrode interface as a result of the exceptional pore filling achieved by iCVD.

As future work in continuing this specific aim, the specific surface area of the electrode can

be enhanced by changing the particle size which by default will enhance the porosity. New

polymer electrolytes can be used and developed and integrated within the nanostructured

electrodes. Another revolutionary way to branch out from this work is to design a polymer as a

photosensitizer within the porous electrode that could replace the current ruthenium based dye.

Other areas to pursue include enhancing electrode charge collection and diffusion for example

by utilizing higher dimension conductors like carbon nanotubes, and using conducting polymers

as solid state hole transfer materials to replace the liquid electrolyte/redox couple combination.

6.3 Specific aim 3

We demonstrated oCVD's unique ability to synthesize unsubstituted polythiophene based on the

direct vapor-to-solid oxidative polymerization of thiophene monomer using antimony

pentachloride oxidant. A stronger oxidant was found to enable a longer conjugation length and

higher conductivity as long as the thiophene ring structure was preserved. Ultrathin

polythiophene films have been successfully integrated within high aspect ratio porous

nanostructures of AAO, TiO2 and activated carbon by operating oCVD under slow kinetics that

favored reactant mass transport. Significantly, ultrathin conformal polythiophene coatings that

preserved the high surface area of the porous nanostructures showed enhanced capacitance over

that of planar films that could be attributed to nanoconfinement effects not accessible in flat

planar geometries. As a result, pseudocapacitors consisting of polythiophene coated activated

carbon electrodes gave considerably higher energy storage capacity compared to bare activated

carbon.

118

To branch out from this work, the base electrode can be changed and carbon nanotube

forests can be used to optimize the charge storage enhancement of the polymer by changing the

underlying dimensions that influence the double layer structure. Other conducting polymers e.g.

poly(3,4-ethylenedioxythiophene) can be synthesized through oCVD and these derivatives of

polythiophene can show enhanced performance in charge storage capacitance and cycling

stability.

6.4 List of publications

R. Bose, S. Nejati and‎ .‎ .‎S.‎ au‎“Grafted‎Poly(ethylene‎oxide)‎(PEO)‎Anti-fouling Surfaces

using‎ nitiated‎Chemical‎Vapor‎Deposition‎(iCVD)”,‎Macromolecules 2012,

S. Nejati and‎ .‎ .‎ S.‎ au‎ “Chemical‎Vapor‎Deposition‎ Synthesis‎ of‎Tunable‎ nsubstituted‎

Polythiophene”, Langmuir 2011, 27(24), 15223-15229.

S. Nejati and K. K. S. au‎“Pore‎Filling‎of‎Nanostructured‎Electrodes‎in‎Dye‎Sensitized‎Solar‎

Cells‎by‎ nitiated‎Chemical‎Vapor‎Deposition”,‎Nano letters 2011, 11 (2), 419–423.

S. Nejati and‎ .‎ .‎S.‎ au‎“ ntegration‎of‎polymer‎electrolytes‎in‎dye‎sensitized‎solar‎cells‎by‎

initiated‎chemical‎vapor‎deposition”,‎Thin Solid Film 2011 519(14), 4151-4155.

R. Bose, S. Nejati,‎ .‎ .‎ S.‎ au‎ “ nitiated‎ Chemical‎Vapor‎Deposition‎ (iCVD)‎ of‎Hydrogel‎

Polymers”‎ECS Transactions 2009 25(8), 1229-1235.

Bavarian, M.; Nejati, S.; Lau, K. K.S.; Lee,‎ D;‎ and‎ Soroush‎ M.,‎ “Effects‎ of‎ Critical‎ Dye‎

Sensitized‎ Solar‎ Cell‎ Parameters‎ on‎ Cell‎ Performance:‎ a‎ Theoretical‎ Study”, to be

submitted 2013

S R. Spurgeon, J. D. Sloppy, C. R. Winkler, M. Jablonski, S. Nejati, K. Jambunathan, A. R.

Damodaran, J. C. Idrobo,‎ .‎ .‎ au,‎S.E.‎ ofland,‎ .‎W.‎Martin‎ and‎M.‎ .‎Taheri‎ “‎

Substrate-Controlled Strain and Polarization Effects on Magnetization and Curie

Temperature‎in‎ SMO‎/‎PZT‎Thin‎Film‎Oxide‎Heterostructures”, to be submitted 2013

Nejati, S.; Carter, Z; Lau, K. K.‎ S.,‎ “ All Solid State Hybrid Solar Cell through oCVD of

Conjugated Polymer”,‎Manuscript in preparation 2013

119

Cody, C.; Nejati, S.; Lau, K. K. S.,‎“Engineering‎ ight‎Absorption‎within Conventional DSSC

Electrodes‎ sing‎a‎ ow‎Temperature‎Process”,‎Manuscript in preparation 2013

Nejati, S.; Minford, T. E.; Lau, K. K. S., Integration of Conjugated Polymers in Supercapacitor

Using Oxidative Chemical Vapor Depositionto be submitted, 2013

120

Vita

Siamak Nejati was born in Kermanshah, Iran in 1980. From an early age, he was fascinated by

science and in high school he was nominated to be a part of the National Physics Olympiad team

to attend the international competition. He entered Sharif University of Technology at Tehran

and gained his B.S. Degree in Chemical Engineering in 2004. He continued his study in

Biotechnology as a M.S. student and graduated with first rank from his program. He joined

Drexel University in 2007. In September 2008, he started his work in thin films and surface

science under the supervision of Professor Kenneth K. S. Lau.