174
DEVELOPMENT OF STABLE METAL OXIDE ELECTRODES FOR THE CONVERSION OF ELECTRICITY TO CHEMICAL FUELS A DISSERTATION SUBMITTED TO THE DEPARTMENT OF CHEMISTRY AND THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY Jonathan David Prange July 2011

DEVELOPMENT OF STABLE METAL OXIDE …xr045qb7231/Jonathan Prange... · DEVELOPMENT OF STABLE METAL OXIDE ELECTRODES FOR THE ... Fourier transform infrared spectroscopy, ... Verma,

  • Upload
    dokiet

  • View
    217

  • Download
    2

Embed Size (px)

Citation preview

DEVELOPMENT OF STABLE METAL OXIDE ELECTRODES FOR THE

CONVERSION OF ELECTRICITY TO CHEMICAL FUELS

A DISSERTATION SUBMITTED TO THE DEPARTMENT OF CHEMISTRY AND

THE COMMITTEE ON GRADUATE STUDIES OF STANFORD UNIVERSITY IN

PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

Jonathan David Prange

July 2011

http://creativecommons.org/licenses/by-nc/3.0/us/

This dissertation is online at: http://purl.stanford.edu/xr045qb7231

© 2011 by Jonathan David Prange. All Rights Reserved.

Re-distributed by Stanford University under license with the author.

This work is licensed under a Creative Commons Attribution-Noncommercial 3.0 United States License.

ii

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Christopher Chidsey, Primary Adviser

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

T Stack

I certify that I have read this dissertation and that, in my opinion, it is fully adequatein scope and quality as a dissertation for the degree of Doctor of Philosophy.

Robert Waymouth

Approved for the Stanford University Committee on Graduate Studies.

Patricia J. Gumport, Vice Provost Graduate Education

This signature page was generated electronically upon submission of this dissertation in electronic format. An original signed hard copy of the signature page is on file inUniversity Archives.

iii

iv

Abstract

The conversion of renewable sources of electricity to molecular fuels is widely

viewed as an important component of our future energy economy. To accomplish this,

stable electrodes are necessary to perform electrochemical reactions of interest for

extended periods of operation with high efficiency. In the first part of this thesis, a

method to immobilize homogeneous catalysts to a metal oxide electrode through click

chemistry on an attached p-azidophenyl phosphonic acid linker is presented. This

strategy allows for a convergent approach to surface modification that results in stable

attachments while allowing for facile charge transfer between the electrode and the

immobilized catalyst. The deposition of p-azidophenyl phosphonic acid to the metal

oxide surface and subsequent click with molecules of interest was investigated with

Fourier transform infrared spectroscopy, electrochemistry and X-ray photoelectron

spectroscopy.

The electrochemical oxidation of water to supply the electrons needed for fuel synthesis

remains a challenge due to the lack of materials which can both efficiently remove four

electrons and withstand the harsh oxidative conditions of the reaction. A novel type of

dimensionally stable anode utilizing silicon as the base substrate and also as an in-situ

photon collector has been developed. It uses a thin layer of titanium dioxide deposited

by atomic layer deposition to protect the silicon. A thin layer of physical vapor

deposited iridium is used as the water oxidation catalyst. Results shown include water

oxidation efficiency in both light and dark conditions and over a range of pH values

with an emphasis on the operational stability and durability of the anodes.

v

Acknowledgements

I would like to first start by thanking my thesis advisor, Professor Christopher E.

D. Chidsey, for all of his training, insight and guidance, and ultimately giving me the

opportunity to prove myself in his laboratory. His suggestions, comments and advice

have shaped the way I approach problems and obstacles, not only in science, but in

every aspect of my life. I would also like to thank my committee chair, Professor Paul

C. McIntyre, whom I consider more of a co-advisor and collaborator. Working with

Professor McIntyre has been a tremendous experience and I truly appreciate all of his

guidance and support as well as his suggestions and insights into our work. I am also

very grateful for my two readers and collaborators, Professor T. Daniel P. Stack and

Professor Robert M. Waymouth. Professor Stack and Professor Waymouth helped me

navigate my graduate career by offering advice and suggestions on anything from my

research project and proposals to the difficult situation of switching research groups in

the middle of my graduate career. I would also like to thank Professor Justin DuBois,

my non-reader, and Professor Richard Zare for their support and helping me through the

transition of switching labs.

I would also like to thank my first research advisor, Professor Dmitry Yandulov.

Professor Yandulov taught me a lot about inorganic reaction mechanisms and kinetics

as well as helping me master some advanced laboratory techniques.

I thank Professor Thomas Jaramillo and Professor Matthew Kanan for the many

fruitful discussions about science and energy. Their suggestions and insights have been

vi

greatly appreciated and I have learned so much from both of them. I look forward to

their contributions to this field in the future. I thank Dr. Todd A. Eberspacher whom is

a great resource in so many ways. His no nonsense approach to problem solving has

made my graduate career that much easier. He is also one of the most knowledgeable

people I have met at Stanford. Dr. Steve Lynch has also been a tremendous resource for

both the optics facility and the NMR facility. Chuck Hitzman has also been great in

helping out with using the XPS and all the software questions I have had.

I am truly grateful for all my collaborators throughout my graduate career.

Randall Lowe first made the p-azidophenyl phosphonic acid and selflessly allowed me

to work with it. I am very grateful for this and the scientific knowledge he has given

me over the years. He is one of the most capable people I have met and I thank him for

his friendship and all his fruitful discussions. I am also thankful for Alissa Sasayama,

who helped tremendously on this project. From the McIntyre Lab, I would like to thank

Yi Wei (Vincent) Chen. Vincent has worked on the solar water splitting project with

me for a major portion of my graduate career. He is one of the smartest, hardest

working and selfless people I have met at Stanford. He is also persistent, as he has

attempted over and over again to teach me solid state physics to no avail. Without him,

this project doesn’t go anywhere, and I am so thankful for his scientific and personal

friendship, he truly is a great scientist and great person. Simon Duehnen has also been a

great pleasure to work with. He worked very hard on the water splitting project and

became a completely independent scientist by the end of his stay at Stanford. I look

forward to visiting him in Germany soon. Ali Hosseini deserves a mention here, as he

vii

has always been helpful with any question I have ever had on any project, he is a great

friend.

I also want to thank all of the current and former Chidsey group members for all

their help and ultimately putting up with me in the office and lab. In no particular

order, Randall Lowe, Ali Hosseini, Simon Duehnen, Charles McCrory, Anando

Devadoss, Josh Ratchford, Alex Neuhausen, Vadim Ziatdinov and the undergrads

Alissa Sasayama, Marty Casey, Jeff Jensen and David Lapham. I want to thank

members of the McIntrye lab as well, for putting up with me and all my requests for

ALD samples. In no particular order: Vincent Chen, Marika Gunji, Jaesoon Ahn,

Rathnait Long, Shu Hu, Rahim Esfandyarpour, Cynthia Ginestra and Andy Lin. I

would also like to mention my former lab mates in the Yandulov lab: Eunsung Lee,

Kendra Kuhl, Jessica DeMott, Sang-won Kwo, Ngon Tran and Georg Platz.

I have had a great relationship with members of both Professor Stack’s group

and Professor Waymouth’s group. I want to thank everyone I have interacted with in

these two labs over the years, but would like to mention, in no particular order, Pratik

Verma, Matt Pellow, Brian Smith, Eric Stenjehem, and Tim Storr from the Stack Lab

and Matt and Liz Kiesewetter, David Pearson, Kristen Brownell and Antoni DeCrisi

from the Waymouth Lab. Thank you again to both of these labs for everything from

letting me ‘borrow’ lab equipment to suggestions and random conversations. I would

also like to thank all those involved in the GCEP collaboration from the labs of

Professor Chidsey, Stack, Waymouth and Jaramillo. There are many graduate students

and post docs that have contributed to my positive experience in graduate school, not to

name them all, but I do thank all of you.

viii

I am also grateful for Roger Kuhn, Patricia Dwyer and everyone else in the

Chemistry department front office. They work so hard and are the best at what they do.

I will miss having them around to make my life easier.

I would also especially thank my close friends here: Kevin and Cara, Brian and

Moria, Colin and Andi, Jason Wagnor and Scott Tabakman. You guys are the best. I

couldn’t have done this without you guys.

My graduate career was full of ups and downs. The one consistency was always

my family. My dad and mom, Dave and Carol Prange, have prepared me my whole life

to meet all the challenges I have ever been presented here in graduate school and in life,

and I am so grateful for them. I would also like to mention all of my brothers and

sisters and their spouses: Malinda and Tim Smith, Scott and Laurie Prange, Vanessa

and Matthew Ewing, and Chris Prange and his girlfriend, Shawna. They have been

very supportive and a great source of encouragement and happiness throughout my time

here. I would also like to thank the best nieces and nephews an uncle could have:

Courtney and Josh, and Asheley and Tyler (and any future ones). You guys are

awesome and I enjoy spending time with all of you, and I am sorry I haven’t been

around much these last few years. Finally, I want to thank my best friend, Sarah

Sherlock. She is the best thing to happen to me at Stanford, and without her

unconditional love and unwavering strength and support, this would not have been

possible. I love her and appreciate her more than she will ever know. This thesis is for

all of you, thank you for everything.

ix

Table of Contents

Abstract .......................................................................................................................... iv

Acknowledgements ......................................................................................................... v

Table of Contents ........................................................................................................... ix

List of Figures .............................................................................................................. xiii

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

1.1: Thesis Objective ............................................................................................... 1

1.2: Background ........................................................................................................ 2

Modification of Metal Oxide Electrodes .................................................. 2

Protecting Photoanodes Used in Solar Water Splitting ............................ 5

1.3: Analytical Methods ............................................................................................ 7

Fourier Transform Infrared Spectroscopy ................................................ 7

X-Ray Photoelectron Spectroscopy .......................................................... 7

Electrochemical Methods ......................................................................... 8

1.4: Methodology ................................................................................................... 11

1.5: Collaborations ................................................................................................. 12

1.6: Figures ............................................................................................................ 13

1.7: References ....................................................................................................... 18

x

Chapter 2: Modification of Indium Tin Oxide Electrodes with p-Azidophenyl

Phosphonic Acid ........................................................................................................... 20

2.1: Preface ............................................................................................................ 20

2.2: Abstract ........................................................................................................... 21

2.3: Introduction ..................................................................................................... 22

2.4: Materials and Methods .................................................................................... 25

2.5: Results and Discussion ................................................................................... 30

2.6: Conclusion ...................................................................................................... 37

2.7: Figures ............................................................................................................. 38

2.8: References and Notes ...................................................................................... 47

2.9: Supporting Information ................................................................................... 52

Chapter 3: Introduction ............................................................................................... 64

3.1: Abstract ........................................................................................................... 64

3.2: The Global Energy Challenge ......................................................................... 65

Renewable and Alternative Sources of Energy ...................................... 67

3.3: Solar Water Splitting ...................................................................................... 69

Photoanodes and Water Oxidation Catalysts ......................................... 74

Protection of Photoanodes Used in Solar Water Splitting ...................... 76

3.4: Figures ............................................................................................................ 79

3.5: References ....................................................................................................... 84

xi

Chapter 4: Stable Si Photoanodes for Water Splitting ............................................ 90

4.1: Preface ............................................................................................................ 90

4.2: Abstract ........................................................................................................... 91

4.3: Introduction ..................................................................................................... 92

4.4: Results and Discussion ................................................................................... 95

4.5: Conclusion .................................................................................................... 101

4.6: Methodology ................................................................................................. 102

4.7: Figures .......................................................................................................... 104

4.8: References ..................................................................................................... 112

4.9: Supporting Materials ..................................................................................... 117

4.10: Supporting Materials Figures ...................................................................... 123

Chapter 5: Effect of TiO2 Thickness and Catalyst Layer on Efficiency and

Stability of Silicon Anodes for Water Oxidation ..................................................... 134

5.1: Preface .......................................................................................................... 134

5.2: Abstract ......................................................................................................... 135

5.3: Introduction ................................................................................................... 136

5.4: Results and Discussion ................................................................................. 137

5.5: Conclusions ................................................................................................... 143

5.6: Figures .......................................................................................................... 144

5.7: References ..................................................................................................... 148

xii

5.8: Supporting Materials ..................................................................................... 151

5.9: Supporting Materials Figures ........................................................................ 156

5.10: Supporting Materials References ................................................................ 160

xiii

List of Figures

Figure 1.1: Illustration of immobilization strategy ......................................................... 13

Figure 1.2: Titanium Pourbaix Diagram. ...................................................................... 14

Figure 1.3: Electrochemical Water Oxidation Setup. .................................................... 15

Figure 1.4: FTIR Total Reflectance Accessory. ............................................................. 16

Figure 1.5: ALD Process. ............................................................................................... 17

Figure 2.1: Immobilization Schematic. ......................................................................... 38

Figure 2.2: XPS Analysis. .............................................................................................. 39

Figure 2.3: Deposition of p-Azidophenyl Phosphonic Acid. ......................................... 40

Figure 2.4: FTIR Spectrum Before and After Click. ...................................................... 41

Figure 2.5: Cyclic Voltammogram of Clicked Ethynylferrocene. ................................. 42

Figure 2.6: Click Results for Azide-terminated Monolayers. ........................................ 43

Figure 2.7: Mixed Monolayers on ITO. ......................................................................... 44

Figure 2.8: Effect of Water on p-Azidophenyl Phosphonic Acid Deposition. ............... 45

Figure 2.9: Stability of Azide on ITO Surface. .............................................................. 46

Figure 3.1: Current and Future Global Energy Demand. ............................................... 79

Figure 3.2: Solar Spectrum. ............................................................................................ 80

Figure 3.3: Band Gaps of Common Semiconductors. .................................................... 81

Figure 3.4: Recombination Pathways for Photogenerated Electron/Hole Pairs. ............ 82

Figure 3.5: Photoelectrochemical Water Splitting Configurations. ............................... 83

xiv

Figure 4.1: Anode Design and Water Oxidation Results. ............................................ 105

Figure 4.2: Stability Tests. ........................................................................................... 106

Figure 4.3: Constant Potential Stability Test. ............................................................... 107

Figure 4.4: XPS Depth Profiling Analysis. .................................................................. 108

Figure 4.5: Anode Electrochemical Performance. ....................................................... 109

Figure 5.1: Electrochemical Results of Anodes with Various TiO2 Thicknesses. ....... 144

Figure 5.2: Anodes Stability Tests. .............................................................................. 145

Figure 5.3: XPS Analysis of Samples Before and After Stability Tests. ..................... 146

1

Chapter 1: Introduction

1.1: Thesis Objective

The specific topics of my thesis are: (a) understanding the attachment and

coverage of p-azidophenyl phosphonic acids on conductive metal oxide electrodes and

(b) developing oxide tunnel barriers to protect low band gap semiconductors employed

for photoelectrochemical oxidation of water. These two topics are related by the overall

objective of understanding how to transfer electrons between a species in solution and

an electrode in order to interconvert electricity and chemical fuels. A fundamental

understanding and control of electron transfer reactions to and from effective catalysts

on corrosion resistant electrode surfaces will enable the construction of more efficient

and robust electrochemical devices important for small molecule oxidation or reduction

necessary for renewable electricity production and storage.

2

1.2: Background

Modification of Metal Oxide Electrodes

The first project in this thesis is an investigation into the modification of a metal

oxide (MO) electrode with compounds of interest from solution. The use of MO

electrodes for applications such as fuel cells and organic photovoltaic devices often

requires modification of the surface to achieve the desired properties. For example,

many reactions of interest in a fuel cell, such as the four electron reduction of dioxygen

to water, are found to be sluggish or inefficient at the MO electrode surface. The ability

to modify the MO surface with an active catalyst for this reaction would be highly

desirable. This strategy also holds true for organic photovoltaic devices, where a light

absorbing organic compound must be attached to a transparent conductive MO

electrode in order to efficiently inject the photogenerated charge into the device. In

these applications, a well-defined and robust attachment of various compounds of

interest, either catalysts or light absorbing molecules, is required for operation.

A variety of attachment functionalities, or anchors, to a MO surface are known

including siloxanes, carboxylic acids and phosphonic acids.1-4

These molecules are all

claimed to form self-assembled monolayers (SAMs) on MO surfaces from a deposition

solution containing the adsorbate and the MO electrode. Each one of these anchors

could potentially form multiple bonds to the surface resulting in a chemically and

thermally stable attachment. Although these functional groups attach to the surface

readily from solution, some challenges must be overcome before implementation in an

operational device. First, charges must be able to transfer between the MO surface and

3

catalyst or light absorbing molecule that is attached to it. Second, synthesizing one of

these attachment functional groups to every catalyst or small molecule of interest

presents a formidable challenge. A method to immobilize a compound and retain its

function on a surface while not having to go through elaborate synthetic procedures

would be highly desirable.

The strategy employed in Chapter Two of this thesis is to synthesize a p-

azidophenyl phosphonic acid that could act as an anchor to the surface and a linker to a

molecule of interest in solution. The phosphonic acid group of this molecule will act as

the anchoring group to a MO surface, attaching in a bi-dentate or tri-dentate mode.5

The azide group of the p-azidophenyl phosphonic acid will allow for the coupling of

various ethynyl-terminated molecules in solution. The pi system of the entire molecule

will allow for facile electron transfer between the MO electrode surface and the redox

species that is to be attached. This method will allow for a convergent approach to

surface modification by utilizing one common immobilized linker for a number of

different compounds to be tethered to a MO electrode surface. An illustration of this

convergent immobilization strategy is shown in Figure 1.1.

The azide group of p-azidophenyl phosphonic acid is known to react selectively

with a terminal alkyne functional group in the copper(I)-catalyzed azide-alkyne

cycloaddition (CuAAC) reaction discovered independently by Sharpless and Meldel in

2002. This reaction is an example of a ‘click’ reaction as described by Sharpless in

2001.6 The ability to choose ligands or small molecules that have this alkyne group and

selectively couple them to the surface affords the ability to screen and study a large

number of combinations. Additionally, the product of the click reaction, a 1,2,3-

4

triazole, is both thermally and chemically stable and allows for facile electron transfer

due to the pi-boding within the molecule. The clicked molecules in Chapter Two were

studied using Fourier transform infrared spectroscopy (FTIR), electrochemistry and X-

ray photoelectron spectroscopy (XPS) to better understand the overall properties of the

clicked phosphonate-MO system. The coverages were found to be low compared to

expected monolayers while water was found to play an important role in monolayer

formation. Some of the properties that were investigated were surface coverage

electron transfer from a one-electron redox active molecule clicked onto the surface and

the oxidative stability of both the arylphosphonate attachment and also the triazole

linkage.

Mixed monolayers of the azide terminated p-azidophenyl phosphonic acid and a

diluent molecule, phenyl phosphonic acid, were also prepared. These mixed

monolayers allow for site isolation of the azide groups which lead to a more complete

click reaction on the surface and proper spacing of a catalyst of interest. This spacing

ensures that the mechanism of any electrocatalytic reaction would occur in a

mononuclear pathway, meaning that one catalyst on a surface could not form a dimer

with a nearby catalyst. The ability to irreversibly attach and site-isolate catalysts on p-

azidophenyl phosphonic acid modified MO electrodes allows for new investigations

into electrocatalysis and electron transfer. It is envisioned that these studies will help

lead to the development of more robust and efficient devices for use in applications

such as fuel cells.

5

Protecting Photoanodes Used in Solar Water Splitting

The second project that was investigated as part of this thesis was the

development of new strategies to protect photoanodes operated in solar water splitting

setups. Solar water splitting has long been viewed as a promising strategy to store solar

energy in the form of chemical fuels.7, 8

In order to accomplish this, however,

photoanodes must be employed that are both efficient and stable under the harsh

operating conditions typically associated with the water oxidation half reaction. A

photoanode typically consists of a solar light absorbing semiconductor substrate that is

modified with a catalyst that will allow for efficient water oxidation at the

semiconductor-solution interface. During operation, most semiconductors of interest

will corrode or oxidize to an insulator form or dissolve into the electrolyte solution.

These deactivation processes will restrict the choice of semiconductors used in

photoelectrochemical devices to highly oxidized types that typically have larger than

optimal band gaps for solar absorption. Providing adequate protection of the

semiconductors with more appropriate band gaps while still allowing for efficient

device operation remains a challenge.

The strategy used in Chapter Four and Five of this thesis was to develop a

dimensionally stable photoanode that utilizes a protective coating imposed between a

silicon substrate and catalyst layer for water oxidation. Atomic Layer Deposition

(ALD) was used to synthesize a thin, conformal coating of titanium dioxide (TiO2) that

would serve as a barrier between the semiconductor and the water-containing

electrolyte. The ALD process allows for thickness control on the atomic scale, as only

a single atomic layer is deposited per cycle, as discussed in the methodology section

6

below. This allows for uniform coatings that are free of pinholes and cracks that would

otherwise allow oxidants to reach and corrode the silicon substrate. The TiO2 layer was

chosen because of its superb stability over a range of oxidizing potentials at all pH’s, as

illustrated in the titanium Pourbaix diagram shown in Figure 1.2.9 This diagram plots

the major species observed at the given electrochemical potential and pH in aqueous

conditions. The catalyst that was used in this work was a physical vapor deposited layer

of iridium. The photoanode was employed in a solar water splitting setup as illustrated

in Figure 1.3. In this setup, a platinum wire was used as both the counter electrode and

water reduction catalyst. Operational stability was investigated in both dark and solar

illumination conditions and monitored by holding the anode at either a constant current

or constant potential until the device stopped working. Once these stability tests were

completed, the anodes were analyzed with XPS and transmission electron microscopy

(TEM) to determine the composition and structure of the photoanode.

Chapter 3 serves as an introduction chapter for the motivation behind this work

as well as offering a more thorough discussion of solar water splitting devices and the

semiconductors and catalysts that have been employed for this purpose.

7

1.3: Analytical Methods

A variety of spectroscopic techniques were used throughout my thesis to

characterize and study the electrode surface and subsequent modifications. To help the

reader better grasp the details of the work presented here, a brief description of the

techniques used will be discussed in this section.

Fourier Transform Infrared Spectroscopy

Fourier transform infrared spectroscopy (FTIR) was used extensively in my

thesis work to characterize and verify the attachment of p-azidophenyl phosphonic acid

to the surface of an ITO electrode (Chapter Two). A total reflectance accessory was

designed and built that would allow for reproducible sample measurements and analysis

between experiments (Figure 1.4). The stretching frequency of an azide functional

group appears centered at ~2100 cm-1

, a relatively silent region of the IR spectrum for

most functional groups. An aryl azide will appear as a doublet feature and the alkyl

azide will appear as a singlet. The peaks can be integrated and used to help track the

amount of p-azidophenyl phosphonic acid deposited and how much is reacted when

clicked.

X-Ray Photoelectron Spectroscopy

X-ray photoelectron spectroscopy (XPS) was used throughout my thesis in order

to analyze the surfaces of electrodes at various points of preparation and operation.

This technique works by monitoring the kinetic energy (KE) of an electron hitting a

detector that was ejected via the absorption of an X-ray by a surface atom of a sample.10

8

The measured KE of the ejected electron is related to the binding energy (BE) of that

electron to the nucleus and the incident X-ray by Equation 1.1.

1.1

The X-ray source used in this spectroscopy is constant which allows for the BE to be

readily calculated. The BE is also unique for each atom, which allows for both

quantitative and qualitative information to be obtained for each experiment. One

drawback, however, is that analysis by XPS is restricted to surface atoms as an ejected

electron will be attenuated by scattering inelastically off of other atoms with an

exponentially increasing likelihood as a function of depth.

XPS was also used to analyze the electrode structure as a function of depth in a

process known as depth profiling (Chapter Four). This technique etches a small amount

of the surface with a stream of argon ions, exposing a fresh portion of the sample to be

analyzed. The fresh spot is then subjected to the X-ray beam for a typical XPS analysis.

Once completed, the sample is etched again, exposing a deeper layer of that spot of the

sample. This method allows for atomic ratios to be compiled as a function of depth into

the sample.

Electrochemical Methods

A central analytical tool used throughout my thesis was electrochemistry.11

In

both projects presented here (Chapters Two, Four and Five), cyclic voltammetry (CV)

was performed in order to investigate how effective modified and unmodified electrode

surfaces were in transferring electrons to and from a redox active species in an

electrolyte solution. A typical CV will monitor the amount of current (i) measured at a

9

working electrode as a function of linearly scanning applied potential (E). The potential

is applied between the working electrode and reference electrode and scanned at a given

rate (in units of V/s). Once the potential reaches the desired final potential, it returns to

the initial potential to complete the cycle. Several of these cycles can be performed

during an experiment. The measured current, or current density when normalized by

the exposed area of the working electrode (in units of A/cm2), can be plotted versus the

ramping potential resulting in a current-potential or i-E curve. The setup also employs a

counter electrode to complete the circuit from the working electrode to the solution.

The counter electrode is typically made of a noble metal, such as platinum, which will

conduct current effectively but not react with the electrolyte solution.

If a CV experiment is performed with a redox active species in the electrolyte

solution a peak will appear corresponding to the oxidation or reduction of that molecule.

As the potential is scanned positively, the measured current will appear as an anodic

peak. In a static solution, the peak would increase, reach a maximum and then

decrease. This peak shape is due to the reduced form of the redox active molecule

being depleted as it is oxidized near the electrode surface. If the redox molecule is

reversible, the cathodic peak corresponding to the reduction current will be observed as

the potential is scanned back negatively. The peak-to-peak separation is an indication

of how effective the electron transfer occurs between the electrode and redox active

molecule either in solution or attached to the surface of the electrode. The reversible

potential of the redox species, E°, is typically centered between the two peaks due to an

additional potential required to overcome analyte concentration polarization effects and

kinetic effects associated with moving an electron between the molecule and the

10

electrode. This additional potential beyond that of the reversible potential is defined as

the overpotential (η) for the oxidation or reduction reaction, and can be used to compare

the efficiency of various electrodes and immobilized catalysts at performing a specific

electrochemical oxidation or reduction.

Another useful outcome of using electrochemistry to study electron transfer

efficiency is the ability to calculate coverage of a molecule immobilized on the surface.

If each attached molecule contains a redox active group, such as ferrocene, the total area

of the peaks can be converted to an electrochemical coverage as an outcome of

Faraday’s law. The integral of the current in a peak can be converted to a total charge,

which can be used to determine the number of electrons transferred in the process. The

number of electrons can be used to determine the amount of the redox species present

depending on the number of electrons that can be transferred per species. In the case of

ferrocene, one electron would equate to one ferrocene molecule, while for water

oxidation, four electrons would have to be counted for each oxygen molecule produced.

This method of calculating electrochemical coverage was used in Chapter Two to

estimate the total coverage of p-azidophenyl phosphonic acid attached to a metal oxide

surface.

Additional electrochemical experiments were performed in Chapter Four and

Five where an anode was held at either a constant current or constant potential for

various lengths of time. These steady-state experiments were used to determination

anode stability during water oxidation as a function of time exposed to the applied

conditions. Failed samples resulted when current was no longer passing under constant

potential or when the potential would increase dramatically to hold a constant current.

11

1.4: Methodology

As mentioned above, ALD was used to deposit thin, conformal layers of TiO2

on silicon substrates for the purpose of protecting the silicon during

photoelectrochemical water splitting in Chapters Four and Five of my thesis. The ALD

process, illustrated in Figure 1.5, builds up TiO2 one atomic layer at a time by exposing

the silicon substrate to first a titanium precursor, tetrakis-(dimethylamido)titanium

(TDMAT), followed by exposure to water vapor. The TDMAT will react with surface

oxides and hydroxides and then with the water vapor that is introduced. The hydroxides

and water vapor react with the ligands of the TDMAT, producing dimethylamine, which

is removed in vacuo leaving behind a single layer of TiO2. The next cycle will be

started and the process will continue until the desired thickness is obtained. The ALD

process will also coat any structure it is exposed to, meaning a number of geometric

shapes and morphologies can be used and the TiO2 will evenly coat all of the exposed

surface area with crack and pinhole-free conformal layers.

12

1.5: Collaborations

The work conducted in Chapter Two was done with the help of Randall D.

Lowe and Alissa F. Sasayama. Randall first synthesized and characterized the p-

azidophenyl phosphonic acid and first demonstrated the attachment of the molecule to a

metal oxide surface. Alissa helped perform a number of the depositions and subsequent

analyses that are reported. Alissa also helped with the design and fabrication of the

total reflectance FTIR accessory, a representation of which is shown in Figure 1.4.

Additionally, I would like to thank members of the Stack, Waymouth and Jaramillo

groups for their scientific discussions during the collaborative Global Climate and

Energy Project here at Stanford.

Chapters Four and Five of this thesis were done in collaboration with Professor

Paul McIntyre and some of his students in the Materials Science and Engineering

Department at Stanford. Vincent Chen was a close collaborator and worked on every

aspect of these two chapters with me. Without him, this work would not have been

completed. Simon Duehnen, a visiting student in the Chidsey lab from Hanover,

Germany, helped by working on the electrochemistry and stability tests that are

reported, as well as some initial ruthenium and platinum work that is shown in Chapter

5. Yohan Park, Jaesoo Ahn and Rathnait Long helped with the TiO2 deposition and

characterization as well as ALD chamber maintenance. Marika Gunji prepared and

collected TEM images that are shown here.

I am grateful for all the scientific interactions and discussions with all of these

individuals and without them this thesis would not have been possible.

13

1.6: Figures

Figure 1.1: Illustration of immobilization strategy. Illustration of the general

strategy used to selectively immobilize a molecule of interest (active group) to the

surface of an electrode. In Chapter Two, the azide group on the surface would act as

the immobilized functional group while the terminal ethnynyl group of the molecule in

solution would act as the linker.

14

Figure 1.2: Titanium Pourbaix Diagram. Representation of the Pourbaix diagram of

titanium in an aqueous solution. These diagrams are often used to determine the

thermodynamically expected species at the given pH and potential. The pH-dependent

water oxidation and proton reduction potentials are also plotted to illustrate the stability

of TiO2 at these conditions.

15

Figure 1.3: Electrochemical Water Oxidation Setup. Illustration showing water

oxidation occurring at the surface of the TiO2 coated anode to produce molecular

oxygen. The four electrons that are removed from water tunnel through the oxide layers

to the backside contact of silicon. They then go to the platinum counter electrode to

reduce four protons to two equivalents of molecular hydrogen.

16

Figure 1.4: FTIR Total Reflectance Accessory. Illustration of total reflectance

accessory used for FTIR experiments in Chapter 2.

17

Figure 1.5: ALD Process. Illustration of ALD process used for fabrication of

photoanodes that were studied in Chapters Four and Five. The process starts at (1) by

initiating a seed layer of the titanium precursor tetrakis-(dimethylamido)-titanium

(TDMAT) onto an oxide surface. This will continue in the sequence until it uniformly

coats the entire surface of the oxide with a monolayer (2). The water vapor is then

introduced which reacts with the ligands of the TDMAT (3). The removal of the

ligands leaves behind an oxide layer of titanium (4) that can allow for another cycle of

the precursor. This process can be repeated until the proper uniform thickness of TiO2

is obtained.

18

1.7: References

1. Ulman, A., Formation and Structure of Self-Assembled Monolayers. Chem. Rev.

1996, 96, (4), 1533-1554.

2. McElwee, J.; Helmy, R.; Fadeev, A., Thermal stability of organic monolayers

chemically grafted to minerals. J. Colloid Interf. Sci. 2005, 285, (2), 551-556.

3. Armstrong, N. R.; Veneman, P. A.; Ratcliff, E.; Placencia, D.; Brumbach, M.,

Oxide Contacts in Organic Photovoltaics: Characterization and Control of Near-

Surface Composition in Indium-Tin Oxide (ITO) Electrodes. Acc. Chem. Res.

2009, 42, (11), 1748-1757.

4. Mingalyov, P. G.; Lisichkin, G. V., Chemical modification of oxide surfaces

with organophosphorus(v) acids and their esters. Rus. Chem. Rev. 2006, 75, (6),

541-557.

5. Paramonov, P. B.; Paniagua, S. A.; Hotchkiss, P. J.; Jones, S. C.; Armstrong, N.

R.; Marder, S. R.; Bredas, J.-L., Theoretical Characterization of the Indium Tin

Oxide Surface and of Its Binding Sites for Adsorption of Phosphonic Acid

Monolayers. Chem. Mater. 2008, 20, (16), 5131-5133.

6. Rostovtsev, V.; Green, L.; Fokin, V. V.; Sharpless, K., A stepwise Huisgen

cycloaddition process: Copper(I)-catalyzed regioselective "ligation" of azides

and terminal alkynes. Angew. Chem. Int. Edit. 2002, 41, (14), 2596-2599.

7. Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Mi, Q.; Santori,

E. A.; Lewis, N. S., Solar Water Splitting Cells. Chem. Rev. 2010, 110, (11),

6446-6473.

19

8. Gratzel, M., Photoelectrochemical cells. Nature 2001, 414, (6861), 338-344.

9. Pourbaix, M. J. N., Atlas of electrochemical equilibria in aqueous solutions. 1st

Edi. ed.; Pergamon Press: Oxford, 1966.

10. Moulder, J. F.; Stickle, W. F.; Sobol, P. E.; Bomben, K. D., Handbook of X-ray

photoelectron spectroscopy. Physical Electronics: 1995.

11. Bard, A. J.; Faulkner, L. R., Electrochemical methods: fundamentals and

applications - 2nd ed. John Wiley & Sons, Inc.: New York, 2001.

20

Chapter 2: Modification of Indium Tin Oxide Electrodes with p-

Azidophenyl Phosphonic Acid

2.1: Preface

This chapter presents an investigation into the deposition and modification of an

indium tin oxide electrode with an azide-terminated self-assembled monolayer attached

through an aryl phosphonate group. This chapter is derived from a manuscript that is to

be submitted soon for publication. Randall Lowe first synthesized the p-azidophenyl

phosphonic acid molecule and attached it to a metal oxide surface. Alissa F. Sasayama

was instrumental running a number of experiments and aiding in the design and

construction of the total reflectance FTIR accessory.

21

2.2: Abstract

Modification of electrode surfaces is a necessary requirement for the

development of electrochemical devices having specific functions. A novel azide-

terminated phenyl phosphonic acid was synthesized and adsorbed onto indium tin-

doped oxide electrodes resulting in a well-defined self-assembled monolayer. The

azide-terminated monolayers were clicked with various ethynyl-terminated molecules in

solution including ethynylferrocene, which was used to calculate an electrochemical

coverage of up to 8 x 1013

molecules/cm2. Inferred coverages of up to 1 x 10

14

molecules/cm2 were obtained after adjusting for the unreacted azide remaining on the

surface. The absence of water from the deposition solution was found to allow for more

densely-packed monolayers to form on the surface, offering coverages as high as 2 x

1014

molecules/cm2. The phosphonate attachment to the surface and 1,2,3-triazole that

resulted from the click reaction were found to be oxidatively stable to a variety of

applied potentials and chemically stable towards various solvents.

22

2.3: Introduction

Self-assembled monolayers (SAMs) are commonly used to modify various

surfaces in order to obtain specific functionalities. Monolayers formed from alkane

thiols on gold are one of the most well-known and widely used examples1-3

, but other

surfaces are known to allow for monolayer formation including silicon, silver, alumina,

and other metal-oxide materials like tin doped indium oxide (ITO).4, 5

ITO has become

popular recently finding use as electrodes in solar cell designs and organic light

emitting devices (OLEDs) due to its high conductivity, transparency and oxidative

stability.6 There are a number of functional groups known to bind to ITO surfaces

including organosilanes and organic acids such as carboxylic acids7, 8

and phosphonic

acids.9-13

Each of these binding molecules could potentially result in multiple points of

attachment to the surface, affording well-defined and robust linkages. For this reason,

much attention has been given to the design of SAMs on ITO surfaces that have specific

chemical functionalities that will allow for coupling with molecules of interest, e.g.

molecular catalysts, biomolecules and nanomaterials.

There have been many different methods used to deposit phosphonic acids onto

metal oxide surfaces. Typically, the conditions employed include exposing a substrate

to a deposition solvent containing some concentration of the phosphonic acid for a

given amount of time. Measured coverages are consistently observed to be less than the

theoretical coverages of 4 x 1014

molecules cm-2

.14, 15

Reported coverages have been

found to be between 2 x 1013

to 3 x 1014

molecules cm-2

as measured by

electrochemistry, indicating that the coverage is unpredictable. The higher coverages

obtained previously have often been in the presence of highly concentrated solutions of

23

the adsorbate in contact with the electrode. It is believed that under most conditions the

binding to the surface is predominantly a bidentate or tridentate orientation giving P—

O—In bonds.16, 17

The multiple bonds formed with the ITO surface result in a well-

defined SAM though the coverages are typically lower than theoretically expected.

Modification of metal oxide electrode surfaces is a necessary component for the

design and use of electrocatalysts. One strategy that has been adopted is the use of

SAMs that feature azide functional groups that will result in azide-terminated

monolayers after deposition onto a surface.18-20

These terminal azide groups are then

reacted with ethynylated substrates through the Cu(I)-catalyzed azide alkyne coupling

reaction (CuAAC), as described by Sharpless.21

One of the advantages of this method

is that the coupling product, a triazole, is chemically stable and will also allow for facile

electron transfer between the substrate and the electrode surface. The click reaction,

typically employed in homogeneous solutions, has been shown to work on various

azide-terminated carbon, metal and metal oxide electrodes.18, 19, 22

The electrode

surfaces that are exposed to the mild conditions of the reaction are typically stable,

showing minimal or no change in structure or electronic properties. Furthermore, this

convergent approach to surface modification and derivitization greatly increases the

number of systems that can be studied experimentally.

In this work, a new p-azidophenyl phosphonic acid molecule was synthesized

and deposited onto an ITO electrode resulting in a well-defined azide-terminated SAM.

The deposition conditions were studied and optimized to obtain the most densely

packed monolayers possible. The optimized conditions were found to be deposition of

p-azidophenyl phosphonic acid onto ITO substrates from a 1 mM ethanolic solution at

24

80°C. Removing water from the ethanol was found to further increase the coverage of

the adsorbed phosphonic acid, highlighting the importance of performing the deposition

under anhydrous conditions. The resulting azide-terminated monolayers were then

reacted with various compounds having terminal alkynes via the copper (I) catalyzed

click reaction. Total reflectance Fourier transform infrared (FTIR) spectroscopy and X-

ray photoelectron spectroscopy (XPS) were used to characterize each step of the process

while cyclic voltammograms (CV) were obtained to calculate electrochemical coverage

when clicking with an electroactive species. The robustness of the adsorbed

phosphonate and clicked molecules under harsh potentials were examined to determine

overall stability of the phosphonate attachment and triazole. These azide-terminated

ITO electrodes are envisioned to be used as robust platforms for clicking various

ethynylated ligands and catalysts for use in electrocatalysis (Figure 2.1).

25

2.4: Materials and Methods

Reagents. Ethanol, acetonitrile (MeCN), dimethyl sulfoxide (DMSO), copper(II)

sulfate, sodium perchlorate, 1-ethynyl-4-(trifluoromethyl)benzene, ethynyl ferrocene,

ascorbic acid, and phenyl phosphonic acid were purchased from commercial sources

and used as received. The Cu(I)-stabilizing ligand TTMA (TTMA = tris-

(ethylacetyltriazolyl)methylamine) was synthesized using published procedures.23

Synthesis of the p-azidophenyl phosphonic acid is outlined below. The click catalyst

solution was made by adding 1 equiv. of CuSO4 to 1.1 equiv. of TTMA in deionized

(DI) water with 2 equiv. of the reductant ascorbic acid.

Synthesis of p-Azidophenyl Phosphonic Acid. The p-azidophenyl phosphonic acid

that was used for these studies was synthesized from 4-bromoacetanalide following

published synthetic procedures (Scheme 1).24-26

Scheme 1. Synthesis of p-azidophenyl phosphonic acid.

Substrates. Indium tin oxide (ITO) deposited onto glass slides was purchased from a

commercial source (Delta Technologies Limited, Stillwater, MN) and had a sheet

resistivity of 4-8 /. The ITO surfaces were cut into the appropriate sizes and rinsed

with ethanol followed by drying in a stream of N2. The ITO substrates were cleaned by

a variety of methods including O2 plasma (Harrick Plasma Expanded Cleaner, PDC-001

26

model) using O2:Ar (10:90) etching gas, sonication in various solvents and acid/base

rinse (aq. NH4OH followed by aq. HCl).11, 27-29

In general, these surface cleaning

techniques did decrease the surface carbon, as measured by XPS, but did not result in

increased N1s peak areas for the azide functional group nor increased electrochemical

coverages when clicked with ethynylferrocene.

Formation of Self Assembled Monolayers (SAMs). Monolayers of p-azidophenyl

phosphonic acid on ITO substrates were prepared by exposing the surface to a 1 mM

ethanolic solution at 80C for one hour. Deposition from other solvents did not result in

increased surface coverage of the p-azidophenyl phosphonic acid (see supporting

information). Mixed monolayers were formed by exposing the surface to an ethanolic

solution with the desired ratio of the p-azidophenyl phosphonic acid and the diluent

molecule, phenyl phosphonic acid. The total phosphonic acid concentration in the

solution was 1 mM. After 1 hour at 80°C, the samples were rinsed with copious

amounts of ethanol at room temperature and dried in a stream of N2 gas to remove

excess physisorbed or unattached phosphonic acid. All samples were analyzed or

clicked shortly after removal from the deposition solution.

Formation of Triazole. The azide-terminated monolayers were coupled with terminal

alkynes in the presence of a copper (I) catalyst. In a typical setup, a sample was placed

in a solution composed of the following: 50 μL of 10 mM ethynyl-terminated

compound, 50 μL of an aqueous 10 mM [Cu(TTMA)2+

](SO42-

) solution, 10 μL of an

aqueous 100 mM ascorbic acid solution, and 1.5 mL of DI water. The reaction vial was

capped and stored in the absence of light for the desired amount of time. After

completion, the substrate was rinsed with ethanol, sonicated for 5-10 min in fresh

27

DMSO, rinsed again with copious amounts of ethanol, and finally dried in a stream of

N2 gas. The click ligand TTMA was used over other ligands because of faster reaction

rates under the conditions employed (see supporting information).

Fourier Transform Infrared (FTIR) Reflection Spectroscopy. The relative amount

of p-azidophenyl phosphonic acid deposited onto ITO substrates was monitored with

total reflectance IR spectroscopy by integrating the area of the azide peak above the

background absorbance at ~2100 cm-1

. The setup for reflectance IR consisted of a p-

polarizer mounted on a reflectance accessory that reflected light off a gold mirror placed

85° from the normal to a 16 mm diameter aperture upon which a sample was placed.

The reflected light off the sample was directed to another 85° gold mirror, and then

refocused onto a deuterated triglycine sulfate (DTGS) detector. A diagram of this setup

is included in the supporting information. A total of 256 scans were taken at 4 cm-1

resolution for each sample. Spectra were taken against a background of bare ITO with

each spectrum baseline corrected with water and CO2 compensation performed. This

technique was used to monitor both the deposition of p-azidophenyl phosphonic acid

onto the surface and the completion of the click reaction on the azide-terminated

monolayers by tracking the disappearance of the azide signal as a function of time

exposed to the click reaction.

Electrochemical Measurements. A bored Teflon cone (0.25” inner diameter, area of

0.3167 cm2) pressed against the sample was used as the cell for electrochemical

measurements. The electrolyte solution was 0.1 M sodium perchlorate in either water

or acetonitrile. A platinum wire counter electrode and a glass frit-isolated

Ag(s)/AgCl(s)/sat. KCl(aq) reference electrode were suspended above the sample. The

28

potential was controlled and scanned using a WaveNow potentiostat (Pine

Instrumentation). All measurements were performed in air at room temperature at 1 V/s

scan rates unless noted otherwise.

X-Ray Photoelectron Spectroscopy (XPS). XPS measurements were conducted on a

PHI-XPS machine using a monochromatic Al Kα (1486.7 eV) X-ray source at an

incident angle of 45.0°. The azide functional group of arylazides is known to slowly

decompose photochemically and, potentially, with X-ray damage. To account for this,

decomposition studies were performed in situ on the deposited p-azidophenyl

phosphonic acid on ITO (see Supporting Information). Ten scans on the nitrogen 1s

(N1s) region resulted in negligible decomposition of the azide peak. High resolution

scans consisted of 10 scans at 0.1 eV resolution with the following elements being

analyzed: N1s, indium 3d5 (In3d5) and phosphorus 3p (P2p) regions. The integrations

of the N1s peaks were normalized to the area of the In3d5 peak of the same sample for

all data shown throughout this study.

Fitting Model for Nitrogen 1s Region for Azide Analysis. High-resolution scans of

the N1s region were fitted using standard graphing and fitting software. Organic azides

are known to give distinct XPS features, namely two peaks with area ratio of 2:1. The

model for fitting the nitrogen region was developed by analyzing a master spectrum

made by the co-addition of many individual spectra that were collected. The peak at

404 eV was fitted with one Gaussian while the peak at 400 eV was fitted with two

Gaussians of equal area and full-width half-max (FWHM). In addition, the peak at 400

eV had two small Gaussians with the same FWHM fitted to account for decomposition

products or other contaminant nitrogen organics (see supporting information). A

29

Shirley baseline was employed for the fitting model. All relative positions and FWHMs

of the peaks in the model were locked while the absolute position was allowed to float

as a set. High resolution scans were normalized to the area of the Indium 3d5 peak,

which was fit with two Gaussians with all parameters floating.

30

2.5: Results and Discussion

Deposition on ITO Surfaces Monitored by XPS. XPS offers a convenient

spectroscopic method that allows for monitoring elemental composition during various

points of the deposition and click reaction on the ITO surfaces. Figure 2.2 shows a

representative XPS spectrum of the N1s region before deposition (Figure 2.2A), after

deposition of p-azidophenyl phosphonic acid (Figure 2.2B), and after clicking with

ethynylferrocene (Figure 2.2C). A typical spectrum of deposited p-azidophenyl

phosphonic acid onto an ITO substrate results in one small peak for the phosphorus

atom at 133 eV (not shown), and two peaks for nitrogen atoms in the N1s region at 400

eV and 404 eV (Figure 2.2B). The peak areas of the nitrogen peaks are found in a ratio

of 2:1, corresponding to the two outer, electron-rich nitrogen atoms, and the one central,

electron-deficient nitrogen atom of the azide. The areas of the nitrogen peaks belonging

to the azide were determined by the fitting model described above, and were normalized

to the area of the In3d5 peak. The deposition of p-azidophenyl phosphonic acid onto

ITO was monitored over time in both heated and room temperature solutions using the

N1s/In3d5 peak area ratios (Figure 2.3). It was found that one hour in the heated

deposition solution was sufficient to obtain the maximum N1s/In3d5 ratio. The

deposition at room temperature took longer, typically overnight, to reach the coverage

obtained from the one hour heated deposition.

The ITO surfaces used for these studies were observed to be free of any nitrogen

contaminants before deposition of the phosphonic acid allowing for monitoring of the

deposition and the subsequent click reaction on the surfaces (Figure 2.2A). Peak fitting

analysis of the N1s region reveals that a minor portion of the peak at 400 eV contained

31

amine-like nitrogen, which could arise from decomposition of the azide in situ or

degradation of the p-azidophenyl phosphonic acid source used during deposition and

sample handling. The N1s region, after clicking with an ethynyl-terminated species,

reveals the loss of the peak at 404 eV and subsequent broadening and enlarging of the

400 eV peak, consistent with conversion of the electron-deficient nitrogen of the azide

to amine-like nitrogen of the resulting triazole (Figure 2.2C). Peak fitting analysis for

the reacted substrates verified that not all of the azide functional groups on the surface

are reacted to completion, typically with 20% unreacted remaining on the surface.

Analysis Using Total Reflection Infrared Spectroscopy. FTIR spectroscopy was

used to monitor the intensity of the azide signal on ITO substrates after deposition and

subsequent click reaction using a total reflection accessory. The aryl azide asymmetric

stretching mode appears at ~2100 cm-1

as a doublet, presumably due to coupling to aryl

ring modes (Figure 2.4). This azide signal was integrated and used to monitor the

relative coverage of the p-azidophenyl phosphonic acid onto the ITO surface. It was

confirmed that deposition from a heated ethanol solution for one hour resulted in the

highest coverages onto ITO substrates. FTIR was then used to monitor the completion

of the click reaction with ethynylferrocene. Full monolayers of p-azidophenyl

phosphonic acid on the surface were observed to not react to completion, typically with

a residual azide signal of ~20% remaining (Figure 2.4). The FTIR result for the

incomplete click reaction on the azide-terminated monolayers is consistent with the

results obtained using XPS that also show incomplete click (Figure 2.2C).

Azide Surface Coverage Measured by Electrochemistry. One of the best methods

available for quantifying the surface coverage of deposited monolayers is the

32

measurement of redox charge by electrochemistry. In this study, ethynylferrocene was

coupled to azide-terminated ITO substrates via the click reaction and used to determine

electrochemical coverage by integration of the current above the background charging

current in cyclic voltammograms (Figure 2.5). The resulting coverage for a full

monolayer of p-azidophenyl phosphonic acid on an ITO substrate clicked with

ethynylferrocene was 7.7(4) x 1013

molecules cm-2

under the deposition conditions

employed. Increasing the exposure time of the azide-terminated ITO substrates to the

click solution to overnight did not result in an increased electrochemical coverage or

further decrease of the azide integration monitored by FTIR. The IR integration of the

azide peak before click was 0.076 cm-1

and after click was 0.016 cm-1

(Figure 2.4).

Thus we conclude that the electrochemical coverage corresponds to (0.076–

0.016)/0.076 80% of the azide groups being clicked with ethynylferrocene. The

expected monolayer coverage for these monolayers would be 7.7 x 1013

/80% ≈ 1.0 x

1014

molecules cm-2

. Previous coverages obtained for similar ferrocene-derived

molecules, Fc(COOH) and Fc(CH2)6PO(OH)2, attached to an ITO substrate were

determined to be ~2 x 1014

molecules cm-2

and 2.7 x 1014

molecules cm-2

,

respectively.30, 31

The lower than expected coverage obtained for p-azidophenyl phosphonic acid

in this study could be a result of any of the following possibilities: (a) rough surface of

ITO, masking some azide sites while exposing others, (b) clustering or islanding of

azide molecules on the surface due to strong interactive forces between neighboring

molecules or inhomogeneities of surface composition, e.g. lack of exposed Lewis acid

binding sites, which does not allow for an even deposition of molecules, (c) the

33

phosphonic acid molecule oligomerizes on the surface, cross-linking with another

phosphonic acid molecule before reacting with surface hydroxyls, or (d) where the

steric demands of an attached ferrocene to one azide group blocks the catalyst approach

to an adjacent azide group, preventing the click reaction from occurring at the adjacent

site. It is evident that steric constraints play a role as mixed monolayers with lower

azide coverages react to completion, as determined by IR. A pre-clicked solution of

ethynylferrocene with p-azidophenyl phosphonic acid was prepared and deposited onto

an ITO substrate resulting in a coverage of 6 x 1013

molecules/cm2. This result suggests

that the low coverage is the result of the lack of Lewis acidic binding sites on the

surface of ITO. Furthermore, the deposition of the p-azidophenyl phosphonic acid onto

a fluorine doped tin oxide substrate (FTO, 4-8 Ω) resulted in lower coverages,

indicating the difficulty in obtaining pure monolayers for species that have to be reacted

to azide-terminated monolayers due to steric constraints.

Mixed Monolayers on ITO. It is common to have some disparity between the mole

fraction of adsorbates in the deposited layer on a substrate and the mole fraction of the

adsorbates in the deposition solution. Each phosphonic acid used could have different

solubilities in the deposition solvent that would potentially affect the kinetics and

thermodynamics of deposition onto surfaces. In this study, IR, XPS and

electrochemistry were used to determine how diluting the p-azidophenyl phosphonic

acid with phenyl phosphonic acid affected adsorption and coverage onto ITO substrates.

Mixtures of the p-azidophenyl phosphonic acid with the diluent phenyl phosphonic acid

were deposited from 1 mM ethanolic solutions that were heated to 80°C for one hour.

The IR, XPS and electrochemical coverage data was plotted against the mole fraction of

34

azide in solution, χazide (Figure 2.7). For each data set, a near linear relationship was

found to exist for the mixed monolayers, indicating deposition of the expected statistical

mixture of the phosphonic acids onto the ITO. This technique could be employed when

a specific coverage of the azide functional group is desired on the surface. Unlike the

pure monolayers of p-azidophenyl phosphonic acid, the samples with 50 % monolayers

were found to react completely with ethynylferrocene resulting in a negligible IR

integration (Figure 2.6A). The coverage of the 50 % monolayer was found to be ~5 x

1013

molecules cm-2

which would correlate well with the inferred full monolayer

coverage of 1.0 x 1014

molecules cm-2

(Figure 2.6B). This result offers further evidence

that the full monolayer of p-azidophenyl phosphonic acid on the surface is densely

packed in some areas of the surface making it difficult for the bulky catalyst/ligand

complex to approach the next free azide on the surface.

Effect of Water on Monolayer Formation. FTIR was used to study the effect of

water on the adsorption of p-azidophenyl phosphonic acid onto ITO in order to optimize

deposition conditions. It was found that using dry, anhydrous ethanol for the deposition

solvent increased the azide peak as measured by FTIR (Figure 2.8A). This peak

decreased when rinsed with copious amounts of ethanol after removal from the

deposition solution, indicating a large fraction of the p-azidophenyl phosphonic acid

was physisorbed to the surface. The effect of the presence of water during deposition is

not unexpected due to the dehydration reaction that must occur for a phosphonic acid to

deposit on a metal oxide surface. To drive the dehydration of the phosphonic acids that

were physisorbed on the surface after deposition, samples were heated to 105°C for 5-

10 minutes followed by an ethanol rinse to remove any physisorbed molecules. The

35

combination of depositing p-azidophenyl phosphonic acid in anhydrous ethanol

followed by heating the samples and rinsing with ethanol before analysis resulted in the

highest IR signal (Figure 2.8A).

Experiments were also performed to removal of p-azidophenyl phosphonic acid

from the ITO surface. The azide peak obtained after the pure monolayer was formed

from dry ethanol solvent followed by heating in an oven and subsequent rinse with

ethanol was exposed to either an 80°C dry ethanol solvent or an 80°C ethanol/water

(50/50) solvent for one hour. The samples exposed to the dry ethanol solvent had

similar IR integrations while the samples exposed to the ethanol/water solvent lost a

significant amount of azide IR signal (Figure 2.8B). Exposure to an aqueous solvent at

room temperature, however, appeared to have little to no effect on the azide IR signal.

These results highlight the importance keeping the monolayers from being exposed to

elevated temperatures when water is present due to desorption of p-azidophenyl

phosphonic acid from ITO. Samples deposited in dry ethanol and subsequently heated

at 105°C did not produce increased electrochemical coverages with ethynylferrocene.

However, assuming that an IR integration of 0.076 cm-1

corresponds to an

electrochemical coverage of 1.0 x 1014

molecules cm-2

, an IR integration of 0.15 cm-1

obtained for these samples deposited in dry ethanol, after heating and rinsing, would

correspond to a coverage of 2.0 x 1014

molecules cm-2

, which is approaching the

expected full monolayer values.14, 15, 32

Stability of the Triazole and Phosphonate/ITO Attachement. A phosphonic acid

molecule was used as the attachment molecule onto ITO due to the stability of the

phosphonate attachment to the surface. Phosphonates attached to metal oxide surfaces

36

are known to attach in a bi- or tri-dentate mode, resulting in a linkage that is stable

under a variety of pH’s, temperatures, and potentials. The exposure of the attachment

to elevated aqueous solutions was found to remove the phosphonate from the ITO

surface (Figure 2.8B). The product after employing the click reaction, a di-substituted

1,2,3-triazole, is thought to be robust while still offering facile electron transfer from an

electroactive species immobilized to the electrode. To test this stability towards

oxidative conditions, p-ethynyl-(trifluoromethyl)benzene was clicked to freshly

prepared azide-terminated monolayers on ITO. Analysis of the N1s, P2p, and F1s

regions by XPS was used to monitor atomic composition before and after being

subjected to an oxidizing potential of 1.5 V vs. Ag/AgCl/KCl in 0.1 M sodium

perchlorate in MeCN. The F1s region of the XPS spectrum indicates that the peak

before and after exposure to the oxidizing conditions were similar (Figure 2.9). Similar

analysis on pure azide-terminated monolayers reveals that the azide signal before and

after exposure to oxidizing conditions remains the same, as measured by FTIR (see

supporting information).

37

2.6: Conclusion

This report highlights the ability to functionalize an ITO surface with p-

azidophenyl phosphonic acid that will allow for coupling of ethynylated compounds

through click chemistry. Deposition of the phosphonic acid from anhydrous ethanol

solutions resulted in the highest inferred coverages. The azide functional group was

clicked with ethynylferrocene resulting in 1,4-disubstituted 1,2,3-triazoles that allowed

for facile electron transfer between the tethered ferrocene molecule and the ITO

electrode. The phosphonate attachment to the surface, and the resulting triazole formed

after the CuAAC reaction, are robust and stable when exposed to a variety of oxidizing

potentials, although the attachment was observed to desorb from the surface when

exposed to elevated temperatures of aqueous solutions. Electrochemical coverages

were found to be short of theoretical monolayers. Despite the lower than expected

coverages obtained, the deposition of p-azidophenyl phosphonic acid resulted in a well-

behaved SAM that is envisioned to provide an oxidatively stable platform for clicking

various ethynyl-terminated electrocatalysts.

38

2.7: Figures

Figure 2.1: Immobilization Schematic. Schematic view of process for immobilizing

the p-azidophenyl phosphonic acid onto an ITO surface. The second step highlights

how the click reaction can be employed to immobilize various ethynyl-terminated

molecules from solution. This strategy allows for a ‘plug and play’ methodology that

will allow various catalysts of interest to be studied electrochemically.

39

Figure 2.2: XPS Analysis. XPS analysis of N1s region of ITO (A) before deposition;

(B) after deposition of p-azidophenyl phosphonic acid; and (C) after clicking the azide

terminated ITO in (B) with ethynylferrocene.

A.

B.

C.

40

Figure 2.3: Deposition of p-Azidophenyl Phosphonic Acid. Deposition of p-

azidophenyl phosphonic acid onto ITO monitored at room temperature () and 80°C ()

by XPS. The areas of the N1s peaks corresponding to the azide were calculated and

normalized to the area of the In3d5 peak using the fitting method described in the text

and supporting materials.

41

Figure 2.4: FTIR Spectrum Before and After Click. IR spectrum of azide region

appearing at ~2100 cm-1

for a monolayer of p-azidophenyl phosphonic acid after

deposition (―) and after clicking with ethynylferrocene (- - -). The remaining azide peak

after click corresponds to ~15-20 % of the unclicked azide integration.

42

Figure 2.5: Cyclic Voltammogram of Clicked Ethynylferrocene. Cyclic

voltammogram of a full monolayer of p-azidophenyl phosphonic acid on ITO clicked

for 3 hours with ethynylferrocene. CV obtained at 1 V/s using a 0.1 M NaClO4 in

MeCN solution. The coverage corresponds to ~8 x 1013

molecules/cm2.

43

Figure 2.6: Click Results for Azide-terminated Monolayers. IR (A) and

Electrochemical (B) results obtained after clicking full () and 50% () azide-

terminated monolayers with ethynylferrocene for the given times with the

CuSO4/TTMA catalyst system. The p-azidophenyl phosphonic acids were deposited

with wet ethanol and heated then rinsed before analysis. Electrochemical coverages in

(B) correspond to the value x 1013

molecules/cm2.

A.

B.

44

Figure 2.7: Mixed Monolayers on ITO. Plots of mixed monolayers deposited onto

ITO surfaces for 1 hr at 80°C in 1 mM (azide + diluent) ethanol solutions monitored by

(A) XPS, (B) IR spectroscopy, and (C) electrochemistry (x 1013

molecules/cm2) after

clicking mixed monolayers with ethynylferrocene for 3 hrs. The XPS integration in (A)

was obtained using the fitting method described in the text and in the supporting

materials and normalizing to the In3d5 peak of each sample. The IR integration in (B)

was obtained by integrating the area of the azide peak appearing at ~2100 cm-1

for each

sample. The diluent molecule for each data set was phenyl phosphonic acid.

A.

B.

C.

45

Figure 2.8: Effect of Water on p-Azidophenyl Phosphonic Acid Deposition. (A) IR

integrations of the azide peak under various deposition conditions. No rinsing is for

samples that were removed from the deposition solution and dried in a stream of N2.

Heated samples were placed in an oven at 105-110°C for 5-10 minutes then rinsed

thoroughly with ethanol and dried in a stream of N2. Samples with no heat were

removed from the deposition solution, rinsed thoroughly with ethanol and dried in a

stream of N2. The inferred coverage corresponds to the value that would be expected if

every azide on the surface was reacted with ethynylferrocene. (B) Removal of azide

from the surface by exposure of as deposited samples to dry ethanol and ethanol/water

A.

B.

46

(50/50) at 80°C for one hour.

Figure 2.9: Stability of Azide on ITO Surface. High-resolution scans of the F1s peak

by XPS before (―) and after (- - -) exposure to 1.5 V vs. Ag/AgCl for 5 minutes using a

solution of 0.1 M NaClO4 in MeCN.

47

2.8: References and Notes

1. Porter, M. D.; Bright, T. B.; Allara, D. L.; Chidsey, C. E. D., Spontaneously

organized molecular assemblies. 4. Structural characterization of n-alkyl thiol

monolayers on gold by optical ellipsometry, infrared spectroscopy, and

electrochemistry. J. Am. Chem. Soc. 1987, 109, (12), 3559-3568.

2. Chidsey, C. E. D.; Bertozzi, C. R.; Putvinski, T. M.; Mujsce, A. M.,

Coadsorption of ferrocene-terminated and unsubstituted alkanethiols on gold:

electroactive self-assembled monolayers. J. Am. Chem. Soc. 1990, 112, (11),

4301-4306.

3. Chidsey, C. E. D.; Loiacono, D. N., Chemical functionality in self-assembled

monolayers: structural and electrochemical properties. Langmuir 1990, 6, (3),

682-691.

4. Ulman, A., Formation and Structure of Self-Assembled Monolayers. Chem. Rev.

1996, 96, (4), 1533-1554.

5. Gardner, T. J.; Frisbie, C. D.; Wrighton, M. S., Systems for orthogonal sefl-

assembly of electroactive monolayers on Au and ITO - an approach to molecular

electronics. J. Am. Chem. Soc. 1995, 117, (26), 6927-6933.

6. Armstrong, N. R.; Veneman, P. A.; Ratcliff, E.; Placencia, D.; Brumbach, M.,

Oxide contacts in organic photovoltaics: characterization and control of near-

surface composition in indium-tin oxide (ITO) electrodes. Acc. Chem. Res.

2009, 42, (11), 1748-1757.

7. Carter, C.; Brumbach, M.; Donley, C.; Hreha, R. D.; Marder, S. R.; Domercq,

B.; Yoo, S.; Kippelen, B.; Armstrong, N. R., Small molecule chemisorption on

48

indium-tin oxide surfaces: enhancing probe molecule electron-transfer rates and

the performance of organic light-emitting diodes. J. Phys. Chem. B 2006, 110,

(50), 25191-25202.

8. Yan, C.; Zharnikov, M.; Golzhauser, A.; Grunze, M., Preparation and

characterization of self-assembled monolayers on indium tin oxide. Langmuir

2000, 16, (15), 6208-6215.

9. Mingalyov, P. G.; Lisichkin, G. V., Chemical modification of oxide surfaces

with organophosphorus(v) acids and their esters. Rus. Chem. Rev. 2006, 75, (6),

541-557.

10. Paniagua, S. A.; Hotchkiss, P. J.; Jones, S. C.; Marder, S. R.; Mudalige, A.;

Marrikar, F. S.; Pemberton, J. E.; Armstrong, N. R., Phosphonic acid

modification of indium-tin oxide electrodes: Combined XPS/UPS/contact angle

studies. J. Phys. Chem. C 2008, 112, (21), 7809-7817.

11. Donley, C.; Dunphy, D.; Paine, D.; Carter, C.; Nebesny, K.; Lee, P.; Alloway,

D.; Armstrong, N. R., Characterization of indium-tin oxide interfaces using X-

ray photoelectron spectroscopy and redox processes of a chemisorbed probe

molecule: Effect of surface pretreatment conditions. Langmuir 2002, 18, (2),

450-457.

12. Koh, S. E.; McDonald, K. D.; Holt, D. H.; Dulcey, C. S.; Chaney, J. A.;

Pehrsson, P. E., Phenylphosphonic acid functionalization of indium tin oxide:

Surface chemistry and work functions. Langmuir 2006, 22, (14), 6249-6255.

49

13. Gao, W.; Dickinson, L.; Grozinger, C.; Morin, F.; Reven, L., Self-assembled

monolayers of alkylphosphonic acids on metal oxides. Langmuir 1996, 12, (26),

6429-6435.

14. Clearfield, A.; Smith, G. D., Crystallography and structure of .alpha.-zirconium

bis(monohydrogen orthophosphate) monohydrate. Inorg. Chem. 1969, 8, (3),

431-436.

15. Cao, G.; Hong, H. G.; Mallouk, T. E., Layered metal phosphates and

phosphonates: from crystals to monolayers. Acc. Chem. Res. 1992, 25, (9), 420-

427.

16. Hotchkiss, P. J.; Li, H.; Paramonov, P. B.; Paniagua, S. A.; Jones, S. C.;

Armstrong, N. R.; Bredas, J.-L.; Marder, S. R., Modification of the Surface

Properties of Indium Tin Oxide with Benzylphosphonic Acids: A Joint

Experimental and Theoretical Study. Adv. Mater. 2009, 21, (44), 4496-4501.

17. Paramonov, P. B.; Paniagua, S. A.; Hotchkiss, P. J.; Jones, S. C.; Armstrong, N.

R.; Marder, S. R.; Bredas, J.-L., Theoretical Characterization of the Indium Tin

Oxide Surface and of Its Binding Sites for Adsorption of Phosphonic Acid

Monolayers. Chem. Mater. 2008, 20, (16), 5131-5133.

18. Collman, J. P.; Devaraj, N. K.; Chidsey, C. E. D., "Clicking" functionality onto

electrode surfaces. Langmuir 2004, 20, (4), 1051-1053.

19. Collman, J. P.; Devaraj, N. K.; Eberspacher, T.; Chidsey, C. E. D., Mixed azide-

terminated monolayers: A platform for modifying electrode surfaces. Langmuir

2006, 22, (6), 2457-2464.

50

20. Devaraj, N. K.; Decreau, R. A.; Ebina, W.; Collman, J. P.; Chidsey, C. E. D.,

Rate of interfacial electron transfer through the 1,2,3-triazole linkage. J. Phys.

Chem. B 2006, 110, (32), 15955-15962.

21. Rostovtsev, V.; Green, L.; Fokin, V. V.; Sharpless, K., A stepwise Huisgen

cycloaddition process: Copper(I)-catalyzed regioselective "ligation" of azides

and terminal alkynes. Angew. Chem. Int. Edit. 2002, 41, (14), 2596-2599.

22. Devadoss, A.; Chidsey, C. E. D., Azide-modified graphitic surfaces for covalent

attachment of alkyne-terminated molecules by "click" chemistry. J. Am. Chem.

Soc. 2007, 129, (17), 5370-5372.

23. Zhou, Z.; Fahrni, C., A fluorogenic probe for the copper(I)-catalyzed azide-

alkyne ligation reaction: Modulation of the fluorescence emission via (3)(n,pi*)-

(1)(pi,pi*) inversion. J. Am. Chem. Soc. 2004, 126, (29), 8862-8863.

24. Cooper, R.; Camp, P.; Gordon, R.; Henderson, D.; Henry, D.; Mcnab, H.; De

Silva, S.; Tackley, D.; Tasker, P.; Wight, P., The assembly of rotaxane-like

dye/cyclodextrin/surface complexes on aluminium trihydroxide or goethite.

Dalton Trans. 2006, (23), 2785-2793.

25. Mohapatra, S.; Pramanik, P., Synthesis and stability of functionalized iron oxide

nanoparticles using organophosphorus coupling agents. Colloid Surface A 2009,

339, (1-3), 35-42.

26. Kim, T.; Kim, K., N-Butyllithium-Mediated Reactions of 1-(2-

Azidoarylmethyl)-1H-benzotriazoles with Alkyl Halides. J. Heterocyclic Chem.

2010, 47, (1), 98-111.

51

27. Besbes, S.; Ben Ouada, H.; Davenas, J.; Ponsonnet, L.; Jaffrezic, N.; Alcouffe,

P., Effect of surface treatment and functionalization on the ITO properties for

OLEDs. Mat. Sci. Eng. C-Bio. 2006, 26, (2-3), 505-510.

28. Brumbach, M.; Veneman, P. A.; Marrikar, F. S.; Schulmeyer, T.; Simmonds, A.;

Xia, W.; Lee, P.; Armstrong, N. R., Surface composition and electrical and

electrochemical properties of freshly deposited and acid-etched indium tin oxide

electrodes. Langmuir 2007, 23, (22), 11089-11099.

29. Chaney, J.; Koh, S.; Dulcey, C. S.; Pehrsson, P. E., Surface chemistry of carbon

removal from indium tin oxide by base and plasma treatment, with implications

on hydroxyl termination. Appl. Surf. Sci. 2003, 218, (1-4), 258-266.

30. Zotti, G.; Schiavon, G.; Zecchin, S.; Berlin, A.; Pagani, G., Adsorption of

ferrocene compounds an indium-tin-oxide electrodes. Enhancement of

adsorption by decomposition of ferrocenium molecules by oxygen. Langmuir

1998, 14, (7), 1728-1733.

31. Vercelli, B.; Zotti, G.; Schiavon, G.; Zecchin, S.; Berlin, A., Adsorption of

hexylferrocene phosphonic acid on indium-tin oxide electrodes. Evidence of

strong interchain interactions in ferrocene self-assembled monolayers. Langmuir

2003, 19, (22), 9351-9356.

32. Gui, J. Y.; Stern, D. A.; Lu, F.; Hubbard, A. T., Surface chemistry of 5-

membered heteroaromatics at Pt(111) electrodes studied by EELS, LEED,

Auger-spectroscopy and electrochemistry - furan, pyrrole and thiophene. J.

Electroanal. Chem. 1991, 305, (1), 37-55.

52

2.9: Supporting Information

Cleaning methods for ITO. A number of different cleaning procedures were

investigated in order to increase the surface coverage of the p-azidophenyl phosphonic

acid. The ITO samples were cut into appropriate sizes and rinsed with ethanol followed

by drying in a stream of N2 gas. Some samples were exposed to an oxygen plasma

(90:10 Argon:Oxygen) for up to 10 minutes to remove any adventitious carbon-

containing molecules that might have been on the surface. Other samples were exposed

to the plasma followed by exposure to a mild acid bath. A solvent rinse and sonication

in either ethanol, chloroform or acetone was also investigated for ability to remove any

contaminants from the ITO surface. X-ray photoelectron spectroscopy (XPS) was

performed to determine the atomic composition on the surface. The samples were

analyzed for carbon 1s (C1s), indium 3d5 (In3d5), tin 3d5/2 (Sn3d5/2) and nitrogen 1s

(N1s) with results shown in Figure S1 below. It was found that although the plasma

cleaning appeared to decrease the amount of adventitious carbon, the samples did not

yield increased p-azidophenyl phosphonic acid coverages as measured by Fourier

transform infrared (FTIR) spectroscopy and electrochemistry after clicking with

ethynylferrocene. Results shown in Figure 2S indicate that the deposition on samples

without pretreatment or cleaning yielded the highest coverages. This result appears to

be consistent with the hypothesis that the deposition conditions, an ethanolic solution at

80°C, acted as the best method of cleaning the surface. For this reason the samples used

throughout this study were not pre-cleaned by any technique except an ethanol rinse to

remove any glass pieces and dust that may have been present after cutting the ITO.

53

Peak fitting analysis. A peak fitting procedure was developed to analyze the peaks

appearing at the N1s region of the XPS spectrum related to the azide functional group

of the p-azidophenyl phosphonic acid. Standard mathematical fitting and analysis

software was used to perform the fits and integrations. The method consisted of first

setting a pure Gaussian peak for each of the three nitrogen atoms of an azide functional

group: one for the signal at a binding energy of 404 eV (peak 1) and two within the

signal at a binding energy of 400 eV (peak 2 and peak 3). The peak areas of these three

Gaussian curves were locked to the same area and relative position to each other. Two

additional smaller Gaussian peaks (peak 4 and peak 5) were also used within the signal

at a binding energy of 400 eV to account for any nitrogen contaminates or azide

decomposition products that may have been introduced or formed in the vacuum

chamber of the XPS instrument. Peak 4 and peak 5 were allowed to float in area with

their positions locked. The full-width half-max values and relative peak locations for

all five peaks were also locked with the locations allowed to move as a set due to drift

in the instrument from day to day use. The optimal peak placement for each of the five

Gaussian peaks was determined by addition of many spectra and fitting of this multiple

sample spectrum until the residuals were optimized. The method was completed by

employing a Shirley baseline through the noise of the spectrum that was to be analyzed.

An example of an azide group being fit by this method is shown in Figure S3. The

integration of the N1s region corresponding to the three azide nitrogen atoms was

normalized to the In3d5 peak integration of each sample, which typically remained

within ~10 % range from sample to sample.

54

X-ray decomposition of azide functional group during XPS analysis. Figure S4

below shows the combined integration for all three of the azide peaks, peak 4 area, and

peak 5 area on the same spot of an ITO substrate as a function of scans taken. This

analysis was used to determine the maximum amount of X-ray damage the samples

could withstand in situ while yielding the best signal-to-noise ratio. A maximum of ten

scans was found to be the optimal number of scans for obtaining data before

decomposition of the azide group by X-ray damage during the experiment. The ten

scans were used for each narrow high resolution azide region analysis throughout this

study.

FTIR total reflection accessory. A total reflection accessory was designed and built

for analysis of ITO slides by absorption FTIR spectroscopy. A schematic of the

accessory is shown in Figure S5. The accessory had two gold mirrors angled at 10° to

normal to allow for reflection of light onto and off the sample at a near grazing angle.

The accessory was also equipped with a p-polarizer to filter out s-polarized light. The

samples were placed ITO side down on a 16 mm diameter aperture and scanned 256

times by a deuterated triglycine sulfate (DTGS) detecotor.

Deposition of p-azidophenyl phosphonic acid. Various deposition conditions of the

p-azidophenyl phosphonic acid were attempted in this study. Figure S6 shows the

comparison of depositing the p-azidophenyl phosphonic acid in both ethanol and

tetrahydrofuran solvents (Figure S6A) and from ethanol in the presence of 1 equivalent

of either acetic acid or pyridine present (Figure S6B). It was found that a pure ethanol

deposition yielded the best coverages as determined by IR integration of azide signal

55

and electrochemically by integrating the peak that appeared after clicking with

ethynylferrocene.

Comparison of click conditions. Three click systems were compared in this study to

determine which was most effective at clicking the p-azidophenyl phosphonic acid that

was deposited onto ITO surfaces. The first was a CuSO4 salt in water with no

accelerating ligand, the second a CuSO4 salt with 1.1 equivalents of tris-

(triazolylbenzyl)methyl amine (TBTA) in water, and the third a CuSO4 salt with 1.1

equivalents of TTMA in water. The disappearance of the azide group at ~2100 cm-1

was followed by IR. The integrations were plotted as a function of time of exposure to

the various click reaction catalyst systems. The results are shown in Figure S7. It was

found that the CuSO4/TTMA click system was the most effective, typically being

completed within an hour. The CuSO4/TBTA and CuSO4/no ligand systems click

slowly over time, eventually reaching the values obtained by CuSO4/TTMA system.

Stability of azide on ITO surface. The azide-terminated monolayer was exposed to a

potential bias of 1.5 V vs. Ag/AgCl for 5 minutes in 0.1 M sodium perchlorate in

MeCN. The azide signal was found to decrease slightly by IR after exposure. The

spectra are shown in Figure S8. This result indicates that the attachment is robust for

extended periods of time when exposed to such oxidative conditions.

56

Figure 2.S1. XPS analysis of atomic species on the surface of ITO after cleaning with

the indicated procedures. The carbon percentage decreases with the cleaning

procedures while the relative indium and tin ratios increase. This cleaning did not lead

to improved coverages of p-azidophenylphosphonic acid on the ITO surface however.

Surface Analysis of Cleaned Samples

0

10

20

30

40

50

60

70

Carbon Indium Tin Nitrogen

Atom

Su

rfa

ce

Ato

mic

Pe

rce

nt

Uncleaned

Plasma Cleaned

Plasma + Acid

57

Figure 2.S2. IR spectra of azide region after p-azidophenyl phosphonic acid was

deposited for one hour from a 1 mM ethanolic solution at 80°C on ITO slides that were

uncleaned (), cleaned by oxygen plasma (), and cleaned with oxygen plasma

followed by mild acidic rinse ().

58

Figure 2.S3. Example of XPS peak fitting analysis of N1s region for azide functional

group in this work. Peaks 1, 2, and 3 all correspond to a nitrogen atom within the azide

group and are locked in peak area to each other. Peak 4 corresponds to nitrogen

contamination and peak 5 to a decomposition product of the azide in the chamber. All

peaks are locked to the same full-width at half-max and relative positions to each other,

although they are allowed to float as a set to account for instrument drift.

59

Figure 2.S4. Decomposition of the azide peaks and overall nitrogen signal as a

function of the number of scans taken during an XPS analysis. It was determined that

ten scans were optimal to obtain N1s data while minimizing in situ X-ray damage.

60

Figure 2.S5. Schematic of total reflection accessory used in absorption FTIR

spectroscopy.

61

Figure 2.S6. Representative cyclic voltammograms of p-azidophenyl phosphonic acid

on ITO clicked with ethynylferrocene. The p-azidophenyl phosphonic acid was

deposited (A) without acid or base in ethanol () and THF () and (B) with either one

equivalent of acetic acid () or one equivalent of pyridine () to the azide in ethanol.

62

Figure 2.S7. Comparison of click catalyst/ligand systems used in this study. The IR

integration of the full azide peak from the azide-terminated monolayer was monitored

as a function of time exposed to the indicated click-ligand system.

63

Figure 2.S8. IR spectra of samples before (―) and after (- - -) exposure to 1.5 V vs.

Ag/AgCl in 0.1 M sodium perchlorate for 5 minutes. The integration was found to

decrease slightly after exposure, indicating the oxidative stability of the p-azidophenyl

phosphonic acid on the ITO surface.

64

Chapter 3: Introduction

3.1: Abstract

This chapter is as an introduction to the work that was done in Chapters Four

and Five. This work was part of a solar water splitting project in collaboration with

Prof. Paul McIntyre and one of his graduate students, Vincent Chen, of the Materials

Science and Engineering department at Stanford. This chapter will open with a

discussion about the motivating factors behind our work, namely the increasing demand

for clean energy for an increasing world population. Following this will be a short

review of how accomplishing solar water splitting can be achieved, with a focus on

water oxidation at a photoanode. The conclusion of this chapter will introduce some

techniques we used in the following two chapters of my thesis.

65

3.2: The Global Energy Challenge

Meeting the ever-growing demand for energy on a global scale while

simultaneously preserving the environment presents a formidable challenge for

scientists. It is predicted, that by the year 2050, our global energy demand could triple

from our 2010 consumption rate of 15 terawatts (TW) to 45 TW.1 This projection could

vary depending on a number of parameters that include population growth rates, wealth

of these populations (GDP), energy consumption and conservation rates as well as

political concerns. Currently on the global scale, most energy consumed comes from

fossil-based fuel sources (~80%) with the balance being supplemented with

unsustainable bio-mass, nuclear and renewable sources (Figure 3.1A).2 One concern as

the year 2050 approaches is how the predicted 45 TW of global energy demand will be

met (Figure 3.1B).

There are a number of strategies that could be employed to address the rising

global energy demand. Some strategies are more feasible than others while all could be

used in any combination to solve the energy challenge. The first is that people could

start to conserve energy at unprecedented rates. Technology could also play a role by

allowing us to use energy more efficiently. Another strategy is to keep burning fossil-

based fuels at alarming rates and hope that either: (1) nothing happens to the

environment from the resulting carbon dioxide emissions or (2) that adequate carbon

capture and sequestration technologies are developed. Another attractive strategy is to

develop a cheap, carbon-free, renewable source of alternative energy. This could

include any or all of the following sources: solar, hydro-electric, wind, biomass,

geothermal, and tidal. Any of the strategies to be used will also have to meet the base

66

requirement of being both cheap and globally scalable, as to adequately compete with

the currently-used fossil fuel based sources of energy.

The three primary types of fossil fuel based energy are oil, coal and natural gas.

Currently, these energy sources make up the majority of the total global energy

consumption (Figure 3.1A). The high energy density, low cost of use and seemingly

endless supply of these sources (see below) are the main reasons they are currently used

to provide most of our energy. For example, coal is widely considered the cheapest

way to produce electricity at about 1-4 cents per kilowatt hour (kWh) on average.2 The

next cheapest source is natural gas, costing about 2-5 cents per kWh. Oil is the next

cheapest, costing between 6-8 cents per kWh on average. For comparison, the cheapest

renewable source of electricity is wind, which at 5-7 cents per kWh compares favorably

to oil, but cannot compete with coal or natural gas when it comes to how cheaply

electricity can be produced from it. These costs are summarized in Table 3.1 below.

Table 3.1: Cost per kWh for various sources of electricity.2, 3

Source Cost (¢/kWhr) Source Cost (¢/kWhr)

Coal 1-4 Wind 5-7

Natural Gas 2-5 Nuclear 6-7

Oil 6-8 Solar 25-50

To go along with the low cost of electricity production using fossil fuels as the

energy source, there appears to be an abundant supply of these sources. For example, it

is estimated that we have a total (proven and unproven reserves) of about 50-150 years

of oil, 200-600 years of natural gas, and about 2000 years of coal, all based on burn

rates from the year 2000.2 The big error bars associated with these values are due to

availability and ability to obtain these sources, some of which are in remote, hard-to-

67

access regions of the planet. One point that is clear, however, is that we will not run out

of these resources anytime in the immediate future.

In summary, it is clear that the planet will provide many years of fossil fuel

energy for human use. These sources in principle can be used to meet our ever-growing

energy requirements; however, careful consideration for the environmental

consequences of using these sources should be exercised. A more responsible strategy

is to take advantage of the renewable alternative sources of energy, which are not only

abundant and relatively untapped, but also are free of any carbon, and thus do not

produce CO2 when being used.

Renewable and Alternative Sources of Energy

As mentioned above, there are a number of various alternative energy sources

available that could be used in hopes of solving the growing energy demand. Any of

these sources could be used in combination with each other, and each has some obvious

strengths and weaknesses that need to be addressed. The one common theme for all of

these energy sources, however, is that they all produce energy with a minimal amount

of carbon dioxide emissions. Some of these sources produce electricity as a result of

their use, e.g. wind, and some produce a fuel, such as ethanol from sustainable biomass.

The only alternative carbon-free source of energy that can meet our entire future

global energy demand by itself is solar. It has been calculated that sunlight provides

120,000 TW of power that strikes the earth, while an estimated total of 800 TW could

feasibly be collected.2, 4

The most common strategy used to harness solar energy is

68

through the use of photovoltaics (PVs), which are semiconductor materials that will

collect photons which in turn will produce electricity in the form of direct current.

There have been many advancements in the solar industry over the years, but despite

these, PVs are still relatively inefficient, and the cost of the materials used to make PVs

and the cost to install them remains high. As a result, some of the best commercially

available PVs collect incident sunlight with about 10-20% efficiency, while still costing

the user around $300 per square meter.

One drawback of using solar energy as an alternative energy source in any

capacity is the fact that sunlight is intermittent while our energy demands are

continuous. As a peak power source, solar is readily available, as PVs will readily

provide vast quantities of electricity. On cloudy days, or at night, however, electricity

will not be available from PVs. Complicating the matter further is that current

technology does not have a great method to store electricity for off peak use. Currently,

one of the best strategies for storing electricity is pumping water uphill using electricity

obtained by a PV during the day and then letting the water run downhill and through a

turbine at night to get the electricity back. The same idea is used in storing energy by

compressing gas in the day and allowing the gas to expand through a turbine at night.

Batteries and capacitors are a well-known source for energy storage and could also find

use for storing solar energy. Another energy storage mechanism is utilizing chemical

bonds of a fuel. A very promising strategy is to use sunlight to drive the splitting of

water into molecular oxygen and molecular hydrogen, a process known as solar water

splitting. The hydrogen could be used as a fuel source directly, or converted into a

liquid fuel that is easier to use and handle through a few synthetic steps.

69

3.3: Solar Water Splitting

In order to take advantage of all the solar energy that strikes the earth, it is

necessary to store some of the collected daytime sunlight energy for use during the

nighttime. The direct conversion of solar energy to chemical fuels would address this

issue. There are many solar fuel syntheses that, in principle, should work. The best

strategy is performing solar water splitting, which is a process that uses solar energy to

split water to produce molecular oxygen and hydrogen, the ladder of which can be used

as a fuel. There are two common methods employed to accomplish this: (1) using

stable semiconductors to absorb light and create free charge carriers (holes and

electrons) which can be used to split water at the surface/liquid interface, and (2) using

a PV to collect light and create a current, which can be used to run an electrolyzer to

split water. Regardless of the approach taken, any solar water splitting scheme or

system would ultimately have to consist of materials that will collect incident solar

irradiation, generate the appropriate charge carriers and catalyze the oxidation and

reduction of water to make oxygen and hydrogen, respectively, as shown in equation

3.1:

Water oxidation occurs at a standard reversible potential (E°) of 1.23 V versus

the Normal Hydrogen Electrode (NHE) at pH 0. Proton reduction occurs at a standard

reversible potential of 0 V at pH 0, which is the definition of NHE. Both of these

reactions are highly pH dependent as the reversible potentials will shift the Nernstian 59

mV per pH unit change. The kinetic bottleneck of the water splitting reaction has long

70

been considered the water oxidation half reaction of equation 3.1 due to large

overpotentials needed for the reaction to occur. This is due to the necessity of having to

ultimately remove four protons and four electrons from two water molecules with the

subsequent formation of an oxygen-oxygen double bond.

The first requirement of any solar water splitting design is the efficient

absorption of solar irradiation. The solar output spectrum is shown in Figure 3.2, and is

commonly referred to as AM1.5G solar irradiation.5 Materials that can be used in

practical devices and will collect solar irradiation most efficiently are semiconductors.

These are materials that will absorb photons and promote electrons across a bandgap,

which is the energy difference between the valence band and conduction band of the

material. In order to split water, a semiconductor with a minimum bandgap of at least

1.23 eV is needed. Quite often, larger band gaps are required to fulfill water splitting.

These larger band gaps are needed to overcome the overpotential (η) of the water

oxidation and proton reduction reactions. Figure 3.3 depicts a number of band gaps for

some of the more commonly used semiconductors used in solar water splitting setups.6

Semiconductors used for solar water splitting must produce photo-generated

holes oxidizing enough to inject into water to make oxygen while the electrons must

have enough energy to combine with protons to make hydrogen. This can be

accomplished by choosing a semiconductor, or multiple semiconductors, that supply the

proper photovoltage to split water, or by using a catalyst layer attached to a small band

gap semiconductor that allows the band edges to move independently from the solution

when a bias is applied. In either case, the semiconductor must also have a geometry

that will allow a photogenerated charge carrier to migrate through the material and

71

perform the oxidation of water and reduction of protons before thermal recombination

occurs. These destructive recombination pathways are either radiative or nonradiative,

with pathways occurring (1) in the bulk of the semiconductor, (2) in the depletion layer

of the semiconductor, (3) by the majority carrier tunneling through the electric field at

the semiconductor/liquid junction, (4) by the majority carrier thermally escaping the

interfacial barrier and (5) at defect sites within the semiconductor.7 These

recombination pathways, illustrated in Figure 3.4, decrease overall photocurrent, and

thus efficiency, that could otherwise be obtained by the device.

The bandgap of a semiconductor must also not be too large in order to split

water with any efficiency. This is because all semiconductors collecting light with

energy greater than the bandgap will experience thermalization losses from carriers

relaxing to the band edge.8 As a general rule of thumb, a good photovoltaic device will

generate at most about two-thirds of its band gap in photovoltage at 1 sun illumination.

The system reported by Honda and Fujishima, which uses titanium dioxide (bandgap =

3.0 eV) as the semiconductor, is restricted to the UV portion of the solar spectrum and

limits the total current density that can be obtained to a maximum of 2 mA/cm2.9 The

maximum efficiency of any photovoltaic design, whether a single semiconductor or

multi-junction setup, is governed by both the fraction of solar energy absorbed and

photovoltage obtained. Both of these parameters are related to the bandgap of the

material(s) used. Even with the best catalysts used for water splitting, the

overpotentials associated with water oxidation and reduction at 1 sun illumination

would still add an additional voltage (~few hundred millivolts, see Table 3.4) to the

potential needed, bringing the total to ~1.7 V to 2.0 V to obtain reasonable current

72

densities. As a result, a single semiconductor used for solar water splitting, without any

additional energy input, must have a bandgap of at least ~1.7 to 2.0 eV with the band

edges straddling the water oxidation and proton reduction potentials plus associated

overpotentials. These larger bandgap requirements result in less absorption of the solar

spectrum, lowering the operating efficiency from about a 30% theoretical maximum to

that of about 10%. The solar-to-hydrogen efficiency (ε) for any photoelectrolysis cell

can be calculated using equation 3.2 below:

(3.2)

where Jm is the measured current density (A/cm2), Vapp is the applied potential (V)

measured between the photoanode and photocathode and Pin (W/cm2) is the input power

density of the solar illumination.11

These calculations should be conducted based off

results obtained when the solar water splitting cell is investigated as part of a two

electrode setup. Band diagrams for several possible water splitting schemes utilizing

solar irradiation as the only energy input are shown in Figure 3.4.

In order to make a useful solar water splitting device to be used on a global

scale, a material must be found, or designed, that simultaneously meets all the following

criteria: (1) efficiently absorbs a large fraction of the solar spectrum, (2) proper bandgap

energy alignment that allows for water oxidation and proton reduction, (3) adequately

allows charge transfer of photogenerated minority carriers to the surface/liquid

interface, (4) efficiently catalyzes water oxidation and proton reduction at the interface,

(5) remains stable during operating conditions at any pH and very oxidizing

environments for water oxidation, (6) made out of abundant and cheap materials, and

73

(7) is non-toxic. To date, there has not been a material discovered or designed that

simultaneously meets all of these criteria. There have only been a few photoanode

unassisted water splitting systems discovered which can use sunlight as the only energy

input to drive the water splitting reaction.12-15

Other promising water splitting designs include multi-junction semiconductors

in ohmic contact with each other. A factor of two hit in theoretical efficiency is

suffered for each junction due to minority charge carrier recombination at the interfaces.

Therefore, for example, a theoretical maximum efficiency of 15% would be obtained

for a two semiconductor, single p-n junction setup used for splitting water.10

The multi-

junction design reported by Turner was shown to catalyze water splitting with an

overall efficiency of 12.4 %.16

Unfortunately, this setup utilizes expensive materials to

operate, which are neither scalable nor robust for any extended periods of operation.

Current research efforts have broadly focused on finding cheap and abundant materials

that can support rapid charge transfer at the semiconductor/liquid interface, have

substantial durability under operating conditions and absorb a larger fraction of the solar

spectrum.

The idea of using sunlight to split water is not a new concept, as plants have

been accomplishing this feat since their existence. As has been shown above, there

remains many challenges to eventually making a commercially viable system that

works with cheap materials over an extended period of time. With the development or

discovery of a new semiconductor material that could be used toward solar water

splitting, an efficient water oxidation catalyst must be coupled either directly or through

a wire to accomplish the production of oxygen and hydrogen. The next topic of this

74

chapter will introduce some water oxidation catalysts that could be employed in a solar

water splitting device.

Photoanodes and Water Oxidation Catalysts

One of the best strategies to address the low efficiency of photoelectrochemical

water splitting cells is through the incorporation of more efficient water oxidation

catalysts. As mentioned above, the water oxidation reaction of equation 3.1 is often the

kinetic bottleneck for water splitting due to the large overpotentials commonly

associated with the reaction. Catalysts that can more efficiently remove four electrons

from water while forming the oxygen-oxygen double bond can be incorporated onto a

semiconductor material which should improve the overall device efficiency. There has

been much focus on the design and development of new water oxidation catalysts,

ranging from discrete transition metal-based homogeneous catalysts17-21

, metal and

metal oxide films22-24

and combinations of immobilized catalysts on

semiconductor/electrode surfaces25, 26

.

This introductory chapter will focus mainly on heterogeneous metal and metal

oxide water splitting catalysts in favor of discrete homogeneous catalysts. The most

active and best performing water oxidation catalysts to date are metal and metal oxides

of iridium and ruthenium. The problem with these metals, however, is their associated

high costs, which limit their applicability for large scale use. Table 3.2 below

summarizes the overpotential (η) required to obtain 1 mA/cm2 current density for some

of the common metal and metal oxide catalysts employed for water oxidation. This

table is not exhaustive for all water oxidation catalysts that are known and the reader

should consult some of the relevant reviews on this topic for a more extensive list.27, 28

75

Table 3.2: Various water oxidation catalysts and their performance at i = 1 mA/cm2.

Catalyst E, V vs. NHE η, V pH Ref

Ru

RuO2 1.43 0.2 0 29, 30

Ru-Ir 1.52 0.294 0 31a

Ru-Pt 1.7 0.474 0 31a

Ir

Ir 1.55 0.324 0 31a

IrO2 (planar) 1.42 0.561 6.3 23

IrO2 (particle) 1.39 0.473 5.3 32

IrO2 (particle) 1.12 ~0.3 7 22, 33

Ir-Pt 1.6 0.374 0 31a

Pt

Pt 1.94 0.714 0 31a

PtO2 1.042 0.638 0 34

Co

Co-Phosphate 1.24 0.41 7 24

Co3O4 0.844 0.44 14 35

Other Metals

NiOx 0.509 0.235 16.2 31a

MnO2 1.37 0.2 13 36

Semiconductors

Fe2O3 (dark) 1.83 0.6 14 37

Fe2O3 (light) 1.1 -0.13 14 37

WO3 (dark) 1.85 1.03 7 38

WO3 (light) 0.9 0.496 0 39 a Measured at 80°C

The overpotentials given in Table 3.4 were calculated by taking the difference from the

measured potential versus NHE and the theoretical reversible potential, E°, which can

be calculated by substituting the given pH into equation 3.3.

E° = 1.23 V – (0.059 V) pH vs. NHE (3.3)

If Table 3.4 is any indication, the amount of materials studied for efficient water

oxidation is quite extensive and thorough. If a new catalyst system cannot be found, the

only option left is to find methods to incorporate minimal amounts of the best

performing catalysts, iridium or ruthenium, onto a stable semiconductor employed in

solar water splitting. Simply putting a good water oxidation catalyst on a

76

semiconductor, however, does not allow for a device to be operational in solar water

splitting. The semiconductor must often be protected from the harsh environments of

the water splitting reaction, namely oxidizing potentials and extreme pH’s. The next

section of this chapter will discuss some methods for protecting semiconductors during

operation in solar water splitting.

Protection of Photoanodes Used in Solar Water Splitting

A critical component of any photoelectrochemical water splitting device that

utilizes a water oxidation catalyst attached to a photon absorbing semiconductor

substrate is operational stability. As has been discussed, the water oxidation reaction is

an oxidatively demanding reaction often occurring in harsh pH and oxidative

environments. The harsh conditions often employed for photoelectrochemical water

oxidation at a photoanode will typically oxidize the photon absorbing semiconductor

before water. In the case of silicon as the photoanode, the silicon will oxidize to silicon

dioxide which is an insulating material that doesn’t conduct. In other cases, the

semiconductor will slowly dissolve into the water over time. Since avoiding the water

oxidation reaction at the anode surface is impossible for a photoanode with small

enough bandgaps to harvest an optimal fraction of the solar spectrum, a material must

be used to protect the underlying base substrate. A good protecting layer must be

utilized that meets the following criteria: (1) adequately protects the semiconductor

from corrosion, or oxidation, (2) allows for efficient transfer of holes and electrons

between the semiconductor and catalyst, and (3) does not absorb much light. If the

protection layer absorbs a large fraction of light it will decrease the amount of photons

able to reach the semiconductor, which will decrease the photocurrent and efficiency

77

that could otherwise be obtained. Previous attempts at protecting semiconductors used

as photoanodes have included depositing thick layers of oxidized protective coatings on

the photoanode that prevent the semiconductor surface from being exposed to oxides

that are generated during water oxidation.40

These thick coatings often decrease device

efficiency, as they would not allow facile electron transport or would absorb solar

irradiation themselves and act as the semiconductor instead of the substrate they were

employed to protect. Some methods also include thin layers of protective coatings.41

These often suffered stability issues as the coatings were typically too thin to be

uniform, which would allow for oxide migration and thus corrosion of the underlying

semiconductors. Other methods have included doping ruthenium into titanium dioxide

electrodes. The ruthenium helps the titanium dioxide become more conducting and acts

as a catalyst, while the titanium dioxide provides stability for the ruthenium. These

devices are known as dimensionally stable anodes, and are commonly employed in the

chloro-alkaline industry for the production of chlorine gas and more recently water

oxidation to produce oxygen; unfortunately these types of coatings are expensive

because the noble metal must be present throughout the thickness of the thick titanium

dioxide films.42-45

In the next two chapters, a new method will be introduced that accomplishes the

protection of a small band gap base semiconductor during photoelectrochemical water

splitting while not decreasing efficiency. We utilized a thin layer of titanium dioxide

(TiO2) that acts as a corrosion barrier to an underlying semiconductor while still

allowing for facile charge transport between it and the catalyst layer on top of the

structure that was used in water oxidation. The method used to deposit thin conformal

78

layers of TiO2 was atomic layer deposition (ALD). This technique will cycle between a

titanium precursor and water to create a single layer of TiO2 on a surface. The cycles

can be repeated until the desired thickness is obtained, resulting in a uniform thickness

that is free of pinholes or cracks.

79

3.4: Figures

Figure 3.1: Current and Future Global Energy Demand. Pie charts depicting global

energy use in terawatts for each energy source in both (A) 2010 at 15 TW total energy

and (B) 2050 at 45 TW total energy. The 2050 pie chart is a projection based off of the

discussion from the text and already has the energy sources from 2010 added in. The

additional 30 TW of energy consumption remains to be filled in.

A. B.

80

Figure 3.2: Solar Spectrum. Solar irradiance spectrum as detected at the Earth’s

surface, referred to as AM1.5G solar irradiation.5

81

Figure 3.3: Band Gaps of Common Semiconductors. Commonly used

semiconductors employed in photoelectrochemical water splitting devices and their

band gaps. Horizontal lines represent the standard reduction potentials for proton

reduction and water oxidation. A photoelectrochemical cell utilizing one

semiconductor material that only uses solar irradiation to drive water splitting must

have a band gap that straddles these horizontal lines, plus any additional overpotentials

necessary to catalyze each reaction.6

82

Figure 3.4: Recombination Pathways for Photogenerated Electron/Hole Pairs. (A)

An electron of a n-type semiconductor absorbing a photon to create an electron/hole

pair within the depletion region. In an ideal case the hole will travel to the

semiconductor/liquid interface and the electron to the backside contact because of the

band bending that occurs due to the dipole created at the surface. (B) Illustration of

possible pathways for loss of photocurrent due to (1) electron/hole recombination

within the bulk of the semiconductor, (2) electron/hole recombination within the

depletion region of the semiconductor, (3) electron tunneling across the junction to

some acceptor in solution, (4) thermal relaxation of electron across barrier due to

excitation above the potential barrier (thermionic emission), and (5) electron/hole

recombination at defect sites within the semiconductor or at the interface.

83

Figure 3.5: Photoelectrochemical Water Splitting Configurations. Possible

configurations of anodes, cathodes and PV’s that will perform water splitting without

the input of external bias. The setup employed in (A) is a single photoanode material

that has a band gap of at least 1.23 eV and can split water. The PV-photoanode in (B)

describes a setup that requires a photovoltaic material in series with the photoanode

material to reach the required 1.23 eV bandgap. This would require 2 photons absorbed

for every electron used to reduce protons. The setup shown in (C) indicates a solar cell

which uses a photoanode and photocathode to reach 1.23 eV. This arrangement would

also require the collection of two photons for each electron used in proton reduction.

A. B.

C.

84

3.5: References

1. Lewis, N. S.; Nocera, D. G., Powering the planet: Chemical challenges in solar

energy utilization. Proc. Natl. Acad. Sci. 2006, 103, (43), 15729-15735.

2. Lewis, N. S., Powering the planet. MRS Bull. 2007, 32, (10), 808-820.

3. World Energy Assessment: Energy and the Challenge of Sustainability.

http://www.undp.org/energy/activities/wea/drafts-frame.html

4. Cook, T. R.; Dogutan, D. K.; Reece, S. Y.; Surendranath, Y.; Teets, T. S.;

Nocera, D. G., Solar Energy Supply and Storage for the Legacy and Non legacy

Worlds. Chem. Rev. 2010, 110, (11), 6474-6502.

5. ERDA/NASA -1022/77/16 1977.

6. Kudo, A.; Miseki, Y., Heterogeneous photocatalyst materials for water splitting.

Chem. Soc. Rev. 2009, 38, (1), 253-278.

7. Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Mi, Q. X.;

Santori, E. A.; Lewis, N. S., Solar Water Splitting Cells. Chem. Rev. 2010, 110,

(11), 6446-6473.

8. Shockley, W.; Queisser, H. J., Detailed Balance Limit of Efficiency of p-n

Junction Solar Cells. J. of Applied Phys. 1961, 32, (3), 510-519.

9. Fujishima, A.; Honda, K., Electrochemical Photolysis of Water at a

Semiconductor Electrode. Nature 1972, 238, (5358), 37-38.

10. Varghese, O. K.; Grimes, C. A., Appropriate strategies for determining the

photoconversion efficiency of water photo electrolysis cells: A review with

examples using titania nanotube array photoanodes. Sol. Energ. Mat. Sol. C

2008, 92, (4), 374-384.

85

11. Chen, Z.; Jaramillo, T. F.; Deutsch, T. G.; Kleiman-Shwarsctein, A.; Forman, A.

J.; Gaillard, N.; Garland, R.; Takanabe, K.; Heske, C.; Sunkara, M.; McFarland,

E. W.; Domen, K.; Miller, E. L.; Turner, J. A.; Dinh, H. N., Accelerating

materials development for photoelectrochemical hydrogen production:

Standards for methods, definitions, and reporting protocols. J. Mater. Res. 2010,

25, (1), 3-16.

12. Morisaki, H.; Watanabe, T.; Iwase, M.; Yazawa, K., Photoelectrolysis of Water

with TiO2 Covered Solar Cell Electrodes. Appl. Phys. Lett. 1976, 29, (6), 338-

340.

13. Wagner, F.; Somorjai, G., Photocatalytic and Photoelectrochemical Hydrogen

Production on Strontium Titanate Single Crystals. J. Am. Chem. Soc. 1980, 102,

(17), 5494-5502.

14. Sato, J.; Saito, N.; Yamada, Y.; Maeda, K.; Takata, T.; Kondo, J.; Hara, M.;

Kobayashi, H.; Domen, K.; Inoue, Y., RuO2-loaded beta-Ge3N4 as a non-oxide

photocatalyst for overall water splitting. J. Am. Chem. Soc. 2005, 127, (12),

4150-4151.

15. Waki, I.; Cohen, D.; Lal, R.; Mishra, U.; DenBaars, S. P.; Nakamura, S., Direct

Water Photoelectrolysis with Patterned n-GaN. Appl. Phys. Lett. 2007, 91, (9),

093519.

16. Khaselev, O.; Turner, J., A monolithic photovoltaic-photoelectrochemical

device for hydrogen production via water splitting. Science 1998, 280, (5362),

425-427.

86

17. Gersten, S. W.; Samuels, G. J.; Meyer, T. J., Catayltic Oxidation of Water by an

Oxo-Bridged Ruthenium Dimer. J. Am. Chem. Soc. 1982, 104, (14), 4029-4030.

18. Lebeau, E. L.; Adeyemi, S. A.; Meyer, T. J., Water Oxidation by

[(tpy)(H2O)2RuIIIORuIII(H2O)2(tpy)]4+. Inorg. Chem. 1998, 37, (25), 6476-

6484.

19. Sala, X.; Romero, I.; Rodriguez, M.; Escriche, L.; LLOBET, A., Molecular

Catalysts that Oxidize Water to Dioxygen. Angew Chem Int Edit 2009, 48, (16),

2842-2852.

20. Zong, R.; Thummel, R., A new family of Ru complexes for water oxidation. J.

Am. Chem. Soc. 2005, 127, (37), 12802-12803.

21. McDaniel, N. D.; Coughlin, F. J.; Tinker, L. L.; Bernhard, S., Cyclometalated

iridium(III) aquo complexes: Efficient and tunable catalysts for the

homogeneous oxidation of water. J. Am. Chem. Soc. 2008, 130, (1), 210-217.

22. Nakagawa, T.; Beasley, C. A.; Murray, R. W., Efficient Electro-Oxidation of

Water near Its Reversible Potential by a Mesoporous IrOx Nanoparticle Film. J.

Phys. Chem. C 2009, 113, (30), 12958-12961.

23. Yagi, M.; Tomita, E.; Kuwabara, T., Remarkably high activity of

electrodeposited IrO2 film for electrocatalytic water oxidation. J. Electroanal.

Chem. 2005, 579, (1), 83-88.

24. Kanan, M. W.; Nocera, D. G., In situ formation of an oxygen-evolving catalyst

in neutral water containing phosphate and Co2+

. Science 2008, 321, (5892),

1072-1075.

87

25. Francas, L.; Sala, X.; Benet-Buchholz, J.; Escriche, L.; LLOBET, A., A Ru-

Hbpp-Based Water-Oxidation Catalyst Anchored on Rutile TiO2. Chemsuschem

2009, 2, (4), 321-329.

26. Chen, Z.; Concepcion, J. J.; Jurss, J. W.; Meyer, T. J., Single-Site, Catalytic

Water Oxidation on Oxide Surfaces. J. Am. Chem. Soc. 2009, 131, (43), 15580-

15582.

27. Matsumoto, Y.; Sato, E., Electrocatalytic properties of transition metal oxides

for oxygen evolution reaction. Materials Chemistry and Physics 1986, 14, (5),

397-426.

28. Brimblecombe, R.; Dismukes, G. C.; Swiegers, G. F.; Spiccia, L., Molecular

water-oxidation catalysts for photoelectrochemical cells. Dalton Trans. 2009,

(43), 9374-9384.

29. Lodi, G.; Sivieri, E.; Debattisti, A.; Trasatti, S., Ruthenium Dioxide Based Film

Electrodes. 3. Effect of Chemical Composition and Surface Morphology on

Oxygen Evolution in Acid Solutions. J. Appl. Electrochem. 1978, 8, (2), 135-

143.

30. Neumannspallart, M.; Kalyanasundaram, K.; Gratzel, C.; Gratzel, M.,

Ruthenium Dioxide Electrodes as Suitable Anodes for Water Photolysis. Helv.

Chim. Acta. 1980, 63, (5), 1111-1118.

31. Miles, M. H.; Klaus, E. A.; Gunn, B. P.; Locker, J. R.; Serafin, W. E., Oxygen

Evolution Reaction on Platinum, Iridium, Ruthenium and their Alloys at 80

Degrees C in Acid Solutions. Electrochimica Acta 1978, 23, (6), 521-526.

88

32. Yagi, M.; Tomita, E.; Sakita, S.; Kuwabara, T.; Nagai, K., Self-assembly of

active IrO2 colloid catalyst on an ITO electrode for efficient electrochemical

water oxidation. J. Phys. Chem. B 2005, 109, (46), 21489-21491.

33. Nakagawa, T.; Bjorge, N. S.; Murray, R. W., Electrogenerated IrOx

Nanoparticles as Dissolved Redox Catalysts for Water Oxidation. J. Am. Chem.

Soc. 2009, 131, (43), 15578-15580.

34. Iwakura, C.; Fukuda, K.; Tamura, H., Anodic Evolution of Oxygen on Platinum

Oxide Electrode in Alkaline Solutions. Electrochimica Acta 1976, 21, (7), 501-

508.

35. Iwakura, C.; Honji, A.; Tamura, H., The Anodic Evolution of Oxygen on

Co3O4 Film Electrodes in Alkaline Solutions. Electrochimica Acta 1981, 26,

(9), 1319-1326.

36. Gorlin, Y.; Jaramillo, T. F., A Bifunctional Nonprecious Metal Catalyst for

Oxygen Reduction and Water Oxidation. J. Am. Chem. Soc. 2010, 132, (39),

13612-13614.

37. Kay, A.; Cesar, I.; Graetzel, M., New benchmark for water photooxidation by

nanostructured alpha-Fe2O3 films. J. Am. Chem. Soc. 2006, 128, (49), 15714-

15721.

38. Seabold, J. A.; Choi, K.-S., Effect of a Cobalt-Based Oxygen Evolution Catalyst

on the Stability and the Selectivity of Photo-Oxidation Reactions of a WO3

Photoanode. Chem. Mater. 2011, 23, (5), 1105-1112.

89

39. Santato, C.; Ulmann, M.; Augustynski, J., Photoelectrochemical properties of

nanostructured tungsten trioxide films. J. Phys. Chem. B 2001, 105, (5), 936-

940.

40. Kohl, P. A.; Frank, S. N.; Bard, A. J., Semiconductor Electrodes. 11. Behavior

of n-Type and p-Type Single Crystal Semiconductors Covered with Thin

Normal TiO2 Films. J. Electrochem. Soc. 1977, 124, (2), 225-229.

41. Decker, F.; Fracastorodecker, M.; Badawy, W.; Doblhofer, K.; Gerischer, H.,

The Photocurrent Voltage Characteristics of the Heterojunction Combination n-

Si/SnO2/Redox Electrolyte. J. Electrochem. Soc. 1983, 130, (11), 2173-2179.

42. Chen, X.; Chen, G., Stable Ti/RuO2-Sb2O5-SnO2 electrodes for O-2 evolution.

Electrochimica Acta 2005, 50, (20), 4155-4159.

43. Cipris, D.; Pouli, D., Oxygen Evolution on Dimensionally Stable Anode

Materials. J. Electroanal. Chem. 1976, 73, (1), 125-128.

44. Cornell, A.; Hakansson, B.; Lindbergh, G., Ruthenium-based dimensionally

stable anode in chlorate electrolysis - Effects of electrolyte composition on the

anode potential. J. Electrochem. Soc. 2003, 150, (1), D6-D12.

45. Ouattara, L.; Diaco, T.; Duo, I.; Panizza, M.; Foti, G.; Comninellis, C.,

Dimensionally stable anode-type anode based on conductive p-silicon substrate.

J. Electrochem. Soc. 2003, 150, (2), D41-D45.

90

Chapter 4: Stable Si Photoanodes for Water Splitting

4.1: Preface

This work was done in collaboration with with Professor Paul C. McIntyre and

his student, Yi Wei (Vincent) Chen, and was recently submitted and accepted for

publication to the journal Nature Materials and reprinted here with permission.1

Vincent and I helped design and perform the experiments and analyzed the results and

prepared the manuscript. I would like to thank Jung-Yong Lee, Rostam Dinyari, and

Prof. Peter Peumans for providing access to the solar simulation setup. I also thank

Prof. Thomas Jaramillo, Dr. Shin-Jung Choi, Kendra Kuhl, and Blaise Pinaud for

impedence spectroscopy and reversible hydrogen electrode measurements, as well as

their many helpful discussions. I also thank Dr. Byungha Shin and Dr. Michael

Shandalov for their help in TiO2 deposition and characterization. This work was

partially supported by the Stanford Global Climate and Energy Project. I would also

like to acknowledge the Center for Integrated Systems and a Precourt Institute for

Energy seed grant for funding for this work.

1. Chen, Y. W.; Prange, J. D.; Duehnen, S.; Park, Y.; Gunji, M.; Chidsey, C. E. D.;

McIntyre, P. C., Atomic layer-deposited tunnel oxide stabilizes silicon photoanodes

for water oxidation. Nature Mat. 2011, 10, 539-544.

91

4.2: Abstract

A leading approach for large-scale electrochemical energy production with

minimal global-warming gas emission is to use a renewable source of electricity, such

as solar energy, to oxidize water, providing an abundant source of electrons needed in

fuel synthesis. We report corrosion-resistant, nanocomposite anodes for the oxidation of

water required to produce renewable fuels. Silicon (Si), an earth-abundant element and

an efficient photovoltaic material, is protected by atomic layer deposition (ALD) of a

highly uniform, 2 nm thick layer of titanium dioxide (TiO2) and then coated with an

optically transmitting layer of a known catalyst (3 nm iridium). Photoelectrochemical

water oxidation was observed to occur below the reversible potential while dark

electrochemical water oxidation was found to have low-to-moderate overpotentials at

all pH values, resulting in an inferred photovoltage of ~550 mV. Water oxidation is

sustained at these anodes for many hours in harsh pH and oxidative environments

whereas comparable silicon anodes without the TiO2 coating quickly fail. The desirable

electrochemical efficiency and corrosion resistance of these anodes is made possible by

the low electron-tunneling resistance (< 0.006 Ω cm2 for p

+-Si) and uniform thickness

of atomic-layer deposited TiO2.

92

4.3: Introduction

Rising global energy demand and a growing concern for the climate promote

interest in new technologies to harness energy from renewable sources while decreasing

dependence on fossil fuels.1 One interesting approach is to produce hydrogen or other

reduced molecular fuels from intermittent energy sources such as solar and wind.1,2

Electrons needed for renewable fuel production at large scale are likely to come from

the oxidation of water:

298Kat NHE vs.pH V 059.0V 23.1 -e 4 HA4 O-A 4 O H22

O22

E

where the base, A-, and reversible electrochemical potential,

2OE , are determined by the

composition and pH of the system. Water oxidation is a demanding electrochemical

reaction requiring oxidatively robust, and yet inexpensive, anodes. One strategy to

reduce costs is to combine the solar harvesting properties of a photovoltaic with water

oxidation in the form of a photoanode. Here, we report on a novel method of

passivating the surface of an otherwise unstable photoanode material (Si) for operation

in an aqueous electrolyte solution. The protected photoanode both absorbs solar

photons and provides holes in its valence band to oxidize water to molecular oxygen, as

illustrated in Figure 1A.

Due to its stability over a range of pH and potentials, titanium dioxide (TiO2) is

a useful photoanode material.3 Although the large bandgap of TiO2 (~3 eV) permits

photooxidation of water in the absence of an applied bias, it limits absorption to the

small fraction of the solar spectrum in the ultraviolet, resulting in low efficiency as a

solar photoanode.4 Photoanodes such as iron oxide (Fe2O3, bandgap 2.3 eV) and

tungsten oxide (WO3, bandgap 2.7 eV) semiconductors have also been shown to

93

perform water oxidation, and are both considered to be promising in dual bandgap water

splitting cells.5 However, these materials are limited to lower saturation current

densities at higher applied biases due to their modest electronic conductivity and

moderate bandgaps.6,7

Many small bandgap semiconductor photoanodes, such as Si

(bandgap 1.1 eV), are capable of absorbing a large portion of the solar spectrum, but are

not stable at the highly oxidative potentials required for water oxidation.8

An inexpensive and easily synthesized corrosion-resistant layer that protects the

underlying semiconductor while not inhibiting charge carrier (electron or hole) transport

and photon absorption would be highly desirable. Previous attempts at passivating Si

photoanodes with protective noble metal and noble metal silicide layers9,10

, or by

doping the SiO2 layer on Si for better conductivity11

, were unsuccessful in obtaining a

high performance, long-lasting water oxidation photoanode. Attempts at protecting Si

photoanodes with TiO2 deposited by chemical vapor deposition (CVD) were also

unsuccessful.12

Titanium oxide layers thick enough to avoid the dissolution of the

underlying substrate through pinholes or cracks were found to inhibit electron transfer

from the electrolyte to the substrate, resulting in poor water oxidation performance.12

Passivation of compound semiconductors, such as CdSe and CuO, with TiO2 or other

materials was also attempted but was also unable to simultaneously achieve long

endurance and good conversion efficiency for water oxidation.13-15

Using methods from state-of-the-art semiconductor electronics technology, we

have combined the best properties of TiO2 and Si to synthesize an oxidatively robust

and efficient photoanode for water oxidation. In this work, as-received Si substrates

with an initial SiO2 surface layer were coated with a pinhole-free layer of TiO2 by

94

atomic layer deposition (ALD).16

A TiO2 thickness of 2 nm was found to prevent

oxidation of the Si while being thin enough to allow facile electron tunneling between

an overlying catalyst layer and the base substrate. Iridium, one of many well-known

metal and metal oxide catalysts that promote efficient water oxidation over a range of

pH17-21

, was deposited on top of the TiO2 layer by physical vapor deposition (PVD)

methods. An Ir film 3 nm thick was found to be a sufficient amount of catalyst while

not blocking the transmission of photons into the Si substrate. A schematic

representation of the nanocomposite structure of the optimized anode is shown in Figure

4.1A. Figures 4.1B and C show a cross-sectional transmission electron micrograph

(TEM) and a schematic band diagram of the photoanode structure, respectively.

95

4.4: Results and Discussion

Lightly phosphorus-doped n-type Si wafers (n-Si, 0.1-0.2 cm, 500 μm

thickness) were used as the base substrate, onto which 24 cycles of ALD-TiO2 (2 nm)

and 3 nm of PVD-Ir were sequentially deposited (Ir/TiO2/n-Si). Water oxidation

currents at these nanocomposite anodes were measured as a function of potential with

and without simulated solar irradiation at 1 sun (AM1.5G) in acidic (1 M H2SO4),

neutral (1 M phosphate-buffered, pH 7) and basic (1 M NaOH) solutions (Figure

4.1D). Electron transfer from Ir to Si through the oxide layers fills photogenerated

holes in the Si valence band. Electrons are re-supplied to the Ir layer via the water

oxidation reaction. With n-type Si substrates in the dark, which have a very low thermal

population of holes, modest overpotentials for water oxidation produced minimal

currents (of order A/cm2). With illumination, which produces photo-generated holes in

the Si substrate, current densities at all pH values increased dramatically. The potential

required to obtain 1 mA/cm2 was observed to be ~200 mV lower than the

thermodynamic water oxidation potential as a result of the additional voltage made

available by the photo-generated holes (Table 4.1). This onset potential is similar to

that of the best reported values for Fe2O3 and WO3 photoanodes.6,7

However, the

smaller bandgap of Si and its relatively large carrier mobilities allow the nanocomposite

anode to achieve much higher current densities at biases exceeding the onset potential.

Tafel slopes for the nanocomposite anode were found to be 60 mV/decade at 0.1

mA/cm2 for each of the solutions.

Solar irradiation at 1 sun provides 2.7 x 1017

photons/cm2 with energies greater

than the Si bandgap, corresponding to a theoretical maximum current density of 43

96

mA/cm2.22

The measured current densities of the illuminated samples in Figure 4.1D

exceed 10 mA/cm2 at moderate overpotentials. At greater applied potentials, the current

density saturates near 30 mA/cm2, as expected for an efficient Si photoanode (see

supplementary materials, section 4.9). The Ir/TiO2/n-Si anode produces significantly

greater current density (mA/cm2 instead of A/cm

2) compared to reported bulk TiO2-

based photoanodes.23

Such high current density is made possible by the intrinsic

decoupling of the electrochemical reaction site (the Ir catalyst) from the photovoltaic

device (the Si substrate) in our nanocomposite anode. The decoupling provided by the

ALD-TiO2 layer allows for further engineering of anode performance by improving

either of the two components.24

It is worth noting that these high current densities were

measured in the absence of surface texturing, which would increase the number of

reactive sites per nominal substrate area of the nanocomposite anode4 and would be

expected to result in even higher current densities. Moreover, the strategy of using a

thin ALD layer to protect the semiconductor is not expected to limit the current density

at 1 sun of illumination. The maximum current density of 43 mA/cm2

is several orders

of magnitude less than the current densities resulting from leakage of charge across the

ultra-thin ALD-grown metal oxide gate dielectrics in state-of-the-art field effect

transistors25

, consistent with the low tunneling resistance and stable photoanode

operation observed in the present experiments.

In order to study the water oxidation reaction in the dark, heavily boron-doped

p-Si substrates (0.001-0.005 cm) were used. Water oxidation currents were measured

as a function of potential for these Ir/TiO2/p-Si anodes in acidic, neutral, and basic

solutions (Figure 4.1E). The anodes were found to be active in all three solutions,

97

requiring overpotentials of ~350 mV to achieve a current density of 1 mA/cm2 in each

solution (Table 4.1). These results are consistent with overpotentials previously

reported for similar current densities19,21,26

using dimensionally stable anodes with thick

(> 1 m) noble metal oxide catalyst coatings. Comparing the overpotentials measured

for n-Si anodes in the light with those for p-Si anodes in the dark, the photovoltage was

calculated to be in the range of 510-570 mV (dark ~+350 mV, light ~–200 mV). This

photovoltage is similar to that of the best Si photoelectrochemical solar cells27

, and

close to the open circuit photovoltage reported for high quality pn junction Si solar cells

(~700 mV).28

The somewhat lower photovoltage observed here compared to Si solar

cells may result from a less than optimal choice of the work function of the catalyst

metal on the TiO229

, and non-idealities such as non-radiative carrier recombination at

defects.30

Because electrodes used for water oxidation are exposed to highly corrosive and

oxidative environments, the endurance of the nanocomposite anodes was investigated in

electrochemical life tests. Figure 4.2 shows the measured potential required for the

illuminated anodes to give a constant current of 1 mA through the 0.196 cm2 sample in

both 1 M acid and 1 M base solutions. The samples without the TiO2 layer failed under

illumination in both solutions within half an hour, reaching the maximum voltage the

potentiostat could supply, while the samples with the TiO2 layer lasted for at least 8

hours, the duration of these endurance tests. Figure 4.3 shows the current obtained on

Ir/TiO2/p-Si in the dark as a function of time while being held at a constant potential of

1.7 V vs. NHE with and without the TiO2 layer for 24 hours in the 1M NaOH solution.

The sample with the corrosion resistant TiO2 layer remained operational for at least 24

98

hours, while the sample without the TiO2 layer failed within 0.5 hours. It is worth

noting that the nominal current densities in these lifetime measurements are achieved at

higher measured or applied biases than in the cyclic voltammograms in Figure 4.1

because of inefficient oxygen removal from the sample surface in the lifetime test cell

(See supplementary materials, section 4.9).

Analysis of samples that gave results in Figure 4.2 by x-ray photoelectron

spectroscopy (XPS) depth profiling after constant current life tests (1 mA on a 0.196

cm2 area) revealed that the discrete layering of the nanocomposite remains intact for the

anode containing TiO2 (Figure 4.4A). Additionally, cross-sectional TEM of similarly

tested samples reveals that the SiO2 thickness was comparable to that of the initial

interfacial SiO2 prior to testing (see supplementary materials, section 4.9). Samples

without the TiO2 layer grew a thick, insulating SiO2 layer after the constant current

experiment (Figure 4.4B). These results indicate that a corrosion resistant TiO2 layer of

only 2-3 nm thickness protects the underlying Si substrate under the conditions

investigated, while still allowing for efficient transport of electrons and holes between

the solution and the Si anode.

The efficiency of electronic transport across this interface was characterized in

greater detail using a benchmark electrolyte solution composed of aqueous

ferri/ferrocyanide ions. The exchange of electrons between this solution and metal

electrodes is fast, allowing for characterization of the electronic transport across the

interposed TiO2 layer by cyclic voltammetry (CV).31

Facile electron transport was

observed for samples coated with the ultrathin Ir layer (Figure 4.5A). Samples without

the Ir layer exhibit orders of magnitude lower current density and no observable

99

Fe(II)/Fe(III) redox waves, indicating the importance of the ultrathin metal layer as a

charge carrier mediator between the substrate and the solution. In addition to the Ir

overlayer, other metals such as Pt and Ru were deposited and found to serve the same

function. The p-Si anodes with ALD-TiO2 exhibit relatively small peak-to-peak

splitting (130 mV) and high current densities, comparable to results obtained with bulk

metal electrodes. Samples with a substantially thicker TiO2 layer (10 nm instead of 2

nm) resulted in increased peak-to-peak splitting (610 mV), indicating the importance of

using a thin TiO2 layer for efficient tunneling-mediated transport of electrons.

The Ir/TiO2/n-Si anodes were analyzed in both the dark and solar simulated light

for electronic transfer efficiency using the ferri/ferrocyanide solution (Figure 4.5A).

The dark CV reveals an asymmetry, with the anodic peak missing and the cathodic peak

remaining. This is consistent with the presence of a sufficient concentration of

electrons in the n-Si to rapidly reduce Fe(III) and the lack of holes required to oxidize

Fe(II). This is also the cause of the low water oxidation current density observed for the

Ir/TiO2/n-Si anode without illumination (Figure 4.1D). As a result, the electrolyte-solid

interface behaves as a Schottky junction, giving rise to the observed CV asymmetry. In

contrast, the CV of the illuminated sample not only recovers its symmetry, but also

displays a negative potential shift and an increase in peak current density. The

recovered anodic peak is due to the photo-generated holes. Comparing the Fe(II)/Fe(III)

redox potentials for the dark p+-Si and light n-Si samples, a negative shift of ~550 mV

is observed, which is consistent with the shift observed for water oxidation

overpotentials at 1 mA/cm2 between the two samples.

100

The intrinsic electron transport properties of the nanocomposite anodes under

conditions similar to those used in dark electrolysis were probed by temperature

dependent, metal contact current-voltage (I-V) measurements. Electron tunneling

through the TiO2 layer on the p+-Si substrate was confirmed by varying the TiO2

thickness. The current density for thin (< 2 nm) TiO2 samples was observed to have

very small temperature dependence, a strong indication of hole transport by tunneling

(Figure 4.5B).32

From the I-V measurement, the resistance of the 2 nm TiO2 was

estimated to be less than 0.006 Ω cm2 on p-Si substrates at room temperature. Thicker

(> 4 nm) TiO2 samples result in an increasingly temperature-dependent current density,

an indication of a more thermally-activated and bulk-limited conduction mechanism

such as trap-assisted tunneling or Frenkel-Poole conduction (Figure 4.5B).33

The

significant thickness dependence of the measured electronic conduction across the TiO2

layer indicates the importance of using a deposition method such as ALD, which

exhibits excellent uniformity and control of thickness at the nanoscale.

The ALD-TiO2 film thickness we have investigated is much smaller than that

described in a recent publication in which conformal deposition of relatively thick (~ 35

nm) TiO2 was performed over n- or p-type Si nanowires.4 These coated nanowire

arrays exhibited improved behavior compared to correspondingly coated planar Si

photoelectrodes; however, the current densities attributed to water oxidation were much

smaller than those that we have observed. This is consistent with greater resistance to

carrier transport associated with the larger TiO2 thickness, and with the absence of the Ir

layer that catalyzes water oxidation and promotes electronic conduction across our

nanocomposite photoanodes.

101

4.5: Conclusion

In this report, we have demonstrated and characterized the operation of an

efficient and dimensionally stable semiconductor anode for photoelectrochemical water

oxidation. This nanocomposite anode uses a pinhole-free, corrosion resistant, ALD-

grown TiO2 tunnel oxide layer that protects an underlying Si substrate during water

oxidation at an overlying catalyst layer in both dark and light conditions. The ultrathin

ALD-TiO2 layer is thick enough to permit hours of continuous operation in corrosive

environments (acidic or basic) without apparent structural change, while being thin

enough to allow facile electronic transport via tunneling. Compared to previously

reported metal oxide photoanodes, this nanocomposite device is capable of reaching

much higher saturation current densities (tens of mA/cm2), and of maintaining low

overpotentials at moderate current densities. The measured photovoltage range of ~510-

570 mV approaches the open-circuit photovoltage for state-of-the-art silicon solar cells.

The reported nanocomposite structure allows for the decoupling of the electrochemical

reaction at the catalyst surface from the underlying photovoltaic substrate, which should

permit future improvements by further optimizing the different components. Therefore,

this approach is quite general, and should have applications in protecting semiconductor

substrates other than silicon, and in integration of other conductive catalyst layers

besides iridium. Additionally, ALD is currently used for industrial processes in

semiconductor device fabrication34,35

, which should allow for this technique to be

employed on a large scale in this application.

102

4.6: Methodology

The Si wafers used in the reported experiments were degenerately doped p+-type

Si (100) wafers (0.001-0.002 Ωcm, 500 m thickness) and n-type Si (100) wafers (0.1-

0.2 Ωcm, 500 m thickness). The wafers were used as received, with a thin (< 2 nm)

SiO2 layer as prepared by the wafer vendor. TiO2 was deposited by ALD at 200°C with

tetrakisdimethylamido titanium (TDMAT) as the titanium source and H2O as the

oxygen source. Metal deposition was performed by e-beam evaporation. The backside

contacts for the n-Si and p+-Si substrates were e-beam evaporated Al and Pt

respectively. The samples were heated to 400°C for 30 minutes in a forming gas

environment (95% N2, 5% H2). 1 sun illumination was provided by a Sciencetech

AM1.5G solar simulator, and the intensity was adjusted to 1 sun with a calibrated

photodiode.

The neutral solution was made with 0.4 M Na2HPO4, and 0.6 M NaH2PO4 with

NaOH used to adjust the pH to 7. The acidic and basic solutions were made with 1 M

H2SO4 and 1 M NaOH respectively. A hydrogen electrode was used to measure

reversible hydrogen potential of the three solutions to calculate the reported

overpotentials. The aqueous ferri/ferrocyanide solution was made to be 10 mM

K3Fe(CN)6, 10 mM K4Fe(CN)6, and 1 M KCl. The CVs were measured at 100 mV/s. A

bored (5 mm diameter, 0.196 cm2 area) Teflon cone was pressed against the sample and

used to contain the electrolyte solution. A Pt wire was used as a counter electrode, and a

glass-frit isolated Ag(s)/AgCl(s)/sat. KCl(aq) was used as a reference electrode.

Potentials measured versus Ag/AgCl/KCl were converted to NHE. Both electrodes were

suspended over the sample in the electrolyte solution. All measurements were

103

conducted on a WaveNow potentiostat in air at room temperature. Stability tests were

performed in the electrochemical cell with a 1 mL/s flow of the appropriate electrolyte

solution, provided by a Cole-Parmer peristaltic pump.

See more details in the supporting materials section of this chapter (section 4.9).

104

4.7: Figures

105

Figure 4.1: Anode Design and Water Oxidation Results. (A) Schematic and (B) TEM

image of the nanocomposite anode. (C) Approximate energy band diagram of

nanocomposite anode at 1 V vs. NHE under illumination in pH 0 solutions. (D) Water

electrolysis using n-Si substrates in the dark for acidic (∙∙∙), neutral (∙∙∙), and basic (∙∙∙)

solutions and 1 sun solar simulated light for acidic (–), neutral (–), and basic (–)

solutions. (E) Water electrolysis using p+-Si substrates in acidic (–), neutral (–), and basic

(–) solutions. Vertical lines in (D) and (E) represent the thermodynamic redox potential for

water oxidation at the appropriate pH. Scan rates were 0.1 V/s and potentials were

corrected for solution resistance as measured by impedance spectroscopy (see

supplementary materials, section 4.9).

106

Figure 4.2: Stability Tests. Constant current stability tests performed on n-Si samples

at 1 mA on a 0.196 cm2 sample area with 1 sun solar illumination in (A) 1 M Acid with

(–) and without (∙∙∙) the TiO2 protection layer and (B) 1 M Base with (–) and without

(∙∙∙) the TiO2 protection layer. Potentials were measured versus Ag/AgCl/KCl and

converted to NHE after correction for solution resistance as measured by impedance

spectroscopy.

107

Figure 4.3: Constant Potential Stability Test. Constant potential lifetime tests at 1.7

V vs. NHE on a 0.196 cm2 sample in 1M NaOH solution circulated at 120 mL/min on

(–) Ir/TiO2/p+-Si and (--) Ir/p

+-Si anodes.

108

Figure 4.4: XPS Depth Profiling Analysis. Analysis of n-Si samples after stability

test under 1 sun illumination that (A) have the TiO2 protection layer and (B) samples

without the TiO2 protection layer. The elements are represented by () for O, () for Si,

() for Ir, and (♦) for Ti.

109

Figure 4.5: Anode Electrochemical Performance. (A) Cyclic voltammogram of the

ferri/ferrocyanide solution for (-) 2 nm TiO2/p+-Si, (-) Ir/2 nm TiO2/p

+-Si, (-) Ir/10 nm

TiO2/p+-Si, (-) Ir/2 nm TiO2/n-Si, and (-) Ir/2 nm TiO2/n-Si with simulated 1 sun

110

illumination. All measurements were done in the dark except otherwise mentioned. (B)

Temperature dependent current density measurement through () 2 nm, () 4 nm, and

() 10 nm TiO2 films on p+-Si substrates. The line (-) indicates the compliance of the

meter. 50 nm thick Ir dots of 100 m diameter were used as the top metal contact.

Current measured in a probe station (no electrolyte present) at a Si substrate voltage vs.

the Ir contact of 0.5 V (flatband voltage=-0.13 V). Current-voltage data are provided in

the supplementary materials (section 4.9).

111

Table 4.1: Water oxidation overpotentials measured at 1 mA/cm2.

Substrate 1M NaOHa (V) pH 7

a (V) 1M H2SO4

a (V)

Ir/TiO2/p+-Si (Dark)

0.384 0.346 0.332

Ir/TiO2/n-Si (Light)

-0.171 -0.219 -0.200

a Calculated E

0 values of 1M NaOH, pH 7 and 1M H2SO4 solutions are +0.417 V,

+0.816 V, +1.206 V vs. NHE respectively.

112

4.8: References

1. Lewis, N. S. & Nocera, D. G. Powering the planet: Chemical challenges in solar

energy utilization. Proc. Natl. Acad. Sci. 103, 15729-15735, (2006).

2. Lewis, N. S. Toward cost-effective solar energy use. Science 315, 798-801,

(2007).

3. Pourbaix, M. J. N. Atlas of electrochemical equilibria in aqueous solutions. 1st

Edi., (Pergamon Press, 1966).

4. Hwang, Y. J., Boukai, A. & Yang, P. D. High Density n-Si/n-TiO2 Core/Shell

Nanowire Arrays with Enhanced Photoactivity. Nano Lett. 9, 410-415, (2009).

5. Walter, M. G. et al. Solar Water Splitting Cells. Chem. Rev. 110, 6446-6473,

(2010).

6. Kay, A., Cesar, I. & Gratzel, M. New benchmark for water photooxidation by

nanostructured alpha-Fe2O3 films. J. Am. Chem. Soc. 128, 15714-15721, (2006).

7. Sivula, K., Le Formal, F. & Gratzel, M. WO3-Fe2O3 Photoanodes for Water

Splitting: A Host Scaffold, Guest Absorber Approach. Chem. Mater. 21, 2862-

2867, (2009).

8. Matsumura, M. & Morrison, S. R. Anodic Properties of N-Si and N-Ge

Electrodes in HF Solution under Illumination and in the Dark. J. Electroanal.

Chem. 147, 157-166, (1983).

113

9. Fan, F. R. F., Keil, R. G. & Bard, A. J. Semiconductor Electrodes .48. Photo-

Oxidation of Halides and Water on N-Silicon Protected with Silicide Layers. J.

Am. Chem. Soc. 105, 220-224, (1983).

10. Howe, A. T., Hawkins, R. T. & Fleisch, T. H. Photoelectrochemical Cells of the

Electrolyte-Metal-Insulator-Semiconductor (Emis) Configuration .1. Metal

Thickness and Coverage Effects in the Pt-Silicon Oxide-N-Si System. J.

Electrochem. Soc. 133, 1369-1375, (1986).

11. Contractor, A. Q. & Bockris, J. O. M. Investigation of a Protective Conducting

Silica Film on N-Silicon. Electrochim. Acta 29, 1427-1434, (1984).

12. Kohl, P. A., Frank, S. N. & Bard, A. J. Semiconductor Electrodes .11. Behavior

of N-Type and P-Type Single-Crystal Semiconductors Covered with Thin

Normal-TiO2 Films. J. Electrochem. Soc. 124, 225-229, (1977).

13. Suleymanov, A. S. On the possibility of the transformation of solar energy to

chemical energy in the electrochemical cell with photoanode CdSe/TiO2. 16,

741-743, (1991).

14. Siripala, W., Ivanovskaya, A., Jaramillo, T. F., Baeck, S. H. & McFarland, E.

W. A CU2O/TiO2 heterojunction thin film cathode for photoelectrocatalysis.

Sol. Energ. Mat. Sol. C 77, 229-237, (2003).

15. Rajeshwar, K., Kaneko, M., Yamada, A. & Noufi, R. N. Photoelectrochemical

Oxidation of Halide-Ions at Naked, Catalytically Modified, and Polymer-Coated

N-Cds Electrodes in Aqueous-Media. J. Phys. Chem. 89, 806-811, (1985).

114

16. Ritala, M. & Leskela, M. Atomic layer epitaxy - a valuable tool for

nanotechnology? Nanotechnology 10, 19-24, (1999).

17. Liu, F. et al. Mechanisms of water oxidation from the blue dimer to photosystem

II. Inorg. Chem. 47, 1727-1752, (2008).

18. Kanan, M. W. & Nocera, D. G. In situ formation of an oxygen-evolving catalyst

in neutral water containing phosphate and Co2+

. Science 321, 1072-1075,

(2008).

19. Yagi, M., Tomita, E., Sakita, S., Kuwabara, T. & Nagai, K. Self-assembly of

active IrO2 colloid catalyst on an ITO electrode for efficient electrochemical

water oxidation. J. Phys. Chem. B 109, 21489-21491, (2005).

20. Comninellis, C. & Vercesi, G. P. Characterization of DSA-Type Oxygen

Evolving Electrodes - Choice of a Coating. J. Appl. Electrochem. 21, 335-345,

(1991).

21. Nakagawa, T., Bjorge, N. S. & Murray, R. W. Electrogenerated IrOx

Nanoparticles as Dissolved Redox Catalysts for Water Oxidation. J. Am. Chem.

Soc. 131, 15578-15579, (2009).

22. ERDA/NASA -1022/77/16, (1977).

23. Fujishima, A. & Honda, K. Electrochemical Photolysis of Water at a

Semiconductor Electrode. Nature 238, 37-38, (1972).

115

24. Howe, A. T. & Fleisch, T. H. Photoelectrochemical Cells of the Electrolyte-

Metal-Insulator-Semiconductor (Emis) Configuration .2. Use of Nonnative

Oxides in Pt/Oxide/N-Si Systems. J. Electrochem. Soc. 134, 72-76, (1987).

25. Process Integration, Devices, and Structures. International Technology

Roadmap for Semiconductors,

<http://www.itrs.net/links/2008ITRS/Update/2008Tables_FOCUS_A.xls>

(2008).

26. Ouattara, L. et al. Dimensionally stable anode-type anode based on conductive

p-silicon substrate. J. Electrochem. Soc. 150, D41-D45, (2003).

27. Switzer, J. A. The N-Silicon Thallium(III) Oxide Heterojunction

Photoelectrochemical Solar-Cell. J. Electrochem. Soc. 133, 722-728, (1986).

28. Green, M. A., Emery, K., Hishikawa, Y. & Warta, W. Solar cell efficiency

tables (version 35). Prog. Photovoltaics 18, 144-150, (2010).

29. Nozik, A. J. P-N Photoelectrolysis Cells. Appl. Phys. Lett. 29, 150-153, (1976).

30. Nelson, J. The physics of solar cells. (Imperial College Press, 2003).

31. Sharpe, A. G. The chemistry of cyano complexes of the transition metals.

(Academic Press, 1976).

32. Schuegraf, K. F. & Hu, C. M. Hole Injection SiO2 Breakdown Model for Very-

Low Voltage Lifetime Extrapolation. IEEE Trans. Electron. Dev. 41, 761-767,

(1994).

116

33. Nelson, J. & Chandler, R. E. Random walk models of charge transfer and

transport in dye sensitized systems. Coord. Chem. Rev. 248, 1181-1194, (2004).

34. Doering, R. & Nishi, Y. in Handbook of Semiconductor Manurfacturing

Technology 14-11 – 14-37 (CRC Press, Boca Raton, 2008).

35. Raaijmakers, I., Soininen, P. T., Granneman, E. H. A. & Haukka, S. P.

Protective Layers prior to Alternating Layer Deposition. USA patent US

2001/0054769A1 (2001).

117

4.9: Supporting Materials

Materials. All chemicals used in this study were purchased from commercial

sources and used as received. Degenerately doped p+-type Si (100) wafers (0.001-0.002

Ωcm, 500 m thickness) and n-type Si (100) wafers (0.1-0.2 Ωcm, 500 m thickness)

were purchased from El-Cat. All electrochemical measurements were performed in

electrolyte solutions prepared with DI water. A ferri/ferrocyanide solution was made to

be 10 mM of both K3Fe(CN)6 and K4Fe(CN)6∙3H2O in 1 M aqueous KCl. Experiments

for water electrolysis were performed in either acidic (1 M H2SO4), neutral (1 M

phosphate-buffered, pH 7) or basic (1 M NaOH) solutions.

Sample Preparation. Atomic layer deposition (ALD) of TiO2 was performed at

200°C on silicon wafers that were coated with a thin (< 2 nm) chemical oxide as a result

of surface preparation by the wafer vendor. Tetrakisdmethylamido titanium (TDMAT)

was used as the titanium source and water vapor as oxygen source. Each source was

heated to a line temperature of 80°C. The system pressure was maintained at 1.1 Torr

nominally, and nitrogen was used as the carrier gas. The pulse and purge durations of

the titanium and water sources were 5 s and 0.75 s respectively. Unless stated

otherwise, a total of 24 cycles of ALD was performed to obtain a TiO2 film of

approximately 2 nm thickness. The noble metal layer was then deposited by e-beam

evaporation with a quartz crystal balance used to monitor the thickness of material

deposited to obtain the nanocomposite electrode. For all samples, a thin layer (3 nm) of

metal was e-beam evaporated onto the backside of the silicon wafer (platinum for p+-Si

samples and aluminum for n-Si samples). This backside metal forms the electrical

contact and eliminates any Schottky junctions that could be formed at the back of the

118

substrate. Samples were annealed in forming gas (95% N2, 5% H2) at 400°C for 30

minutes in a quartz tube furnace before experiments were performed.

Electronic tunneling mediated by the metallic overlayer. Efficient direct

tunneling from the Si substrate through a thin TiO2 layer requires sufficient density of

states in the layer above TiO2. A metal layer (e.g. Pt, Ir, Ru, etc…) has abundant density

of states, and therefore, could sustain large tunneling current. The liquid electrolytes,

however, have limited density of states, resulting in the low tunneling current when they

are in direct contact with TiO2. Therefore, the metal layer effectively mediates the

charge transfer from Si and the reaction with the electrolyte. Other conductive catalysts,

such as IrO2, are also capable of serving the same role as the demonstrated metal

catalyst layers, with details to be reported in future publications.

Electrochemical Methods. All electrochemical experiments were performed

on either a WaveNow or WaveNano potentiostat (Pine Research Instrumentation) in air

at room temperature. A Pt wire was used as the counter electrode and a glass frit-

isolated Ag/AgCl/sat. KCl electrode as the reference electrode. A 5 mm bored Teflon

cone (area 0.196 cm2) was pressed against the DI water-rinsed nanocomposite working

electrode. All measured potentials were converted to the NHE reference scale using

E(NHE) = E(Ag/AgCl/KCl) + 0.197 V. The pH values for the electrolyte solutions used

in this study were determined by utilizing a reversible hydrogen electrode and

measuring the open circuit potential for each solution and adjusting the pH dependent

water oxidation potentials accordingly. Hydrogen gas was sparged into the electrolyte

solutions for ten minutes, with a platinum rotating disc electrode as the working

electrode and platinum mesh as the counter electrode.

119

Cyclic voltammetry (CV). All CVs were measured, unless stated otherwise, at

100 mV/s in the indicated electrolyte solution. The open circuit potential (OCP) was

measured before each experiment and used as the starting and ending potential for each

CV. A peristaltic pump (Cole Parmer, Norprene tubing, ID = 1.6 mm) was used to

continuously circulate electrolyte solution at a flow rate of 1 mL/s impinging on the

sample in the bore of the Teflon cone for all dark and illuminated water electrolysis

experiments (Figure 4.S1 for dark, Figure 4.S2 for illuminated). CV’s obtained with

solar illumination on n-Si anodes were performed in static solutions without pumping.

A modified pumping configuration with the electrolyte flow entering through an angled

hole in the side of the bore of the Teflon cone was used for stability tests conducted

under solar illumination so as not to block the light with the delivery tube (Figure 4.S2).

Table S1 lists the peak-to-peak splitting for the ferri/ferrocyanide solution for different

anode structures.

Impedance spectroscopy measurements. Impedance spectroscopy was

performed on a Princeton Applied Science impedance spectrometer to determine the

solution resistance of the electrolytes used in this study. The Nyquist plots for the

acidic, neutral and basic electrolyte solutions are shown in Figure 4.S3. Extrapolation

of the Nyquist plots to large frequencies reveals series resistance to be 7.5 Ω, 33.5 Ω

and 15 Ω for the acidic, neutral and basic solutions respectively. The obtained series

resistance values correspond to theoretical calculations very well. As an example, the

resistivity of 1 M NaOH is 5.9 cm (by linear extrapolation). By using a cylindrical

geometry of 5 mm height and 5 mm diameter, the calculated resistance is 15 .

120

Chronoamperometry. Chronoamperometry experiments were performed by

holding the potential at 1.5 V vs. Ag/AgCl/KCl reference electrode and monitoring the

current over time in 10 mL of the specified electrolyte solution. The electrolyte

solution was circulated with the peristaltic pump as described above. Results for a 18

hour experiment in neutral solution without illumination are shown in Figure 4.S4

below.

Chronopotentiometry (CP). All CP experiments were performed by holding

the current constant at 1 mA across a 0.196 cm2 anode surface area while monitoring

the potential over time in 10 mL of the specified electrolyte solution. The electrolyte

solution was circulated with the peristaltic pump using the procedure described above.

Diffusion-limited current. To understand the discrepancy between the

expected current density values obtained by extrapolation of the measured Tafel slopes

at low current density and the measured values obtained during CP and CA

experiments, diffusion-limited currents were estimated by determining the limiting

current for ferrocyanide oxidation under identical flow conditions and then adjusting

that value for the concentrations and the published diffusion constants of the different

species that might limit mass transfer during water oxidation: either transport of the

reactants (H2O or OH-) to the surface or transport of the products (O2 and H

+ or H2O)

away from the surface. The analysis showed that with a flow rate of 1 mL/s, mass

transfer of the reactants to the surface should not be limiting, with achievable current

densities beyond the solar flux limitations at 1 sun illumination. However, a major

discrepancy was found for mass transport of the sparingly soluble product of water

oxidation, O2, away from the electrode surface. Under the fastest flow rates possible in

121

our cell, 2 mL/s, oxygen bubbles would be expected to form at current densities of a

few mA/cm2. Because the current distribution is not expected to be uniform when mass

transport is limiting in this flow arrangement, we expect that bubbles form when the

average current density is of the order 1 mA/cm2. These bubbles should effectively

decrease the catalytic surface area and thus reduce the total current at a specific

overpotential relative to that expected by extrapolation of the Tafel slope measured at

low current density.

Temperature dependent tunneling experiments. Temperature dependent

tunneling experiments were performed on Ir/TiO2/p+-Si electrodes. Three thicknesses (2

nm, 4 nm, and 10 nm) of TiO2 were chosen to observe the thickness-dependent

tunneling behavior. A shadow mask is used to define the circular top metal contacts of

100 m diameter. The thickness of Ir top metal contact is 50 nm. The substrate bias was

scanned from 0 to 1 V while monitoring the current density, as shown in Figure 4.S5.

Transmission electron microscopy of Ir/TiO2/p+-Si anode after stability test.

Cross-sectional TEM was conducted after the 3 hr CP stability test (constant current at 5

mA/cm2) for the Ir/TiO2/p

+-Si sample (Figure S6). Comparing Figure 4.S6 with Figure

4.1B, it is evident that cross-sectional TEM images detect no apparent structural change

in the TiO2-protected anodes after the life test.

X-ray photoelectron spectroscopy (XPS). The XPS measurements were

performed using a PHI VersaProbe system with a 100 W Al-Kα X-ray source on a spot

size of 100 µm at a 45° incident angle. The binding energy scan range was 0-1200 eV

in 1 eV steps, and the pass energy was 117.4 eV, which provides the optimal balance

122

between scan resolution and counting statistics in our system. A dual beam neutralizer

(7 V Ar+ and 30 V electron beam) was used to neutralize sample charging. XPS depth

profiling was performed using an Ar+ ion beam at 2 kV and 1 A with an area of 2x2

cm. Spectra were collected at intervals of 3 s during sputtering.

Atomic Force Microscopy (AFM). The AFM images for these samples were

taken on a Park XE-70 instrument set on a non-contact tapping mode with a scan rate of

1 Hz. A representative AFM image of the nanocomposite anode after synthesis is

shown in Figure 4.S9.

123

4.10: Supporting Materials Figures

Figure 4.S1. Cross-sectional schematic of experimental setup utilized for dark

electrochemical experiments on Ir/TiO2/p+-Si anodes.

124

Figure 4.S2. Cross-sectional schematic of experimental setup utilized for stability tests

with n-Si/TiO2/Ir under solar illumination. This geometry allows for unobstructed

illumination of the sample.

125

Figure 4.S3. Nyquist plots from series resistance measurements from (A) acidic, (B)

neutral and (C) basic solutions measured from 300 kHz to 0.1 Hz.

126

Figure 4.S4. Constant potential measurement at 1.5 V vs. Ag/AgCl/sat. KCl in pH 7

buffered solution without illumination for 18 hours for Ir/p+-Si (∙∙∙) and Ir/TiO2/p

+-Si

(-).

127

Figure 4.S5. Temperature dependent tunneling current measurement for the Ir/TiO2/p+-

Si electrode. The ALD-grown TiO2 thicknesses were (A) 2 nm, (B) 4 nm, and (C) 10

128

nm. Measurements were taken at (-) 23oC, (-) 50

oC, and (-) 75

oC. The line (-)

indicates the compliance of the meter.

129

Figure 4.S6. Cross-sectional TEM image of Ir/TiO2/p+-Si anode taken after 3 hr

constant current (5 mA on a sample area of 0.196 cm2) stability test.

130

Figure 4.S7. Equilibrium band diagrams. (A) Ir/TiO2/p+-Si and (B) Ir/TiO2/n-Si anode

in contact with ferrocyanide solution. It is assumed that the redox level of the solution

equilibrates with the Ir metal catalyst.

131

Figure 4.S8. Light saturated current density. Cyclic voltammogram of Ir/TiO2/n-Si in

1M H2SO4 under simulated solar illumination at 1 sun.

132

Figure 4.S9. AFM image of Ir covered ALD-TiO2 sample. Root-mean-square

roughness ~0.2 nm

133

Table S1. Summary of peak-to-peak potential separation for anodes measured in

ferri/ferrocyanide solution

Substrate Eanodic to Ecathodic (mV)

TiO2/Si No observable peaks

Ir/TiO2/Si

130

Indium Tin Oxidea

220

a Delta Technologies, Product # CB-40IN-S211, 4-8 Ω with a small area, front-side

electrical contact outside the Teflon cone perhaps inducing a series resistance that

accounts for the larger peak-to-peak separation.

134

Chapter 5: Effect of TiO2 Thickness and Catalyst Layer on Efficiency and

Stability of Silicon Anodes for Water Oxidation

5.1: Preface

This chapter is an extension of the work that was done in Chapter. This work

explores what effects various thicknesses of the TiO2 protective layer have on charge

transfer efficiency and stability of the catalyst/TiO2/p+-Si anode. Additionally, the

catalyst layer was changed from iridium to other metals to examine the functionality

and utility of the TiO2/p+-Si base substrate. Understanding how the catalyst layer and

protective coating thickness effects both charge transfer efficiency and stability will

enable the development of more versatile anodes to use electricity for fuel synthesis.

I collected all of the data in this chapter while Vincent Chen and Rathnait Long

helped with the anode fabrication. The results reported here are being expanded upon

and will be incorporated into a manuscript to be submitted for publication.

135

5.2: Abstract

Recently, we reported on a silicon-based photoanode structure for

photoelectrochemical water oxidation stabilized by a thin layer of titanium dioxide

grown by atomic layer deposition. The catalyst layer was chosen to be iridium due to

its efficiency in catalyzing the water oxidation reaction and relative stability in a range

of pH values. Herein, we report on silicon/titanium dioxide/catalyst structures with

various thicknesses of titanium dioxide and the effect these different thicknesses have

on overall anode efficiency and the stability during water oxidation. Additionally, the

titanium dioxide was coated with different catalyst layers in order to study the

versatility of the anode for supporting these catalysts as well as the performance of each

layer for the oxidation of water.

136

5.3: Introduction

Electrochemical water splitting has long been considered a promising strategy to

store renewable electricity in the form of molecular fuels.1, 2

Of the two half reactions

necessary for water splitting (Scheme 1), the water oxidation half-reaction has been

considered the kinetic bottleneck due to the large overpotentials commonly associated

with the reaction. The water reduction half reaction (E° = 0 V vs. NHE at pH 0)

produces molecular hydrogen which can be used as a fuel.

Scheme 1: Water oxidation and proton reduction half reactions and pH-dependent

reversible potentials.

In order to be considered for large scale commercial use, a photoelectrochemical

water splitting device must utilize a cheap catalyst that minimizes the overpotential and

increases the efficiency of the water-oxidation half reaction.3, 4

Recently, we have

reported on a nanocomposite photoanode protected by a thin layer of titanium dioxide

(TiO2) that was deposited by atomic layer deposition (ALD) onto a photoactive silicon

substrate. An iridium layer was used as the water oxidation catalyst.5 This structure

takes advantage of a minimal amount of precious metal catalyst in order to decrease the

overall cost of the device while minimizing the absorption of sunlight prior to its

absorption by the photoactive silicon. Here, we report on how altering the thickness of

the protecting TiO2 layer and how changing the water oxidation catalyst layer effects

the resulting charge transfer efficiency and stability of the silicon-based anode toward

water oxidation.

137

5.4: Results and Discussion

The original photoanode structure was protected with a thin, conformal TiO2

protection layer deposited by ALD that was made to be 2 nm thick. In this study, we

employ otherwise similar anodes that can function in the dark by use of heavily

phosphorous-doped silicon substrates (p+-Si, 0.001-0.005 Ω·cm, 500 μm thick) coated

with ALD-TiO2 by exposing the surface to a titanium precursor, tetrakis-

(dimethylamido)titanium (TDMAT), and water vapor in alternating cycles until

thicknesses of 2, 5 and 10 nm were obtained. After e-beam depositing 2 nm of iridium

on each of these samples, the electron transport efficiency as a function of TiO2

thickness was investigated by monitoring the Fe2+/3+

couple in a 1 M aqueous KCl

solution containing 10 mM of both K4Fe(CN)6 and K3Fe(CN)6 (FFC solution) using

cyclic voltammetry (CV). The results in Figure 5.1A show that an increasing TiO2

thickness increases the peak-to-peak splitting of the oxidation and reduction peaks and

decreases the overall current density obtained. This effect is more dramatic with the

sample having a 10 nm TiO2 coating, indicating charge transfer efficiency through the

structure decreases rapidly as a function of increasing TiO2 thickness.

The peak-to-peak splitting values are plotted as a function of TiO2 thickness in

Figure 5.1B. These results show an exponential-type relationship which indicates that

charge transport through the structure occurs via a tunneling mechanism. This result is

not unexpected as current densities of up to a few A/cm2 are known to occur through

ultrathin oxide layers deposited by ALD in state-of-the-art field effect transistors.6 We

believe this conduction occurs by tunneling through the TiO2. The layer was previously

found to be uniform and continuous by TEM with no evidence of the catalyst layer

138

intercalating into the TiO2 layer to provide an alternative mechanism to transfer charge

through the insulating oxide layers of the anode. To date, ALD is the only method

capable of depositing such thin conformal layers free of any pinholes or cracks that

would otherwise allow for metal intercalation or oxidation of the underlying silicon

substrate by the electrolyte.7, 8

The charge transfer efficiency through the anodes with different TiO2

thicknesses can be further demonstrated by performing water oxidation. Figure 5.1C

shows water oxidation CVs that were performed in a 1 M NaOH solution.9 The

samples with thicker TiO2 layers had decreased performance for water oxidation when

compared to samples with thinner TiO2 layers. Figure 5.1D shows the water oxidation

overpotentials for each TiO2 thickness required to obtain a current density of 1 mA/cm2.

The current density obtained at any applied electrochemical potential decreases as the

thickness of the TiO2 protective coating is increased, consistent with the increasing

peak-to-peak splitting results obtained for the FFC solution shown in Figure 5.1B. The

maximum current density obtained at a given applied electrochemical potential was

observed to decrease as the TiO2 thickness was increased. The TiO2 layer can be

viewed as a resistor, where an increasing TiO2 thickness increases the series resistance

through which the current flows. This increasing resistance will decrease the current

density at a given overpotential, and thus the efficiency, that can be obtained by the

anode.

Once the charge transfer behavior was established, the stability of the interface

toward water oxidation was investigated as a function of the TiO2 thickness. Each

sample was subjected to a constant potential endurance test by holding the

139

electrochemical potential at 1.65 V vs. NHE in a 1 M H2SO4 solution for up to 48 hours

(Figure 5.2A). This applied electrochemical potential corresponds to a 444 mV

overpotential at this pH, which represents a value at which most water oxidation

catalysts of interest could operate at reasonable current density (see Table 5.1). CVs of

water oxidation in 1 M H2SO4 solution were performed before the stability test for each

sample (Figure 5.2B). The current density at the beginning of the stability test was

observed to be comparable to the current density obtained at the same overpotential in

these CVs. The current density initially decreases to reach a steady-state current for

samples that have TiO2 protective coatings. This diffusion-limited current density of 1-

2 mA/cm2 is due to the inefficient removal of molecular oxygen bubbles away from the

anode surface (see supporting materials). The samples containing the TiO2 protective

coating were found to remain operational throughout the entirety of the stability test of

48 hours in this report. The sample without the TiO2 layer was found to decrease to a

current density below that of the sample with 5 nm of TiO2 within half an hour and

decrease steadily to a minimal value after 25 hours. XPS analysis of this sample after it

was stopped reveals a loss of Ir and increase in Si peaks attributed to SiO2 (Figure 5.3A)

The loss of Ir from the surface is most likely due to either oxidative stress or poor

adhesion of the Ir layer to TiO2 with each removal mechanism accelerated by pumping

the electrolyte solution. Ir loss was also observed for samples with the TiO2 protective

coating, indicating the failure mechanism is due to SiO2 growth, which is consistent

with previously published results.5 It is worth noting that the current densities

obtained for the 10 nm TiO2 sample are consistent with the model of thicker TiO2 layers

having more resistance and less measured current density. This observation highlights

140

the importance of balancing stability with efficiency for these anodes to be considered

for commercial, large scale use. Currently, new photoelectrochemical water splitting

flow cell devices are being designed that will more efficiently remove oxygen bubbles

from the anode while allowing for light absorption.

The thickness of the Ir catalyst layer on top of the TiO2/p+-Si anode was also

varied in order to determine how this affected the anode charge transfer efficiency. The

thickness of the Ir layer from the standard 2 nm was changed to smaller and larger

thicknesses of 1 nm and 3 nm, each deposited onto samples with 2 nm of TiO2. The Ir

layer thickness was found to have a negligible effect on the FFC peak-to-peak splitting

and water oxidation in acidic (1 M H2SO4), neutral (1 M phosphate-buffered) and basic

(1 M NaOH) solutions (Table 5.1). This result indicates that thinner Ir layers could be

employed in photoelectrochemical water splitting devices as efficiency would not suffer

and the amount of precious metal catalyst, and thus cost, would be minimized. The

catalyst layer is vital, however, as no charge will pass without the layer present to act as

a tunnel mediator between the silicon base substrate and redox species in solution. It

has yet to be determined how the catalyst thickness will be affected by the loss of Ir

observed during stability tests.

In addition to varying the thickness of the Ir catalyst, the type of catalyst was

also varied. Different catalysts may be optimal for different electrochemical reactions.

These catalyst layers were deposited onto TiO2/p+-Si anodes with a 2 nm TiO2 layer.

The results of these studies are tabulated in Table 5.1. A 3 nm layer of ruthenium was

found to be a competent water oxidation catalyst in the acidic solution. The active form

of the catalyst was likely a ruthenium oxide formed in situ as evidenced by a pre-

141

oxidation peak near the onset of the water oxidation wave (see supporting materials).

The overpotentials required to obtain 1 mA/cm2 were comparable to the Ir catalyst in

the acidic solution, but the catalyst dissolved off the anode with successive CV scans

and increasing pH. A 2 nm layer of platinum was also deposited and found to have

comparable charge transfer efficiency in FFC but displayed larger overpotentials for

water oxidation at all pH values. This is due to the fact Pt is not as good as a water

oxidation catalyst as Ir or Ru.10-13

Despite the inefficiency at catalyzing water

oxidation, the Pt layer still demonstrates the comparable ability to mediate charge

transfer to and from the silicon substrate as evidenced by the FFC peak-to-peak splitting

results. A 2 nm gold layer deposited onto the anode structure was found to have a peak-

to-peak splitting value of 120 mV in the FFC solution which was as efficient as the 3

nm of Ir catalytic layer. The water oxidation behavior of the gold layer resulted in large

overpotentials, consistent with the fact that gold is not a good water oxidation catalyst.

A 2 nm Co catalyst layer was also deposited and found to have large peak-to-peak

splitting in the FFC solution. This result indicates that the Co layer is not a good

electron transfer mediator possibly due to poor adhesion of the layer to the TiO2.

The search for a cheap, efficient and robust water oxidation catalyst made from

earth-abundant elements has led to the discovery of a cobalt-phosphate (Co-Pi) catalyst

found to oxidize water in neutral pH.14

Although the oxidized form of cobalt, Co3O4, is

known to oxidize water15

, the observed overpotentials for the planar Co layer were

found to be considerably greater than the Ir and Ru catalyst layers observed in this

report. To improve on the water oxidation overpotentials, the Co-Pi catalyst was

deposited in situ following published procedures.14

The deposited Co-Pi catalyst did

142

decrease the overpotential by 236 mV for water oxidation in the pH 7 solution but had a

minor effect in the acidic and basic solutions. Furthermore, the overpotential necessary

to obtain a current density of 1 mA/cm2 was found to be 400 mV worse than the Ir and

Ru catalyst layers at pH 7. This result is consistent with the thin layer of Co deposited

exhibiting poor electron transfer efficiency with the FFC solution. The Co-Pi catalyst

deposited here does not compare well to the results of the original study14

which reports

an overpotential of around 400 mV for 1 mA/cm2 at pH 7, not 730 mV as obtained here.

It is worth noting, however, that in this work the catalyst was deposited onto a Co layer

instead of an indium tin-doped oxide (ITO) electrode as was originally published. A

recent report has illustrated the ability to deposit the CoPi catalyst onto a thin layer of

ITO that was coating a pn junction Si wafer to be used for photoelectrochemical water

splitting.16

143

5.5: Conclusions

This report has highlighted the effect that changing the TiO2 thickness of the

nanocomposite catalyst/TiO2/p+-Si anode has on charge transport behavior and water

oxidation efficiency. The charge transport efficiency through the structure decreases as

the TiO2 thickness is increased. This is consistent with the model of the TiO2 layer

acting as a resistor in series with the electrochemical process and charge transfer though

the structure. The catalyst layer on the TiO2 was also varied in order to study the effect

different metal layers have on electron transfer and water oxidation in a range of pH

solutions. The Ir layer was found to be the best performing at all pH values overall, but

other catalyst layers have been shown to be capable catalyst layers that may allow for

greater functionality of the anode for other electrochemical reactions.

144

5.6: Figures

Figure 5.1: Electrochemical Results of Anodes with Various TiO2 Thicknesses. (A)

CVs in the FFC solution using 2 nm of Ir deposited onto various thicknesses of TiO2 on

p-Si anodes. (B) Plot of peak-to-peak splitting values in panel (A) as a function of TiO2

thickness. (C) Water oxidation CVs in 1 M NaOH solution using 2 nm of Ir deposited

onto various thicknesses of TiO2 on p-Si anodes. (D) Calculated overpotentials for the

samples in panel (C) to obtain a current density of 1 mA/cm2. The pH value for the 1 M

NaOH solution was found to be 13.7 as measured against the reversible hydrogen

electrode leading to a reversible potential of 0.418 V versus NHE. The corrected

solution resistance was 15 Ω as measured by electrochemical impedance spectroscopy.

145

Figure 5.2: Anodes Stability Tests. (A) Constant potential stability tests performed on

anodes with 2 nm of Ir on various thicknesses of TiO2 on p-Si in 1 M H2SO4 solution at

an applied electrochemical potential of 1.65 V vs. NHE. This applied potential

corresponds to an overpotential of 444 mV in this solution. (B) Water oxidation CVs

in 1 M H2SO4 solution for the samples used in panel (A). The current densities

obtained by CV at 1.65 V indicate much greater values when compared to the steady

state current densities in (A) due to diffusion of the product, molecular oxygen, away

from the surface.

146

Figure 5.3: XPS Analysis of Samples Before and After Stability Tests. (A) X-ray

photoelectron (XPS) spectra of Ir, Si and Ti regions for a 2 nm Ir layer deposited onto a

p-Si substrate before () and after () the stability test in 1 M H2SO4 solution. The

binding energy for atomic Si in the bulk is at 100 eV, while the binding energy for Si in

SiO2 is 104 eV, both shown to be increasing after the stability test was performed. This

sample lost Ir as the stability test was conducted and is absent of Ti. (B) XPS spectra

for Ir, Si and Ti regions for 2 nm Ir layer deposited onto the 5 nm TiO2/p-Si substrate

before () and after () the stability test in 1 M H2SO4 solution. This sample was

operational after the stability test, indicating that the loss of Ir is not the reason for

failure of the sample without the TiO2 layer. The Ti peak is shown to increase, while

the underlying Si was not observed.

147

Table 5.1. Results obtained for peak-to-peak splitting of ferro/ferri cyanide (FFC)

peaks and water oxidation overpotentials at 1 mA cm-2

for different catalyst layers on

2 nm of TiO2 on p+-Si.

Catalyst layer FFC splitting (mV) 1 M Acid (mV) 1 M Neutral (mV) 1 M Base (mV)

1 nm Ir 160 300 311 340

2 nm Ir 140 282 319 337

3 nm Ir 120 265 316 343

3 nm Rua 195 265 383 -

2 nm Pt 175 469 628 575

2 nm Aub 120 1160 1120 790

2 nm Co > 1000 885 966 937

2 nm Co + CoPic > 1000 817 730 1061

a Catalyst layer found to dissolve off electrode with increasing cycle numbers and pH

b Deposited onto 1 nm Ti adhesion layer on TiO2/p

+-Si

c Co-Pi catalyst deposited according to published procedures

14

148

5.7: References

1. Lewis, N. S., Powering the planet. MRS Bull. 2007, 32, (10), 808-820.

2. Walter, M. G.; Warren, E. L.; McKone, J. R.; Boettcher, S. W.; Mi, Q.; Santori,

E. A.; Lewis, N. S., Solar Water Splitting Cells. Chem. Rev. 2010, 110, (11),

6446-6473.

3. Gratzel, M., Photoelectrochemical cells. Nature 2001, 414, (6861), 338-344.

4. Brimblecombe, R.; Dismukes, G. C.; Swiegers, G. F.; Spiccia, L., Molecular

water-oxidation catalysts for photoelectrochemical cells. Dalton Trans. 2009,

(43), 9374-9384.

5. Chen, Y. W.; Prange, J. D.; Duehnen, S.; Park, Y.; Gunji, M.; Chidsey, C. E. D.;

McIntyre, P. C., Atomic layer-deposited tunnel oxide stabilizes silicon

photoanodes for water oxidation. Nature Mat. 2011, 10, 539-544.

6. Process Integration, Devices, and Structures. International Technology Roadmap

for Semiconductors.

http://www.itrs.net/links/2008ITRS/Update/2008Tables_FOCUS_A.xls

7. Decker, F.; Fracastorodecker, M.; Badawy, W.; Doblhofer, K.; Gerischer, H.,

The Photocurrent Voltage Characteristics of the Heterojunction Combination n-

Si/SnO2/Redox Electrolyte. J. Electrochem. Soc. 1983, 130, (11), 2173-2179.

8. Kohl, P. A.; Frank, S. N.; Bard, A. J., Semiconductor Electrodes. 11. Behavior

of n-Type and p-Type Single Crystal Semiconductors Covered with Thin

Normal TiO2 Films. J. Electrochem. Soc. 1977, 124, (2), 225-229.

9. The pH value for each of the 1 M H2SO4, 1 M phosphate buffered and 1 M

NaOH solutions were obtained by measuring the reversible hydrogen electrode

149

values in each solution. Hydrogen was sparged into the electrochemical cell and

a Pt disc was used as the working electrode with a Pt mesh as the counter

electrode. The values obtained show a pH of 0.4, 6.99 and 13.97 for the acidic,

neutral and basic solutions, respectively.

10. Yagi, M.; Tomita, E.; Kuwabara, T., Remarkably high activity of

electrodeposited IrO2 film for electrocatalytic water oxidation. J. Electroanal.

Chem. 2005, 579, (1), 83-88.

11. Nakagawa, T.; Beasley, C. A.; Murray, R. W., Efficient Electro-Oxidation of

Water near Its Reversible Potential by a Mesoporous IrOx Nanoparticle Film. J.

Phys. Chem. C 2009, 113, (30), 12958-12961.

12. Neumannspallart, M.; Kalyanasundaram, K.; Gratzel, C.; Gratzel, M.,

Ruthenium Dioxide Electrodes as Suitable Anodes for Water Photolysis. Helv.

Chim. Acta. 1980, 63, (5), 1111-1118.

13. Miles, M. H.; Klaus, E. A.; Gunn, B. P.; Locker, J. R.; Serafin, W. E., Oxygen

Evolution Reaction on Platinum, Iridium, Ruthenium and their Alloys at 80

Degrees C in Acid Solutions. Electrochimica Acta 1978, 23, (6), 521-526.

14. Kanan, M. W.; Nocera, D. G., In situ formation of an oxygen-evolving catalyst

in neutral water containing phosphate and Co2+

. Science 2008, 321, (5892),

1072-1075.

15. Iwakura, C.; Honji, A.; Tamura, H., The Anodic Evolution of Oxygen on Co3O4

Film Electrodes in Alkaline Solutions. Electrochimica Acta 1981, 26, (9), 1319-

1326.

150

16. Pijpers, J. J. H.; Winkler, M. T.; Surendranath, Y.; Buonassisi, T.; Nocera, D.

G., Light-induced water oxidation at silicon electrodes functionalized with a

cobalt oxygen-evolving catalyst. Proc. Nat. Acad. Sci., 2011, 108, (25) 10056-

10061.

151

5.8: Supporting Materials

Materials. All chemicals used in this study were purchased from commercial

sources and used as received without further purification. The Si wafers used were

degenerately doped p-type Si (100) wafers (p-Si, 0.001-0.002 Ωcm, 500 m thickness)

obtained from El-Cat. The wafers were used as received, with a thin (< 2 nm) SiO2

layer as prepared by the wafer vendor. Metal deposition was performed by e-beam

evaporation of all the catalyst layers used in this study, with the Co-Pi system deposited

according to published methods.1 The backside contact for the p-Si substrates was 2 nm

of Pt deposited by e-beam evaporation which was used to avoid any Schottky Junctions

that might form. After fabrication, the samples were heated to 400°C for 30 minutes in

a forming gas environment (95% N2, 5% H2). The electrolyte solutions were made by

dissolving the appropriate reagent into Millipore filtered water (10 MΩ resistance)

obtained on a Millipore filtration system. A ferri/ferrocyanide solution was made to be

10 mM of both K3Fe(CN)6 and K4Fe(CN)6∙3H2O in 1 M aqueous KCl. Experiments for

water electrolysis were performed in either acidic (1 M H2SO4), neutral (1 M

phosphate-buffered, pH 7) or basic (1 M NaOH) solutions. The pH of the neutral

solution was made by dissolving Na2HPO4 and NaH2PO4 in water with the pH adjusted

until a pH of 7 was achieved. All pH values were calculated by measuring the

reversible hydrogen potential in each solution and all solution resistance measurements

obtained by electrochemical impedance spectroscopy.

Sample Preparation. Atomic layer deposition (ALD) of TiO2 was performed

at 200°C on silicon wafers that were coated with a thin (< 2 nm) chemical oxide as a

result of surface preparation by the wafer vendor. Nitrogen gas was the carrier gas and

152

tetrakis-(dimethylamido)titanium (TDMAT) was used as the titanium source and water

vapor as the oxygen source. Each reagent was heated to a line temperature of 80°C

with the system pressure maintained at 1.1 Torr. The pulse and purge durations of the

titanium and water sources were 5 s and 0.75 s respectively. A total of 24 cycles of

ALD was performed to obtain a TiO2 film of approximately 2 nm thickness while 48

cycles and 120 cycles of ALD were performed to obtain TiO2 thicknesses of 5 and 10

nm, respectively. The thicknesses were obtained using ellipsometry that was calibrated

with thicknesses measured by transmission electron microscopy (TEM). The noble

metal layers were deposited by e-beam evaporation with a quartz crystal balance used to

monitor the thickness of material deposited to obtain the final anode. All samples were

annealed in forming gas (95% N2, 5% H2) at 400°C for 30 minutes in a quartz tube

furnace before experiments were performed.

Electrochemical Methods. All electrochemical experiments were performed on a

WaveNow potentiostat (Pine Research Instrumentation) in air at room temperature. A

Pt wire was used as the counter electrode and a glass frit-isolated Ag/AgCl/sat. KCl

electrode as the reference electrode. A 5 mm bored Teflon cone (area 0.196 cm2) was

pressed against the catalyst/TiO2/p-Si sample that was used as the working electrode.

The pH values for the electrolyte solutions used in this study were determined by

utilizing a reversible hydrogen electrode (RHE) and measuring the open circuit

potential for each solution and adjusting the water oxidation potentials accordingly.

RHE measurements were conducted by sparging hydrogen gas into the electrolyte

solutions for ten minutes, with a platinum rotating disc electrode used as the working

electrode and platinum mesh as the counter electrode. All measured potentials in this

153

study were collected using the Ag/AgCl reference electrode and then converting to the

NHE reference scale using E(NHE) = E(Ag/AgCl/KCl) + 0.197 V. All cyclic

voltammograms (CVs) were measured at 100 mV/s in the indicated electrolyte solution.

A peristaltic pump (Cole Parmer, Norprene tubing, ID = 1.6 mm) was used to

continuously circulate electrolyte solution at a flow rate of 2 mL/s impinging on the

sample in the bore of the Teflon cone for all dark water electrolysis experiments.

Impedance spectroscopy measurements. Impedance spectroscopy was

performed on a Princeton Applied Science impedance spectrometer to determine the

solution resistance of the electrolytes used in this study. Extrapolation of the Nyquist

plots to large frequencies reveals series resistance to be 7.5 Ω, 33.5 Ω and 15 Ω for the

acidic, neutral and basic solutions, respectively.

X-ray photoelectron spectroscopy (XPS). The XPS measurements were

performed using a PHI VersaProbe system with a 100 W Al-Kα X-ray source on a spot

size of 100 µm at a 45° incident angle. The binding energy scan range was 0-1000 eV

in 1 eV steps with a pass energy of 117.4 eV for survey scans. High resolution scans

were taken in the appropriate ranges for each atom in 0.1 eV steps with a pass energy of

23.1 eV. A dual beam neutralizer (7 V Ar+ and 30 V electron beam) was used to

neutralize sample charging.

Platinum Catalyst Layer. The Pt catalyst layer was deposited onto the TiO2/p-

Si anode by physical vapor deposition methods and found to be 2 nm thick. CVs for the

Pt coated sample in the FFC solution and 1 M H2SO4, 1 M phosphate buffered and 1 M

NaOH solutions are shown in Figure S1A and S1B.

154

Ruthenium Catalyst Layer. The Ru catalyst layer was deposited onto the

TiO2/p-Si anode via e-beam physical vapor deposition. The thickness was found to be 3

nm. The CV of the Ru layer in the FFC solution shows normal behavior for a catalyst

that allows for facile charge transport through the structure (Figure S2A). For water

oxidation, however, the Ru on the first CV cycle was found to have an irreversible

oxidation peak before onset of water oxidation (Figure S2B). On the second cycle, the

pre-water oxidation peak was not observed, and the water oxidation efficiency was

measured. With each successive cycle, the water oxidation peak was found to decrease

until ultimately the peak became too small to measure at 1 mA/cm2 current density.

The Ru sample was analyzed by XPS after analysis and was found to have a sharp

decrease in the amount of Ru on the surface when compared to the freshly prepared

portion of the sample (Figure S2C). The Ru layer most likely oxidized to RuO2, or

possibly RuO4, and dissolved off of the surface and into the electrolyte solution.

Cobalt Catalyst Layer. The Co catalyst layer was deposited by e-beam

physical vapor deposition and found to be 2 nm thick. Curiously, CVs performed in the

FFC solution yielded no peaks in the potential windows they were scanned. Activity

for water oxidation on this catalyst was found to be sluggish requiring greater

overpotentials to reach 1 mA/cm2 (Figure S3A). We then decided to deposit the CoPi

water oxidation catalyst onto this Co catalyst layer to see if the water oxidation

performance improved. After performing a CA at 1.29 V for up to 3 hours from a 0.5

mM CoSO4 solution in 100 mM phosphate buffered solution, the current measured

reached a maximum and the CA was stopped and the deposition solution was removed

and replace with each a 1 M H2SO4, 1 M phosphate buffered and 1 M NaOH solution.

155

The water oxidation overpotential was decreased dramatically in the neutral solution

(Figure S3B). However, the water oxidation behavior in both acid and base seemed to

have negligible effect with the CoPi catalyst deposited.

Gold Catalyst Layer. A 2 nm thick layer of gold was deposited by physical

vapor deposition methods. The layer was found to have small peak-to-peak splitting for

the FFC solution indicating gold being a good electron transfer catalyst layer (Figure

S4A). The overpotentials for water oxidation were found to large which is consistent

with the fact that gold is a poor water oxidation catalyst.

Chronoamperometry. Chronoamperometry (CA) experiments were performed

by holding the applied electrochemical potential at a constant value while monitoring

the current over time. The electrolyte solution was circulated with the peristaltic pump

as described above. The current density for each experiment was found to peak at time

zero and decrease to a steady state current value below the expected current density

value obtained by CV in the same solution at the same applied electrochemical

potential. The cause of this decrease is due to the inefficient removal of molecular

oxygen, a sparingly soluble product of water oxidation. The oxygen bubbles block the

surface with a diffuse layer, preventing reactants from reaching the surface and getting

oxidized. The steady state current density that results is due to an equilibrium of

reactants to the surface and removal of the oxygen away from the surface.

156

5.9: Supporting Materials Figures

Figure S1. (A) CV with 2 nm of Pt on TiO2/p-Si in the FFC solution. (B) Water

oxidation CVs for Pt sample in acidic, neutral and basic solutions. The reduction wave

in the acidic solution is attributed to the reduction of a Pt-O layer formed during water

oxidation.

157

Figure S2. (A) CV with 3 nm Ru deposited onto TiO2/p-Si in the FFC solution

showing oxidation and reduction peaks. (B) Water oxidation CVs in 1 M Acid ()

and 1 M phosphate buffered () solutions showing the decreasing current density with

increasing cycle number, indicating that the Ru catalyst is coming off the surface. (C)

XPS narrow scans in the Ru region showing the amount of Ru on the electrode after

preparation is much greater than the amount of Ru on the electrode after analysis.

158

Figure S3. (A) Water oxidation CVs in acidic (), neutral () and basic ()

solutions for a 2 nm Co catalyst layer deposited onto 2 nm TiO2/p-Si. (B) Water

oxidation CVs of the Co layer from (A) with the Co-Pi catalyst deposited in situ.

159

Figure S4. (A) CV of 2 nm of gold deposited onto 2 nm of TiO2/p-Si in FFC solution.

(B) CVs of the gold sample in basic (),neutral () and acidic () solutions for water

oxidation. The results show that the obtained current density is considerably lower than

the values obtained using other water oxidation catalysts.

160

5.10: Supporting Materials References

1. Kanan, M. W.; Nocera, D. G., In situ formation of an oxygen-evolving catalyst

in neutral water containing phosphate and Co2+

. Science 2008, 321, (5892),

1072-1075.