conjugated organic polymers as photocathode materials in...

132
Conjugated Organic Polymers as Photocathode Materials in Organic Photoelectrochemical Cells by Patrick Fortin B.Sc., University of Alberta, 2013 Thesis Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in the Department of Chemistry Faculty of Science © Patrick Fortin 2019 SIMON FRASER UNIVERSITY Spring 2019 Copyright in this work rests with the author. Please ensure that any reproduction or re-use is done in accordance with the relevant national copyright legislation.

Upload: others

Post on 08-Jun-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

Conjugated Organic Polymers as Photocathode

Materials in Organic Photoelectrochemical Cells

by

Patrick Fortin

B.Sc., University of Alberta, 2013

Thesis Submitted in Partial Fulfillment of the

Requirements for the Degree of

Doctor of Philosophy

in the

Department of Chemistry

Faculty of Science

© Patrick Fortin 2019

SIMON FRASER UNIVERSITY

Spring 2019

Copyright in this work rests with the author. Please ensure that any reproduction or re-use is done in accordance with the relevant national copyright legislation.

Page 2: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

ii

Approval

Name:

Degree:

Title:

Examining Committee:

Date Defended/Approved:

Patrick Fortin

Doctor of Philosophy (Chemistry)

Conjugated Organic Polymers as Photocathode Materials in Organic Photoelectrochemical Cells

Chair: Corina Andreoiu Associate Professor

Steven Holdcroft Senior Supervisor Professor

Byron D. Gates Supervisor Associate Professor

Loren Kaake Supervisor Assistant Professor

Vance E. Williams Internal Examiner Associate Professor

Timothy L. Kelly External Examiner Associate Professor Department of Chemistry University of Saskatchewan

March 29, 2019

Page 3: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

iii

Abstract

The work presented in this thesis aims at studying the photoelectrochemical

properties of conjugated organic polymers, with a particular interest in their application as

photocathode materials to perform the hydrogen evolution reaction. Whereas inorganic

semiconductors have been the primary focus in the area of solar water splitting, organic

semiconductors have only recently emerged as a new class of photoelectrode material.

Furthermore, the research emphasis on organic semiconductor systems used for water

splitting applications has largely been focused on improving performance through device

engineering. Reports concerning the interfacial thermodynamic and electronic processes

are less frequent. Given the difference in photophysical properties compared to inorganic

semiconductors and the change in chemical environment compared to solid-state organic

electronics, the characterization and elucidation of interfacial processes at the organic

semiconductor-electrolyte remains an area of need.

The ability of uncatalyzed P3HT photocathodes to perform the hydrogen evolution

reaction in acidic aqueous media is first investigated. Whereas previous reports in the

literature have attributed the aqueous photoactivity of P3HT to the hydrogen evolution

reaction, this study reveals that residual dissolved oxygen is largely responsible for the

observed photocurrents and that, despite favorable thermodynamics, hydrogen is not

evolved at P3HT photocathodes in the absence of a catalyst.

Following the initial investigation of P3HT photocathode performance, strategies

to improve photocurrent densities at organic photocathodes are explored. In these studies,

nanostructured photocathodes are prepared from P3HT:PCBM nanoparticles. The

optoelectronic and morphological properties, as well as the photoelectrochemical

performance of the nanostructured photocathodes are compared to planar P3HT:PCBM

photocathodes. To achieve hydrogen production, platinum nanoparticle catalysts are

deposited onto the organic layer photoelectrochemically, where the increased surface

area of the nanostructured electrodes leads to enhanced catalyst loadings and increased

photocurrent densities compared to the planar photocathodes.

Finally, the influence of PCBM on interfacial energy alignment of the redox couple

at the semiconductor-electrolyte interface is investigated in a non-aqueous electrolyte

using a benzoquinone redox couple. Photoelectrochemical measurements show that the

presence of PCBM at the semiconductor-electrolyte interface leads to the formation of an

interfacial dipole layer and decreases the built-in potential developed at the

semiconductor-electrolyte interface.

Page 4: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

iv

Keywords: conjugated organic polymers, photoelectrochemistry, water splitting,

hydrogen evolution, nanoparticles, P3HT, P3HT:PCBM

Page 5: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

v

Dedication

To Kristine, Shadow and Belle

Page 6: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

vi

Acknowledgements

I would like to thank Prof. Steven Holdcroft for the opportunity to be a part of his

research group. His guidance has been instrumental in my growth as a scientist during

my time at SFU. I would also like to thank my supervisory committee members, Prof. Loren

Kaake and Prof. Byron Gates, for their insightful comments and constructive feedback

over the course of my project.

I would like to thank all the present and former members of the Holdcroft group

that I have worked with over the years. Special thanks need to be given to Dr. Graeme

Suppes for all the knowledge and technical expertise he has shared with me, as well as

Dr. Pankaj Chowdhury and Mr. Subash Rajasekar for all the great technical discussions

we have had.

I would like to thank the SFU Department of Chemistry and 4D Labs staff who have

helped me along the way: everyone at the electronics and machine shop for helping to

keep my experiments running, Mr. Bruce Harwood for fabricating the glass

photoelectrochemical cell, and Dr. Xin Zhang for the extensive training and wealth of

knowledge regarding electron microscopy.

Page 7: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

vii

Table of Contents

Approval .......................................................................................................................... ii

Abstract .......................................................................................................................... iii

Dedication ....................................................................................................................... v

Acknowledgements ........................................................................................................ vi

Table of Contents .......................................................................................................... vii

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

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

List of Acronyms, Symbols and Abbreviations .............................................................. xvi

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

1.1. Hydrogen as an Alternative Fuel ............................................................................ 1

1.2. Photoelectrolysis of Water ..................................................................................... 4

1.2.1. Fundamentals of Semiconductor Photoelectrochemistry ............................... 4

1.2.2. The Hydrogen Evolution and Oxygen Evolution Reactions ............................ 9

1.2.3. Device Architecture and Design Principles .................................................. 10

Semiconductor Photocatalyst Systems .................................................................. 11

Photoelectrochemical Cells .................................................................................... 13

Strategies to Increase Performance ....................................................................... 18

1.3. Conjugated Organic Polymers ............................................................................. 21

1.3.1. Overview ..................................................................................................... 21

1.3.2. Electronic Processes in Organic Semiconductors ........................................ 23

1.3.3. History ......................................................................................................... 25

1.3.4. P3HT ........................................................................................................... 28

1.4. Organic Photoelectrochemical Cells .................................................................... 31

1.5. Photoelectrode Performance and Experimental Methods .................................... 34

1.5.1. Benchmark and Diagnostic Efficiencies ....................................................... 34

Solar-to-Hydrogen Conversion Efficiency ............................................................... 34

Applied Bias Photon-to-Current Efficiency (ABPE) ................................................. 35

Incident Photon-to-Current Efficiency (IPCE) ......................................................... 36

Absorbed Photon-to-Current Efficiency (APCE) ..................................................... 37

1.5.2. Experimental Methods ................................................................................. 38

Photoelectrochemical Cell ...................................................................................... 38

Photocurrent .......................................................................................................... 39

Faradaic Efficiency ................................................................................................. 40

Built-In Potential ..................................................................................................... 40

Ultraviolet-Visible Spectroscopy ............................................................................. 41

Photoluminescence Spectroscopy ......................................................................... 42

Grazing Incidence Wide Angle X-Ray Scattering ................................................... 42

Dynamic Light Scattering ....................................................................................... 44

1.6. Thesis Scope ....................................................................................................... 45

Page 8: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

viii

Chapter 2. On the Origin of Photocurrent at Uncatalyzed P3HT Films in Aqueous Acidic Media ...................................................................................................... 47

2.1. Introduction .......................................................................................................... 47

2.2. Experimental ....................................................................................................... 48

2.2.1. Materials ...................................................................................................... 48

2.2.2. Film Preparation .......................................................................................... 49

2.2.3. Photoelectrochemistry ................................................................................. 49

2.2.4. Hydrogen Gas Detection ............................................................................. 50

2.2.5. Dissolved Oxygen Detection ........................................................................ 50

2.3. Results ................................................................................................................ 50

2.3.1. P3HT photocurrent vs. O2 concentration ...................................................... 50

2.4. Conclusion........................................................................................................... 53

Chapter 3. Hydrogen Evolution at Conjugated Polymer Nanoparticle Electrodes 55

3.1. Introduction .......................................................................................................... 55

3.2. Experimental ....................................................................................................... 58

3.2.1. Materials ...................................................................................................... 58

3.2.2. P3HT:PCBM Nanoparticle Films .................................................................. 58

3.2.3. Deposition of Pt ........................................................................................... 59

3.2.4. Photoelectrochemistry ................................................................................. 59

3.2.5. Gas Chromatography .................................................................................. 60

3.2.6. Electron Microscopy .................................................................................... 60

3.2.7. Spectroscopy ............................................................................................... 60

3.2.8. Grazing Incidence Wide Angle X-ray Scattering .......................................... 61

3.2.9. Dynamic Light Scattering ............................................................................. 61

3.3. Results and Discussion ....................................................................................... 61

3.3.1. Nanoparticle Synthesis ................................................................................ 61

3.3.2. Characterization of Nanoparticles ................................................................ 63

3.3.3. Photoelectrochemistry of Nanoparticle-based Photocathodes ..................... 68

3.3.4. Conclusion ................................................................................................... 71

Chapter 4. Energy Level Alignment and Interfacial Dipole Layer Formation at the P3HT:PCBM-Electrolyte Interface in Organic Photoelectrochemical Cells ... 74

4.1. Introduction .......................................................................................................... 74

4.2. Experimental ....................................................................................................... 76

4.2.1. Materials ...................................................................................................... 76

4.2.2. P3HT:PCBM Films ...................................................................................... 76

4.2.3. Photoelectrochemistry ................................................................................. 76

4.3. Results and Discussion ....................................................................................... 77

4.3.1. Open Circuit Potential & Onset Potential for Photocurrent ........................... 77

4.3.2. Role of a PEDOT:PSS Charge Selective Layer ........................................... 80

4.4. Conclusion........................................................................................................... 84

Chapter 5. Conclusions and Future Directions ...................................................... 86

Page 9: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

ix

5.1. Conclusions ......................................................................................................... 86

5.2. Future Directions ................................................................................................. 87

5.2.1. Rational Polymer Design ............................................................................. 87

Conjugated Organic Polymers with High Dielectric Constants ................................ 88

Band-Edge Engineering ......................................................................................... 90

5.2.2. Device Engineering ..................................................................................... 93

Polymer Orientation ............................................................................................... 93

Nanostructured Photoelectrodes ............................................................................ 95

References ................................................................................................................... 98

Page 10: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

x

List of Tables

Table 3-1 Solubility of P3HT & PCBM in various solvents36 and solvent properties. .............................................................................................. 62

Page 11: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

xi

List of Figures

Figure 1.1 Relative Fermi level (EF) positions of a) intrinsic, b) p-type and c) n-type semiconductors. Also shown are the valence band (VB) and conduction band (CB) levels..................................................................... 6

Figure 1.2 The relative energy levels at the semiconductor-electrolyte interface before (left) and after (right) equilibration for a) p-type and b) n-type semiconductors. ECB, EVB, EF and Eredox represent the positions of the conduction band, valence band, Fermi level and redox potential, respectively. EF, equil represents the equilibrium level attained between EF and Eredox after contact. Vbi and W represent the built-in potential and the depletion layer width, respectively. .............................................. 7

Figure 1.3 Schematic illustration of the semiconductor-electrolyte interface and the Helmholtz layer. The Gouy layer has been omitted for simplicity as it can be ignored for the applications discussed in this thesis. ............. 8

Figure 1.4 Diagram showing photon absorption, charge migration and subsequent redox reactions at a) one-step and b) two-step photocatalyst systems. ........................................................................... 13

Figure 1.5 Photoelectrochemical water splitting configurations for a) p-type photocathodes, b) n-type photoanodes and c) tandem photoelectrode systems. ........................................................................ 14

Figure 1.6 a) Schematic depiction of nanoparticle, nanorod and branched nanorod electrodes. b-d) Cross-sectional view of the nanoparticle, nanorod and branched nanorod electrodes. e-g) Top-down view of the nanoparticle, nanorod and branched nanorod electrodes. The insets of c) and d) represent magnified SEM images of their respective electrodes.60 Reprinted with permission from Cho, I. S.; Chen, Z.; Forman, A. J.; Kim, D. R.; Rao, P. M.; Jaramillo, T. F.; Zheng, X. Nano Lett. 2011, 11, 4978. Copyright 2011 American Chemical Society. .................................................................................. 20

Figure 1.7 Orbital configurations of the ground state (S0), the first excited singlet state (S1) and the first exctied triplet state (T1), where the arrows represent the electron spin. .................................................................... 24

Figure 1.8 Jablonski diagram indicating the electronic and vibrational levels of singlet and triplet states, as well as radiative and non-radiative transitions. Excitation processes are represented by straight, solid arrows; radiative recombination processes are represented by straight, dashed arrows; and non-radiative transitions are represented by undulated arrows. .......................................................... 25

Figure 1.9 Chemical structures of a few select conjugated organic polymers .......... 27

Figure 1.10 Chemical structure of two D-A polymers. Donor units are colored red and acceptor units are colored blue. ....................................................... 28

Figure 1.11 A representation of the three possible coupling configurations of 3-hexylthipohene monomers: head-to-head, head-to-tail, and tail-to-tail. ‘Hex’ is used to represent the hexyl functional group (-C6H13). ......... 29

Page 12: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

xii

Figure 1.12 Diagram depicting the bulk heterojunction morphology achieved when a P3HT:PCBM film is cast on an ITO substrate. ........................... 30

Figure 1.13 Energy level diagram of a P3HT photocathode on ITO and H+/H2 redox couple in solution. ......................................................................... 32

Figure 1.14 Organic photoelectrochemical cell with a device architechture of ITO/PEDOT:PSS/P3HT:PCBM/MoS3.114 Reprinted from Queyriaux, N.; Kaeffer, N.; Morozan, A.; Chavarot-Kerlidou, M.; Artero, V. J. Photochem. Photobiol. C Photochem. Rev. 2015, 25, 1, Copyright 2015, with permission from Elevier. ........................................................ 33

Figure 1.15 Diagram of the glass cell used for photoelectrochemical measurements. ...................................................................................... 38

Figure 1.16 Example of a typical electorolysis trace. The three important regions are labelled, where i. represents the dark region, ii. the photocurrent transient region, and iii. the steady-state photocurrent region................. 39

Figure 1.17 a) An idealized example of a typical OCP measurement, where regions i and ii represent the OCP measured under dark and illuminated conditions, respectively. b) A linear sweep voltammogram under chopped illumination, a vertical dashed line has been drawn from the onset of photocurrent to highlight Vonset..................................... 41

Figure 1.18 Diagram showing the scattering geometry of a typical GIWAXS measurement. ........................................................................................ 43

Figure 1.19 Diagram of an idealized P3HT unit cell of where the (100) and (010) orientations are shown in the lamellar and π-stacking directions, respectively. ........................................................................................... 44

Figure 1.20 Outline of this thesis organized in a flow chart. ....................................... 46

Figure 2.1 Electrolysis of P3HT-coated ITO electrodes at -0.24 V vs. SCE (0 V vs. RHE) performed in 0.1 M H2SO4 (black), 0.1 M NaCl (dotted), 0.1 M NaOH (dashed), and 0.5 M Na2SO3 oxygen scavenger (dot-dashed). Light intensity: 100 mW cm-2 (300-700 nm), under chopped illumination cycles of 30 s dark / 1 min light. .................. 51

Figure 2.2 Gas chromatographic analysis of H2 gas in the photoelectrochemistry cell after 202.5 mC of charge over 60 s (dot-dashed) and after passing 32.5 mC of electrolytic charge over 10 s (dashed) at a Pt foil (i.e., conventional electrochemical HER); 6 h photoelectrolysis at a P3HT-ITO electrode (dotted) after passing 56.6 mC of cathodic charge. The solid black signal represents a sample of air (i.e, blank signal). Signals are offset for clarity. ....................................................... 52

Figure 2.3 Photoelectrolysis of P3HT coated ITO electrodes in 0.1 M H2SO4 with different dissolved oxygen concentrations. Oxygen deficient (<0.01 ppm) conditions (red), ambient (7.80 ppm) conditions (green) and oxygen saturated (> 20.0 ppm) conditions (blue). Light intensity: 100 mW cm-2 (300-700 nm). The electrodes were biased at -0.24 V vs. SCE (0 V vs. RHE). Chopped illumination cycles consist of 30 s dark / 1 min light. ............................................................ 53

Figure 3.1 General schemes showing the formation of organic polymer nanoparticles by (a) the mini-emulsion method and (b) the precipitation method. .............................................................................. 57

Page 13: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

xiii

Figure 3.2 Size distribution of P3HT:PCBM nanoparticle diameter determined by (a) dynamic light scattering (DLS) and (b) transmission electron spectroscopy (TEM). For DLS measurements, 1 drop of nanoparticle dispersion was diluted in 2 mL of IPA. TEM samples were prepared by drop casting a diluted nanoparticle dispersion in IPA on a lacey carbon TEM grid. .................................................................................... 64

Figure 3.3 Normalized UV-vis spectra of a P3HT:PCBM nanoparticle film (solid black), a solution cast P3HT:PCBM bulk heterjunction film (solid red), a dispersion of P3HT:PCBM nanoparticles in IPA (dashed black) and a P3HT:PCBM solution in o-DCB (dashed red). nanoparticle and solution cast films were ~90 nm thick. .................................................... 65

Figure 3.4 (a) GIWAXS data obtained for P3HT:PCBM nanoparticle (black) and solution-cast (red) films. The angle of incidence was 0.16°. Inset: Lamellar stacking of P3HT chains in an edge-on fashion along the (100) direction. 2D GIWAXS plots of the P3HT:PCBM nanoparticle and solution-cast electrodes are shown in (b) and (c), respectively. ....... 66

Figure 3.5 Steady-state fluorescence emission spectra of a solution-cast P3HT film (dashed), a solution cast P3HT:PCBM film (red) and a P3HT:PCBM nanoparticle film (black), all on ITO. An excitation wavelength of 515 nm was used. ........................................................... 67

Figure 3.6 SEM images of a P3HT:PCBM nanoparticles on ITO electrodes. Images (a) and (b) depict the nanoparticle films at magnification values of 60k and 250k, respectively. ..................................................... 67

Figure 3.7 Platinum deposition by electrolysis under chopped illumination for P3HT:PCBM nanoparticle (black) and solution-cast (red) films. The potential at the photoelectrodes was held at -0.24 V vs. SCE (0 V vs. RHE). ........................................................................................ 68

Figure 3.8 SEM images, collected using the backscattered electron detector, of Pt deposited after a deposition period of 100 s on (a) a P3HT:PCBM nanoparticle electrode and (b) a solution-cast P3HT:PCBM thin film electrode. ............................................................................................... 69

Figure 3.9 Linear sweep voltammetry (a) and electrolysis (b) under chopped illumination comparing P3HT:PCBM nanoparticle (black) and solution-cast (red) photocathodes in 0.1 M H2SO4. Experiments were performed both with (solid) and without (dashed) Pt nanoparticle catalysts. Linear sweep voltammetry of Pt deposited on bare ITO is shown in (c). Linear sweep voltammograms were measured at a scan rate of 5 mV/s and electrolysis measurements were performed at -0.24 V vs. SCE (0 V vs. RHE). .......................................................... 70

Figure 3.10 Electrolysis under chopped illumination comparing P3HT:PCBM solution-cast photocathodes prepared using Pt deposition periods of 100 s (solid) and 250 s (dashed). Measurements were performed at - 0.24 V vs. SCE (0 V vs. RHE). ............................................................. 71

Figure 4.1 a) OCP vs. light intensity plot for P3HT (circles), 10 wt% PCBM (squares), 25 wt% PCBM (diamonds) and 50 wt% PCBM (triangles) and b) linear sweep voltammetry results for pristine P3HT (black), 10 wt% PCBM (red), 25 wt% PCBM (green) and 50 wt% PCBM (blue). Device architecture for all devices is ITO/P3HT:PCBM.

Page 14: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

xiv

Measurements were carried out in 0.1 M Bu4N PF6 in ACN with 1 mM BZQ. Potentials measured vs. Ag/Ag+ reference (0.01 M AgNO3). For OCP measurements, the points plotted, and error bars represent the average OCP values measured at each light intensity and their standard deviation, respectively. For LSV measurements, the Y-axis is offset for clarity, vertical dashed lines are drawn from the onset potential of their respective measurements. ........................................... 78

Figure 4.2 Energy level diagram of a) P3HT and b) P3HT:PCBM (50 wt% PCBM) photocathodes on ITO, in contact with the BZQ redox couple. ............................................................................................................... 79

Figure 4.3 Figure 3. a) OCP vs. light intensity plot for P3HT (circles), 10 wt% PCBM (squares), 25 wt% PCBM (diamonds) and 50 wt% PCBM (triangles) in P3HT:PCBM and b) linear sweep voltammetry results for pristine P3HT (black), 10 wt% PCBM (red), 25 wt% PCBM (green) and 50 wt% PCBM (blue). Device architecture for all devices is ITO/PEDOT:PSS/P3HT:PCBM. Measurements were carried out in 0.1 M Bu4N PF6 in ACN with 1 mM BZQ. Potentials measured vs. Ag/Ag+ reference (0.01 M AgNO3). For OCP measurements, the points plotted, and error bars represent the average OCP values measured at each light intensity and their standard deviation, respectively. For LSV measurements, the Y-axis is offset for clarity, vertical dashed lines are drawn from the onset potential of their respective measurements. ..................................................................... 81

Figure 4.4 Photoelectrolysis using pristine P3HT (black), 10 wt% PCBM (red), 25 wt% PCBM (green) and 50 wt% PCBM (blue) in P3HT:PCBM films. A bias of -0.8 V (vs. Ag/Ag+) was applied for all electrolysis measurements. All electrodes have an active layer thickness of 220 nm and a device architecture of ITO/PEDOT:PSS/P3HT:PCBM. Measurements were carried out in 0.1 M Bu4N PF6 in ACN with 1 mM BZQ. ...................................................................................................... 83

Figure 4.5 Photoelectrolysis at pristine P3HT (black), 10 wt% PCBM (red), and 25 wt% PCBM (green). A bias of -0.8 V (vs. Ag/Ag+) was applied for all electrolysis measurements. All electrodes have a device architecture of ITO/PEDOT:PSS/P3HT:PCBM. Measurements were carried out in 0.1 M Bu4N PF6 in ACN with 1 mM BZQ. .......................... 84

Figure 5.1 Chemical structures of MDMO-PPV, PBD and their oligo(ethylene glycol)-bearing analogues. ..................................................................... 89

Figure 5.2 Chemical structures of the polymers used to study the effect of different conjugated units223 and fluorinated backbone substituents225 on molecular energy levels. .................................................................... 92

Figure 5.3 Chemical structures of PCDTBT and APFO-Green 1 ............................. 93

Figure 5.4 Schematic illustration of the a) rubbing and b) friction transfer processes used to create thin polymer films with controlled polymer chain orientation.236 Reprinted from Brinkmann, M.; Hartmann, L.; Biniek, L.; Tremel, K.; Kayunkid, N. Macromol. Rapid Commun. 2014, 35, 9, Copyright 2013, with permission from WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim. .............................................................. 95

Page 15: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

xv

Figure 5.5 Scematic illustration of the nanoimpronting method used to create nanostructured polymer films. ................................................................ 96

Figure 5.6 Schematic illustration of the formation of polymer nanostructures using the templating method.242 Copyright 2011, reproduced with permission from the Royal Society of Chemistry. ................................... 97

Page 16: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

xvi

List of Acronyms, Symbols and Abbreviations

A Absorbance

ABPE Applied bias photon-to-current efficiency

APCE Absorbed photon-to-current efficiency

B-NR Branched nanorod

BHJ Bulk heterojunction

CDL Depletion layer capacitance

CH Helmholtz capacitance

Ctot Total capacitance

CE Counter electrode

COP Conjugated organic polymer

CB Conduction band

D Diffusion coefficient

D-A Donor-acceptor

DLS Dynamic light scattering

EF Fermi level

Eredox Redox potential

GC Gas chromatography

GIWAXS Grazing incidence wide angle x-ray scattering

HH Head-to-tail

HER Hydrogen evolution reaction

HT Head-to-tail

IPCE Incident photon-to-current efficiency

ITO Indium-doped tin oxide

HOMO Highest occupied molecular orbital

I Intensity of transmitted light

Io Intensity of incident light

jph Photocurrent density

jsc Short circuit current density

kB Boltzmann constant

LSV Linear sweep voltammetry

LUMO Lowest unoccupied molecular orbital

NA Density of acceptor dopants

Page 17: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

xvii

NP Nanoparticle

NR Nanorod

OCP Open circuit potential

OER Oxygen evolution reaction

OFET Organic field effect transistor

OPEC Organic photoelectrochemical cell

OPV Organic photovoltaic

ORR Oxygen reduction reaction

P Power density

P3HT Poly(3-hexylthiophene)

PCBM [6,6]phenyl-C61-butyric acid methyl ester

PCE Power conversion efficiency

PL Photoluminescence

q Elementary charge

RHE Reversible hydrogen electrode

S0 Singlet ground state

S1 First excited singlet state

SCE Saturated calomel electrode

SEM Scanning electron microscopy

SHE Standard hydrogen electrode

STH Solar-to-hydrogen efficiency

T Temperature

TEM Transmission electron microscopy

TT Tail-to-tail

T1 First excited triplet state

Vbi Built-in potential

Vph Photovoltage

VOC Open circuit voltage

Vonset Onset potential

VB Valence band

W Depletion layer width

WE Working electrode

XRD X-ray diffraction

εo Permittivity of vacuum

Page 18: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

xviii

εr Relative dielectric constant

η Viscosity

ηe-/h+ Absorptance

ηF Faradaic efficiency

ηinterface Interfacial charge transfer efficiency

ηtransport Charge transport efficiency

λ Wavelength

Page 19: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

1

Chapter 1. Introduction

The introduction of this thesis will serve to highlight the importance of renewable

energy in the world today, with an emphasis on solar fuels research in general, and

gradually narrowing the scope to discuss the potential role of organic

photoelectrochemical cells (OPECs). Here, fundamental concepts of semiconductor

physics and photoelectrochemistry will be introduced, along with key concepts and

characteristics pertaining to conjugated organic polymers that are central to the research

undertaken. Various photoelectrochemical technologies and design principles will be

described, along with important characterization techniques and their underlying

concepts. Finally, a thorough literature review and historical overview of organic

photoelectrochemical cells is provided to demonstrate the rapid progress in the field,

establish the current state of the field and provide insight into where future research in the

field may be headed.

1.1. Hydrogen as an Alternative Fuel

A rapidly increasing global population and urbanization of developing countries

has resulted in rising global energy demand, predicted to grow from 18.5 TW in 20171 to

30 TW in 2050.2 A serious concern associated with the increasing energy demand is the

increase in CO2 emissions from the consumption of carbon-based energy sources. The

implementation of renewable energy technologies is, therefore, crucial to limiting the rise

in CO2 emissions, and consequently their contributions to increasing global temperatures.

Current renewable energy sources include wind, hydroelectric, tidal, geothermal, biomass,

nuclear and solar. Of these sources, however, only solar energy has the potential to meet

the global energy demand, as it is estimated that 120,000 TW of power reaches the earth’s

surface every year from the sun.

Although the prospect of meeting the energy demands of the entire planet using

direct solar energy is appealing, the intermittent nature of sunlight and the limited capacity

of the electrical grid infrastructure bring about the need for large-scale energy storage

solutions. While various energy storage technologies exist (e.g., mechanical,

Page 20: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

2

electrochemical, superconducting magnetic, and cryogenic), chemical energy storage

offers the advantage of high energy density and facile transportation for on demand use.3

Hydrogen, specifically, is an attractive fuel as it can be produced directly through the

photoelectrolysis of water, where solar energy is used to drive redox reactions at

semiconductor photoelectrodes immersed in aqueous solutions, resulting in the evolution

of hydrogen and oxygen gas, as shown in Equation 1.

2H2O (l) + hν → 2H2 (g) + O2 (g) (1)

The hydrogen produced can then be stored, transported and used on demand in

fuel cells or combustion engines, while emitting only water vapour as a by-product.

Considering the above, it is understandable that the use of abundant renewable resources

(i.e., water and sunlight) to produce a fuel with no carbon emissions upon consumption is

often referred to as the “Holy Grail” of renewable energy.

Although hydrogen represents an attractive alternative to fossil fuels, the lack of

economically viable methods to produce clean and sustainable hydrogen at large-scales

represents a major challenge towards widespread adoption. Currently, 80-85% of the

hydrogen produced globally is derived via steam methane reforming, while the remainder

is produced via coal gasification and the electrolysis of water.4 The major drawback of

hydrogen production from hydrocarbon feedstocks is the production of CO2 as a by-

product. While downstream CO2 sequestration and/or utilization technologies may provide

short term relief regarding carbon emissions, ultimately, carbon-free hydrogen production

is required to sustain any sort of “hydrogen economy”. Although CO2 is not a direct by-

product of water electrolysis, it is an energy intensive process due to the large operating

voltages (~1.9 V) required to operate electrolyzers at their optimal current densities

(~1 A/cm2). Additionally, the electrode materials employed are generally expensive and

not appropriate for the large-scale production of hydrogen. However, efforts are being

made in the area of water electrolysis to develop new materials and technologies in hopes

of lowering operating voltages and material costs.5

Although great progress towards the ultimate goal of solar hydrogen production

has been made over the years, a number of challenges remain to be addressed.

Photoelectrochemical water splitting represents a complex technology where devices

must be able to efficiently absorb sunlight, possess appropriate thermodynamics for the

Page 21: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

3

desired hydrogen and oxygen evolution reactions to take place, generate sufficient

photovoltage to drive these reactions, and exhibit sufficiently fast kinetics at the

semiconductor-electrolyte interface. Additionally, these devices must exhibit long-term

stability under operating conditions and must not be prohibitively expensive. These

requirements determine whether a material can be considered for use in

photoelectrochemical water splitting technologies and are generally divided into three

broad categories: efficiency, stability and cost. The baseline targets for the economical

production of solar hydrogen established by the US Department of Energy are 10% solar-

to-hydrogen efficiency, an operating lifetime of 5 years and a cost of $2-4 per kilogram of

H2 produced. Although it is not uncommon to find materials that exceed one, or even two,

of the three requirements, a material that meets all three requirements has yet to be

discovered. For example, many metal oxide semiconductors represent inexpensive

materials that produce stable photocurrents for extended periods of time, however they

suffer from poor efficiencies.6,7 On the opposite end of the spectrum, multijunction

photoelectrodes based on III-V compound semiconductors are capable of reaching solar-

to-hydrogen efficiencies over 10%, but suffer from long-term stability issues and use

expensive materials.8,9

In addition to the scientific challenges, the future of solar driven hydrogen

production is further complicated by technical obstacles such as storage, infrastructure,

and, depending on the type of reactor used, the separation of hydrogen and oxygen gases

from the potentially explosive gas mixture that is created. Storing the generated hydrogen

is an issue because, although the gravimetric energy density of hydrogen is very high, its

volumetric energy density is rather low. To circumvent the volumetric energy density

issues, various storage technologies are being explored. These technologies range from

simple solutions such as high pressure storage containers10 and liquid cryo-storage,11 to

more sophisticated approaches like high surface area metal organic frameworks that are

capable of storing hydrogen through physisorption.12 Additionally, the lack of infrastructure

is often cited as a major technical and economical barrier to the widespread adoption of

hydrogen as an energy source. Currently, there is little infrastructure in place to support a

hydrogen economy and would require significant capital investment to construct the large-

scale hydrogen production, storage, long-distance transmission, local distribution and

dispensing systems needed. Although the seemingly insurmountable task of creating the

Page 22: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

4

infrastructure for a new primary energy source from the ground up looms large, several

reports and proposals exist that argue its viability.13,14

1.2. Photoelectrolysis of Water

1.2.1. Fundamentals of Semiconductor Photoelectrochemistry

Semiconductors represent the photoabsorbing component of

photoelectrochemical devices and are responsible for the generation of charge carriers

that will ultimately be used in the electrochemical redox reactions that take place at the

electrolyte interface. The understanding of the complex interactions of semiconductors

with incoming solar radiation and the liquid electrolyte is crucial to furthering the area of

photoelectrolysis. This section outlines the fundamental properties of semiconductors, as

well as the important processes that occur when semiconductors interact with light and at

the semiconductor-electrolyte interface.

Semiconductors are a class of materials whose conductive properties can be

manipulated so as to behave as either insulators or conductors under certain conditions.

The conductivity of a semiconductor is best described by the band model. As

semiconductors are made up of many covalently bonded atoms, the energy levels become

so dense that broad energy bands are formed as opposed to the discrete energy levels

observed for single atoms or molecular systems.15 When an electron resides in the lower

energy band, referred to as the valence band, no conduction may occur and the

semiconductor acts as an insulator,* however, when an electron gains enough energy to

populate the higher energy band, known as the conduction band, it is now free to

participate in conduction. The excitation of an electron to the conduction band leaves

behind a vacancy in the valence band, providing an opportunity for other valence band

electrons to move and occupy the newly formed vacancy. For simplicity, this vacancy is

often thought of as a particle with a positive charge that moves through the valence band,

known as a hole. The minimum energy needed to promote an electron from the valence

band to the conduction band is referred to as the band gap energy.

* This assumes that the valence band is full.

Page 23: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

5

Exposing a semiconductor to solar radiation is a common way to excite a

semiconductor electron from the valence band to the conduction band, where

electromagnetic radiation with energy equal to or greater than the semiconductor band

gap will be absorbed. The formation of photogenerated electron-hole pairs, often referred

to as excitons, is the fundamental process of solar energy production and conversion.

These photoexcited electrons, however, exist in a meta-stable state where they will

eventually recombine with a hole. This charge carrier recombination represents a major

loss mechanism in photoelectrochemical applications.16

Also central to semiconductor photoelectrochemistry is the concept of the Fermi

level, which represents a hypothetical energy level where, at absolute zero, all energy

states below the Fermi level are occupied and all energy states above the Fermi level are

unoccupied. At finite temperatures it is convenient to describe the Fermi level as an energy

level that, at thermodynamic equilibrium, has a 50% chance of being occupied by an

electron at a given temperature. For intrinsic (i.e., undoped) semiconductors, the Fermi-

level lies in the middle of the band gap. For extrinsic (i.e., doped) semiconductors, charge

carrier densities can be manipulated by introducing either donor or acceptor atoms. Take,

for example, a defect-free crystal lattice made up of silicon atoms, a group IV element with

four valence electrons. Each silicon atom will share one valence electron with each of its

four neighbouring silicon atoms, who also share one of their electrons, forming a

covalently bonded network where all available valence electrons are participating in

bonding, i.e., no free electrons are available to participate in conduction. By replacing one

silicon atom with a group V element (e.g., phosphorous) that contains five valence

electrons, the group V atom will share one valence electron with each of its four

neighbouring silicon atoms and one electron remains “free” to participate in conduction.

As a result, the number of negative charge carriers (electrons), is greater than the number

of positive charge carriers (holes) and is referred to as n-type doping. The opposite is true

if a silicon atom is replaced by a group III element (e.g., boron) that contains three valence

electrons. The lack of a shared electron to one of the surrounding silicon atoms will result

in a vacancy (hole) that is free to participate in conduction. In this case, holes will

outnumber electrons and is referred to as p-type doping. A major consequence of doping

is the shift in Fermi level within the semiconductor, as n-type doping results in the Fermi

level shifting upwards towards the bottom of the conduction band and p-type doping will

shift the Fermi level downwards towards the top of the valence band. As the relative ratio

Page 24: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

6

of positive and negative charge carriers can be controlled through doping, the concept of

majority and minority carriers can be applied. In p-type materials, holes act as majority

carriers while electrons act as minority carriers; the opposite is true for n-type materials,

i.e., electrons are the majority carriers and holes are the minority carriers.

Figure 1.1 Relative Fermi level (EF) positions of a) intrinsic, b) p-type and c) n-type semiconductors. Also shown are the valence band (VB) and conduction band (CB) levels.

The semiconductor-electrolyte interface plays an important role in semiconductor

photoelectrochemistry and the ability of a semiconductor to perform a

photoelectrochemical reaction is highly dependent on both the interfacial thermodynamics

and kinetics. The relative positions of the semiconductor Fermi level and the

electrochemical potential of the redox couple in solution are crucial to device performance,

as are the electron transfer kinetics across the semiconductor-electrolyte interface. For a

semiconductor in contact with electrolyte, the electrochemical potential of the redox couple

must lie between the valence and conduction band edges (i.e., within the band gap) of the

semiconductor and above (below) the Fermi level for p-type (n-type) semiconductors.

The difference in energy between the semiconductor Fermi level and

electrochemical potential of the redox couple leads to spontaneous charge transfer at the

interface, where majority carriers will be transferred from the semiconductor to the

electrolyte until an equilibrium is reached. As a result, there will be a region that extends

from the semiconductor surface into the bulk that is depleted of majority carriers, aptly

referred to as the depletion layer. Consequently, an internal electric field that extends the

width of the depletion layer is developed. The band bending diagrams depicting these

phenomena are shown below in Figure 1.2.

The depletion layer plays an important role in semiconductor

photoelectrochemistry as any photogenerated charges that originate within the depletion

layer can be effectively separated by the internal electric field, which will influence the

Page 25: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

7

overall performance of the photoelectrochemical device. The depletion layer is determined

by several semiconductor properties and can be calculated by Equation 2, where W is the

Figure 1.2 The relative energy levels at the semiconductor-electrolyte interface before (left) and after (right) equilibration for a) p-type and b) n-type semiconductors. ECB, EVB, EF and Eredox represent the positions of the conduction band, valence band, Fermi level and redox potential, respectively. EF, equil represents the equilibrium level attained between EF and Eredox after contact. Vbi and W represent the built-in potential and the depletion layer width, respectively.

depletion layer width, εr is the relative dielectric constant of the semiconductor, εo is the

permittivity of vacuum, Vbi is the built-in potential, q is the elementary charge and NA is the

doping density.

W = √2εrεoVbi

qNA (2)

The transfer of majority carriers across the semiconductor-electrolyte interface

during the equilibration process results in a net charge at the semiconductor surface. For

p-type semiconductors, the transfer of positively charged majority carriers (i.e., holes)

across the interface leads to a negatively charged surface. To counterbalance this surface

charge, a redistribution of ions occurs within the electrolyte near the semiconductor

surface, forming two oppositely charged layers at the semiconductor-electrolyte interface.

The resulting formation is termed the electrochemical double layers and is analogous to

an electrical capacitor consisting of two oppositely charged plates separated by some

distance. As such, a potential drop exists across the electrochemical double layer. The

Page 26: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

8

electrolyte component of the double layer consists of a compact layer of solvated ions at

the immediate semiconductor-electrolyte boundary, known as the Helmholtz layer, and a

diffuse layer that extends into the bulk of the electrolyte, known as the Gouy layer. When

the concentration of the electrolyte is sufficiently high (≥0.01 M), the contribution from the

diffuse layer can be neglected. As a result, the potential drop in solution will be confined

to the Helmholtz layer.

Figure 1.3 Schematic illustration of the semiconductor-electrolyte interface and the Helmholtz layer. The Gouy layer has been omitted for simplicity as it can be ignored for the applications discussed in this thesis.

The formation of the electrochemical double layer becomes important when

considering the application of an external bias potential to the photoelectrochemical cell,

which is common in solar water splitting applications. When an external bias is applied,

the potential difference will be distributed over both the depletion layer within the

semiconductor and the Helmholtz layer in solution. The total capacitance can then be

modeled as two capacitors in series:

1

𝐶tot=

1

𝐶DL+

1

𝐶H (3)

where Ctot is the total capacitance, CDL is the depletion layer capacitance and CH is the

Helmholtz capacitance. Given that the charge, Q, in both layers are equal and the

relationship between capacitance, charge and potential is given as:

𝐶 =𝑄

𝑉 (4)

Page 27: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

9

the potential distribution across the both the depletion (ΔVDL) and the Helmholtz (ΔVH)

layers can be described by:

∆𝑉DL

∆𝑉H=

𝐶H

𝐶DL (5)

Because of the inverse relationship between capacitance and the distance

between two oppositely charged plates in the classical capacitor model, or the thickness

of each layer for the electrochemical double layer, the Helmholtz capacitance will be much

larger than the depletion layer capacitance. The typical width of a Helmholtz layer is

~2-5 Å, whereas depletion layer widths typically range from 5-500 nm.2 As CH>>CDL, any

applied potential will fall only across the depletion layer of the semiconductor. The

application of a bias potential to a semiconductor working electrode will, therefore,

increase both the magnitude of band bending and depletion layer width within the

semiconductor. This is an important property of the semiconductor-electrolyte interface

that becomes useful for the characterization of photoelectrochemical cells, as will be

further discussed in Section 1.5.2.

1.2.2. The Hydrogen Evolution and Oxygen Evolution Reactions

Solar-driven water splitting takes advantage of the photogenerated charges

produced within the semiconductor layer to perform complementary redox reactions at the

electrolyte interface, where the reduction and oxidation reactions are responsible for the

evolution of hydrogen and oxygen gases, respectively. Under acidic conditions, the

reduction and oxidation reactions can be written as

4 H+ + 4e− → 2 H2 𝐸° = 0.00 V vs. SHE

2 H2O → O2 + 4H+ + 4e− 𝐸° = 1.23 V vs. SHE

2 H2O → 2H2 + O2 𝐸cell° = −1.23 V vs. SHE

In theory, the thermodynamic driving force required for water splitting to occur is

1.23 V. This suggests that a single semiconductor material that possesses a band gap

>1.23 eV and whose conduction band and valence band energies straddle the redox

potentials of the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER),

respectively, can be used to successfully drive both the HER and OER at the

semiconductor-electrolyte interface. In practice, however, electron transfer at the

Page 28: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

10

semiconductor-electrolyte interface produces losses due to the concentration and kinetic

overpotentials required to perform the HER and OER. As a result, the energy needed for

the successful photoelectrolysis of water is often cited as 1.6-2.4 V.17

The large overpotentials required for photoelectrolysis effectively limits the

materials that are capable of “overall water splitting” to semiconductors with large band

gaps. Because of their large band gaps, most of these materials require photons with

sub-400 nm wavelengths (i.e., ultraviolet radiation) to generate photoexcited electrons.

The dependence on ultraviolet radiation reduces the maximum achievable efficiencies for

these materials, as it represents only a fraction of the solar spectrum. Even if all UV light

up to 400 nm was utilized without any losses, the resulting solar-to-hydrogen efficiency

would only be 2%.18 As photons with 400-800 nm wavelengths (i.e., visible light) represent

nearly half of the incoming solar radiation, the efficient use of photons in the visible region

is required for highly efficient photoelectrolysis. For example, if all photons up to 600 nm

were utilized, an efficiency of 16% could be realized. Further increasing the wavelength of

utilized photons to 800 nm would result in an efficiency of 32%.18

The limitations brought about by the overpotentials required for the

photoelectrolysis of water have motivated researchers to develop photoelectrochemical

water splitting systems with configurations that take advantage of semiconductor

properties targeted towards either the HER or OER, multijunction configurations that use

semiconductors with complementary band gaps to maximize absorption across the solar

spectrum and the implementation of cocatalysts to improve charge transfer kinetics. The

following section will explore the development of device architectures and design

principles commonly used in photoelectrolysis.

1.2.3. Device Architecture and Design Principles

As the field of solar water splitting matures, device architectures and design

principles are continually improving to push performance higher. Nevertheless, the

majority of water splitting systems can be classified into two major categories, namely

photocatalytic and photoelectrochemical systems.

Page 29: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

11

Semiconductor Photocatalyst Systems

Photocatalytic systems consist of semiconductor particles suspended in an

electrolyte and are capable of performing photocatalytic redox reactions in the absence of

any externally applied bias. The simple design allows for relatively inexpensive and facile

scale-up of reactor technologies.19 Two strategies exist for achieving photocatalytic water

splitting using heterogenous semiconductor catalysts; the first method is a one-step

system, where a single photoabsorber possessing a conduction band edge more negative

than the hydrogen evolution reaction (HER) redox potential and valence band edge more

positive than the oxygen evolution reaction (OER) redox potential is employed to

simultaneously generate hydrogen and oxygen gas. An efficient one-step photocatalytic

system must satisfy the following requirements: (i) have a low band gap for the efficient

use of visible light, (ii) be resistant to corrosion under photoirradiation and (iii) possess

suitable band edge energetics for both hydrogen and oxygen evolution.

While metal oxides are generally stable against photocorrosion, it is difficult to

develop a metal oxide photocatalyst that has both sufficiently negative conduction band

to perform the HER and a sufficiently narrow band gap to ensure good utilization of the

solar spectrum due to the highly positive valence band formed by the O 2p orbital.20

Alternatively, non-oxide materials, such as sulfide and nitride semiconductors, possess

appropriate band levels for overall water splitting using visible light, however, these

materials are generally plagued by stability issues as the self-oxidation process competes

with oxygen evolution. This has been shown for CdS photocatalysts, where the S2- anions

are preferentially oxidized over water molecules in the absence of a sacrificial electron

donor.21 One class of semiconductors that has been shown to be an effective one-step

water splitting system are metal oxynitrides. When (Ga1-xZnx)(N1-xOx) is modified with a

hydrogen evolution cocatalyst, Rh2-yCryO3, stoichiometric amounts of H2 and O2 are

evolved, and an apparent quantum yield of 5.9% at 420-440 nm has been achieved.22

As it is difficult to find a material that meets all the above requirements, a more

commonly applied strategy is the two-step, or z-scheme, water splitting system. The two-

step method allows for the use of two separate semiconductor materials, each optimized

for either the water reduction or oxidation processes and couples them with a redox shuttle

that will react with the photogenerated holes produced at the HER photocatalyst and

photogenerated electrons produced at the OER photocatalyst. The two-step system also

Page 30: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

12

allows for the product gases to be generated separately, which is not possible for a one-

step system. Although the two-step system has significant advantages over the one-step

system, it requires twice as many photons to achieve water splitting and faces challenges

with regards to the backwards reaction of the redox shuttle at the photocatalyst materials.23

The two-step water splitting scheme was first conceptualized in 1979 by Allen J.

Bard,24 but the successful stoichiometric evolution of H2 and O2 in a two-step

photocatalytic system was not demonstrated until 2001.25 This was accomplished using a

Pt-loaded anatase TiO2 photocatalyst to perform the HER, bare rutile TiO2 photocatalyst

to perform the OER and an iodate/iodide (IO3-/I-) redox mediator. Because of the large

bandgap of TiO2, this system is, however, only capable of operating under UV irradiation.

A visible light driven, two-step photocatalyst system was developed later that same year,

employing a Pt-loaded SrTiO3 photocatalyst doped with Cr and Ta as the hydrogen

evolution photocatalyst, a Pt-loaded WO3 oxygen evolution photocatalyst, and an

iodate/iodide (IO3-/I-) redox mediator.26 Although the quantum efficiency for this system

was rather low (0.1% at 420.7 nm), it represented an important milestone in visible light

driven photocatalytic water splitting. More recently, La- and Rh-codoped SrTiO3

(SrTiO3:La,Rh) and Mo-doped BiVO4 (BiVO4:Mo) photocatalysts were embedded onto a

gold layer and modified with a Ru species co-catalyst.27 The underlying gold layer allows

for direct charge transfer between the SrTiO3:La,Rh hydrogen evolution photocatalyst and

the BiVO4:Mo oxygen evolution photocatalyst, effectively removing the need for a redox

mediator in solution and, hence, any competing backwards reactions caused by the redox

mediator. Under visible light irradiation, the Ru-modified SrTiO3:La,Rh/Au/BiVO4:Mo sheet

was able to split pure water, evolving stoichiometric amounts of H2 and O2 with a solar-to-

hydrogen energy conversion efficiency of 1.1% and apparent quantum yield of 33% at

419 nm.

Although photocatalytic systems represent a viable water splitting system, the

remainder of this thesis will focus on photoelectrochemical cells. Comprehensive reviews

of heterogenous photocatalysts can be found elsewhere.18,28,29

Page 31: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

13

Figure 1.4 Diagram showing photon absorption, charge migration and subsequent redox reactions at a) one-step and b) two-step photocatalyst systems.

Photoelectrochemical Cells

Photoelectrochemical systems consist of either a two- or three-electrode setup,

where the electrodes are connected to a potentiostat, allowing for the control over

variables such as applied potential or current. The minimum requirements for a

photoelectrochemical cell are a semiconductor working electrode, where photogenerated

charges are used to perform a redox reaction at the semiconductor-electrolyte interface,

and a counter electrode, at which the complementary redox reaction occurs.

Photoelectrochemical systems can be further broken down into photoanodes, where a

n-type semiconductor is used as the working electrode to perform photooxidation

reactions; photocathodes, where a p-type semiconductor is used as the working electrode

to perform photoreduction reactions; and tandem cells, where a n-type photoanode and

p-type photocathode are simultaneously employed to perform the complementary

photooxidation and photoreduction reactions without any external bias.

While single photoelectrode systems allow for the use of low band gap materials

with properties tailored for either the reduction or oxidation of water, they generally require

an external bias to generate a potential large enough to split water. The use of tandem

photoelectrochemical cells offers the advantage of exploring various combinations of

semiconductors with complementary absorption and stability characteristics, leading to

systems capable of achieving higher photovoltages and overall water splitting in the

absence of an external bias.17

Page 32: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

14

Since the first report of photoelectrochemical water splitting, which employed a

TiO2 photoanode and a Pt counter electrode,30 a multitude of semiconductor materials

have been developed and used as photoelectrodes for solar water splitting applications.

A few of the most promising materials are highlighted below.

Figure 1.5 Photoelectrochemical water splitting configurations for a) p-type photocathodes, b) n-type photoanodes and c) tandem photoelectrode systems.

Photoanode Materials

TiO2-based photoanodes have been extensively studied since the inaugural work

of Fujishima and Honda in 1972,31 as TiO2 is an earth-abundant, non-toxic material that

exhibits good stability. Due to its large band gap (3.2 eV for anatase and 3.0 eV for rutile

TiO2), however, the maximum theoretical solar-to-hydrogen (STH) efficiency for these

materials is very low (1.3% for anatase and 2.2% for rutile TiO2).32 As a result, various

metal oxide semiconductors have been investigated over the years in an attempt to

increase the utilization of visible light (i.e., decrease the band gap of the semiconductor)

while maintaining the desirable properties associated with TiO2, such as good stability,

abundance and low-toxicity. A promising metal oxide that has emerged is α-Fe2O3

(hematite).

Hematite is a highly abundant material that exhibits good stability and low toxicity.

Additionally, it has a favorable band gap, reported between 1.9 and 2.2 eV. The reduction

in band gap, compared to TiO2, increases the maximum theoretical STH efficiency of

hematite to 12.9% for a band gap value of 2.2 eV.32 Despite these promising properties,

hematite also possess significant drawbacks, including: short diffusion lengths (2-4 nm)

and short charge carrier lifetimes (on the order of picoseconds), leading to large

recombination rates; relatively low absorption coefficient (~103 cm-1), resulting in the need

for relatively thick films (400-500 nm) to ensure efficient photon collection; and poor water

Page 33: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

15

oxidation kinetics.33 To overcome these limitations, strategies including nanostructuring

and cocatalyst modification have been applied to hematite photoanodes.† For example,

photoanodes consisting of hematite nanosheets modified with Ag and cobalt phosphate

nanoparticle cocatalysts have reportedly achieved a photocurrent density of 4.7 mA/cm2

at 1.23 V vs. RHE and a STH efficiency of 0.55%.34

In addition to metal oxides, n-type group IV semiconductors (e.g., n-Si), group III-V

compound semiconductors (e.g., n-GaAs and n-InP) and group II-VI compound

semiconductors (e.g., n-CdTe) have been the subject of investigation for

photoelectrochemical water splitting applications. In general, these materials possess

narrow band gaps (e.g., 1.1 eV for Si, 1.42 eV for GaAs and 1.5 eV for CdTe) and high

charge carrier mobilities, allowing them to achieve large current densities and high STH

efficiencies. Unfortunately, these materials are also very susceptible to photocorrosion or

photopassivation, making them highly unstable under operating conditions.35,36 Efforts in

this area have, therefore, been focused on improving the stability of these materials. The

use of a thin TiO2 layer grown by atomic layer deposition has been shown to prevent the

corrosion of Si, GaAs and GaP photoanodes.37 When modified with a NiO cocatalyst, Si

photoanodes exhibited photocurrent densities >30 mA/cm2 at 1.17 V vs. RHE for over 100

hours of continuous oxygen evolution, while the analogous GaAs and GaP photoanodes

exhibited operational lifetimes of over 25 hours. In a separate study, CdTe photoanodes

that had a TiO2 protecting layer decorated with NiO cocatalyst exhibited a steady

photocurrent density of 21.5 mA/cm2 at 2.07 V over four days of operation.38

Photocathode Materials

Cu2O is an attractive material for use as a photocathode as it is abundant, has low

toxicity and possesses a band gap of 2 eV, which translates to a theoretical maximum

STH efficiency of 18%.39 In contrast to metal oxides used for photoanodes, Cu2O suffers

from poor stability and is susceptible to photocorrosion. As the redox potentials of the

reduction and oxidation of monovalent copper oxide lie within the band gap, parasitic

self-reduction and self-oxidation pathways to Cu and CuO, respectively, are kinetically

and thermodynamically possible.40 The incorporation of protective layers have been

shown to improve stability, for example, an aluminium-doped zinc oxide (AZO) layer and

† An in-depth discussion regarding cocatalyst modification and nanostructuring is presented in the proceeding sections: Cocatalysts and Nanostructures

Page 34: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

16

TiO2 layer were sequentially deposited on top of a Cu2O photocathode by atomic layer

deposition, followed by the photoelectrochemical deposition of a MoS2+x cocatalyst.41 The

n-type AZO creates a p-n junction with the p-type Cu2O, promoting efficient transfer of

photogenerated electrons from Cu2O into the AZO layer and the TiO2 layer protects the

buried Cu2O-AZO junction. When used in conjunction with a MoS2+x cocatalyst, the Cu2O-

based photocathode exhibited stable photocurrents of 5.7 mA/cm2 at 0 V vs. RHE over a

continuous operating period of 10 hours.

More recently, I-III-VI2 chalcopyrites, a class of p-type semiconductors where I =

Cu, Ag; III = Ga, In; VI = S, Se, have emerged as promising photocathode materials. These

materials possess large minority-carrier diffusion lengths (up to a few micrometers),

leading to enhanced charge separation; high absorption coefficients (~105 cm-1), allowing

for efficient photon collection using thin films; and tunable band gap energies (1.0-2.4 eV),

achieved by adjusting the alloy composition.42 Photocurrent densities as high as

16 mA/cm2 at 0 V vs. RHE were achieved using a Cu(In,Ga)(Se,S)2 photocathode

passivated with a ZnS layer, and further modified with a Pt hydrogen evolution

cocatalyst.43 The stability of these materials has also been demonstrated using a CuGaSe2

photocathode modified with a thin CdS layer and Pt cocatalyst.44 Here, stable

photocurrents of ~3.5 mA/cm2 at 0 V vs. RHE were measured under constant operation

for more than 10 days.

Finally, p-type group IV semiconductors (e.g., p-Si) and group III-V compound

semiconductors (e.g., p-GaN, p-InP and p-GaAs) have been studied as photocathodes for

the production. Similar to their n-type analogues used as photoanodes, these materials

exhibit large photocurrents and STH efficiencies, but suffer from poor stability under

operating conditions. Examples of these materials used for photoelectrochemical

hydrogen evolution have been published for decades, but more recently, the incorporation

of passivation layers and nanostructured architectures have been reported as strategies

to improve device stability and performance.45–49

Tandem Cell Configurations

Tandem photoelectrochemical cells incorporate both a n-type photoanode and

p-type photocathode to create a two-absorber system capable of unassisted, solar-driven

water splitting. Semiconductors with complementary spectral absorption profiles are used

to ensure efficient utilization of incoming solar radiation. For example, a wider band gap

Page 35: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

17

semiconductor would be used as the front photoelectrode, absorbing the high energy

photons while the lower energy photons would be transmitted to, and subsequently

absorbed by, the narrow band gap semiconductor used to form the back photoelectrode.

By matching the band edge positions of the photoanode and photocathode materials, the

combined photovoltages produced at each photoelectrode is sufficient to overcome the

thermodynamic threshold to perform the hydrogen and oxygen evolution reactions and

their associated overpotentials, removing the need for any external bias.

The combination of efficient photon utilization and large achievable photovoltages

allows for tandem cells to be highly efficient. Calculations done by the Lewis group have

shown that the theoretical maximum achievable STH efficiency for a tandem

photoelectrochemical cell with an optimal band gap combination is 29.7%.50 This model

assumes optimal band gaps of 1.60 eV and 0.95 eV for the front and back absorbers,

respectively, a fill factor of 0.85, a solution resistance of 5 Ω/cm2 and the use of low

overpotential Pt and RuO2 coctalysts. Using the same assumptions, theoretical STH

efficiencies greater than 25% can be achieved by pairing semiconductor materials with

band gaps between 1.6-1.8 eV with Si, which has a band gap of 1.1 eV.

Practically, however, measured STH efficiencies have been much lower. The

current benchmark for unassisted solar water system using a tandem structure is a 14%

solar-to-hydrogen efficiency produced by a III-V photovoltaic tandem absorber.51 Here, the

front layer consists of GaInP (Eg=1.78 eV) with a modified AlInP overlayer and Rh

cocatalyst, while the back layer consists of a GaInAs (Eg=1.26 eV) photoabsorber, where

the photogenerated minority carriers are shuttled to a separate RuO2 electrode via

external circuit, where they are used to perform the oxygen evolution reaction. Although

this configuration exhibits high performance, the materials and preparation methods are

prohibitive to large scale applicability. Ultimately, cheap and abundant materials will be

necessary to meet the technical requirements for large-scale hydrogen production.

In an effort to develop tandem photoelectrochemical cells from earth abundant

materials, Jang et al. combined a hematite photoanode modified with a NiFeOx oxygen

evolution cocatalyst and a Si photocathode modified with a TiO2/Pt overlayer in a tandem

configuration.52 This configuration was able to achieve unassisted solar water splitting with

a STH efficiency of 0.91%, however the authors noted that the photocurrents achieved at

Page 36: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

18

the back Si photocathode were reduced five-fold (compared to photocurrents measured

as a standalone photocathode) due to the initial absorption of hematite.

To rectify the issue of parasitic photon absorption by the front absorber in tandem

cells, the photoelectrodes can be placed in a parallel (i.e., side-by-side) configuration. Ding

et al. developed a tandem photoelectrochemical cell based on a Ni-modifed Si solar cell

as the photocathode and a Mo-doped BiVO4 photoanode decorated with a FeOOH oxygen

evolution cocatalyst, where a higher STH efficiency was achieved when the

photoelectrodes were employed in a side-by-side configuration (2.21%) over a stacked

configuration (0.77%).53 Although the stacked photoelectrode configuration allows for

more efficient utilization of incoming solar irradiation and wireless configurations that are

compatible with convenient assembly, the increased efficiencies of the side-to-side

configuration cannot be overlooked.

Strategies to Increase Performance

As evidenced by the previous sections, the efficient production of hydrogen and

oxygen rarely occurs at bare semiconductor electrodes and several strategies have been

developed to improve the efficiency of these photoelectrochemical systems. This section

will cover two general strategies to improve water splitting systems: (i) using HER/OER

cocatalysts to promote the desired water splitting reactions and (ii) creating

nanostructured photoelectrodes to take advantage of specific size effects at the

nanoscale.

Cocatalysts

The generation of intermediate species during the hydrogen and oxygen evolution

reactions, coupled with the multielectron nature of these reactions leads to kinetic barriers

that must be overcome in addition to the 1.23 V thermodynamic requirement. In most

cases, the formation of intermediate species at bare semiconductor surfaces presents a

large energy barrier, and thus large overpotentials are required to drive the HER/OER. By

loading a cocatalyst onto the semiconductor surface, the overpotentials required to

perform the HER/OER can be reduced drastically.

While noble metal catalysts are generally thought of as the best performing

catalysts, e.g., IrO2 and RuO2 as oxygen evolution catalysts and Pt as a hydrogen

evolution catalyst, their high cost and low abundance make them unsuitable for large scale

Page 37: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

19

water splitting applications. Similar to the semiconductor candidates for solar-driven water

splitting, the cocatalysts employed should be inexpensive and readily available. For this

reason, non-precious metal oxygen and hydrogen evolution catalysts have been the

subject of many investigations. For example, the Jaramillo group has compared the

overpotentials needed to achieve a current density of 10 mA/cm2, the estimated current

density needed to reach 10% STH efficiency, for twenty-six oxygen evolution catalysts

and eighteen hydrogen evolution catalysts.54 From this benchmarking study, a few

promising non-precious OER catalysts were identified. While twelve non-precious oxygen

evolution catalysts were able to achieve current densities of 10 mA/cm2 at overpotentials

between 0.35-0.45 V (compared to 0.29 V for Ru), only six showed promising long-term

stability where the overpotential required to maintain the current density of 10 mA/cm2

remained constant, or even decreased slightly in some cases, after twenty-four hours. The

six promising candidates were Co/P, CoFe, NiCo, NiFe, NiMoFe and NiZn.

In the same study, two hydrogen evolution catalysts, NiMo and NiMoCo, were

shown to exhibit comparable overpotentials to polished polycrystalline Pt and platinized

Pt electrodes. The overpotentials associated with both Pt electrodes is 0.04 V, while NiMo

and NiMoCo exhibit overpotentials of 0.045 and 0.05 V, respectively. CoMo also exhibited

a relatively low overpotential of 0.1 V. Moreover, the overpotentials required remained

stable over twenty-four hours for all three of these catalysts (i.e., NiMo, NiMoCo and

CoMo).

The cocatalysts identified here represent only a small subset of materials that have

been developed as oxygen and hydrogen evolution catalysts. Many other OER cocatalysts

exist, such as perovskites, spinels, metal hydroxide and oxyhydroxides, metal

chalcogenides and metal pnictides, and have been summarized in a recent review by

Suen et al.55 Similarly, other non-precious HER cocatalysts such as, metal sulfides,

selenides, carbides, nitrides, and phosphides are reviewed by Zou and Zang.56

Nanostructures

As the area of nanoscience has experienced rapid advancement, nanostructuring

of existing semiconductor materials has become a popular method of improving water

splitting devices and offers distinct advantages over bulk semiconductor properties. These

advantages include high aspect ratios, leading to shorter path lengths that photogenerated

charge carriers must travel before reaching the semiconductor-electrolyte interface;

Page 38: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

20

improved light distribution due to the scattering of light at nanoscale features; tunability of

energy levels due to quantum confinement effects; and enhanced charge transfer across

the semiconductor-electrolyte interface due to increased surface area. A multitude of

nanostructured photoelectrodes exist, including nanoparticles,57 nanorods,58 nanotubes,59

and branched nanorods.60 Although the performance of a material for water splitting

applications will ultimately depend on its intrinsic properties, nanostructuring has been

beneficial to the development of improved photoelectrode materials.

Figure 1.6 a) Schematic depiction of nanoparticle, nanorod and branched nanorod electrodes. b-d) Cross-sectional view of the nanoparticle, nanorod and branched nanorod electrodes. e-g) Top-down view of the nanoparticle, nanorod and branched nanorod electrodes. The insets of c) and d) represent magnified SEM images of their respective electrodes.60 Reprinted with permission from Cho, I. S.; Chen, Z.; Forman, A. J.; Kim, D. R.; Rao, P. M.; Jaramillo, T. F.; Zheng, X. Nano Lett. 2011, 11, 4978. Copyright 2011 American Chemical Society.

The ability of nanoparticle (NP), nanorod (NR) and branched nanorod (B-NR) TiO2

photoanodes to perform the oxygen evolution reaction was compared by Cho et al, where

it was shown that the branched nanorods exhibited the highest performance.60 The

photocurrent densities measured for the NP, NR and B-NR photoanodes were 0.01, 0.31

Page 39: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

21

and 0.83 mA/cm2, respectively, at 0.8 V vs. RHE. Additionally, the incident photon to

current efficiencies of the NP, NR and B-NR photoanodes were 6% at 370 nm, 42% at

380 nm and, 67% at 380 nm respectively. The improved efficiency at the branched

nanorods is attributed to the improved charge carrier separation and collection due to the

small diameters of the branches, leading to a reduced pathlength for photogenerated

charges to reach the semiconductor-electrolyte interface. Moreover, the authors report a

4-fold increase in surface area for the branched nanorods over bare nanorods, leading to

enhanced hole transfer at the semiconductor-electrolyte interface.

Branched nanorods offer better charge transport properties over nanoparticles, as

it has been reported that electron mobilities in nanoparticle films are hindered by trap

states that exist at grain boundaries between particles.61 Additionally, branched nanorods

exhibit improved absorption over nanoparticles as the absorption properties of

nanoparticle films are often limited by reflection losses at the surface.62 Although the

charge transport and absorption properties of nanorods and branched nanorods are

similar, the increased surface area of branched nanorods leads to increased charge

transfer at the semiconductor-electrolyte interface. For these reasons, branched nanorods

represent a promising nanostructure geometry that can be applied to other

semiconductors that exhibit better water splitting capabilities than TiO2.

1.3. Conjugated Organic Polymers

1.3.1. Overview

Conjugated organic polymers (COPs) are macromolecules made up entirely of

main-group elements, such as hydrogen, carbon, nitrogen, oxygen and sulfur.

Conjugation in organic molecules refers to the overlap of adjacent atomic p-orbitals,

forming molecular π-orbitals. This orbital overlap leads to the delocalization of π-electrons

within the molecule. In COPs, conjugation is brought about by alternating single and

double bonds that make up the polymer backbone. If heteroatoms are present, the lone

pair electrons located within the pz orbital of the heteroatom will also contribute to the

conjugation of the polymer backbone. As a result, COPs exhibit semiconducting

properties, lending them to a larger classification of materials referred to as organic

semiconductors.

Page 40: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

22

Organic semiconductors look to combine the optoelectronic properties of

traditional inorganic semiconducting materials with the desirable properties associated

with organic molecules, namely mechanical flexibility; synthetic tunability, allowing the

optoelectronic properties of a material to be tailored through synthetic organic chemistry;

and solubility, allowing these materials to be compatible with large scale, roll-to-roll,

solution-processing techniques (e.g., slot die coating, spray coating, ink-jet printing). Along

with COPs, two other main subcategories of organic semiconductors exist: (i) amorphous

molecular films, formed by the deposition of organic small molecules and (ii) molecular

crystals, where conjugated molecules form highly ordered structures, held together

through van-der-Waals interactions. For this thesis, only COPs will be considered further.

Amorphous molecular films and molecular crystals have been reviewed elsewhere.63,64

Although COPs are semiconducting materials, important distinctions regarding

their semiconducting properties exist when compared to their inorganic counterparts. The

first consideration is the nature of conductivity in inorganic semiconductors and COPs.

Intrinsic conductivity at room temperature is attainable for certain inorganic

semiconductors due to free charge carrier generation by thermal excitation. This is due,

in part, to the low band gap energies of inorganic semiconductors (e.g., 0.67 eV for Ge,

1.1 eV for Si and 1.4 eV for GaAs). Additionally, the relative dielectric constant, εr, of

inorganic semiconductors is large (e.g., 16.0 for Ge, 11.7 for Si and 13.1 for GaAs),

resulting in minimal coulombic attraction between photogenerated electron-hole pairs,

allowing charge carriers to conduct freely within the valence and conduction bands upon

excitation. In contrast, εr values of COPs are typically between 2 and 4, leading to large

coulombic attraction between photogenerated electron-hole pairs. As a result, COPs

exhibit exciton diffusion lengths on the order of 10–20 nm, resulting in high recombination

rates. Further discussion on the short diffusion lengths and high recombination rates of

COPs is undertaken in section 1.3.4., along with commonly employed strategies to

overcome these limitations.

As mentioned previously, COPs are characterized by their extended molecular

π-orbitals and delocalized π-electrons. As the number of overlapping π-orbitals increases

with increasing monomer units added to the polymer chain, the density of discrete

molecular orbitals increases until a continuous band is formed. In the ground state, the

lower band is comprised of occupied π-orbitals, while the upper band is made up of

unoccupied π*-orbitals. In organic semiconductors, the highest occupied molecular orbital

Page 41: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

23

(HOMO) and lowest unoccupied molecular orbital (LUMO) are analogous to the valence

band maximum and conduction band minimum in inorganic semiconductors, respectively.

Additionally, the overlapping π-orbitals are instrumental in charge transport

through polymer films. High charge carrier mobilities are observed along the polymer

backbone, where π-orbital overlap is strong, and moderate charge carrier mobilities are

observed in the π-π stacking direction for cofacially aligned polymer chains, where

moderate π-orbital overlap exists. The dependence of charge carrier mobility on orbital

overlap results in significantly enhanced mobilities in polymer films that form ordered films.

More recently however, indacenodithiophene-based polymers have been shown

to exhibit high charge carrier mobilities, outperforming highly ordered, semicrystalline

polymers, despite having a nearly amorphous microstructure.65 It should be noted that

although these polymers are often cited as being amorphous, there exists enough small

range order to produce detectable diffraction peaks by synchrotron XRD experiments. To

explain this deviation from the longstanding view that high charge carrier mobilities are

only achievable for polymers with highly crystalline microstructures, Noriega et al. have

proposed a model where the small crystalline domains that exist within the amorphous

polymer matrix are linked together by a common polymer chain extending from one

ordered domain, through the amorphous region, and into the second ordered domain.66

These bridging polymer chains allow charges to be transported between ordered polymer

domains throughout the film. Studies involving these indacenodithiophene polymers have

led researchers to postulate that an increased tolerance to disorder may represent a more

significant factor in designing high mobility polymers than increased crystallinity. Indeed,

further studies have shown that indacenodithiophene polymers display a low degree of

energetic disorder, owing to their planar, torsion-free backbone configuration.67

1.3.2. Electronic Processes in Organic Semiconductors

The following section will provide a brief overview of the ground/excited states and

the vibronic structure of organic semiconductors. The ground state of a molecule refers to

an electron configuration with the lowest possible electronic energy. For conjugated

organic polymers (COPs), this corresponds to a configuration in which electrons occupy

only the molecular orbitals up to and including the HOMO level, leaving the LUMO

unoccupied. When an electron is promoted from the HOMO to the LUMO, the molecule

Page 42: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

24

transitions from the ground state to an excited state configuration and is referred to as an

electronic transition.

In addition to electronic energies, vibrational energies must be considered as

molecules are not stationary, but oscillate around their equilibrium position.68 Each

electronic state will, therefore, have multiple vibrational energies associated with it. As a

result, transitions from the ground state to any vibrational level in the excited state can

occur, where the resulting absorption spectrum represents a superposition of the

transitions to different excited state vibrational levels. The simultaneous change in

vibrational and electronic energy levels is referred to as a vibronic transition.

Finally, to obtain a complete description of the polymer’s state, the spin of the

electrons must be taken into consideration, where electron spins may only have a value

of either +½ or -½. If the spins of the unpaired electrons located in the π- and π*-orbitals

of the excited state are antiparallel (i.e., add up to a total spin of zero), the molecule is

said to be in the singlet state. Conversely, if the excited state electron spins are parallel

(i.e., add up to a total spin of one), the molecule is said to be in the triplet state.

Figure 1.7 Orbital configurations of the ground state (S0), the first excited singlet state (S1) and the first exctied triplet state (T1), where the arrows represent the electron spin.

In addition to optical excitations of an electron from the ground to an excited state,

the process of recombination represents an important phenomenon in organic electronics.

During recombination, an electron residing in a π*-orbital of an excited state molecule will

relax down to the ground state, emitting a photon with energy equal to that of the transition.

Kasha’s rule states that the emitting level of a given multiplicity will be the lowest level of

that multiplicity.69 Simply put, for any radiative relaxation process, the transition must

Page 43: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

25

originate from the lowest vibrational level for a given excited state. As a result, any electron

excited to a higher vibrational level within the excited state must first lose its vibrational

energy by vibrational relaxation, returning to the zeroth vibrational level of the excited state

before emission can occur. The emissive transition may, however, take place to any

vibrational level in the ground state. Because of Kasha’s rule, the emission spectra of a

given molecule will therefore be red-shifted relative to its absorption spectra, referred to

as a Stokes shift.

Figure 1.8 Jablonski diagram indicating the electronic and vibrational levels of singlet and triplet states, as well as radiative and non-radiative transitions. Excitation processes are represented by straight, solid arrows; radiative recombination processes are represented by straight, dashed arrows; and non-radiative transitions are represented by undulated arrows.

1.3.3. History

The earliest reported conjugated organic polymer appeared in the literature in

1862, where polyaniline was presumably formed by the electrochemical oxidation of

aniline.70 Despite this early discovery, it would take nearly 100 years for the area of

conducting and semiconducting polymers to gain significant traction, beginning with the

Page 44: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

26

synthesis of polyacetylene in 1958.71 The development of conjugated polymers continued

into the 1960s, most notably with the synthesis of polypyrrole in 1968.72 Although the

advent of electroactive plastics was exciting, the usefulness of polyaniline, polyacetylene

and polypyrrole remained limited as they were synthesized as insoluble powders. A

tremendous breakthrough occurred in 1974 when researchers were able to successfully

synthesize polyacetylene directly into thin films.73 Just a few years later, a discovery with

even greater significance was made when the collaborative efforts of Heeger, MacDiarmid

and Shirakawa showed that through the doping of polyacetylene films with Cl2, Br2, I2 and

AsF5, their conductivity could be enhanced by a factor of 107.74 In fact, Heeger,

MacDiarmid and Shirakawa were awarded the 2000 Nobel Prize in Chemistry for their

work on electrically conductive plastics. The area of conjugated organic polymers saw an

influx of researchers following the work on chemically doped polyacetylene films, leading

to the discovery of numerous polymers and their applications in organic electronics such

as organic field effect transistors, organic light emitting diodes and organic photovoltaics.

Of the polymers discovered, polyphenylene, polythiophene and polyphenylene

vinylene were especially well studied, however, much like the unsubstituted conjugated

polymers that came before them, these polymers were largely insoluble. In an effort to

synthesize soluble conjugated polymers, it was found that the introduction of alkyl or

alkoxy sidechains to the conjugated backbone afforded solubility to the polymers. Early

examples of soluble conjugated organic polymers are poly([2-methoxy-5-(3,7-

dimethyloctyloxy)]-1,4-phenylene-vinylene) (MDMO-PPV) and poly(3-hexylthiophene)

(P3HT). When combined with a fullerene electron acceptor, these polymers showed

reasonable performance when employed as active layers in photovoltaic devices, e.g., up

to 3% and 6.5% power conversion efficiencies (PCE) for MDMO-PPV75 and P3HT76

respectively. Despite the reasonable PCEs achieved with these polymers, MDMO-PPV

has been plagued by poor morphological control, leading to increased recombination

rates, and P3HT is limited by its rather large band gap of 2 eV and high lying HOMO level,

resulting in poor utilization of the solar spectrum and a small open circuit voltage (VOC),

respectively.77

Page 45: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

27

Figure 1.9 Chemical structures of a few select conjugated organic polymers

Synthetic design strategies have led to a class of conjugated organic polymers

known as donor-acceptor (D-A) polymers, where the conjugated backbone is made up of

alternating electron donating and electron accepting units.78 In general, D-A polymers

exhibit smaller band gap energies, owing to the hybridization of the energy levels of the

donor and acceptor moieties. Additionally, the donor-acceptor configuration leads to the

HOMO being located mainly on the donor moiety, while the LUMO is located mainly on

the acceptor moiety. As a result, a partial positive charge resides on the donor unit and a

partial negative charge resides on the acceptor unit upon absorption of a photon, this aids

in stabilizing the photogenerated exciton. While ensuring that a polymer has a sufficiently

low band gap will ensure efficient absorption of the solar spectrum, the HOMO position

will also influence the VOC, where a lower lying HOMO will lead to a higher VOC.‡ Therefore,

a balance between lowering the band gap while maintaining a sufficiently high VOC is

required. Many donor and acceptor building blocks have been synthesized and their

effects on the optoelectronic properties of polymer systems investigated;78–80 and

advances in donor-acceptor polymers have led to single-junction organic solar cells with

PCEs above 10%.81–84

‡ This assumes that the energetics of the acceptor (fullerene or non-fullerene) remain unchanged.

Page 46: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

28

Figure 1.10 Chemical structure of two D-A polymers. Donor units are colored red and acceptor units are colored blue.

Although these D-A polymers offer significant advantages over the second-

generation conjugated polymers, like P3HT, they are synthetically challenging to produce,

making them prohibitively expensive to employ in exploratory areas of research, such as

organic photoelectrochemical cells. The research undertaken in this thesis, therefore,

focuses on P3HT, where advances in synthetic methodology, reasonable optoelectronic

properties and extensive literature reports make it an ideal candidate for a proof-of-

concept material in organic photoelectrochemical cells.

1.3.4. P3HT

Poly(3-hexylthiophene) (P3HT) was first developed in 198585 to address the

solubility issues of unsubstituted polythiophene, itself developed only a few years earlier

in 1980.86,87 While the conjugated thiophene backbone is maintained, the hexyl chains

afford solubility to the polymer, allowing for solution processing of polymer films. In

addition to solubility, the incorporation of alkane side chains has significant influence over

the structural, electronic and optical properties of the polymer.88 Due to the asymmetry of

the 3-hexylthiophene monomer, three different monomer coupling configurations exist:

head-to-tail, head-to-head, and tail-to-tail. The term regioregularity is used to describe the

ratio of head-to-tail couplings within the polymer chain. High regioregularity is desired for

producing films with a high degree of order,89 leading to higher charge carrier mobilities

Page 47: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

29

compared to regiorandom polymer configurations.90 When regioregularity is sufficiently

high, backbone planarization and aggregation of P3HT chains in the solid state leads to

the formation of crystalline domains. P3HT films are, therefore, characterized as being

semi-crystalline; as such, the degree of crystallinity within P3HT films can measured using

x-ray scattering techniques.91

Figure 1.11 A representation of the three possible coupling configurations of 3-hexylthipohene monomers: head-to-head, head-to-tail, and tail-to-tail. ‘Hex’ is used to represent the hexyl functional group (-C6H13).

P3HT has a band gap of ~2 eV, providing adequate absorption of the solar

spectrum, with P3HT films exhibiting an absorption maximum situated at 515 nm.

Additionally, the large absorption coefficient of P3HT leads to shallow absorption depths,

allowing for efficient collection of solar radiation at thin films, on the order of 100 nm.92 The

small relative dielectric constant of P3HT (εr = 3.5), however, results in exciton diffusion

lengths on the order of 10–20 nm and increased recombination rates. To overcome the

challenges associated with P3HT’s small εr value, electron accepting materials, such as

fullerenes (e.g., [6,6]phenyl-C61-butyric acid methyl ester (PCBM)), or n-type organic

semiconductors, are routinely employed to enhance exciton separation.

The implementation of a bilayer configuration in organic photovoltaics to improve

charge separation, and consequently device efficiency, was first reported by Tang, where

a perylene electron acceptor was deposited on top of a copper pthalocyanine donor,

leading to power conversion efficiencies of 1%.93 Later, the discovery of photoinduced

Page 48: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

30

electron transfer from conjugated organic polymers to fullerene (C60) would lead to

fullerenes becoming the electron accepting material of choice in organic photovoltaics.94

Subsequent research in the area of organic photovoltaics focused on designing

polymer/fullerene bilayers to enhance charge separation at the donor-acceptor interface,

however, charge separation in this bilayer configuration is limited to photogenerated

excitons formed within 10-20 nm of the donor-acceptor interface. Any excitons generated

further than one diffusion length away from the interface will recombine, leading to high

recombination rates in the bulk of the semiconductor layer. By increasing the interfacial

area between electron donating and electron accepting domains, the pathlength from the

exciton generation site to the donor-acceptor interface is minimized. This is generally

accomplished by employing a bulk heterojunction (BHJ) morphology, where phase

separation between the donor and acceptor materials is controlled, resulting in an

interpenetrating donor-acceptor blend.95 If the exciton can reach the donor-acceptor

interface before recombination occurs, effective electron transfer from donor to the

acceptor material can occur, while the hole remains in the donor phase. The separated

charges are now free to percolate through the donor and acceptor materials and be

collected at their respective electrodes. A prevailing hypothesis in the area of organic

photovoltaics is that the formation of highly ordered active layer morphologies consisting

of alternating donor and acceptor domains that are twice the width of the exciton diffusion

length should lead to efficient charge separation and excellent charge transport. This has

spurred research efforts to fabricate nanostructured active layers for use in organic solar

cells.96 To date, however, no reports of nanostructured active layers outperforming the

more traditional BHJ approach to device fabrication exist.

Figure 1.12 Diagram depicting the bulk heterojunction morphology achieved when a P3HT:PCBM film is cast on an ITO substrate.

In general, the operation of organic photovoltaic cells and organic

photoelectrochemical cells can be broken down into three major steps: (i) photon

Page 49: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

31

absorption, where a photon with energy equal to or greater than the band gap of the

organic semiconductor is absorbed, resulting in the excitation of a ground state electron

in the HOMO to the LUMO, leaving behind a hole. (ii) exciton migration, where the

photogenerated electron-hole pair migrates through the system. During the exciton

migration step, if the electron-hole pair can reach the donor-acceptor interface it can be

separated into mobile charge carriers. If the electron-hole pair is not able to reach the

donor-acceptor interface within the excited state lifetime of the material, radiative

recombination can occur, resulting in the emission of a photon. (iii) Charge collection,

where any separated mobile charge carriers that have made their way through the system

can be extracted at their respective contacts.

1.4. Organic Photoelectrochemical Cells

The first literature example where researchers took advantage of organic

semiconductor properties for solar fuel generation was the demonstration of the aqueous

photoactivity of polyacetylene films in 1981.97 Only a few years later, in 1985, poly(p-

phenylene) was investigated as a heterogenous photocatalyst for the photoreduction of

water to hydrogen gas, in the presence of a sacrificial electron donor.98 The area of organic

photocatalysts for water splitting remains an active area of research, where various

materials such as graphitic carbon nitride polymers,99,100 poly(azomethine) networks,101

pyrene-based conjugated microporous polymers,102 and carbazole-phenylene-based

polymers,103 to highlight a few, are being studied. Although interesting and important to

the development of organic semiconductor electrochemistry, further discussion of organic

heterogeneous photocatalysts will be forgone for the sake of conciseness. The focus of

this section will, therefore, be directed towards organic photoelectrochemical cells

(OPECs) exclusively.

Until recently, the area of OPECs has remained relatively dormant since the

seminal work on polyacetylene photoelectrodes. The only other example of an organic

photoelectrochemical cell for solar fuels generation surfaced in 1996, where El-Rashiedy

established the aqueous photoactivity of P3HT thin film electrodes.104 Since then, P3HT,

aptly referred to as the fruit fly of conjugated organic polymers, has been the photocathode

material of choice for the budding area of OPECs. Appropriate thermodynamics (i.e., the

redox potential of the hydrogen evolution reaction (HER) lies above P3HT’s Fermi level,

but below its LUMO level) and abundant literature studies have made P3HT an obvious

Page 50: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

32

choice as a proof-of-concept material from which researchers can better understand the

fundamental principles that govern water splitting processes at organic photoelectrodes.

Figure 1.13 Energy level diagram of a P3HT photocathode on ITO and H+/H2 redox couple in solution.

A renewed interest in OPEC research began in 2012 when Lanzarini et al.

demonstrated the production of stable photocurrent densities at P3HT:PCBM

photocathodes in aqueous NaCl electrolyte.105 Since then, publications in the area of

OPECs have been released regularly, and their numbers steadily increasing. In only a few

years, organic photoelectrode performance for water splitting reactions has increased

dramatically, from current densities of sub-μA/cm2 to milliamps/cm2. These improvements

have come largely from the incorporation of hydrogen evolution catalysts, such as Pt106,107

and MoS3,108 as well as incorporating interfacial charge selective layers. Interfacial layers

are routinely employed in organic electronic devices and serve multiple purposes including

tuning the energy level alignment at the electrode-active layer interface,109 facilitating

ohmic contact; defining the polarity of the electrodes, where either contact can serve as

the anode or cathode depending on the materials employed, leading to device

architectures with increased stability110 and better compatibility with roll-to-roll processing

techniques;111 improving charge selectivity by preventing the flow of unwanted charge

carriers across the interfacial layer (i.e., holes to cathode and electrons to anode) that

would result in recombination and lower device efficiency;112 and controlling the

morphology, where vertical stratification of the donor and acceptor materials of the BHJ

can be controlled through surface energy considerations, leading to enhanced charge

collection at both the top and bottom electrodes.113

Page 51: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

33

Figure 1.14 Organic photoelectrochemical cell with a device architechture of ITO/PEDOT:PSS/P3HT:PCBM/MoS3.114 Reprinted from Queyriaux, N.; Kaeffer, N.; Morozan, A.; Chavarot-Kerlidou, M.; Artero, V. J. Photochem. Photobiol. C Photochem. Rev. 2015, 25, 1, Copyright 2015, with permission from Elsevier.

In the context of OPECs, numerous materials have been investigated as hole

transport layers, such as graphene derivatives (e.g., graphene oxide and reduced

graphene oxide),115 amorphous WO3,116 CuI,117 MoO3,118 polyaniline,119 2D MoS2 flakes120

and 3-D MoO3 nanolamellar scaffolds.121 In addition to hole transport layers, electron

transport layers have been used to improve charge transfer between the organic layer and

HER catalyst,122 as well as protect the underlying photoactive organic layer from

degradation.123 These advances in device engineering have led to OPECs with

photocurrent densities as high as 8 mA/cm2 at 0 V vs. RHE, through the incorporation of

a CuI hole transport layer, P3HT:PCBM BHJ active layer, nanostructured TiO2 electron

transport layer and Pt HER catalyst.117 Although these photocurrents are outperformed by

those of p-type silicon photocathodes (27 mA/cm2 at 0 V vs. RHE for silicon nanowire

arrays)124 and copper-based chalcogenides (>20 mA/cm2 at 0 V vs. RHE for CdS and Pt

modified CIGS photocathodes),31 they are comparable to top performing metal-oxide

photocathodes (7.6 mA/cm2 at 0 V vs. RHE for Cu2O with an Al:Zn/TiO2 layer),125 and are

approaching those of group III-V compound materials (11 mA/cm2 at 0 V vs. RHE for

GaInP2 with a graded TiO2/MoOx/MoSx layer).126

Although large photocurrents are impressive on the laboratory scale, ultimately a

technology will have to be compatible with large-scale, cost-effective manufacturing

methods to be commercially relevant. The advancements in device engineering have

Page 52: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

34

indeed allowed for all solution processed OPECs to be assembled. The first example of

an all solution processed photocathode employed a device architecture of

ITO/MoOx/P3HT:PCBM/MoS3 and produced photocurrent densities of 2 mA/cm2 at

0 V vs. RHE for one hour.126 Recently, an all solution processed photocathode with a

device architecture of FTO/CuI/P3HT:PCBM/TiO2/Pt/PEI was produced, achieving

photocurrent densities as high as 5.25 mA/cm2 at 0 V vs. RHE.127 This is approaching

champion all-solution processed photocathodes based on inorganic devices capable of

achieving 8 mA/cm2 at 0 V vs. RHE using Bi-modified CuInS2 (CIS(Bi)) in a device

architecture of Mo/CIS(Bi)/CdS/TiO2/Pt photocathodes.128

Demonstrations of OPECs have expanded beyond the hydrogen evolution

reaction (HER) to include the oxygen reduction reaction (ORR), for both dissolved oxygen

detection129 and as an intermediate step in the production of H2O2;130 organic photoanodes

for water oxidation;131,132 and hybrid inorganic/organic tandem cells for unassisted, overall

water splitting.133,134 Additionally, other research areas related to OPECs are emerging,

such as modified fullerenes for both HER135 and ORR,136 as well as the fabrication of

photoelectrodes with hybrid organic/inorganic bulk heterojunctions for use in

photoelectrochemical energy conversion and storage, e.g., P3HT nanofibers with CdS

nanoparticles137 and P3HT with CdSe nanocrystals.138 However, in addition to device

engineering, a fundamental understanding of the energetics and electronic processes at

the organic-electrolyte interface is necessary for future progress, but reports in these

areas are less common.and the few publications that exist focus on studies performed in

the absence of PCBM, i.e., simplified P3HT-electrolyte interfaces.139–143

1.5. Photoelectrode Performance and Experimental Methods

1.5.1. Benchmark and Diagnostic Efficiencies

Solar-to-Hydrogen Conversion Efficiency

Overall solar-to-hydrogen (STH) conversion efficiency is the most important

parameter used to characterize PECs and is used to compare the performance of various

materials against one another, analogous to power conversion efficiency in photovoltaics.

As the most significant and only acceptable global comparison standard, STH efficiency

is regarded as the benchmark efficiency in solar water splitting.144

Page 53: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

35

STH efficiency describes the overall efficiency of a photoelectrochemical water

splitting device exposed to broadband solar irradiation under zero-bias conditions, i.e., the

working electrode (WE) and counter electrode (CE) are operated under short circuited

conditions. Additionally, for true STH measurements, the WE and CE must be immersed

in identical electrolytes with the same pH due to the Nernstian bias of 59 mV per pH unit

that would arise from the chemical bias between the two solutions were they to be

contained in separate compartments.145 The electrolyte must also be free from any

sacrificial donor or acceptor species that would otherwise provide additional reaction

pathways. Incident radiation should first pass through an AM1.5G filter so as to simulate

true sunlight reaching the terrestrial surface. The AM1.5G filter provides incident light that

is a good representation of the solar spectrum on a clear day, around noon, for a tilted PV

unit situated at the mid latitudes.146

With the above definitions and conditions in mind, STH efficiency can be

expressed as:

STH = [|jsc|×1.23 V × ηF

Ptotal]

AM1.5G (6)

Where jsc is the short circuit photocurrent density normalized to the electrode area

(in mA/cm2), 1.23 V represents the thermodynamic water splitting potential, ηF is the

faradaic efficiency for hydrogen evolution and Ptotal is the power density of the incident

illumination (in mW/cm2).

In addition to the benchmark STH efficiency measurements, multiple diagnostic

efficiencies exist that provide valuable information about the materials being investigated.

These include the applied bias photon-to-current efficiency (ABPE); the incident photon-

to-current efficiency (IPCE), also referred to as the external quantum efficiency (EQE);

and the absorbed photon-to-current efficiency (APCE), also referred to as the internal

quantum efficiency (IQE). These three major diagnostic efficiencies are discussed below.

Applied Bias Photon-to-Current Efficiency (ABPE)

During the operation of PECs, an external voltage bias may be applied to

compensate for any voltage deficiency or to enhance photogenerated charge carrier

separation by increasing the magnitude of band bending. The application of an external

Page 54: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

36

bias, therefore, requires a separate definition from that of the STH efficiency. This is

accomplished by introducing an applied bias term, Vb, to the STH equation:

ABPE = [|jsc|×(1.23 V− |Vb|) × ηF

Ptotal]

AM1.5G (7)

Incident Photon-to-Current Efficiency (IPCE)

IPCE can be summed up concisely as “the number of electrons out versus the

number of photons in”. Practically, photocurrent at the semiconductor working electrode

is measured under monochromatic irradiation and this is compared against the number of

incident photons for the same wavelength of light, calculated using a calibrated photodiode

detector. The photocurrent and number of incident photons can be converted to electron

transfer rate and incident photon flux, respectively. Thus, IPCE is the ratio of the rate of

electron transfer at the semiconductor-electrolyte interface to the incident photon flux,

represented as a function of wavelength.

IPCE (λ) =electrons/cm2/s

photons/cm2/s =

|jph|

Pmono×

1240

λ (8)

Where jph is the photocurrent density (in mA/cm2), Pmono is the calibrated power

density of the incident monochromatic irradiation, λ is the wavelength of the

monochromatic incident irradiation and 1240 represents the multiplication of Planck’s

constant, h, and the speed of light, c.

IPCE represents an important diagnostic as it considers the efficiencies of three

fundamental processes involved during photoelectrochemical water splitting. The first

being the fraction of excitons (i.e., electron-hole pairs) generated per incident photon flux,

ηe−/h+, the second is charge transport of the photogenerated charges to the

semiconductor-electrolyte interface, ηtransport, and the third is interfacial charge transfer,

ηinterface.

IPCE ∝ ηe−/h+ ηtransport ηinterface (9)

Theoretically, integration of IPCE over the entire solar spectrum can be used to

estimate STH if collected under zero-bias conditions, however, this assumes 100%

Page 55: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

37

Faradaic efficiency, which may not always be the case if other by products are formed or

corrosion occurs.

Absorbed Photon-to-Current Efficiency (APCE)

Both IPCE and STH implicitly include incident photons that are lost due to

reflectance or transmission. APCE is often used to subtract those lost photons and

measure efficiency based only on photons which are successfully absorbed. This is

accomplished by introducing absorptance, ηe−/h+, defined as the fraction of excitons

generated per incident photon flux, i.e., the first fundamental process described above.

Therefore:

APCE ∝ IPCE

ηe−/h+

∝ ηtransport ηinterface (10)

If a material exhibits high ηe−/h+, but a low IPCE, it can therefore be inferred that

the limiting factor will be either poor charge transport efficiency or sluggish charge transfer

kinetics at the semiconductor-electrolyte interface (or both). By re-running IPCE

measurements with a redox couple possessing better interfacial kinetics, the two limiting

factors can be decoupled as only ηinterface should be drastically affected by the nature of

the redox couple.

Absorptance can be determined by performing UV-Vis measurements to obtain

the absorbance, where absorbance is proportional to the thickness of the semiconductor

film and the concentration of the absorbing chromophore. From Beer’s law:

A = − log (𝐼

𝐼𝑜) (11)

Where A is the absorbance, Io is the incident light intensity and I is the intensity of

the transmitted light. Absorptance is expressed as:

ηe−/h+ = 𝐼𝑜−𝐼

𝐼𝑜 = 1 −

𝐼

𝐼𝑜 = 1 − 10−A (12)

APCE can therefore be expressed as:

Page 56: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

38

APCE (λ) = |jph|

Pmono×

1240

λ×

1

1−10−A (13)

1.5.2. Experimental Methods

Many of the parameters discussed in previous sections, such as photocurrent,

faradaic efficiency and built-in potential represent some of the most significant parameters

pertaining to the characterization of PECs. The following section will provide an overview

of the experimental methods used to measure these parameters.

Photoelectrochemical Cell

Photoelectrochemical measurements require a specialized cell that allows for

multiple considerations, such as allowing for irradiation of the semiconductor working

electrode, the ability to maintain an inert atmosphere, sampling of the headspace for

analysis by gas chromatography, and the cell must be outfitted to accommodate the use

of multiple electrodes. The cylindrical body of the photoelectrochemical cell, with flat

windows on either end, allows incident irradiation to pass through the cell; the multiple

ports allow for the use of an inert gas line and various electrodes; and a septum allows for

sampling of the headspace. Typically, photoelectrochemical cells are fabricated from

quartz, to allow high energy UV irradiation to reach the working electrode, however, for

this work the photoelectrochemical cell was fabricated in house and made of glass, which

is opaque to UV light. A diagram of the photoelectrochemical cell used in this work is

presented below in Figure 1.15.

Figure 1.15 Diagram of the glass cell used for photoelectrochemical measurements.

Page 57: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

39

Photocurrent

Arguably the most important parameter in solar water splitting, photocurrent is a

direct measure of device performance and appears in all the efficiency calculations

mentioned above. Photocurrent is measured by potentiostatic coulometry, also referred to

as electrolysis, under chopped illumination. Electrolysis measurements are performed

using a conventional three-electrode setup, where the working electrode is held at

constant potential and current is measured as a function of time. In photoelectrochemical

cells, a light source is used to illuminate the photoelectrode intermittently, resulting in three

distinct regions. The first region is the dark-current, which represents any current

measured while the photoelectrode is not being illuminated. As the redox processes

occurring at the semiconductor-electrolyte interface are driven by photogenerated

charges, dark current in photoelectrochemical cells is generally negligible. The second

region of importance is the steady-state photocurrent, which is the current measured while

the photoelectrode is under illumination. Finally, photocurrent transients appear during the

initial illumination due to charging processes that occur at the semiconductor-electrolyte

interface.147 The value of photocurrent reported is the difference between the steady-state

photocurrent and the dark current. A typical electrolysis plot is shown in Figure 1.16, with

the regions of interest highlighted.

Figure 1.16 Example of a typical electorolysis trace. The three important regions are labelled, where i. represents the dark region, ii. the photocurrent transient region, and iii. the steady-state photocurrent region.

Page 58: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

40

Faradaic Efficiency

Although photocurrent is largely considered the best metric to determine overall

performance of PECs, bulk electrolysis is incapable of differentiating photocurrents due to

desirable water splitting reactions and other parasitic currents, such as the formation of

by-products or corrosion. Differentiation of the origin of photocurrent is crucial when

determining overall STH efficiency as the inclusion of any parasitic photocurrents will

artificially inflate the STH efficiency. The Faradaic efficiency is the fraction of photocurrent

that is used for water splitting reactions versus the total measured photocurrent. To

determine the Faradaic efficiency of a PEC, electrolysis and gas chromatography are

combined, where the theoretical amount of hydrogen gas produced at the cathode is

calculated from the amount of charge passed during electrolysis and the real amount of

hydrogen gas present after electrolysis is measured by gas chromatography. The ratio of

the measured hydrogen to the theoretical amount of hydrogen produced yields the

Faradaic efficiency. Faradaic efficiencies can be maximized, and often approach unity, by

limiting the number of competing charge transfer pathways. This can be accomplished

through purging of the electrolyte with inert gas to remove any dissolved gases and using

high purity electrolytes.

Built-In Potential

The built-in potential (Vbi) is one of several factors that determines the depletion

layer width at the semiconductor surface and will consequently influence charge

separation within the semiconductor. Both open circuit potential (OCP) and linear sweep

voltammetry (LSV) can be used to measure Vbi. OCP measurements are performed using

a conventional three-electrode setup under zero-current conditions (i.e., at open circuit).

The potential is measured as a function of time, beginning under dark conditions and then

under illumination. The difference between the OCP measured under dark and illuminated

conditions is known as the photovoltage (Vph) of the system. At sufficiently high light

intensity, all pre-existing band bending at the semiconductor surface is removed and Vph

is equal to Vbi.

Vbi can also be estimated by LSV, where LSV measurements are also carried out

using a three-electrode setup and current density is measured as a function of potential

under chopped illumination. For p-type semiconductor photoelectrodes, potentials are

swept from more anodic (i.e., positive) potentials to more cathodic (i.e., negative)

Page 59: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

41

potentials. The potential corresponding to the onset of photocurrent (Vonset) represents the

point at which minority carriers are transferred from the semiconductor to the electrolyte,

i.e., the onset of band bending. The difference between Vonset at the illuminated

semiconductor and the reduction potential (Eredox) of the redox couple will therefore be

equal to Vbi when kinetic overpotentials are minimized. Eredox can be found experimentally

by performing cyclic voltammetry measurements, employing a working electrode with low

polarization overpotentials, such as Pt.

Figure 1.17 a) An idealized example of a typical OCP measurement, where regions i and ii represent the OCP measured under dark and illuminated conditions, respectively. b) A linear sweep voltammogram under chopped illumination, a vertical dashed line has been drawn from the onset of photocurrent to highlight Vonset.

Ultraviolet-Visible Spectroscopy

Ultraviolet-Visible (UV-vis) spectroscopy is a form of absorption spectroscopy,

where the sample is irradiated with light in the ultraviolet-visible region of the solar

spectrum. UV-vis spectroscopy measures the absorbance of the sample as a function of

wavelength, where absorbance is determined by measuring the ratio of the intensities of

the incident and transmitted light at a given wavelength. Where A is the absorbance, Io is

the intensity of incident light and I is the intensity of the transmitted light.

A = log𝐼o

𝐼 (14)

In addition to measuring how strongly a material absorbs light of a certain

wavelength, UV-vis can provide information regarding the optoelectronic properties of

conjugated organic polymers, such as the optical band gap, and insight into polymer

Page 60: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

42

aggregation and film morphology. The optical band gap of COPs can be determined from

the wavelength at which the onset of absorption occurs by converting the onset

wavelength into energy. Although UV-vis can be used as a simple method to estimate the

band gap of a material, it should be noted that the optical band gap is not a precise

measurement of the difference between the HOMO and LUMO levels of organic

semiconductors, due to poorly screened electron-electron interactions.148 Backbone

planarization and formation of semi-crystalline domains can be observed in the P3HT UV-

vis spectra through the appearance of the S1 ← S0 0-0 transition, manifesting itself as a

vibronic shoulder at lower energy wavelengths.149 The planarization of the P3HT backbone

is associated with aggregation, where the relative peak ratio of the 0-0 and 0-1 transitions

can provide information on whether H- and J-type aggregation is dominant, as proposed

by Spano.150,151

Photoluminescence Spectroscopy

Photoluminescence (PL) spectroscopy is a form of emission spectroscopy, where

a sample is excited using a light source (typically using monochromatic light corresponding

to the maximum absorption wavelength of the sample) and the intensity of the emitted

radiation is measured as a function of wavelength. PL spectroscopy can provide important

insights into the radiative recombination rates within COPs as successful electron transfer

from the donor to acceptor phase results in the quenching of photoluminescence intensity.

PL peak ratios can also be used in accordance with the H-aggregate model to provide

information regarding polymer chain aggregation.

Grazing Incidence Wide Angle X-Ray Scattering

Grazing incidence wide angle x-ray scattering (GIWAXS) is a subset of x-ray

scattering techniques, where grazing incidence refers to the operation of x-ray scattering

in the reflection scattering geometry, as opposed to the traditional transmission scattering

geometry. In GIWAXS measurements, the incident x-ray beam approaches the sample at

a very shallow angle, that is less than the critical angle of the sample material, to ensure

all external reflection (e.g., P3HT has a critical angle of 0.178° for x-rays with a wavelength

of 1.55 Å) and scattering is measured as a function of exit angle (yz plane) and in-plane

angle (xy plane). The size of the angles measured is dependent on the sample-to-detector

distance, where wide angles (i.e. GIWAXS) are measured when the detector is situated

10-50 cm from the sample and small angles (i.e., GISAXS) at 130-300 cm.152

Page 61: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

43

Consideration of Bragg’s Law dictates that larger scattering angles allow for the resolution

of smaller feature sizes.153 Therefore, GIWAXS is typically employed for the

characterization of P3HT thin film morphology, due to its ability to probe the atomic and

molecular distances within the crystalline domains.

Figure 1.18 Diagram showing the scattering geometry of a typical GIWAXS measurement.

Specifically, GIWAXS can be used to quantify the distribution of polymer chain

orientation within P3HT thin films. Within the crystalline domains, P3HT polymer chains

form a unit cell comprised of two main orientations: (i) Lamellar stacking, where polymer

chains stack in the direction of the hexyl sidechains. This is taken as the (100) direction.

(ii) π-stacking, where the planarized, conjugated backbones of adjacent polymer chains

align. This is considered the (010) direction. Any alignment of the polymer chains in the

direction of conjugation, is generally too disordered and therefore no observable (001)

peaks appear in P3HT scattering profiles.

Page 62: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

44

Figure 1.19 Diagram of an idealized P3HT unit cell of where the (100) and (010) orientations are shown in the lamellar and π-stacking directions, respectively.

When cast as a thin film, the P3HT polymer chains adopt two major orientations

with respect to the substrate, the edge-on orientation, where the lamellar stacking direction

is normal to the substrate and the face-on orientation, where the π-stacking direction is

normal to the substrate. Depending on the application, different orientations will be favored

as charge transport within P3HT films has been shown to be much higher in the π-stacking

direction compared to the lamellar stacking direction.154 In organic field-effect transistors,

for example, an edge-on orientation with in-plane π-stacking is desired, whereas for

devices with a vertical architecture, such as organic photovoltaics (OPVs) and organic

photoelectrochemical cells (OPECs), the face-on orientation with π-stacking normal to the

substrate is preferred. GIWAXS provides quantitative information on the orientation of the

polymer chains in the solid state and is therefore an important measurement when

characterizing the morphology of polymeric thin films.

Dynamic Light Scattering

Dynamic light scattering (DLS) is a common method used to determine the size of

nanoparticles, where the Stokes-Einstein relationship is used to relate particle motion and

particle size. DLS functions by passing a monochromatic laser through a diluted

nanoparticle solution, where light is scattered and collected at a detector situated at 90

degrees to the incident laser beam. As the nanoparticles move due to Brownian motion,

the scattering pattern at the detector also changes. An auto-correlation function tracks the

fluctuation of the scattering intensity at the detector and provides insight into the diffusion

Page 63: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

45

coefficient of the nanoparticles. The hydrodynamic diameter of the nanoparticles can then

be calculated using the Stokes-Einstein relationship:155

dh =kBT

3πηD (15)

where dh is the hydrodynamic diameter, kB is the Boltzmann constant, T is the temperature,

η is the viscosity of the solvent and D is the diffusion coefficient.

1.6. Thesis Scope

The following chapters will explore the photoelectrochemical properties of P3HT,

with a particular interest in its application as a photocathode material to perform the

hydrogen evolution reaction. Although solar water splitting technologies represent a

potential source of abundant renewable energy in the future, the objective of this thesis is

not necessarily to engineer the highest performing photocathodes from conjugated

organic polymers, but rather to gain a better understanding of the thermodynamic and

electronic processes that take place at the organic semiconductor-electrolyte interface.

Chapter two explores the origin of photocurrent at single layer P3HT films

deposited on ITO-coated glass in acidic aqueous solutions. Where it was previously

thought that the sub-μA/cm2 photocurrents observed at the illuminated P3HT-electrolyte

interface was due to the reduction of protons to hydrogen gas, quantitative evidence is

provided that suggests any photocurrents measured at uncatalyzed P3HT photocathodes

are largely due to the reduction of dissolved oxygen. Despite favorable thermodynamics,

this work demonstrates the need for hydrogen evolution catalysts to be employed in

concert with the P3HT photoabsorbing layer to successfully achieve hydrogen evolution.

In chapter three, P3HT:PCBM nanoparticles are synthesized and used to fabricate

nanostructured organic photocathodes. The morphology and optoelectronic properties of

the nanostructured photocathodes are compared with conventional, solution-cast thin

films of P3HT:PCBM, where control of the nanoscale morphology within each nanoparticle

leads to enhanced P3HT:PCBM phase segregation. The photo-assisted deposition of

platinum nanoparticles as hydrogen evolution reaction catalysts onto the P3HT:PCBM

photocathodes facilitates the photoreduction of protons to hydrogen gas and increased

hydrogen production is demonstrated at the nanostructured electrodes.

Page 64: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

46

The role of PCBM at the semiconductor-electrolyte interface is discussed in

chapter four. Specifically, the influence of PCBM on the energy level alignment of the

redox couple in solution and the formation of an interfacial dipole layer are explored in

detail. It is found that the presence of PCBM results in a decreased built-in potential and

depletion layer thickness in P3HT photocathodes. This was shown to influence

photocathode performance in thinner photoelectrodes, where efficient charge separation

was achieved in the absence of a PCBM acceptor phase by tuning the semiconductor

thickness to coincide with depletion layer width.

Final remarks, followed by recommendations for future directions in the field of

organic photoelectrochemical cells mark the end of this thesis. The need for rational

polymer design and organic nanostructures highlights the future directions section.

Figure 1.20 Outline of this thesis organized in a flow chart.

Page 65: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

47

Chapter 2. On the Origin of Photocurrent at Uncatalyzed P3HT Films in Aqueous Acidic Media

Sections of this chapter have been reproduced in part with permission from:

Suppes, G. M.; Fortin, P. J.; Holdcroft, S. J. Electrochem. Soc. 2015, 162, H55. Copyright

2015, The Electrochemical Society. The work presented in this chapter are the

contributions made to the paper by P. Fortin.

2.1. Introduction

At the onset of this project, very few reports concerning the use of conjugated

organic polymers as photocathodes for the hydrogen evolution reaction existed. At the

time, the goal of the project was to determine whether it was possible to evolve hydrogen

at uncatalyzed P3HT photocathodes. Combing through the literature, it seemed as though

it was possible. Examples of conjugated organic polymers used in sacrificial water splitting

systems can be found as early as 1985 wherein it was reported that irradiation of a

suspension of poly(p-phenylene) in the presence of a diethylamine hole scavenger led to

the photoreduction of protons to hydrogen gas.98 It was subsequently reported, using a

photoelectrochemical system, that P3HT exhibited photocathodic activity in aqueous

solutions and that photocurrent could be increased by incorporation of electron acceptors,

such as fullerene.104 The photocurrents measured at these organic photocathodes were

interpreted as being commensurate with the hydrogen evolution reaction. Additionally,

Lanzarini et al. demonstrated that P3HT:PCBM films on ITO generates stable

photocurrent in the sub-µA/cm2 range in aqueous NaCl,105 concluding that the

photocurrent was due to proton reduction at the P3HT:PCBM electrode and oxidation of

Cl- at the counter electrode.

However, careful examination of the literature reports concerning

photoelectrochemical hydrogen evolution at uncatalyzed P3HT surfaces revealed that, in

fact, no quantitative analysis (e.g., gas chromatography) had been performed to confirm

the presence of hydrogen gas. Given this revelation, along with the advancements in the

field that indicate an apparent need for catalysts to promote hydrogen evolution at

Page 66: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

48

irradiated conjugated polymer films, the origin of photocurrent at single layer P3HT films

in aqueous solution was re-examined.

One possible reaction pathway competing with the hydrogen evolution reaction is

the one electron reduction of oxygen to the superoxide radical anion, O2•-. In aqueous

acidic media, the superoxide radical anion undergoes the following disproportionation

reaction to produce hydrogen peroxide (H2O2) and molecular oxygen (O2):156

𝑂2 + 𝑒− → 𝑂2•−

(16)

O2•− + H+ → HO2

• (17)

HO2• + O2

•− + H+ → H2O2 + O2 (18)

where the oxidation state of oxygen changes from -½ in the superoxide radical anion to -

1 in hydrogen peroxide and 0 in molecular oxygen.

In this chapter, the propensity of uncatalyzed P3HT photocathodes to perform the

hydrogen evolution reaction (HER) and oxygen reduction reaction (ORR) is investigated.

Headspace analysis, by gas chromatography, is performed after photoelectrolysis to gain

insight into whether hydrogen is created at irradiated P3HT photocathodes. Additionally,

the relationship between photocurrent density and dissolved oxygen concentration is

examined to determine additional reduction pathways that may occur at irradiated P3HT

photocathodes.

2.2. Experimental

2.2.1. Materials

Regioregular poly(3-hexylthiophene) used for the P3HT-O2 photoelectrochemical

studies was synthesized according to a literature procedure.157 The number-average

molecular weight of the polymer was 35,600 kDa with a polydispersity index (PDI) of 1.66

as determined by gel permeation chromatography using tetrahydrofuran (Sigma-Aldrich,

HPLC grade) eluant, three Waters µStyragel columns at 40°C, a Waters 1515 isocratic

HPLC pump, a Waters 2414 differential refractometer, and a Waters 2487 dual UV

absorbance detector (λ = 254 nm). Polystyrene standards (Waters) were used for

calibration. The degree of head-to-tail regioregularity was determined to be 94% by 1H

Page 67: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

49

NMR spectroscopic analysis, according to standard literature procedure.158,159

Chlorobenzene (Sigma-Aldrich), H2SO4 (Anachmia), NaOH (BDH chemicals), NaCl

(Fisher), Na2SO3 (Fisher) were all used as received. All aqueous solutions were prepared

using water obtained from a Milli-Q water purification system (18 MΩ, EMD Millipore).

Indium tin oxide (ITO) coated glass slides (15 Ω/□) (Colorado Concept Coatings LLC) were

cleaned by sequential sonication in each of the following solutions: isopropyl alcohol,

acetone, H2O: H2O2: NH4OH (5:1:1) solution and distilled water.

2.2.2. Film Preparation

Films for P3HT-O2 photoelectrochemistry were prepared from a 25 mg/mL

chlorobenzene solution that had been stirred overnight at 65˚C. The solution was spin-

cast onto chemically cleaned ITO in a glove box (H2O < 0.1 ppm, O2 < 1 ppm) at 1000 rpm

for 1 minute, affording polymer films ~100 nm thick. Film thicknesses were determined

using an Alpha-Step IQ profilometer (KLA-Tencor). The edges of each electrode were

covered with chemically-resistant and electrically insulating epoxy purchased from

Epoxies Inc. (20-3004 HV) to prevent exposure of ITO to the electrolyte.

2.2.3. Photoelectrochemistry

The light source was a 200 W Xe/Hg lamp (Uhsio America, inc.), used in

combination with a 300-700 nm band pass filter (FSQ-KG3, Newport Corp.) and neutral

density filters (Thorlabs Inc.), to achieve 100 mW cm-2 irradiation, as measured using a

broadband power meter (841-PE, Newport Corporation) equipped with an Ophir thermal

detector head (3A-P-SH-V1). The cell configuration was designed to allow irradiation of

the polymer-electrolyte interface through the electrolyte. A water filter was placed in front

of the electrochemical cell to remove excess heat. Electrochemical measurements were

performed using a Pine Bipotentiostat (AFC-BP1). Photoelectrochemical measurements

were performed in a 3-electrode configuration using a saturated calomel electrode (SCE)

(+0.24 V vs. RHE) and a Pt wire as the reference and counter electrodes, respectively.

Solutions were sparged with argon (Praxair, purity 99.999 %) for 1.5 to 2 hours in a glass

PEC cell and sealed under a flow of argon. Pt was photoelectrochemically-deposited onto

P3HT-coated ITO electrodes from 0.1 mM K2PtCl6 in 1 M H2SO4 at -0.1 V vs. SCE

(0.14 V vs. RHE) for 60 s under 100 mW cm-2 irradiation.

Page 68: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

50

2.2.4. Hydrogen Gas Detection

The headspace in the electrochemical cell was sampled using a 5 mL syringe,

fitted with a gas tight valve (Series A-2, VICI Precision Sampling), and analyzed using an

Agilent Technologies 6890N GC system equipped with a thermal conductivity detector. A

2.13 m Agilent J&W GC packed column in stainless steel tubing was used (inner diameter

of 2 mm, HayeSep N packing material, 60/80 mesh size). Argon was used as a carrier gas

at a flow rate of 30 mL/min1 under 46.2 psi.

2.2.5. Dissolved Oxygen Detection

An oxygen probe (Orion 850 polarographic probe) was used to measure dissolved

molecular oxygen in solution. Unless specifically stated, solutions were sparged with

argon (Praxair, purity 99.999 %) for 1.5 to 2 h prior to use and sealed under a flow of

argon, resulting in a dissolved oxygen concentration <0.01 ppm as measured by the

dissolved oxygen probe. For other experiments, the dissolved oxygen content of the

electrolyte was controlled either by using solutions exposed to ambient air, referred to as

“ambient air” solutions, having a dissolved O2 concentration of 7.80 ppm, or by bubbling

the solutions with oxygen for ten minutes, resulting in oxygen saturated solutions having

a dissolved oxygen concentration >20 ppm.

2.3. Results

2.3.1. P3HT photocurrent vs. O2 concentration

Sub-µA/cm2 photocathodic current is observed under potentiostatic conditions

when a P3HT-coated ITO electrode is immersed into 0.1 M H2SO4 solution purged with

argon, as illustrated in Figure 2.1. In the dark, no cathodic current is observed. As shown

in Figure 2.1, the initial photocurrent observed upon irradiation decays to a pseudo-steady

state. Upon removing the light source and re-illuminating the electrode, the initial current

spike decreased in magnitude and eventually disappeared, rendering a steady state

current of 200-300 nA/cm2. A similar photocurrent profile is observed in 0.1 M NaCl and

0.1 M NaOH, but the magnitudes of the photocurrents are smaller. It was considered that

the photocurrent was due to the reduction of protons, leading to the hydrogen evolution

Page 69: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

51

reaction (HER), because the conduction band electrons of P3HT are sufficiently reductive

to reduce protons in solution.

Figure 2.1 Electrolysis of P3HT-coated ITO electrodes at -0.24 V vs. SCE (0 V vs. RHE) performed in 0.1 M H2SO4 (black), 0.1 M NaCl (dotted), 0.1 M NaOH (dashed), and 0.5 M Na2SO3 oxygen scavenger (dot-dashed). Light intensity: 100 mW cm-2 (300-700 nm), under chopped illumination cycles of 30 s dark / 1 min light.

To examine whether the photocurrent was indeed due to HER, the headspace of

the electrolysis cell was sampled after 6 h of photoelectrolysis, after passing ~56 mC

at -0.24 V vs. SCE (0 V vs. RHE) of cathode charge—equivalent to 0.29 μmol of H2 gas

(assuming 100% current-to-gas efficiency). However, the quantity of H2 gas measured by

GC, if at all produced, was below the detection limit of the instrument. In a series of control

experiments, hydrogen was evolved at a Pt foil working electrode by application

of -0.4 V vs. SCE (-0.16 V vs. RHE) for 10 s, passing 32.5 mC of charge, equivalent to

0.17 μmol of H2; and for 60 s, passing 202.5 mC of charge, equivalent to 1.05 μmol of H2.

Electrolytically-produced H2 was readily detectable by GC, as shown in Figure 2.2,

providing compelling evidence that the photocurrent observed at P3HT-coated ITO

electrodes is not commensurate with hydrogen evolution.

Page 70: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

52

Figure 2.2 Gas chromatographic analysis of H2 gas in the photoelectrochemistry cell after 202.5 mC of charge over 60 s (dot-dashed) and after passing 32.5 mC of electrolytic charge over 10 s (dashed) at a Pt foil (i.e., conventional electrochemical HER); 6 h photoelectrolysis at a P3HT-ITO electrode (dotted) after passing 56.6 mC of cathodic charge. The solid black signal represents a sample of air (i.e, blank signal). Signals are offset for clarity.

The influence of trace oxygen was investigated as the source for photocathodic

current, as oxygen has been shown to act as an electron acceptor in solution.156,160,161

After purging the solutions with inert gas, the concentration measured by the oxygen probe

was below the limits of detection, <0.1 ppm. However, the addition of 0.5 M sodium sulfite,

a known oxygen scavenger,162 reduced the observed photocurrents shown in Figure 2.1

to non-detectable values. This confirms that trace oxygen remaining after purging with

argon is likely responsible for the photocurrent. Moreover, potentiostatic photoelectrolysis

(Figure 2.3) measurements in 0.1 M H2SO4 with increasing dissolved oxygen contents

(<0.01, 7.80, and >20.0 ppm) led to increasing photocurrent and no dark current.

Potentiostatic measurements revealed that, in the presence of a high concentration of

dissolved oxygen, the magnitude of the transient current spike observed immediately

following illumination returns upon repeated on-off cycles. In contrast, the cathodic current

spike fades after the first on-off cycles in the absence of a replenishing amount of oxygen.

The origin of this is uncertain but may involve the well-known P3HT-O2 charge transfer

complex that forms upon their contact and which serves as an electron acceptor,163,164 and

which has been observed for P3HT films in contact with water.165

Page 71: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

53

Figure 2.3 Photoelectrolysis of P3HT coated ITO electrodes in 0.1 M H2SO4 with different dissolved oxygen concentrations. Oxygen deficient (<0.01 ppm) conditions (red), ambient (7.80 ppm) conditions (green) and oxygen saturated (> 20.0 ppm) conditions (blue). Light intensity: 100 mW cm-2 (300-700 nm). The electrodes were biased at -0.24 V vs. SCE (0 V vs. RHE). Chopped illumination cycles consist of 30 s dark / 1 min light.

That P3HT is capable of photoelectrochemically-reducing oxygen is not

unexpected since other PEC reactions are also possible if the appropriate redox couple is

present and provided the redox energy level lies below the energy of the conduction band

of P3HT. The overriding feature of these experiments is that the measured photocurrent

at the P3HT-coated photocathodes in pure aqueous or acidic solution appears largely to

be due to the reduction of dissolved oxygen in solution, even when the electrolyte is

flushed with argon for extended periods of time. Going to greater lengths to exclude

oxygen, such as the addition of oxygen scavengers, reduces the photocurrent to negligible

values. Thus, despite favourable energy level positions of the P3HT LUMO in relation to

the H+/H2 redox couple, and the observation of pH dependence on photocurrent, the HER

is not the predominant mechanism.

2.4. Conclusion

Despite the conduction band electrons of P3HT possessing sufficient redox

potential to reduce protons, no hydrogen gas was detected at irradiated, single layer,

pristine P3HT films. Additionally, when the concentration of dissolved oxygen was

Page 72: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

54

controlled, it was observed that photocurrent density was proportional to dissolved oxygen

concentration. Together, these results suggest that the origin of photocurrent at

uncatalyzed P3HT photocathodes is largely due to the reduction of residual O2 in solution.

In the presence of an oxygen scavenger, photocurrents are reduced to negligible amounts.

This work provides quantitative evidence that a catalyst layer must be used in

conjunction with the organic photoabsorbing layer to promote photoelectrochemical

hydrogen generation. This work provided the foundation for further research carried out in

the Holdcroft lab, where platinum nanoparticle catalysts were deposited

photoelectrochemically onto P3HT films and hydrogen gas was successfully evolved by

photoelectrolysis.107

Page 73: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

55

Chapter 3. Hydrogen Evolution at Conjugated Polymer Nanoparticle Electrodes

Sections of this chapter have been reproduced in part with permission from:

Fortin, P. J.; Rajasekar, S.; Chowdhury, P.; Holdcroft, S. Can. J. Chem. 2018, 96, 148.

Copyright 2018, NRC Research Press. The work presented in this chapter are the

contributions made to the paper by P. Fortin.

3.1. Introduction

With the establishment that hydrogen evolution does not occur at uncatalyzed

P3HT photocathodes (Section 2.4) and a method to incorporate a hydrogen evolution

catalyst onto the photocathodes has been developed, strategies to further improve

photocurrent densities at organic photocathodes were explored.

A strategy often adopted in inorganic photoelectrode materials chemistry for

increasing photocurrent density is to increase the surface area of the electrode by way of

forming a nanostructured electrode surface, as demonstrated with a range of

nanostructured semiconductor photoelectrodes, each of which offer distinct

improvements. For instance, high aspect ratio nanowire electrodes are able to overcome

charge transport constraints of planar electrodes by minimizing the exciton diffusion path

length to the semiconductor-electrolyte interface, so that photogenerated holes and

electrons can be efficiently separated prior to their recombination.166 Nanoholes have been

integrated into 20 µm-thin Si electrodes and decorated with Pt nanoparticle catalysts.167

The high photocurrent densities achieved using these structures (23 mA/cm2 at

0 V vs. RHE) is attributed to the antireflective properties of the nanoholes, which increases

light absorption in the thin electrodes, as well as increasing the charge transfer kinetics

due to suppression of surface recombination. Recently, nonprecious-metal catalysts were

deposited onto pyramidally, nanotextured silicon substrates,168 and the enhanced

interfacial electrode area used to increase the amount of catalytic sites accessible for the

hydrogen evolution reaction, yielding photocurrents of 12 mA/cm2.

Page 74: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

56

Organic semiconducting polymers have the distinct advantage over inorganic

counterparts in that they can be cast from simply-prepared polymer solutions. However,

when cast, they typically form smooth, uniform films leading to an electrode surface area

that is similar to the geometric area of the substrate onto which the film is cast. Drawing

from strategies used in the PEC studies of inorganic films it ought to be possible to

increase the surface area of the polymer film by adopting a nanoparticle approach to film

formation.

Nanoparticles of P3HT and P3HT:PCBM have indeed been the subject of

extensive characterization in recent years as a means to control the morphology of

photoactive layers in organic photovoltaic light harvesting systems in order to offset the

relatively short diffusion length of photogenerated excitons in organic materials.169–174

Nanoparticulate systems also allow deposition of photoactive components from water and

alcohols rather than from traditional chlorinated solvents, as demonstrated for organic

photovoltaic (OPV)175–177 and organic field effect transistor (OFET)178 devices. Organic

polymer and polymer:fullerene nanoparticle dispersions may be prepared by one of two

methods: (i) a mini-emulsion method, wherein a polymer or polymer:fullerene blend is

dissolved in a solvent and subsequently added to a surfactant-containing non-solvent,

where the solvent and non-solvent are immiscible and (ii) a surfactant-free precipitation

method, in which a polymer or polymer:fullerene blend is dissolved in a solvent and directly

injected into a non-solvent, for which the solvent and non-solvent are miscible. Both

methods of nanoparticle formation are illustrated in Figure 3.1.

Although surfactant-stabilized nanoparticles of P3HT:PCBM have been shown to

be smaller than the surfactant-free nanoparticles,179 they form core-shell structures, with

a PCBM-rich core, which may inhibit effective photogenerated charge separation and

collection, whereas the surfactant-free precipitation method forms nanoparticles that

possess a relatively uniform, blended morphology.180 With the surfactant-stabilized

nanoparticle approach, trace amounts of surfactant may remain incorporated in the active

material, potentially interfering with the formation of the bulk-heterojunction, and thus

impinging upon device performance and device lifetime.177 The surfactant-free approach

to nanoparticles mitigates these potential issues. In this method, the polymer of

Page 75: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

57

Figure 3.1 General schemes showing the formation of organic polymer nanoparticles by (a) the mini-emulsion method and (b) the precipitation method.

choice is dissolved in a “good” solvent and the solution injected, under control, into a “poor”

solvent (the non-solvent) so that upon contact with the non-solvent the polymer chains

collapse to form spherical-like nanoparticles.

The concentration of the original polymer solution affects the size of the

nanoparticles formed. As the concentration is increased, the number of polymer chains

per unit volume is increased, leading to larger nanoparticles. In considering the formation

of bulk heterojunction films of donor and acceptor molecules, nanoparticles with the

smallest possible diameters are assumed to be preferred for photovoltaic and

photoelectrochemical applications in order to minimize the donor and acceptor domain

sizes, thereby reducing the exciton diffusion length required to reach the heterojunction

interface. Two key variables essential to controlling nanoparticle preparation are: (i) the

nature of the solvent and non-solvent, which must be miscible to ensure formation of a

Page 76: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

58

homogenous colloidal solution of the polymer; (ii) the concentration of the initial polymer

solution, which needs to be low enough so as to form nanoparticles, yet high enough to

ensure the colloidal solution formed provides films sufficiently thick so as to be examinable

as photocathodes when cast on a substrate. Herein, we report the vacuum-free

preparation of high surface area organic photoelectrodes using surfactant-free

P3HT:PCBM nanoparticles and the subsequent investigation of their morphological and

optoelectronic properties. The influence of these properties on their performance as

organic photocathodes, when coupled with a Pt hydrogen evolution catalyst, is

investigated and compared to photoelectrodes prepared from solution-cast P3HT:PCBM

films.

3.2. Experimental

3.2.1. Materials

P3HT was purchased from Rieke Metals (MW = 50-70 kg/mol, regioregularity: 91-

94%, <0.01% metal impurities). PC61BM (American Dye Sources, Inc.), K2PtCl6 (Johnson

Matthey) and all solvents were all used as-received. All aqueous solutions were prepared

using DI water obtained from a Milli-Q water purification system(18MΩ, EMD Millipore).

Indium tin oxide (ITO) coated glass slides (15 Ω/□) (Colorado Concept Coatings LLC) were

cleaned by sequential sonication for 10 minutes in each of the following solutions:

dichloromethane, deionized water and isopropyl alcohol. Chemically-cleaned ITO slides

were subjected to a plasma treatment using 25% v/v O2 in Ar for 10 minutes (Fischione

Instruments Model 1020 plasma cleaner).

3.2.2. P3HT:PCBM Nanoparticle Films

Solutions of P3HT:PCBM was prepared by first dissolving 10 mg of P3HT in 1 mL

CHCl3 and 10 mg of PCBM in 1 mL CHCl3 , stirring each overnight at 50°C. The solutions

of P3HT and PCBM were mixed together and 1.5 mL of this solution was added by syringe

to 4 mL of isopropyl alcohol (IPA) under sonication at room temperature for 1 minute

(Branson 1510 ultrasonic cleaner). A dispersion of nanoparticles of P3HT:PCBM was

obtained, having a final concentration of 2.7 mg/mL. Next, 100 µL aliquots of the

nanoparticle dispersion were spin-cast onto ITO-coated glass slides at 2000 rpm multiple

times to achieve the desired thickness. A film thickness of ~90 nm required 10 spin-coats.

Page 77: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

59

3.2.3. Deposition of Pt

Platinum nanoparticles were photoelectrochemically-deposited onto films of

P3HT:PCBM nanoparticles from 0.1 mM K2PtCl6 in 1 M H2SO4 by application of an

electrode potential of −0.1 V vs. SCE under chopped (100 mW/cm2) illumination using

5 second dark and 15 second light intervals for a total irradiation time of 60 seconds. The

amount of Pt deposited on the polymer films was calculated using Equation 19:

mPt = Q

nF∗ MWPt (19)

where mPt is the mass of Pt deposited via photoelectrochemical deposition, Q is the charge

passed from the working electrode to the electrolyte and is measured by taking the area

above the Pt deposition curves, F is Faraday’s constant, MWPt it the molecular weight of

Pt and n, the number of electrons, is four.

3.2.4. Photoelectrochemistry

A 200 W Xe/Hg lamp (Uhsio America, inc.) was used in combination with a

300-700 nm band pass filter (FSQ-KG3, Newport Corp.) and neutral density filters

(Thorlabs Inc.), to achieve 100 mW/cm2 irradiation, as measured using a broadband

power meter (841-PE, Newport Corporation) equipped with an Ophir thermal detector

head (3A-P-SHV1). A glass photoelectrochemical cell with two large flat windows and five

accessory ports was used and the cell configuration was designed to allow irradiation of

the polymer–electrolyte interface through the electrolyte. A water filter was placed in front

of the electrochemical cell to remove excess heat. Electrochemical measurements were

performed using a Pine Bipotentiostat (AFC-BP1). PEC measurements were performed

in a 3-electrode configuration using a saturated calomel electrode (SCE)

(+0.24 V vs RHE) and a Pt wire as the reference and counter electrodes, respectively.

Solutions were purged with nitrogen (Praxair, purity 99.999%) for 1 hour in a glass PEC

cell and a positive pressure N2 blanket maintained in the overhead volume of the cell. For

H2 evolution measurements, the cell was sealed under N2 to avoid the loss of any H2 in

the head space above the working electrode.

Page 78: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

60

3.2.5. Gas Chromatography

The headspace in the electrochemical cell was sampled using a 5 mL syringe,

fitted with a gastight valve (Series A-2, VICI Precision Sampling), and analyzed using an

Agilent Technologies 6890N GC system equipped with a thermal conductivity detector. A

2.13 m Agilent J&W GC packed column in stainless steel tubing was used (inner diameter

2 mm, HayeSep N packing material, 60/80 mesh size). Argon (99.999%) was used as a

carrier gas at a flow rate of 30 mL/min under 46.2 psi.

3.2.6. Electron Microscopy

Scanning electron microscopy (SEM) was carried out using a Nova NanoSEM

electron microscope at an accelerating voltage of 5 kV and a working distance of 5 mm.

Films were prepared on ITO as described above and fixed to an aluminum stub using

carbon tape (SPI supplies). A secondary electron (SE) detector was used to obtain images

of the fine nanostructured features of the nanoparticle electrodes and a back scattered

electron (BSE) detector was used to visualize the Pt nanoparticle catalyst on the

nanostructured electrodes. BSE images are contrasted by atomic number, where the

bright phases correspond to the species with a higher atomic number. This is especially

useful in this work as it allows to distinguish between nanoparticle catalysts from

nanoparticle support. Bright field transmission electron microscopy (TEM) images were

acquired with a FEI Tecnai Osiris scanning transmission electron microscope (FEI

Company, Hillsboro, OR, USA) operated at 200 kV. TEM samples were prepared by

dropping an IPA solution of the nanoparticle dispersion on a TEM grid having an ultrathin

carbon film on a lacey carbon support (Prod #01824, Ted Pella Inc).

3.2.7. Spectroscopy

UV Absorption spectra were recorded on a Cary 300 Bio UV-visible

Spectrophotometer (Agilent Technologies). All solution samples were diluted to an

appropriate concentration and measurements were performed in quartz cuvettes. All thin

film samples were ~90 nm thick and prepared on ITO-coated glass substrates. Absorption

intensities of each spectrum were normalized to their respective λmax values. Steady-state

fluorescence emission spectra were performed using a PTI Quantamaster

spectrofluorometer with a xenon short arc lamp (Ushio Inc.) as the excitation source. All

Page 79: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

61

thin film samples were mounted 22.5° normal to the incident light. All spectra were

recorded using an excitation wavelength of 515 nm, coinciding with their λmax values as

determined by UV-Vis.

3.2.8. Grazing Incidence Wide Angle X-ray Scattering

GIWAXS measurements were performed on a SAXSLAB Ganesha 300XL

instrument using a photon wavelength of 1.54 Å (Cu K-α). Measurements were taken

under vacuum at an angle of incidence of 0.16° and the scattered intensity was detected

on a Pilatus3 R 300K detector. Data analysis was performed using the saxsgui version

2.15.02 software package.

3.2.9. Dynamic Light Scattering

DLS measurements were performed at room temperature using a Zetasizer Nano

ZS (Malvern Instrument). 1 drop of the nanoparticle dispersion (2.7 mg/mL) was diluted in

2 mL of the appropriate solvent and measurements were performed in plastic cuvettes

with a 1 cm path length.

3.3. Results and Discussion

3.3.1. Nanoparticle Synthesis

P3HT:PCBM nanoparticles were synthesized via surfactant-free precipitation

method, as described above. Various combinations of solvent (in which polymer:fullerene

solutions were prepared) and non-solvent (in which the polymer:fullerene solutions were

added) were investigated in order to study the synergistic effect on both the size of

nanoparticles and nature of aggregation of the polymer nanoparticles. Solvents/non-

solvents were chosen by considering the solubility data (reported by Machui et al.)181 of

both P3HT and PCBM (Table 3-1).

While poor solubility in the non-solvent ensures that the blend of P3HT and PCBM

will co-precipitate upon the addition of the P3HT:PCBM solution to the non-solvent, the

physical properties of both the solvent and non-solvent are important to the formation of

nanoparticle films. Employing a solvent which possesses a lower boiling point and higher

Page 80: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

62

vapour pressure than the non-solvent ensures that the good solvent of the solution

evaporates prior to evaporation of the non-solvent during the spin coating process. If the

non-solvent evaporates prior to the evaporation of the good solvent, the nanoparticles are

at risk of being re-dissolved by residual solvent during spin casting, which may result in

P3HT:PCBM films more akin to those cast from homogenous solutions.

Table 3-1 Solubility of P3HT & PCBM in various solvents181 and solvent properties

P3HT solubility (mg/mL)

PCBM solubility (mg/mL)

Boiling Point (°C)

Vapor Pressure (mmHg)

Solvents

o-dichlorobenzene 15 42 180 1

chloroform 14 29 61 158

tetrahydrofuran

1 2 65 145

Non-solvents

1-butanol < 0.1 < 0.1 118 5

2-propanol < 0.1 < 0.1 83 33

methanol < 0.1 < 0.1 65 92

ethanol < 0.1 < 0.1 78 43

dimethylsulfoxide < 0.1 0.2 189 1

water < 0.1 < 0.1 100 17

Based on the above criteria, of those listed only CHCl3 and THF were considered

for further use, as o-DCB is not removed prior to the non-solvent during the spin coating

process due to its higher boiling point and lower vapor pressure. The physical properties

of the non-solvent are equally as important to consider, ensuring good film formation

during the spin coating process. Although many non-solvents meet the required criteria of

low P3HT and PCBM solubility, high boiling point and low vapor pressure, those with very

high boiling points and very low volatility are not completely removed during the spin

casting process, leading to excessively long drying times. For this reason, DMSO, BuOH

and water were rejected as non-solvents and not considered further.

In the surfactant-free precipitation method employed, the stability of the

nanoparticles in the non-solvent is also important as there is no surfactant present to

mitigate nanoparticle aggregation. It was found that non-solvents possessing more

Page 81: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

63

amphiphilic character were able to act as both non-solvent and surfactant, leading to

stable nanoparticle dispersions. This was apparent in the series of alcohols where

nanoparticles prepared in MeOH aggregated within a few hours, whereas nanoparticles

prepared in BuOH were stable over several days. As the alkane chain length of the alcohol

is increased, the stability increases due to their increasing amphiphilic nature; however,

boiling point and vapor pressure rapidly increase and decrease, respectively, with higher

carbon homologues, rendering alcohols with aliphatic chains longer than propanol

impractical for the formation of nanoparticle films.

3.3.2. Characterization of Nanoparticles

Dynamic light scattering (DLS) and transmission electron microscopy (TEM) were

used to determine the size of nanoparticles formed. The DLS data shown in Figure 3.2a

implies that a distribution of nanoparticle sizes exists in solution with a z-average diameter

of 140 nm. TEM images in Figure 3.2b indicate smaller nanoparticle sizes than those

obtained by DLS, but a distribution of sizes is observed nonetheless, with larger

nanoparticles having diameters of ~100 nm and smaller nanoparticles having diameters

of ~50 nm. DLS measures the hydrodynamic radius, but the measurements rely upon

Rayleigh scattering, where intensity of Rayleigh scattering scales with the 6th power of

particle diameter. Therefore the presence of larger aggregates skews the average size of

the nanoparticles to larger than actual values.178

As shown in Figure 3.3, UV-Vis absorption spectra of the nanoparticle dispersions

exhibit an absorption maximum situated at 515 nm, with a vibronic shoulder at 600 nm.

The red shift in the nanoparticle absorption spectra, with respect to the absorption of P3HT

dissolved in solution, as well as the existence of a vibronic shoulder are indicative of inter-

chain interactions due to π-π stacking. These features suggests that semi-crystalline

domains of P3HT are formed during formation of nanoparticles.26 The absorption spectra

of the nanoparticle dispersions possess similar absorption spectra to nanoparticle films

prepared therefrom, in contrast to films cast from P3HT polymer solutions where

significant interchain interactions are absent in solution but evolve during film casting.

Comparing the absorption spectra of nanoparticle and polymer solution-cast films, similar

absorption maxima are observed, however, the solution-cast film exhibits a weaker

vibronic shoulder intensity. Considering the H-aggregate model,150,151 it may be concluded

Page 82: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

64

Figure 3.2 Size distribution of P3HT:PCBM nanoparticle diameter determined by (a) dynamic light scattering (DLS) and (b) transmission electron spectroscopy (TEM). For DLS measurements, 1 drop of nanoparticle dispersion was diluted in 2 mL of IPA. TEM samples were prepared by drop casting a diluted nanoparticle dispersion in IPA on a lacey carbon TEM grid.

that the ratio of aggregated to non-aggregated polymer domains is larger for the

nanoparticle film than solution-cast films. However, limitations on over-interpreting the

relevance of the H-aggregate model from UV-Vis spectroscopy have recently been

raised.183 These limitations arise due to the sensitivity of UV-Vis spectra to reflections at

polymer boundary layers and the influence of film thickness and refractive index on

reflectance. The latter is especially relevant when comparing nanoparticle-cast and

solution-cast films, as the surface morphologies are expected to differ significantly.

Further information on the comparative aggregation within the nanoparticle and

solution-cast films was obtained by grazing incidence wide-angle x-ray scattering

(GIWAXS). GIWAXS plots are shown in Figure 3.4. The scattering peak situated at

q= 0.4 Å-1 (d-spacing of 15.7 Å) arises from the lamellar stacking of P3HT crystallites,

considered as the (100) plane in P3HT films. The relative intensity of the scattering peak

at q= 0.4 Å-1 is proportional to the degree of ordering within the P3HT nanodomains of the

P3HT:PCBM films.184 The similar relative scattering intensity of the nanoparticle and thin

film samples suggests a similar extent of ordering between the two. The 2D plot shown

for the nanoparticle electrodes, Figure 3.4b, shows polymer chains with nearly isotropic

orientation as characterized by the broad ditribution of intensity for all peaks, whereas the

2D plot of the

Page 83: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

65

Figure 3.3 Normalized UV-vis spectra of a P3HT:PCBM nanoparticle film (solid black), a solution cast P3HT:PCBM bulk heterjunction film (solid red), a dispersion of P3HT:PCBM nanoparticles in IPA (dashed black) and a P3HT:PCBM solution in o-DCB (dashed red). nanoparticle and solution cast films were ~90 nm thick.

solution-cast electrode, Figure 3.4c, shows only the (100) peak along qy with a faint trace

of the (200) peak, also along qy. These results suggest that the polymer chains within the

nanoparticles do not align themselves in any particular orientation with respect to the

substrate, while the polymer chains of the solution-cast film are preferentially aligned in

an edge-on orientation.

Due to high excitonic binding energies of conjugated polymers, the exciton

diffusion lengths are on the order of 10-20 nm.185 To compensate for the short exciton

diffusion lengths, electron acceptors, such as PCBM, have been introduced to promote

efficient charge separation. If the exciton recombines prior to reaching the donor-acceptor

interface, radiative emission may occur in the form of fluorescence. Ultafast photoinduced

electron transfer from a conjugated polymer to fullerene in bulk heterojunction films were

first reported by Sariciftci et al.94 Photinduced electron transfer manifests itself in

quenching of the photoluminescence intensity of the conjugated polymer film. Improving

the bulk heterojunction, by increasing the interfacial surface area between the donor and

acceptor phases, leads to enhanced quenching of the photoluminescence. Figure 3.5

shows the steady-state fluorescence spectra for both the nanoparticle and solution cast

P3HT:PCBM films, as well as a pristine P3HT film for reference. A clear increase in

quenching of the fluorescence is observed for both the nanoparticle and

Page 84: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

66

Figure 3.4 (a) GIWAXS data obtained for P3HT:PCBM nanoparticle (black) and solution-cast (red) films. The angle of incidence was 0.16°. Inset: Lamellar stacking of P3HT chains in an edge-on fashion along the (100) direction. 2D GIWAXS plots of the P3HT:PCBM nanoparticle and solution-cast electrodes are shown in (b) and (c), respectively.

solution-cast films due to the presence of PCBM. A possible advantage which

semiconducting nanoparticles offer is greater control of the nanoscale morphology by

controlling the extent of phase segregation within each nanoparticle rather than over the

entire active layer.175 The enhanced quenching observed for the P3HT:PCBM

nanoparticle films, with respect to the solution-cast film may be due to enhanced control

of the donor-acceptor phase segregation on the nanoscale.

SEM images of the P3HT:PCBM nanoparticles cast on ITO electrodes are shown

in Figure 3.6, from which the diameter of the nanoparticles is estimated to be between 50

and 100 nm, coinciding with the TEM results described above. Under lower magnification,

Figure 3.6a, a homogeneous distribution of P3HT:PCBM nanoparticles is observed across

the electrode, free from the formation of large aggregates. Under higher magnifications,

Figure 3.6b, layers of partially coalesced spherical nanoparticles are observed, confirming

the formation of nanostructured electrodes. Partial coalescence may be caused by trace

CHCl3 remaining in the nanoparticle dispersion. Although the CHCl3 solvent evaporates

Page 85: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

67

Figure 3.5 Steady-state fluorescence emission spectra of a solution-cast P3HT film (dashed), a solution cast P3HT:PCBM film (red) and a P3HT:PCBM nanoparticle film (black), all on ITO. An excitation wavelength of 515 nm was used.

prior to the alcohol during the spin casting process, its presence during each subsequent

coating may be sufficient to promote plasticization and partial coalescence of the

nanoparticles. Partial coalescence of the nanoparticles may, in fact, be beneficial to

photocathode activity, as previous reports have shown that coalescence of P3HT:ICBA

(indene-C60 bisadduct) nanoparticles, by thermal annealing, increases the performance of

organic photovoltaic devices due to the increase in the short circuit current density brought

about by the connectivity of nanoparticles.182

Figure 3.6 SEM images of a P3HT:PCBM nanoparticles on ITO electrodes. Images (a) and (b) depict the nanoparticle films at magnification values of 60k and 250k, respectively.

Page 86: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

68

3.3.3. Photoelectrochemistry of Nanoparticle-based Photocathodes

Photocathodes were prepared by depositing both P3HT:PCBM nanoparticles and

solution-cast P3HT:PCBM films onto ITO, followed by the photoelectrochemical

deposition of platinum nanoparticles by potentiostatic coulometry (Figure 3.7). By taking

advantage of the photoreductive activity of P3HT and the position of the redox potential

associated with the PtCl62- anion (0.51 V vs. SCE or 0.75 V vs. RHE), which is positive of

both P3HT and PCBM LUMO energy levels, it is possible to reduce PtCl62- to metallic Pt

on the electrode surface by irradiation of the P3HT:PCBM photocathode under a bias

of -0.1 V vs. SCE (0.14 V vs. RHE). The photocurrents observed are attributed to the

reduction of PtCl62- to Pt0.

Figure 3.7 Platinum deposition by electrolysis under chopped illumination for P3HT:PCBM nanoparticle (black) and solution-cast (red) films. The potential at the photoelectrodes was held at -0.24 V vs. SCE (0 V vs. RHE).

SEM images of the electrodes obtained using a back scattered electron detector

illustrate that Pt nanoparticle catalysts (bright spots) are deposited in uniform size and

distribution on the P3HT:PCBM nanoparticle and solution-cast films (Figure 3.8).

Linear sweep voltammetry of the photoelectrodes in aqueous solution, Figure 3.9a,

confirm the role of Pt catalyst on the nature of the electrochemical reactions under

irradiation as there is no appreciable photocurrent generated in their absence. After

deposition of Pt electrocatalysts, however, a cathodic current is observed having an onset

potential at 0 V vs. SCE (0.24 V vs. RHE), which is slightly more negative than the

reported flat-band potential of P3HT 0.1 V vs. SCE (0.34 V vs. RHE) in similar

solutions.186 It is not uncommon for the onset potential to appear up to a few hundred

millivolts more negative than the flat-band potential due to recombination in the space

Page 87: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

69

Figure 3.8 SEM images, collected using the backscattered electron detector, of Pt deposited after a deposition period of 100 s on (a) a P3HT:PCBM nanoparticle electrode and (b) a solution-cast P3HT:PCBM thin film electrode.

charge layer near flat band potential, charge trapping at surface defects and/or poor

charge transfer kinetics which leads to charge accumulation at the surface.187 Negligible

dark currents are observed for both the P3HT:PCBM nanoparticle and P3HT:PCBM

solution-cast photocathodes, however, the nanoparticle photoelectrodes exhibit a small

dark current at potentials more negative than -0.3 V vs. SCE (-0.06 V vs. RHE) which may

be due to electrocatalytic evolution of H2 at Pt which has been inadvertently deposited

onto exposed ITO sites during the Pt deposition process. Voltammograms of Pt deposited

on blank ITO were recorded (Figure 3.9c) and show the onset of H2 at the ITO/Pt

electrodes correlates well with the onset of dark current at the ITO/Pt electrodes.

To minimize the influence of H2 directly evolved at ITO/Pt sites during the photo-

assisted hydrogen evolution experiments, all subsequent photo-electrolysis experiments

were performed at sufficiently positive potentials (-240 mV vs. SCE or 0 V vs. RHE) where

the dark current was negligible. Potentiostatic coulometry data of the P3HT:PCBM

nanoparticle and of the solution-cast films, both with and without Pt catalyst, are presented

in Figure 3.9b. The experiments were performed under chopped illumination, where

significant currents are only observed during the illumination of the photoelectrodes and

return to negligible current in the dark, confirming that photogenerated charges within the

polymer films are responsible for the redox reactions which occur at the interfaces of the

working and counter electrodes within the photoelectrochemical cell.

Page 88: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

70

Photocurrents obtained during electrolysis are increased from sub-µA values at

uncatalyzed photoelectrodes to 35 µA/cm2 and 7 µA/cm2 for the Pt-deposited,

P3HT:PCBM nanoparticle and the Pt-deposited solution-cast photoelectrodes,

respectively. To confirm hydrogen evolution at the nanoparticle photocathodes,

electrolysis was performed for 1 hour and the headspace was positively identified by gas

chromatography as containing hydrogen. The rate of hydrogen generation was calculated

to be 0.1 µmol/h.

Figure 3.9 Linear sweep voltammetry (a) and electrolysis (b) under chopped illumination comparing P3HT:PCBM nanoparticle (black) and solution-cast (red) photocathodes in 0.1 M H2SO4. Experiments were performed both with (solid) and without (dashed) Pt nanoparticle catalysts. Linear sweep voltammetry of Pt deposited on bare ITO is shown in (c). Linear sweep voltammograms were measured at a scan rate of 5 mV/s and electrolysis measurements were performed at -0.24 V vs. SCE (0 V vs. RHE).

Although larger photocurrents are observed at the nanoparticle photoelectrodes

compared to their solution-cast film counterparts, it is noted that the increased surface

Page 89: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

71

area of the nanoparticle electrodes facilitates a larger quantity of Pt electrocatalyst to be

deposited when subjected to identical Pt deposition times under identical conditions. The

mass of Pt deposited was calculated using the total charge passed during

photoelectrochemical deposition, the details of which are provided in the experimental

details, which is a reasonable assumption as the solution had been thoroughly purged

with N2 to remove any dissolved oxygen. For the voltammetric and coulometric curves

provided, the loading of Pt catalyst deposited on the nanoparticle and solution-cast

electrodes films were 1.11 and 0.21 µg/cm2 respectively. To obtain comparable catalyst

loadings to the nanoparticle electrodes, longer Pt deposition periods were used for the

planar films. The longer deposition periods, however, lead to decreased photocathode

performance. The electrolysis results of solution-cast electrodes for which Pt has been

deposited for periods of 100 and 250 seconds are shown in Figure 3.10. These results

show that it is detrimental to extend Pt deposition for long periods of time. As deposition

time is extended, the size of the catalyst particles increases,188 resulting in a decreased

catalytic surface area and thus decreased catalytic activity.139

Figure 3.10 Electrolysis under chopped illumination comparing P3HT:PCBM solution-cast photocathodes prepared using Pt deposition periods of 100 s (solid) and 250 s (dashed). Measurements were performed at - 0.24 V vs. SCE (0 V vs. RHE).

3.3.4. Conclusion

P3HT:PCBM nanoparticles were prepared via simple precipitation method and

used to fabricate nanostructured organic photocathodes. The nanoscale morphology,

polymer crystallinity and photoelectrochemical performance of the nanostructured

Page 90: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

72

P3HT:PCBM electrodes were compared with planar P3HT:PCBM electrodes to better

understand the effects of nanostructuring on organic photocathodes.

It was postulated that control of the nanoscale morphology within individual

nanoparticles may lead to increased donor-acceptor interfacial area, and thus enhanced

charge separation compared to the bulk phase separation of donor-acceptor domains that

occurs in planar electrodes. Steady-state fluorescence spectra were collected for both the

nanostructured and planar P3HT:PCBM as a qualitative measure of photogenerated

charge separation. Although the nanostructured electrodes exhibit a slightly lower

emission signal compared to the planar electrodes, suggesting greater charge separation,

the presence of PCBM is of much greater importance than whether the particular donor-

acceptor morphology is controlled over a 90 nm thin film or within nanoparticles with

diameters of 50-100 nm.

The effect of nanoparticle formation on polymer aggregation and crystallinity was

studied through UV-Vis and GIWAXS measurements. The similarities between the

absorption spectra of P3HT:PCBM nanoparticles suspended in solution and P3HT:PCBM

nanoparticles cast as films suggests that polymer aggregation and crystallinity is

determined during nanoparticle formation. The red-shift, relative to the P3HT:PCBM

solution, and appearance of a vibronic shoulder at longer wavelengths are indicative that

semi-crystalline polymer domains are present within the nanoparticles immediately after

their synthesis and remain largely unchanged through the spin coating process. GIWAXS

measurements were performed for quantitative comparison of the polymer crystallinities

between the nanostructured and planar P3HT:PCBM electrodes. The GIWAXS

measurements revealed similar polymer crystallite sizes for the nanostructured and planar

electrodes, however, the orientation of the polymer chains with respect to the ITO

substrate differed. The polymer chains within the nanostructured electrodes do not adopt

a preferred orientation, whereas the polymer chains in the planar electrode are

preferentially aligned in an edge-on orientation.

Finally, a platinum hydrogen evolution catalyst was deposited onto both the

nanostructured and planar P3HT:PCBM photocathodes and their photoelectrochemical

performances compared in acidic aqueous solution. The nanostructured electrodes

exhibited a five-fold increase in photocurrent density during photoelectrolysis

measurements over the planar electrodes. While the fluorescence, UV-Vis and GIWAXS

Page 91: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

73

measurements suggest that morphological differences between the nanostructured and

planar electrodes exist, the most significant factor in photoelectrochemical performance

appears to be surface area. The larger surface area of the nanoparticle electrodes allows

for greater amounts of platinum to be deposited during the catalyst deposition process,

consequently leading to enhanced photoelectrochemical performance. Pt-loading studies

at planar electrodes show that increasing the Pt deposition period past a certain point will

begin to have adverse effects on photoelectrochemical performance. The growth of

catalyst particles with increasing deposition time, results in decreased catalytic surface

area.

For photoelectrochemical reactions plagued by slow electron transfer kinetics at

the polymer-electrolyte interface, such as the HER, the performance of the

photoelectrodes will ultimately be determined by the amount of catalytic surface area

available. There are, however, photoelectrochemical reactions which have been shown to

occur directly at the polymer-electrolyte interface in the absence of an electrocatalyst,

such as the reduction of O2189 and organic redox couples, such as anthraquinone-2,7-

disulfonate (AQDS).190 The reduction of O2 and AQDS are of interest in the application of

dissolved oxygen sensors and redox flow batteries, respectively. Further study of the

photoelectrochemical reduction of various redox couples at nanostructured conjugated

polymer electrodes is warranted.

Page 92: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

74

Chapter 4. Energy Level Alignment and Interfacial Dipole Layer Formation at the P3HT:PCBM-Electrolyte Interface in Organic Photoelectrochemical Cells

4.1. Introduction

It is well documented that in the field of organic photovoltaics, the formation of a

bulk heterojunction between an electron donating material (e.g. conjugated organic

polymer) and an electron accepting material (e.g., fullerene) is necessary to achieve high

power conversion efficiencies.191 Although the influence of PCBM at the organic-metal

interface in solid state organic electronics has been the subject of many studies, there

have been no studies concerning the influence of PCBM at the organic-electrolyte

interface in organic photoelectrochemical cells.

Previous reports, in the area of organic electronics, have shown that the interfacial

dipole layer formed at organic-metal interfaces alters the energy level alignment between

the Fermi level of the organic layer and the metal work function,192–195 and studies specific

to P3HT:PCBM bulk heterojunction systems have shown that the presence of PCBM at

the interface between the organic layer and the metal cathode leads to band unpinning

and a subsequent decrease in the magnitude of band bending in the organic layer at the

metal cathode contact.196,197 At the organic-metal interface, the wave function of the metal

extends into the organic layer, where overlap with the organic molecular states causes a

shift and broadening of the molecular levels, creating an induced density of interface

states within the band gap. Within the density of states there exists a charge neutrality

level - a level at which filling all states up to that level with electrons results in a surface

without a net charge. As a result, when the Fermi level of the organic layer lies below the

charge neutrality level, the surface will be positively charged. The spontaneous electron

transfer process which occurs between the organic and metal layers is determined by the

relative positions of the charge neutrality level and the metal work function. When the work

function of the metal lies above the charge neutrality level of the organic layer, electrons

are transferred from the metal to the interface states, creating an interfacial dipole layer

and alignment of the metal work function with the charge neutrality level of the organic

Page 93: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

75

layer (i.e., Fermi level pinning). Guerrero et al. found that the band-bending and strength

of the dipole layer at the P3HT:PCBM-metal cathode interface are highly dependent on

the number and nature of the fullerene molecules in contact with the metal, indicating that

PCBM molecules are largely responsible for controlling the density of the interface

states.197 Simply put, when a large number of PCBM molecules are present at the

interface, a large number of induced interfacial states exist. This is consistent with reports

that the extent to which the metal work function aligns with the charge neutrality level is

dependent on the density of interface states in the vicinity of the charge neutrality

level.195,198,199 As the analogy of the semiconductor-metal Schottky barrier is routinely

applied to semiconductor-electrolyte interfaces,15,200–202 applying the above findings to the

P3HT:PCBM photoelectrochemical system may prove useful in understanding the role of

PCBM in organic photoelectrodes.

Any energy level alignment that occurs at the semiconductor-electrolyte interface

will affect the magnitude of band-bending, i.e., the built-in potential (Vbi), and the depletion

layer width. Equation 2 shows the dependence of the depletion layer width, W, on the

built-in potential, Vbi, the relative dielectric constant of the organic layer, εr, the permittivity

of vacuum, εo, the elementary charge, q, and the doping density, NA.

The depletion layer plays an important role in photoelectrochemistry as any

photoinduced charges generated within this region may be separated by the built-in

electric field. It is therefore expected that the influence of PCBM on the energy level

alignment (i.e., Vbi) at the organic semiconductor interface will have a direct impact on

exciton separation within the semiconductor photoelectrodes and subsequent

photoelectrochemical performance. This work aims to elucidate the role of PCBM at the

organic-electrolyte interface of organic photoelectrochemical cells. More specifically, the

influence of PCBM on interfacial energy alignment is investigated through

photoelectrochemical techniques and the relationship between the built-in potential,

depletion layer width and photocathode performance is discussed.

Page 94: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

76

4.2. Experimental

4.2.1. Materials

P3HT was purchased from Rieke Metals (MW = 50-70 kg/mol, regioregularity: 91-

94%, < 0.01% metal impurities). PCBM (American Dye Sources, Inc.), and all solvents

were all used as-received. Indium tin oxide (ITO) coated glass slides (15 Ω/□) (Colorado

Concept Coatings LLC) were cleaned by sequential sonication for 10 minutes in each of

the following solutions: dichloromethane, deionized water and isopropyl alcohol.

Chemically-cleaned ITO slides were subjected to a plasma treatment in air for 10 minutes

(Harrick Plasma PDC-32G).

4.2.2. P3HT:PCBM Films

The appropriate ratios of P3HT and PCBM were weighed out depending on the

sample being prepared and dissolved by stirring in chloroform overnight at 60 °C. Thin

film electrodes were prepared by spincoating 0.4 µL aliquots of the corresponding solution

(filtered through 0.45 µm PTFE filter) onto plasma-cleaned ITO (Harrick Plasma PDC-

32G, 10 min with air plasma) and spun at 2000 rpm. The resulting P3HT and P3HT:PCBM

films were ~220 nm thick, as determined by profilometry (Alpha-Step IQ profilometer from

KLA-Tencor). For samples prepared with a PEDOT:PSS charge selective layer, a

PEDOT:PSS solution was filtered through a 0.45 µm Acrodisk filter and spin-cast onto

plasma cleaned ITO (Harrick Plasma PDC-32G, 10 min with air plasma) at 2000 rpm. The

resulting PEDOT:PSS films were placed in an oven at 120 °C for 10 mins to remove

moisture. The appropriate P3HT or P3HT:PCBM solution was spin-cast on top of the

PEDOT:PSS layer. All electrodes were coated with an epoxy coating (Epoxies Inc.) to

prevent exposure of ITO to electrolyte at the electrode edges.

4.2.3. Photoelectrochemistry

A 200 W Xe/Hg lamp (Uhsio America, inc.) was used in combination with a

300-700 nm band pass filter (FSQ-KG3, Newport Corp.) and neutral density filters

(Thorlabs Inc.), to achieve 100 mW/cm2 irradiation, as measured using a broadband

power meter (841-PE, Newport Corporation) equipped with an Ophir thermal detector

head (3A-P-SHV1). A glass photoelectrochemical cell with two large flat windows and five

Page 95: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

77

accessory ports was used and the cell configuration was designed to allow irradiation of

the polymer–electrolyte interface through the electrolyte. A water filter was placed in front

of the electrochemical cell to remove heat. Photoelectrochemical measurements were

performed using a Pine Bipotentiostat (AFC-BP1). Non-aqueous photoelectrochemical

measurements were performed in a 3-electrode configuration using a silver/silver ion

reference electrode (0.01 M AgNO3 in 0.1 M Bu4N PF6) and a Pt wire as the reference and

counter electrodes, respectively. Solutions were purged with nitrogen (Praxair, purity

99.999%) for 1 hour in a glass PEC cell and a positive pressure N2 blanket maintained in

the overhead volume of the cell.

4.3. Results and Discussion

4.3.1. Open Circuit Potential & Onset Potential for Photocurrent

Open circuit potential and linear sweep voltammetry measurements provide useful

information on the photovoltage (Vph) and onset potential (Vonset) of a semiconductor,

respectively. Through careful consideration of experimental conditions, Vph and Vonset can

be used to extract Vbi, thereby providing quantitative information about energy level

alignment and depletion layer width. The major conditions for meaningful open circuit

potential measurements is that the illumination must be sufficiently intense to completely

remove pre-existing band bending at the surface and that the recombination rates of the

photogenerated charges in the semiconductor must not be so fast to nullify the effect of

illumination.203 A common method to determine whether the light intensity is sufficiently

intense to flatten the bands is to plot photovoltage vs. light intensity, as illustrated in Figure

4.1a. At higher intensities the photovoltage becomes saturated and when the above

conditions are met, Vph approximates Vbi. It should be noted here that fast recombination

rates associated with organic semiconductors results in the incomplete removal of band

bending at the semiconductor-electrolyte interface, consequently leading to observed

photovoltage values that are smaller than the built-in potential. Although quantitative

values of Vbi cannot be extracted from the photovoltage measurements performed, the

results remain significant as a clear trend of decreasing photovoltage is observed as

PCBM content is increased.

Through linear sweep voltammetry, Vbi can be estimated from the difference

between the onset potential (Vonset) at the organic photocathode under illumination and the

Page 96: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

78

thermodynamic reduction potential of the redox couple, when kinetic overpotentials are

minimized. Kinetic overpotentials can be minimized either by introducing catalyst layers to

the semiconductor surface or by choosing appropriate redox couples that possess facile

electrochemical redox kinetics at the semiconductor electrode of interest. In this work,

benzoquinone (BZQ), a reversible redox couple possessing a low redox overpotential at

carbon-based electrodes, is used.

Photovoltage measurements were performed on P3HT:PCBM photocathodes with

varying PCBM content and the results plotted in Figure 4.1a. A general trend of decreasing

photovoltage is observed with increasing PCBM content. At higher PCBM ratios a negative

photovoltage is measured, signaling a reversal in photocathode behavior from p-type to

n-type behavior.

Figure 4.1 a) OCP vs. light intensity plot for P3HT (circles), 10 wt% PCBM (squares), 25 wt% PCBM (diamonds) and 50 wt% PCBM (triangles) and b) linear sweep voltammetry results for pristine P3HT (black), 10 wt% PCBM (red), 25 wt% PCBM (green) and 50 wt% PCBM (blue). Device architecture for all devices is ITO/P3HT:PCBM. Measurements were carried out in 0.1 M Bu4N PF6 in ACN with 1 mM BZQ. Potentials measured vs. Ag/Ag+ reference (0.01 M AgNO3). For OCP measurements, the points plotted, and error bars represent the average OCP values measured at each light intensity and their standard deviation, respectively. For LSV measurements, the Y-axis is offset for clarity, vertical dashed lines are drawn from the onset potential of their respective measurements.

The trend of decreasing photovoltage with increasing PCBM content is explained

by considering, firstly, that the extent of energy level alignment experienced by the BZQ

Page 97: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

79

redox potential is dictated by the density of interface states at the organic layer, and

secondly, that the density of interface states increases as PCBM content is increased.

Therefore, at high PCBM content, the BZQ redox potential will align with the charge

neutrality level, resulting in a deviation from ideal Mott-Schottky behaviour, and thereby

reducing the magnitude of band bending. As photovoltage is proportional to band bending,

Vph (and subsequently Vbi) will be reduced at high PCBM concentration. An energy level

diagram depicting the effect of energy level alignment on Vbi is shown in Figure 4.2.

Figure 4.2 Energy level diagram of a) P3HT and b) P3HT:PCBM (50 wt% PCBM) photocathodes on ITO, in contact with the BZQ redox couple.

The reversal of the photoresponse from n-type to p-type for films possessing high

PCBM content is attributed to a non-Ohmic contact between the ITO and organic film

interfaces. This non-Ohmic contact gives rise to a Schottky barrier and results in band

bending at both the ITO-organic interface, as well as the organic-electrolyte interface. In

the case where Schottky barriers are formed at both sides of a semiconductor layer, the

dominant space charge layer will determine the overall photoresponse of the

photoelectrode.204,205 For pristine P3HT films, the Schottky barrier at the organic-

electrolyte interface is approximated by the difference between the Fermi level of P3HT

and the BZQ redox potential. Similarly, the ITO-organic Schottky barrier is approximated

by the difference between the P3HT Fermi level and the ITO work function. For pristine

P3HT films, the larger Schottky barrier at the organic-electrolyte interface dominates,

sweeping holes (majority carriers) towards the back contact and electrons (minority

carriers) towards the electrolyte, resulting in a p-type photovoltage. For P3HT:PCBM films

with high PCBM content, the Schottky barrier at the organic-electrolyte interface is

Page 98: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

80

diminished as the BZQ redox potential is pinned to the charge neutrality level, so much so

that the ITO-organic barrier dominates the photoresponse and an n-type photovoltage is

observed.

Linear sweep voltammograms of P3HT:PCBM photocathodes with varying PCBM

concentrations are shown in Figure 4.1b; vertical lines have been overlaid on the figure

for interpretation of the onset potential at each photocathode. Comparing the

voltammograms, it is apparent that increasing the PCBM content of the photocathode

results in a negative (cathodic) shift in the onset potential for P3HT:PCBM films on ITO in

the presence of BZQ redox couple under chopped illumination. Onset potentials shift from

-0.36 V for P3HT photocathodes to -0.39 V, -0.46 V and -0.51 V for P3HT:PCBM

photocathodes containing 10, 25 and 50 wt% PCBM, respectively. All onset potential

values are compared against an Ag/Ag+ non-aqueous reference electrode.

The combined results of the photovoltage and onset potential measurements show

there is a negative shift in the flat-band potential as PCBM content in the P3HT:PCBM

films is increased, which is consistent with previous reports of P3HT:PCBM films in

solid-state organic electronic devices.197 The addition of PCBM and introduction of

interface states gives rise to an interfacial dipole layer between the organic layer and the

electrolyte, resulting in the alignment of the BZQ redox couple and the charge neutrality

level. The resulting deviation from the Mott-Schottky limit at the semiconductor-electrolyte

interface produces a measurable decrease in Vbi.

4.3.2. Role of a PEDOT:PSS Charge Selective Layer

The incorporation of interfacial materials that promote an Ohmic contact and high

charge selectivity of holes and electrons at their corresponding electrode contacts have

been shown to be important for optimal device performance in organic electronics.206

PEDOT:PSS is commonly employed as a hole transport layer in organic electronics due

to its high conductivity and optical transparency,207 as well as its high work function, which

facilitates Fermi level pinning of the anode to the positive integer charge-transfer state

(EICT+) of the organic active layer.208,209 Through the incorporation of PEDOT:PSS as a

hole transport layer between the ITO and the organic layer, Ohmic contact between the

ITO and organic layer is achieved, eliminating the Schottky barrier previously created at

the ITO-organic interface. This is supported by the photovoltage measurements using

Page 99: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

81

PEDOT:PSS as a hole transport layer, shown in Figure 4.3a, as a p-type photoresponse

is exhibited across all concentrations of PCBM, i.e., no reversal of photoresponse to n-type

is observed. Although the incorporation of a hole transport layer will eliminate any

competing space charge layers at the ITO-organic interface, the formation of the interfacial

dipole layer at the organic-electrolyte interface still occurs when PCBM is present in the

P3HT:PCBM layer. As a result, the photovoltage measured by open circuit potential is

reduced from 0.26 V to 0.16 V and the onset potential shifts cathodically from -0.27 V

to -0.34 V (vs. Ag/Ag+) as PCBM content is increased from 0 wt% through 50 wt%.

Figure 4.3 Figure 3. a) OCP vs. light intensity plot for P3HT (circles), 10 wt% PCBM (squares), 25 wt% PCBM (diamonds) and 50 wt% PCBM (triangles) in P3HT:PCBM and b) linear sweep voltammetry results for pristine P3HT (black), 10 wt% PCBM (red), 25 wt% PCBM (green) and 50 wt% PCBM (blue). Device architecture for all devices is ITO/PEDOT:PSS/P3HT:PCBM. Measurements were carried out in 0.1 M Bu4N PF6 in ACN with 1 mM BZQ. Potentials measured vs. Ag/Ag+ reference (0.01 M AgNO3). For OCP measurements, the points plotted, and error bars represent the average OCP values measured at each light intensity and their standard deviation, respectively. For LSV measurements, the Y-axis is offset for clarity, vertical dashed lines are drawn from the onset potential of their respective measurements.

With the establishment that PCBM exerts a measurable effect on the interfacial

thermodynamics of P3HT:PCBM photoelectrochemical cells, an investigation concerning

the effect of PCBM on the magnitude of photocurrent generated during photoelectrolysis

was warranted. The consequence of any shift in Vbi caused by the presence of PCBM will

Page 100: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

82

manifest itself in a change in the width of the space charge depletion layer, as shown by

Equation 2.

The depletion layer plays an important role in photoelectrochemistry as any

photoinduced charges generated within this region may be separated by the built-in

electric field. If the depletion layer is extended over the entire width of the organic layer it

could eliminate the need for an electron acceptor and formation of a bulk heterojunction

structure. However, the combination of a small relative dielectric constant (typically 2 to 4)

and dopant concentrations on the order of 1016 reported for organic semiconductors

generally lead to depletion layer widths on the order of ~100 nm.196,197,210 As the film

thickness of organic semiconductor photoelectrodes is typically >200 nm, a significant

portion of the active layer does not experience the effects of the built-in electric field of the

depletion layer. Additionally, as the extent of energy alignment between the BZQ redox

couple and the charge neutrality level is increased with increasing PCBM content, the

decrease in Vbi will lead to a further reduction in depletion layer width.

Using Equation 2, the depletion layer widths of the organic photocathodes can be

estimated for the extreme cases where, (i) no energy level pinning occurs (i.e., for pristine

P3HT films), and (ii) where the redox potential of the BZQ couple is fully pinned at the

charge neutrality level (i.e., for P3HT:PCBM films with high PCBM content). In both cases,

we assume a relative dielectric constant, εr, and dopant density, NA, of 3.5 and 5x1016

cm-3, respectively.196,211 The built-in potential, Vbi, at a semiconductor-electrolyte interface

is defined as the difference between the semiconductor Fermi level and the

electrochemical potential of the redox couple in solution. For P3HT films, Vbi is

approximated as the difference between the P3HT HOMO (5.0 eV) and the BZQ redox

potential (4.0 eV). As P3HT is a p-type semiconductor, its Fermi level will reside close to

the HOMO level, allowing the use of the more well-defined HOMO level as a reasonable

approximation. For P3HT:PCBM films with high PCBM content, the BZQ redox potential

is pinned at the charge neutrality level, and as result, Vbi is approximated as the difference

between the P3HT HOMO and the charge neutrality level (4.7 eV).197 Note that using the

P3HT HOMO level to approximate Vbi will result in slightly larger values than if the P3HT

Fermi level were used due to the small energy offset between the HOMO and Fermi levels.

Nevertheless, this method represents a valid approximation of the upper-limit of depletion

layer width that is useful in predicting photocathode behaviour within comparative studies.

Page 101: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

83

The depletion layer widths calculated for P3HT and P3HT:PCBM (50 wt%) films are 88 nm

and 48 nm, respectively.

In thick films, where the depletion layer extends only partially through the organic

layer, it is reasonable to assume that the bulk heterojunction structure will remain an

important factor in the overall device performance, as a large portion of charge generation

and separation will occur in the field-free region of the film. The photoelectrolysis results

of P3HT:PCBM films with different PCBM concentrations, shown in Figure 4.4, corroborate

this hypothesis, as the photocurrents measured at 220 nm thick P3HT:PCBM

photocathodes with a PEDOT:PSS hole transport layer increase with increasing PCBM

content, despite the decrease in Vbi and depletion layer width.

Figure 4.4 Photoelectrolysis using pristine P3HT (black), 10 wt% PCBM (red), 25 wt% PCBM (green) and 50 wt% PCBM (blue) in P3HT:PCBM films. A bias of -0.8 V (vs. Ag/Ag+) was applied for all electrolysis measurements. All electrodes have an active layer thickness of 220 nm and a device architecture of ITO/PEDOT:PSS/P3HT:PCBM. Measurements were carried out in 0.1 M Bu4N PF6 in ACN with 1 mM BZQ.

To probe the effect of the relative thickness the depletion layer extending into the

organic layer, the thickness of the P3HT:PCBM photocathodes was reduced to ~90 nm

so that the depletion layer extends the entire width of the organic film. As a result, the

majority of photogenerated charges should be efficiently separated, regardless of where

they are absorbed, thus photocurrent should be proportional to Vbi, which has been shown

to decrease with increasing PCBM concentration. Figure 4.5 shows electrolysis results for

90 nm thick organic layers, where the measured photocurrent densities are highest for

Page 102: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

84

pristine P3HT films and decrease as PCBM content is increased. These results show that

for thin organic semiconductor films, where the depletion layer width extends through the

majority of the active layer, Vbi plays a dominant role in the generation of photocurrent.

These results are significant as they show that efficient charge separation is achieved in

organic photoelectrodes in the absence of an electron acceptor if the width of the depletion

layer and width of the film thickness are carefully considered.

Figure 4.5 Photoelectrolysis at pristine P3HT (black), 10 wt% PCBM (red), 25 wt% PCBM (green), and 50 wt% PCBM (blue). A bias of -0.8 V (vs. Ag/Ag+) was applied for all electrolysis measurements. All electrodes have a device architecture of ITO/PEDOT:PSS/P3HT:PCBM. Measurements were carried out in 0.1 M Bu4N PF6 in ACN with 1 mM BZQ.

4.4. Conclusion

Photoelectrochemical techniques were carried out using a reversible redox couple

with low kinetic overpotential, BZQ, where a decrease in photovoltage from 0.26 V to

0.16 V and cathodic shift in onset potential from -0.27 V to -0.34 V (vs. Ag/Ag+) is observed

as PCBM content of the P3HT:PCBM photocathodes is increased from 0 through 50 wt%.

The photovoltage and onset potential measurements reveal that the built-in potential of

the photocathodes decreases upon addition of PCBM. The observed decrease in Vbi

appears to be due to interfacial states introduced by PCBM at the semiconductor-

electrolyte interface, forming an interfacial dipole layer that aligns the BZQ redox potential

with the charge neutrality level of the organic layer.

Page 103: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

85

As Vbi is decreased, the depletion layer width within the organic active layer, which

is important for effective charge separation, is also decreased. This has been shown to

influence photocathode performance in thinner photoelectrodes, where efficient charge

separation was achieved in the absence of a PCBM acceptor phase by tuning the

semiconductor thickness to coincide with depletion layer width. For thicker photocathodes,

where the field-free region of the semiconductor layer is significantly larger than the

depletion layer width, the bulk heterojunction donor-acceptor morphology remains crucial

for the separation of photogenerated charges.

The importance of Ohmic contact at the ITO-organic interface is evident when the

photovoltage measurements at P3HT:PCBM photoelectrodes with and without a hole

transport layer are compared. It was discovered that the photoresponse of the electrodes

changed from p-type to n-type at high PCBM content when no hole transport layer was

present. As the redox potential of the electrolyte becomes pinned to the charge neutrality

level of the organic layer at high PCBM content, the Schottky barrier at the organic-

electrolyte interface is reduced. Consequently, the Schottky barrier present at the ITO-

organic interface becomes dominant, causing minority carriers (i.e., electrons) to now be

swept towards the ITO-organic interface and majority carriers (i.e., holes) towards the

organic-electrolyte interface. The incorporation of a hole transport layer between the ITO

and organic layers promotes Ohmic contact, resulting in a p-type photoresponse across

all PCBM concentrations.

Page 104: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

86

Chapter 5. Conclusions and Future Directions

5.1. Conclusions

Although conjugated organic polymers have been the subject of extensive

investigation over the last few decades, they represent an emerging class of materials for

photoelectrochemical applications. As most reports concerning the study of organic

photoelectrochemical cells have focused on improving performance through device

engineering, the fundamental processes that occur at the semiconductor-electrolyte

interface have remained an underreported field of research. Given the difference in

chemical environment compared to their solid-state counterparts, the elucidation of

underlying thermodynamic and electronic processes at organic photoelectrodes

warranted further investigation. The emphasis of the research conducted in this thesis

was, therefore, placed on gaining a better understanding of the thermodynamic and

electronic processes that occur at the semiconductor-electrolyte interface.

In chapter two, an initial investigation into whether P3HT photocathodes were

capable of performing the hydrogen evolution reaction in the absence of catalyst revealed

that despite appropriate thermodynamics, no hydrogen evolution occurs at the P3HT

photocathodes. It was shown quantitatively that any photocurrents measured during

photoelectrolysis was largely due to the presence of trace amounts of dissolved oxygen.

In addition to investigations concerning fundamental interfacial processes, the

advantages of nanostructured electrodes were explored in chapter three, where

P3HT:PCBM nanoparticles were synthesized and used to prepare nanostructured

photocathodes. When coupled with a photoelectrochemically deposited platinum catalyst,

the nanostructured P3HT:PCBM photoelectrodes show enhanced hydrogen evolution

compared to their planar counterparts. The improved performance is largely attributed to

the greater surface area of the nanostructured electrodes, allowing for increased catalyst

loadings.

Although fullerenes have been shown to be an important component in solid-state

organic electronics, little is known about their influence on the semiconductor-electrolyte

Page 105: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

87

interface in photoelectrochemical applications. As such, the influence of fullerenes,

specifically PCBM, at the semiconductor-electrolyte interface was the subject of

investigation in chapter four. It was found that the presence of PCBM at the

semiconductor-electrolyte interface influences the energy level alignment of the redox

couple in solution, leading to the formation of an interfacial dipole layer. As the redox

potential of the electrolyte species aligns with the interfacial density of states brought

about by PCBM molecules at the photocathode surface, the built-in potential and depletion

layer width of the organic semiconductor are diminished. A major consequence of

decreasing depletion layer width is a subsequent decrease in photogenerated charge

separation. When the depletion layer width of the P3HT active layer is controlled to

coincide with the total polymer film thickness, efficient charge separation is possible in the

absence of a PCBM acceptor phase, and the presence of PCBM is, in fact, detrimental to

performance of photocathodes with thin active layers.

Through fundamental studies, it is often possible to gain insights that can be used

as a guideline for the selection of materials or the development of novel materials with

targeted properties for specific applications. Considering the results discussed in chapters

two and four, it should be possible to engineer fullerene-free organic photocathodes

capable of performing the hydrogen evolution reaction through the combination of a hole

transport layer, ensuring Ohmic contact between the conductive substrate and the organic

semiconductor layer; an organic semiconductor with appropriate HOMO and LUMO

energetics; and a hydrogen evolution catalyst. Additionally, further exploration of organic

nanostructures as a strategy to achieve higher photocurrent densities is warranted, based

on the results from chapter three.

5.2. Future Directions

Based on the findings of this thesis, two general avenues are recommended for

the furtherment of organic photoelectrochemical cells; the first being rational polymer

design and the second being further engineering of current polymer systems.

5.2.1. Rational Polymer Design

A significant advantage of conjugated organic polymers is their synthetic tunability,

which can be exploited to tune the optoelectronic properties of materials. This section will

Page 106: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

88

focus on potential synthetic strategies to improve the charge extraction efficiencies of

organic photocathodes. The dielectric constant and built-in potential are of particular

interest, as these properties have been shown to influence the depletion layer width of

semiconductor photoelectrodes significantly. Through the implementation of high

dielectric constant materials and/or maximizing the built-in potential through band-edge

engineering, it may be possible to enhance charge carrier extraction within organic

photocathodes.

Conjugated Organic Polymers with High Dielectric Constants

The implementation of materials with high dielectric constants has the potential to

produce single junction polymer photoelectrodes, capable of overcoming the large binding

energies associated with traditional conjugated organic polymers while avoiding the

drawbacks of the incorporation of fullerene acceptors discussed in Chapter 4, namely

interfacial dipole formation and energy level alignment leading to decreased built-in

potential and depletion layer width. In general, increasing the relative dielectric constant

of conjugated organic polymers can be achieved through the introduction of polar or

polarizable substituents to the polymer structure.212 One such strategy that has been used

to increase the relative dielectric constant of conjugated organic polymers is the

integration of polar oligo(ethylene glycol) side chains into the polymer structure.213

Initial studies of oligo(ethylene glycol) side chain substitution were performed using

2,5-substituted poly(p-phenylene vinylene) (PPV).214 A relative dielectric constant as high

as 5.5 was achieved when tri(ethylene glycol) (TEG) sidechains were introduced at both

the 2 and 5 positions along the polymer backbone of PPV, yielding poly(2,5-bis-

(triethoxymethoxy)-1,4-phenylene vinylene) (diPEO-PPV). Poly[2-methoxy-5-(3′,7′-

dimethyloctyloxy)-1,4-phenylenevinylene] (MDMO-PPV) was used as a reference

material, where a relative dielectric constant of 3 was obtained. The hole mobility values

of diPEO-PPV and MDMO-PPV were found to be similar, in the range of 10-4 cm2/V s. The

comparable hole mobilities of diPEO-PPV and MDMO-PPV is not unexpected as the

conjugated backbone of the polymer, which is responsible for charge transport through

the film, remains largely unaffected by sidechain substitution. Despite the increased

relative dielectric constant, the performance of photovoltaic devices prepared from diPEO-

PPV blended with a PCBM acceptor were very poor. The poor photovoltaic performance

Page 107: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

89

was attributed to poor active layer morphology, specifically the large amount of polymer

aggregation and unfavorable phase separation that occurs at high ethylene glycol content.

Chang et al. were able to show an increase in power conversion efficiency from

6.2% to 7% in OPV devices by introducing TEG functionalized diketopyrrolopyrrole

(TEGDPP) into a triple component copolymer made up of thienylbenzodithiophene

(BDTT), diketopyrrolopyrrole (DPP) and TEGDPP units.215 The TEGDPP/DPP ratio was

gradually increased and it was found that 10% TEGDPP content exhibited optimal

photovoltaic performance when blended with a PC71BM acceptor. The enhancement in

photovoltaic performance is attributed to the ability of the TEG sidechains to promote

structural ordering (i.e., enhanced π-π stacking) within the polymer domains, while precise

control over the amount of TEG incorporated into the polymer allows for control over phase

separation and hence donor-acceptor interfacial area can be maximized. Corroborating

previous reports, it was also found that large ratios of TEG sidechains led to highly

aggregated polymer chains and large phase separation that was detrimental to

photovoltaic performance.

Figure 5.1 Chemical structures of MDMO-PPV, PBD and their oligo(ethylene glycol)-bearing analogues.

Although the aforementioned studies illustrate that uncontrolled aggregation and

phase separation due to high ethylene glycol content along the polymer backbone is

detrimental to photovoltaic performance when using polymer:fullerene blends, these

issues should not be present if fullerene-free photoelectrodes can be successfully

employed in organic photoelectrochemical cells. The combined benefits of greater

Page 108: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

90

structural ordering and increased dielectric constant by way of sidechain modification may

be an effective strategy to enhance charge separation within organic photocathodes;

especially when considering the possibility that the importance of fullerenes may be

reduced when the built-in electric field at the organic-electrolyte interface is sufficiently

strong (see Chapter 4).

Additional synthetic strategies that incorporate the use of polar/polarizable

substituents include backbone fluorination216 and the use of heavier group 16 atoms in

sulfur containing polymers (e.g., selenium or tellurium).217,218

Band-Edge Engineering

Predictive models have previously been employed to illustrate the relationship

between the power conversion efficiency of organic photovoltaics (OPVs), the bandgap of

the conjugated organic polymer, and the LUMO level of the conjugated polymer.219

Maximizing the energy difference between the polymer donor HOMO and the fullerene

acceptor LUMO will maximize the open circuit potential (VOC) in OPVs, however a balance

between maximizing VOC while maintaining a low enough bandgap to ensure adequate

absorption of the solar spectrum is required. Additionally, the LUMO levels of donor

polymers employed in OPVs are constrained by the fullerene acceptor LUMO level (4.3 eV

for PCBM), as the polymer LUMO must lie ~0.3 eV above the fullerene LUMO to ensure

efficient charge separation at the donor-acceptor interface.220 By following the design rules

developed to ensure optimal energetics in OPVs, it should be possible to develop

materials with optimal energetics to carry out the hydrogen evolution reaction (HER) in

organic photoelectrochemical cells. While the design rules should remain largely

unchanged, one notable difference is that the LUMO level of the polymer donor is now

limited by the redox potential of the HER (4.5 eV) rather than the fullerene LUMO. In

organic photoelectrochemical cells, the energy difference between the polymer HOMO

and the redox potential of the electrolyte species is known as the built-in potential (Vbi) of

the system and represents the maximum achievable driving force for photoelectrolysis.221

Maximizing Vbi is therefore desirable for optimal photoectrochemical performance,

however, similar to OPVs, a balance between maximizing Vbi and ensuring low bandgap

is required.

Subsequent work on controlling molecular energy levels through rational design

has successfully been carried out to produce high performance conjugated organic

Page 109: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

91

polymers.222 In general, two synthetic methods have been adopted to tune the energetics

of conjugated polymers: i) through the modification of the backbone structure, achieved

via the copolymerization of various molecular building blocks and ii) through changing the

backbone substituents along the polymer chain.

Hou et. al. incorporated different conjugated units into the backbone structure of

benzodithiophene (BDT)-based polymers and studied the effect of the various units on the

molecular energy levels and bandgap of the synthesized polymers.223 It was found that

the bandgaps of the polymers could be tuned to fall between 1 and 2 eV, depending on

the conjugated unit employed. The way in which the bandgap was lowered, however, was

not uniform across all conjugated units. Polymers containing a thienopyrazine (TPZ) unit,

for example, lowered the optical bandgap value to 1.05 eV, from 2.13 eV for the BDT

homopolymer, by simultaneously lowering the LUMO and raising the HOMO level.

Although the low bandgap is desirable for greater spectral absorption, raising the HOMO

level, of course, reduces VOC which is undesirable for photovoltaic performance. When

benzothiadiazole (BT) units were introduced into the BDT polymers, the optical bandgap

was lowered to 1.70 eV by only depressing the LUMO level, leaving the HOMO

unchanged. By lowering only the LUMO level, VOC is preserved while more efficient

absorption of the solar spectrum is achieved by the lower bandgap.

Alternatively, the use of electron-withdrawing groups as substituents on the

conjugated backbone have been used to effectively lower the HOMO and LUMO levels

with minimal effect on the optical and morphological properties of the polymer.224,225 A

series of BDT-based polymers containing thienothiophene (TT) units and additional

thiophenes as a pi-bridge were synthesized with an increasing number of fluorine atoms

attached to the BDT and/or TT units.225 By attaching two fluorine atoms on the BDT unit

and one fluorine atom on the TT unit (PBT-3F), it was possible to lower the HOMO and

LUMO levels to 5.20 eV and 3.30 eV, respectively, compared to 4.90 eV and 3.10 eV for

the polymer containing no fluorine substituents (PBT-0F). As a result, the VOC values

increased from 0.56 V to 0.78 V and the average power conversion efficiencies from 4.4%

to 8.3% for the fluorine-free and fluorine containing polymers, respectively.

Page 110: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

92

Figure 5.2 Chemical structures of the polymers used to study the effect of different conjugated units223 and fluorinated backbone substituents225 on molecular energy levels.

Backbone fluorination represents an interesting strategy to maximize the built-in

potential of organic photoelectrochemical cells, as not only does the incorporation of

fluorine as a backbone substituent lower the HOMO level but an additional increase in the

relative dielectric constant of the conjugated polymer should be observed. Unfortunately,

the dielectric constants of the materials mentioned here were not measured, however the

synergistic effect of lowering the molecular energy levels and increasing the dielectric

constant may prove to be beneficial for photoelectrochemical applications.

Going forward, the simplest route to employing low bandgap polymers with deep

HOMO levels as photocathode materials is to use commercially available

polymers, such as poly[N-9′-heptadecanyl-2,7-carbazole-alt-5,5-(4′,7′-di-2-thienyl-2′,1′,3′-

benzothiadiazole)] (PCDTBT).226 PCDTBT has HOMO and LUMO levels situated at 5.5 eV

and 3.6 eV, respectively. The next progression in employing materials with optimal

energetics would be to synthesize conjugated polymers that have previously been studied

but are not available commercially. APFO-Green 1 is an example of a polyfluorene-based

material boasting HOMO and LUMO levels of 5.8 eV and 4.1 eV, respectively.227 The

fluorinated BDT-based polymers synthesized in the Hou group represent interesting

materials for use in organic photoelectrochemical cells as they not only possess adequate

Page 111: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

93

energetics, but the fluorinated backbone should increase the relative dielectric constant of

the materials.224,225

Figure 5.3 Chemical structures of PCDTBT and APFO-Green 1

Finally, an entirely new design space can be explored to synthesize conjugated

organic polymers that possess optimal energetics for fullerene-free photoelectrochemical

applications, where the conjugated polymer LUMO levels are no longer constrained by the

fullerene LUMO. Additionally, if the built-in electric field formed within the depletion layer

is sufficiently strong to promote exciton dissociation, the LUMO offset of 0.3 eV previously

needed to ensure efficient charge dissociation at the donor-acceptor interface is no longer

required. As a result, the utilization of conjugated polymers with LUMO levels approaching

the 4.5 eV redox potential of the hydrogen evolution reaction should be possible. The

resulting bandgaps of materials that possess both low-lying HOMO and LUMO levels

could theoretically be comparable to those of inorganic materials (e.g., 1.1 for Si, 1.4 for

GaAs and 1.5 eV for CdTe).

5.2.2. Device Engineering

Polymer Orientation

Polymer orientation has been shown to influence charge carrier mobility in P3HT-

based organic electronics.91,228,229 Charge carrier mobility is an important parameter in

organic electronics as it influences charge extraction and recombination dynamics in

OPVs230 and allows for fast device operation of OFETs.231 It would, therefore, be

reasonable to expect studies dedicated to charge carrier mobility effects in the context of

Page 112: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

94

organic photoelectrochemical cells (OPECs) will lead to the elucidation of additional

underlying processes for optimal performance of organic photoelectrochemical devices.

As mentioned previously, charge carrier mobility within P3HT films has been

shown to be much higher in the π-stacking direction compared to the lamellar stacking

direction.154 The device geometry of OPECs dictates that controlling the polymer

orientation to maximize face-on orientation of the polymer chains would be beneficial, as

charges are collected in the direction normal to the substrate. A few simple methods to

control polymer orientation have been demonstrated, namely by varying the solvent

volatility,228 and by controlling the spin-coating speed.232 The edge-on orientation is

thought to be the thermodynamically favored orientation of P3HT films, therefore

controlling the rate solvent evaporation has a direct correlation to the orientation of

polymer chains during solidification. Spin-coating P3HT from a less volatile solvent, such

as 1,2-dichlorobenzene, will result in a preferentially edge-on orientation of the polymer

chains, whereas spin-coating from a more volatile solvent, such as chloroform, will result

in the polymer chains preferentially aligning in the face-on orientation with respect to the

substrate. As solvent evaporation rate has been shown to be proportional to spin speed,233

faster spin speeds will also result in enhanced face-on orientation.

Although casting from a more volatile solvent at higher speeds promotes face-on

orientation of the polymer chains, a subsequent decrease in crystallinity within the polymer

layer is also observed. Shioya et al. have used polarized multiple-angle incidence

resolution spectrometry to show that as the face-on component increases in a P3HT thin

film, the crystallinity decreases in a linear fashion.234 There are, however, techniques

which use mechanical shear forces to achieve a face-on oriented film with high

crystallinity, such as rubbing and friction transfer. Rubbing is a simple technique which

uses a cylinder covered by a velvet or microfiber cloth being rolled along the polymer-

coated substrate at constant speed and pressure, and friction transfer utilizes a pneumatic

piston to press a polymer pellet against a heated, moving substrate to deposit an oriented

polymer film. Kajiya et al. have reported an 8-fold increase in the out-of-plane hole mobility

in face-on oriented P3HT films after rubbing.235

Page 113: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

95

Figure 5.4 Schematic illustration of the a) rubbing and b) friction transfer processes used to create thin polymer films with controlled polymer chain orientation.236 Reprinted from Brinkmann, M.; Hartmann, L.; Biniek, L.; Tremel, K.; Kayunkid, N. Macromol. Rapid Commun. 2014, 35, 9, Copyright 2013, with permission from WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

Nanostructured Photoelectrodes

Nanostructured organic photoelectrodes have the potential to increase OPEC

performance due to their large surface areas and narrow feature widths that coincide with

the short exciton diffusion lengths of organic polymers. The use of nanostructured organic

photocathodes to perform the hydrogen evolution reaction was introduced in Chapter 3,

and future directions in this area will be discussed below.

The strategy of employing nanostructures to control the morphology of conjugated

organic polymers has indeed been the subject of previous studies, although these have

focused on solid-state organic electronics.96 Nevertheless, the techniques developed for

the fabrication of nanostructured solid-state organic electronics may be used to produce

potentially useful nanostructures for use as photocathodes in organic

photoelectrochemical cells. Given their desirable properties of large surface area and

narrow feature widths, nanowires and nanorods represent the most promising

nanostructures for use as organic photoelectrodes. Nanoimprinting lithography and

templating represent two methods that have been developed to produce organic

nanostructures.

Nanoimprinting lithography utilizes a prefabricated, nanostructured mold to

introduce nanostructures into already-cast, planar polymer films by imprinting the mold

onto the polymer film under pressure. Although the initial fabrication of the nanostructured

Page 114: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

96

mold is time-consuming and expensive, it can be reused to replicate the desired

nanostructure for multiple samples. Nanoarrays of polymethylmethacrylate (PMMA) with

10 nm diameters have been fabricated via nanoimprinting lithography,237 illustrating the

potential of this technology to produce conjugated polymer nanostructures tailored to

overcome their short exciton diffusion lengths. Organic photovoltaic devices were

fabricated using a double nanoimprinting method the create 2D dot patterned

P3HT:PCBM active layers with feature sizes varying from 200-25 nm wide and 55-80 nm

in height.238 It was shown that charge transport and separation is facilitated by the smaller

feature sizes, as evidenced by the increasing short circuit current densities and

photoluminescence quenching as feature size is decreased. Additionally, fullerene-free

polymer solar cells based on P3HT:PTCDI-C13 were fabricated using nanoimprinted P3HT

nanotrenches with feature widths of 50 nm and a depth of 100 nm.239 The authors were

able to enhance the short circuit densities by a factor of 2.5 using the nanoimprinted

devices compared to planar junction devices. From these select publications, it would

seem feasible to fabricate conjugated organic polymer nanostructures with an aspect ratio

between two and three, that also possess feature widths matching the exciton diffusion

lengths of the polymer.

Figure 5.5 Schematic illustration of the nanoimpronting method used to create nanostructured polymer films.

In general, the templating method consists of filling a nanoporous scaffold with

polymer, whereby the scaffold is subsequently removed through selective etching and

freestanding polymer nanostructures remain. This method has been used to produce

P3HT nanorods oriented perpendicularly on ITO with diameters of 50 nm and heights of

150 nm.240 Here, the authors place an anodized aluminum oxide (AAO) template on a

planar P3HT film, and the nanoholes of the template are filled by P3HT via capillary action

after heating the substrate above the melting point of P3HT, under vacuum. Enhanced

photoluminescence quenching over planar P3HT films was observed for the nanorods,

owing to the enhanced exciton separation as the nanofeatures match the exciton diffusion

Page 115: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

97

length of P3HT. Similarly, power conversion efficiencies were enhanced by a factor of

seven when combined with a C60 accepting phase to form organic photovoltaic devices.

This enhancement is largely credited to the increased short circuit current densities. In

addition to capillary action, polymer nanotubes can be created through

electropolymerization, where the nanostructured scaffold is immersed in an electrolyte

containing the desired monomer and subsequent electropolymerization is carried

out.241,242

Figure 5.6 Schematic illustration of the formation of polymer nanostructures using the templating method.242 Copyright 2011, reproduced with permission from the Royal Society of Chemistry.

Page 116: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

98

References

(1) World Energy Outlook 2018; International Energy Agency, 2018.

(2) Krol, R. van de; Grätzel, M. Photoelectrochemical Hydrogen Production; Springer:

New York, 2012.

(3) Aneke, M.; Wang, M. Appl. Energy 2016, 179, 350.

(4) Simpson, A. P.; Lutz, A. E. Int. J. Hydrogen Energy 2007, 32, 4811.

(5) Zeng, K.; Zhang, D. Prog. Energy Combust. Sci. 2010, 36, 307.

(6) Miller, E. L.; Paluselli, D.; Marsen, B.; Rocheleau, R. E. Sol. Energy Mater. Sol.

Cells 2005, 88, 131.

(7) Van De Krol, R.; Liang, Y.; Schoonman, J. J. Mater. Chem. 2008, 18, 2311.

(8) Khaselev, O.; Turner, J. A. Science 1998, 280, 425.

(9) Cheng, W. H.; Richter, M. H.; May, M. M.; Ohlmann, J.; Lackner, D.; Dimroth, F.;

Hannappel, T.; Atwater, H. A.; Lewerenz, H.-J. ACS Energy Lett. 2018, 3, 1795.

(10) Zheng, J.; Liu, X.; Xu, P.; Liu, P.; Yongzhi, Z.; Yang, J. Int. J. Hydrogen Energy

2012, 37, 1048.

(11) Aceves, S. M.; Espinosa-Loza, F.; Ledesma-Orozco, E.; Ross, T. O.; Weisberg, A.

H.; Brunner, T. C.; Kircher, O. Int. J. Hydrogen Energy 2010, 35, 1219.

(12) Suh, M. P.; Park, H. J.; Prasad, T. K.; Lim, D.-W. Chem. Rev. 2012, 112, 782.

(13) Ogden, J. M. Annu. Rev. Energy Environ. 1999, 24, 227.

(14) James, B. D.; Baum, G. N.; Perez, J.; Baum, K. N. Technoeconomic Analysis of

Photoelectrochemical (PEC) Hydrogen Production; Directed Technologies, Inc.:

Arlington, VA, 2009.

(15) Krishnan Rajeshwar. Encycl. Electrochem. 2007, 1.

Page 117: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

99

(16) Fountaine, K. T.; Lewerenz, H. J.; Atwater, H. A. Nat. Commun. 2016, 7, 1.

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

A.; Lewis, N. S. Chem. Rev. 2010, 110, 6446.

(18) Abe, R. J. Photochem. Photobiol. C Photochem. Rev. 2010, 11, 179.

(19) Goto, Y.; Hisatomi, T.; Wang, Q.; Higashi, T.; Ishikiriyama, K.; Maeda, T.; Sakata,

Y.; Okunaka, S.; Tokudome, H.; Katayama, M.; Akiyama, S.; Nishiyama, H.; Inoue,

Y.; Takewaki, T.; Setoyama, T.; Minegishi, T.; Takata, T.; Yamada, T.; Domen, K.

Joule 2018, 2, 509.

(20) Scaife, D. E. Sol. Energy 1980, 25, 41.

(21) Hsieh, Y. H.; Huang, C. P. Colloids and Surfaces 1991, 53, 275.

(22) Maeda, K.; Teramura, K.; Domen, K. J. Catal. 2008, 254, 198.

(23) Minggu, L. J.; Wan Daud, W. R.; Kassim, M. B. Int. J. Hydrogen Energy 2010, 35,

5233.

(24) Bard, A. J. J. Photochem. 1979, 10, 59.

(25) Abe, R.; Sayama, K.; Domen, K.; Arakawa, H. Chem. Phys. Lett. 2001, 344, 339.

(26) Sayama, K.; Mukasa, K.; Abe, R.; Abe, Y.; Arakawa, H. Chem. Commun. 2001,

2416.

(27) Wang, Q.; Hisatomi, T.; Jia, Q.; Tokudome, H.; Zhong, M.; Wang, C.; Pan, Z.;

Takata, T.; Nakabayashi, M.; Shibata, N.; Li, Y.; Sharp, I. D.; Kudo, A.; Yamada, T.;

Domen, K. Nat. Mater. 2016, 15, 611.

(28) Kudo, A.; Miseki, Y. Chem. Soc. Rev. 2009, 38, 253.

(29) Osterloh, F. E. Chem. Mater. 2008, 20, 35.

(30) Fujishima, A.; Honda, K. Nature 1972, 238, 37.

(31) Kumagai, H.; Minegishi, T.; Sato, N.; Yamada, T.; Kubota, J.; Domen, K. J. Mater.

Page 118: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

100

Chem. A 2015, 3, 8300.

(32) Jiang, C.; Moniz, S. J. A.; Wang, A.; Zhang, T.; Tang, J. Chem. Soc. Rev. 2017, 46,

4645.

(33) Sivula, K.; Le Formal, F.; Grätzel, M. ChemSusChem 2011, 4, 432.

(34) Peerakiatkhajohn, P.; Yun, J. H.; Chen, H.; Lyu, M.; Butburee, T.; Wang, L. Adv.

Mater. 2016, 28, 6405.

(35) Nakato, Y.; Ohnishi, T.; Tsubomura, H. Chem. Lett. 1975, 4, 883.

(36) Yamamoto, A.; Yano, S. J. Electrochem. Soc. 1975, 122, 260.

(37) S. Hu; M. R. Shaner; J. A. Beardslee; M. Lichterman; B. S. Brunschwig; N. S. Lewis.

Science 2014, 344, 1005.

(38) Lichterman, M. F.; Carim, A. I.; McDowell, M. T.; Hu, S.; Gray, H. B.; Brunschwig,

B. S.; Lewis, N. S. Energy Environ. Sci. 2014, 7, 3334.

(39) Moniz, S. J. A.; Shevlin, S. A.; Martin, D. J.; Guo, Z. X.; Tang, J. Energy Environ.

Sci. 2015, 8, 731.

(40) Toe, C. Y.; Zheng, Z.; Wu, H.; Scott, J.; Amal, R.; Ng, Y. H. Angew. Chemie - Int.

Ed. 2018, 57, 13613.

(41) Morales-Guio, C. G.; Tilley, S. D.; Vrubel, H.; Grätzel, M.; Hu, X. Nat. Commun.

2014, 5, 3059.

(42) Kaneshiro, J.; Gaillard, N.; Rocheleau, R.; Miller, E. Sol. Energy Mater. Sol. Cells

2010, 94, 12.

(43) Chae, S. Y.; Park, S. J.; Han, S. G.; Jung, H.; Kim, C.-W.; Jeong, C.; Joo, O.-S.;

Min, B. K.; Hwang, Y. J. J. Am. Chem. Soc. 2016, 138, 15673.

(44) Moriya, M.; Minegishi, T.; Kumagai, H.; Katayama, M.; Kubota, J.; Domen, K. J.

Am. Chem. Soc. 2013, 135, 3733.

(45) Kamimura, J.; Bogdanoff, P.; Lähnemann, J.; Hauswald, C.; Geelhaar, L.; Fiechter,

Page 119: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

101

S.; Riechert, H. J. Am. Chem. Soc. 2013, 135, 10242.

(46) Wang, D.; Pierre, A.; Kibria, M. G.; Cui, K.; Han, X.; Bevan, K. H.; Guo, H.; Paradis,

S.; Hakima, A. R.; Mi, Z. Nano Lett. 2011, 11, 2353.

(47) Lee, M. H.; Takei, K.; Zhang, J.; Kapadia, R.; Zheng, M.; Chen, Y. Z.; Nah, J.;

Matthews, T. S.; Chueh, Y. L.; Ager, J. W.; Javey, A. Angew. Chemie - Int. Ed.

2012, 51, 10760.

(48) Oh, I.; Kye, J.; Hwang, S. Nano Lett. 2012, 12, 298.

(49) Gao, L.; Cui, Y.; Wang, J.; Cavalli, A.; Standing, A.; Vu, T. T. T.; Verheijen, M. A.;

Haverkort, J. E. M.; Bakkers, E. P. A. M.; Notten, P. H. L. Nano Lett. 2014, 14, 3715.

(50) Hu, S.; Xiang, C.; Haussener, S.; Berger, A. D.; Lewis, N. S. Energy Environ. Sci.

2013, 6, 2984.

(51) May, M. M.; Lewerenz, H.-J.; Lackner, D.; Dimroth, F.; Hannappel, T. Nat.

Commun. 2015, 6, 8286.

(52) Jang, J. W.; Du, C.; Ye, Y.; Lin, Y.; Yao, X.; Thorne, J.; Liu, E.; McMahon, G.; Zhu,

J.; Javey, A.; Guo, J.; Wang, D. Nat. Commun. 2015, 6, 7447.

(53) Ding, C.; Qin, W.; Wang, N.; Liu, G.; Wang, Z.; Yan, P.; Shi, J.; Li, C. Phys. Chem.

Chem. Phys. 2014, 16, 15608.

(54) McCrory, C. C. L.; Jung, S.; Ferrer, I. M.; Chatman, S. M.; Peters, J. C.; Jaramillo,

T. F. J. Am. Chem. Soc. 2015, 137, 4347.

(55) Suen, N. T.; Hung, S. F.; Quan, Q.; Zhang, N.; Xu, Y. J.; Chen, H. M. Chem. Soc.

Rev. 2017, 46, 337.

(56) Zou, X.; Zhang, Y. Chem. Soc. Rev. 2015, 44, 5148.

(57) Wahl, A.; Augustynski, J. J. Phys. Chem. B 1998, 102, 7820.

(58) Feng, X.; Zhu, K.; Frank, A. J.; Grimes, C. A.; Mallouk, T. E. Angew. Chemie - Int.

Ed. 2012, 51, 2727.

Page 120: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

102

(59) Shankar, K.; Basham, J. I.; Allam, N. K.; Varghese, O. K.; Mor, G. K.; Feng, X.;

Paulose, M.; Seabold, J. A.; Choi, K.-S.; Grimes, C. A. J. Phys. Chem. C 2009, 113,

6327.

(60) Cho, I. S.; Chen, Z.; Forman, A. J.; Kim, D. R.; Rao, P. M.; Jaramillo, T. F.; Zheng,

X. Nano Lett. 2011, 11, 4978.

(61) Björkstén, U.; Moser, J.; Grätzel, M. Chem. Mater. 1994, 6, 858.

(62) Kavan, L.; Grätzel, M. Electrochim. Acta 1995, 40, 64.

(63) Schwoerer, M.; Wolf, H. C. Organic Molecular Solids; Wiley-VCH Verlag GmbH &

Co. KGaA: Weinheim, 2007.

(64) Shirota, Y. J. Mater. Chem. 2005, 15, 75.

(65) Zhang, W.; Smith, J.; Watkins, S. E.; Gysel, R.; McGehee, M.; Salleo, A.;

Kirkpatrick, J.; Ashraf, S.; Anthopoulos, T.; Heeney, M.; McCulloch, I. J. Am. Chem.

Soc. 2010, 132, 11437.

(66) Noriega, R.; Rivnay, J.; Vandewal, K.; Koch, F. P. V; Stingelin, N.; Smith, P.; Toney,

M. F.; Salleo, A. Nat. Mater. 2013, 12, 1038.

(67) Venkateshvaran, D.; Nikolka, M.; Sadhanala, A.; Lemaur, V.; Zelazny, M.; Kepa,

M.; Hurhangee, M.; Kronemeijer, A. J.; Pecunia, V.; Nasrallah, I.; Romanov, I.;

Broch, K.; McCulloch, I.; Emin, D.; Olivier, Y.; Cornil, J.; Beljonne, D.; Sirringhaus,

H. Nature 2014, 515, 384.

(68) Köhler, A.; Bässler, H. Electronic Processes in Organic Semiconductors; Wiley-

VCH Verlag GmbH & Co. KGaA: Weinheim, 2015.

(69) Kasha, M. Discuss. Faraday Soc. 1950, 9, 14.

(70) Letheby, H. Chem. Soc. 1862, 15, 161.

(71) Natta, G.; Mazzanti, G.; Corradini, P. Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Nat.

Rend. 1958, 25, 3.

(72) Dall’Olio, A.; Dascola, Y.; Varacco, V.; Bocchi, V. C. R. Acad. Sci. Ser. C. 1968,

Page 121: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

103

267, 433.

(73) Ito, T.; Shirakawa, H.; Ikeda, S. J. Polym. Sci. Part A Polym. Chem. 1996, 34, 2533.

(74) Chiang, C. K.; Fincher, C. B.; Park, Y. W.; Heeger, A. J.; Shirakawa, H.; Louis, E.

J.; Gau, S. C.; MacDiarmid, A. G. Phys. Rev. Lett 1977, 39, 1098.

(75) Wienk, M. M.; Kroon, J. M.; Verhees, W. J. H.; Knol, J.; Hummelen, J. C.; Van Hal,

P. A.; Janssen, R. A. J. Angew. Chemie - Int. Ed. 2003, 42, 3371.

(76) Zhao, G. J.; He, Y. J.; Li, Y. Adv. Mater. 2010, 22, 4355.

(77) Li, G.; Zhu, R.; Yang, Y. Nat. Photonics 2012, 6, 153.

(78) Kularatne, R. S.; Magurudeniya, H. D.; Sista, P.; Biewer, M. C.; Stefan, M. C. J.

Polym. Sci. Part A Polym. Chem. 2013, 51, 743.

(79) Marszalek, T.; Li, M.; Pisula, W. Chem. Commun 2016, 52, 10938.

(80) Müllen, K.; Pisula, W. J. Am. Chem. Soc. 2015, 137, 9503.

(81) Zhao, J.; Li, Y.; Yang, G.; Jiang, K.; Lin, H.; Ade, H.; Ma, W.; Yan, H. Nat. Energy

2016, 1, 15027.

(82) Liu, Y.; Zhao, J.; Li, Z.; Mu, C.; Ma, W.; Hu, H.; Jiang, K.; Lin, H. Nat. Commun.

2014, 5, 5293.

(83) He, Z.; Xiao, B.; Liu, F.; Wu, H.; Yang, Y.; Xiao, S.; Wang, C.; Russell, T. P.; Cao,

Y. Nat. Photonics 2015, 9, 174.

(84) Chen, J.; Cui, C.; Li, Y.; Zhou, L.; Ou, Q. Adv. Funct. Mater. 2015, 27, 1035.

(85) Elsenbaumer, R. L.; Jen, K. Y.; Oboodi, R. Synth. Met. 1986, 15, 169.

(86) Yamamoto, T.; Sanechika, K.; Yamamoto, A. J. Polym. Sci. Polym. Lett. Ed. 1980,

18, 9.

(87) Lin, J. W.-P.; Dudek, L. P. J. Polym. Sci. Polym. Chem. Ed. 1980, 18, 2869.

(88) Friedel, B.; McNeill, C. R.; Greenham, N. C. Chem. Mater. 2010, 22, 3389.

Page 122: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

104

(89) Kim, Y.; Cook, S.; Tuladhar, S. M.; Choulis, S. A.; Nelson, J.; Durrant, J. R.; Bradley,

D. D. C.; Giles, M.; McCulloch, I.; Ha, C.-S.; Ree, M. Nat. Mater. 2006, 5, 197.

(90) Jiang, X.; Patil, R.; Harima, Y.; Ohshita, J.; Kunai, A. J. Phys. Chem. B 2005, 109,

221.

(91) Sirringhaus, H.; Brown, P. J.; Friend, R. H.; Nielsen, M. M.; Bechgaard, K.; Spiering,

A. J. H. Nature 1999, 401, 685.

(92) Ohkita, H.; Ito, S. Polymer 2011, 52, 4397.

(93) Tang, C. W. Appl. Phys. Lett. 1986, 48, 183.

(94) Sariciftci, N. S.; Smilowitz, L.; Heeger, a. J.; Wudl, F. Science 1992, 258, 1474.

(95) Yu, G.; Gao, J.; Hummelen, J. C.; Wudl, F.; Heeger, A. J. Science 1995, 270, 1789.

(96) Chen, J. T.; Hsu, C. S. Polym. Chem. 2011, 2, 2707.

(97) Shirakawa, H.; Ikeda, S.; Aizawa, M.; Yoshitake, J.; Suzuki, S. Synth. Met. 1981, 4,

43.

(98) Yanagida, S.; Kabumoto, A.; Mizumoto, K.; Pac, C.; Yoshino, K. J. Chem. Soc.

Chem. Commun. 1985, 8, 474.

(99) Zhao, Z.; Sun, Y.; Dong, F. Nanoscale 2015, 7, 15.

(100) Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J. M.; Domen,

K.; Antonietti, M. Nat. Mater. 2009, 8, 76.

(101) Schwab, M. G.; Hamburger, M.; Feng, X.; Shu, J.; Spiess, H. W.; Wang, X.;

Antonietti, M.; Müllen, K. Chem. Commun. 2010, 46, 8932.

(102) Sprick, R. S.; Jiang, J.-X.; Bonillo, B.; Ren, S.; Ratvijitvech, T.; Guiglion, P.;

Zwijnenburg, M. a.; Adams, D. J.; Cooper, A. I. J. Am. Chem. Soc. 2015, 137, 3265.

(103) Woods, D. J.; Sprick, R. S.; Smith, C. L.; Cowan, A. J.; Cooper, A. I. Adv. Energy

Mater. 2017, 7, 1700479.

Page 123: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

105

(104) El-Rashiedy, O. A.; Holdcroft, S. J. Phys. Chem. 1996, 100, 5481.

(105) Lanzarini, E.; Antognazza, M. R.; Biso, M.; Ansaldo, A.; Laudato, L.; Bruno, P.;

Metrangolo, P.; Resnati, G.; Ricci, D.; Lanzani, G. J. Phys. Chem. C 2012, 116,

10944.

(106) Gustafson, M. P.; Clark, N.; Winther-Jensen, B.; MacFarlane, D. R. Electrochim.

Acta 2014, 140, 309.

(107) Suppes, G. M.; Fortin, P. J.; Holdcroft, S. J. Electrochem. Soc. 2015, 162, H551.

(108) Bourgeteau, T.; Tondelier, D.; Geffroy, B.; Brisse, R.; Laberty-Robert, C.;

Campidelli, S.; de Bettignies, R.; Artero, V.; Palacin, S.; Jousselme, B. Energy

Environ. Sci. 2013, 6, 2706.

(109) Mihailetchi, V. D.; Blom, P. W. M.; Hummelen, J. C.; Rispens, M. T. J. Appl. Phys.

2003, 94, 6849.

(110) White, M. S.; Olson, D. C.; Shaheen, S. E.; Kopidakis, N.; Ginley, D. S. Appl. Phys.

Lett. 2006, 89, 143517.

(111) Hau, S. K.; Yip, H.-L.; Leong, K.; Jen, A. K.-Y. Org. Electron. 2009, 10, 719.

(112) Peumans, P.; Forrest, S. R. Appl. Phys. Lett. 2001, 79, 126.

(113) Germack, D. S.; Chan, C. K.; Joseph Kline, R.; Fischer, D. A.; Gundlach, D. J.;

Toney, M. F.; Richter, L. J.; Delongchamp, D. M. Macromolecules 2010, 43, 3828.

(114) Queyriaux, N.; Kaeffer, N.; Morozan, A.; Chavarot-Kerlidou, M.; Artero, V. J.

Photochem. Photobiol. C Photochem. Rev. 2015, 25, 90.

(115) Bellani, S.; Najafi, L.; Martín-García, B.; Ansaldo, A.; Del Rio Castillo, A. E.; Prato,

M.; Moreels, I.; Bonaccorso, F. J. Phys. Chem. C 2017, 121, 21887.

(116) Mezzetti, A.; Fumagalli, F.; Alfano, A.; Iadicicco, D.; Antognazza, M. R.; di Fonzo,

F. Faraday Discuss. 2017, 198, 433.

(117) Rojas, H. C.; Bellani, S.; Fumagalli, F.; Tullii, G.; Leonardi, S.; Mayer, M. T.;

Schreier, M.; Gra, M.; Lanzani, G.; Fonzo, F. Di; Antognazza, M. R. Energy Environ.

Page 124: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

106

Sci. 2016, 9, 3710.

(118) Rojas, H. C.; Marra, G.; Meda, L.; Lanzani, G.; Mayer, M. T.; Antognazza, M. R.;

Fonzo, F. Di. J. Mater. Chem. A 2016, 4, 2178.

(119) Belarb, E.; Blas-ferrando, V. M.; Haro, M.; Maghraoui-meherzi, H. Chem. Eng. Sci.

2016, 154, 143.

(120) Bellani, S.; Najafi, L.; Capasso, A.; Del Rio Castillo, A. E.; Antognazza, M. R.;

Bonaccorso, F. J. Mater. Chem. A 2017, 5, 4384.

(121) Ghadirzadeh, A.; Fumagalli, F.; Mezzetti, A.; Bellani, S.; Meda, L.; Antognazza, M.

R.; Di Fonzo, F. ChemPhotoChem 2018, 283.

(122) Bourgeteau, T.; Tondelier, D.; Geffroy, B.; Brisse, R.; Cornut, R.; Artero, V.;

Jousselme, B. ACS Appl. Mater. Interfaces 2015, 7, 16395.

(123) Haro, M.; Solis, C.; Molina, G.; Otero, L. A.; Bisquert, J.; Gimenez, S.; Guerrero, A.

J. Phys. Chem. C 2015, 119, 6488.

(124) Huang, S.; Zhang, H.; Wu, Z.; Kong, D.; Lin, D.; Fan, Y.; Yang, X.; Zhong, Z.;

Huang, S.; Jiang, Z.; Cheng, C. ACS Appl. Mater. Interfaces 2014, 6, 12111.

(125) Paracchino, A.; Laporte, V.; Sivula, K.; Grätzel, M.; Thimsen, E. Nat. Mater. 2011,

10, 456.

(126) Gu, J.; Aguiar, J. A.; Ferrere, S.; Steirer, K. X.; Yan, Y.; Xiao, C.; Young, J. L.; Al-

Jassim, M.; Neale, N. R.; Turner, J. A. Nat. Energy 2017, 2, 16192.

(127) Rojas, H. C.; Bellani, S.; Sarduy, E. A.; Fumagalli, F.; Mayer, M. T.; Schreier, M.;

Grätzel, M.; Di Fonzo, F.; Antognazza, M. R. ACS Omega 2017, 2, 3424.

(128) Guijarro, N.; Prévot, M. S.; Yu, X.; Jeanbourquin, X. A.; Bornoz, P.; Bourée, W.;

Johnson, M.; Le Formal, F.; Sivula, K. Adv. Energy Mater. 2016, 6, 1501949.

(129) Bellani, S.; Ghadirzadeh, A.; Meda, L.; Savoini, A.; Tacca, A.; Marra, G.; Meira, R.;

Morgado, J.; Di Fonzo, F.; Antognazza, M. R. Adv. Funct. Mater. 2015, 25, 4531.

(130) Jakešová, M.; Apaydin, D. H.; Sytnyk, M.; Oppelt, K.; Heiss, W.; Sariciftci, N. S.;

Page 125: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

107

Głowacki, E. D. Adv. Funct. Mater. 2016, 26, 5248.

(131) Park, S.-Y.; Kim, M.; Jung, J.; Heo, J.; Hong, E. M.; Choi, S. M.; Lee, J.-Y.; Cho,

S.; Hong, K.; Lim, D. C. J. Power Sources 2017, 341, 411.

(132) Bornoz, P.; Prévot, M. S.; Yu, X.; Guijarro, N.; Sivula, K. J. Am. Chem. Soc. 2015,

137, 15338.

(133) Shao, D.; Cheng, Y.; He, J.; Feng, D.; Zheng, L.; Zheng, L.; Zhang, X.; Xu, J.; Wang,

W.; Wang, W.; Lu, F.; Dong, H.; Li, L.; Liu, H.; Zheng, R.; Liu, H. ACS Catal. 2017,

7, 5308.

(134) Shao, D.; Zheng, L.; Feng, D.; He, J.; Zhang, R.; Liu, H.; Zhang, X.; Lu, Z.; Wang,

W.; Wang, W.; Lu, F.; Dong, H.; Cheng, Y.; Liiu, H.; Zheng, R. J. Mater. Chem. A

2018, 6, 4032.

(135) Chen, Y.; Chen, H.; Tian, H. Chem. Commun. 2015, 51, 11508.

(136) María Girón, R.; Marco-Martínez, J.; Bellani, S.; Insuasty, A.; Comas Rojas, H.;

Tullii, G.; Antognazza, M. R.; Filippone, S.; Martín, N. J. Mater. Chem. A 2016, 4,

14284.

(137) Varga, A.; Endrödi, B.; Hornok, V.; Visy, C.; Janáky, C. J. Phys. Chem. C 2015,

119, 28020.

(138) Lai, L.-H.; Gomulya, W.; Berghuis, M.; Protesescu, L.; Detz, R. J.; Reek, J. N. H.;

Kovalenko, M. V.; Loi, M. A. ACS Appl. Mater. Interfaces 2015, 7, 19083.

(139) Lin, Z.; Ji, L.; Zhang, X. Mater. Lett. 2009, 63, 2115.

(140) Mosconi, E.; Salvatori, P.; Saba, M. I.; Mattoni, A.; Bellani, S.; Bruni, F.; Santiago

Gonzalez, B.; Antognazza, M. R.; Brovelli, S.; Lanzani, G.; Li, H.; Brédas, J. L.; De

Angelis, F. ACS Energy Lett. 2016, 1, 454.

(141) Bellani, S.; Porro, M.; Caddeo, C.; Saba, M. I.; Miranda, P. B.; Mattoni, A.; Lanzani,

G.; Antognazza, M. R. J. Mater. Chem. B 2015, 3, 6429.

(142) Guerrero, A.; Haro, M.; Bellani, S.; Antognazza, M. R.; Meda, L.; Gimenez, S.;

Page 126: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

108

Bisquert, J. Energy Environ. Sci. 2014, 7, 3666.

(143) Tullii, G.; Desii, A.; Bossio, C.; Bellani, S.; Colombo, M.; Martino, N.; Antognazza,

M. R.; Lanzani, G. Org. Electron. 2017, 46, 88.

(144) Chen, Z.; Dinh, H. N.; Miller, E. Photoelectrochemical Water Splitting Standards,

Experimental Methods, and Protocols; Springer: New York, 2013.

(145) Nozik, A. J. Nature 1975, 257, 383.

(146) Bak, T.; Nowotny, J.; Rekas, M.; Sorrell, C. . Int. J. Hydrogen Energy 2002, 27, 991.

(147) Gasiorowski, J.; Kollender, J. P.; Hingerl, K.; Sariciftci, N. S.; Mardare, A. I.; Hassel,

A. W. Phys. Chem. Chem. Phys. 2014, 16, 3739.

(148) Deibel, C.; MacK, D.; Gorenflot, J.; Schöll, A.; Krause, S.; Reinert, F.; Rauh, D.;

Dyakonov, V. Phys. Rev. B 2010, 81, 085202.

(149) Brown, P. J.; Thomas, D. S.; Köhler, A.; Wilson, J. S.; Kim, J. S.; Ramsdale, C. M.;

Sirringhaus, H.; Friend, R. H. Phys. Rev. B 2003, 67, 064203.

(150) Spano, F. C.; Silva, C. Annu. Rev. Phys. Chem. 2014, 65, 477.

(151) Spano, F. C. Acc. Chem. Res. 2010, 43, 429.

(152) Hexemer, A.; Müller-Buschbaum, P. IUCrJ 2015, 2, 106.

(153) Bragg, W. L. Proc. Cambridge Philos. Soc. 1913, 17, 43.

(154) O’Connor, B.; Kline, R. J.; Conrad, B. R.; Richter, L. J.; Gundlach, D.; Toney, M. F.;

DeLongchamp, D. M. Adv. Funct. Mater. 2011, 21, 3697.

(155) Li, Z.; Wang, Y.; Shen, J.; Liu, W.; Sun, X. Opt. Lasers Eng. 2014, 56, 94.

(156) Chin, D. H.; Chiericato, G.; Nanni, E. J.; Sawyer, D. T. J. Am. Chem. Soc. 1982,

104, 1296.

(157) Loewe, R. S.; Khersonsky, S. M.; McCullough, R. D. Adv. Mater. 1999, 11, 250.

(158) Chen, T.; Wu, X.; Rieke, R. J. Am. Chem. Soc. 1995, 117, 233.

Page 127: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

109

(159) Mao, H.; Xu, B.; Holdcroft, S. Macromolecules 1993, 26, 1163.

(160) Costentin, C.; Evans, D. H.; Robert, M.; Savéant, J. M.; Singh, P. S. J. Am. Chem.

Soc. 2005, 127, 12490.

(161) Pozo-gonzalo, C.; Torriero, A. A. J.; Forsyth, M.; Macfarlane, D. R.; Howlett, P. C.

J. Phys. Chem. Lett. 2013, 4, 4.

(162) Choudhury, B.; Weedon, A.; Bolton, J. Langmuir 1998, 14, 6199.

(163) Abdou, M. S. A.; Holdcroft, S. Macromolecules 1993, 26, 2954.

(164) Abdou, M. S. a; Orfino, F. P.; Son, Y.; Holdcroft, S. J. Am. Chem. Soc. 1997, 119,

4518.

(165) Bellani, S.; Fazzi, D.; Bruno, P.; Giussani, E.; Canesi, E. V.; Lanzani, G.;

Antognazza, M. R. J. Phys. Chem. C 2014, 118, 6291.

(166) Foley, J. M.; Price, M. J.; Feldblyum, J. I.; Maldonado, S. Energy Environ. Sci. 2012,

5, 5203.

(167) Jung, J.-Y.; Park, M.-J.; Li, X.; Kim, J.-H.; Wehrspohn, R. B.; Lee, J.-H. J. Mater.

Chem. A 2015, 3, 9456.

(168) Chen, F.; Zhu, Q.; Wang, Y.; Cui, W.; Su, X.; Li, Y. ACS 2016, 8, 31025.

(169) Clafton, S. N.; Huang, D. M.; Massey, W. R.; Kee, T. W. J. Phys. Chem. B 2013,

117, 4626.

(170) Holmes, N. P.; Burke, K. B.; Sista, P.; Barr, M.; Magurudeniya, H. D.; Stefan, M. C.;

Kilcoyne, a. L. D.; Zhou, X.; Dastoor, P. C.; Belcher, W. J. Sol. Energy Mater. Sol.

Cells 2013, 117, 437.

(171) Labastide, J. A.; Baghgar, M.; Dujovne, I.; Venkatraman, B. H.; Ramsdell, D. C.;

Venkataraman, D.; Barnes, M. D. J. Phys. Chem. Lett. 2011, 2, 2089.

(172) Pedersen, E. B. L.; Pedersen, M. C.; Simonsen, S. B.; Brandt, R. G.; Böttiger, a.

P. L.; Andersen, T. R.; Jiang, W.; Xie, Z. Y.; Krebs, F. C.; Arleth, L.; Andreasen, J.

W. J. Mater. Chem. A 2015, 3, 17022.

Page 128: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

110

(173) Nagarjuna, G.; Baghgar, M.; Labastide, J. A.; Algaier, D. D.; Barnes, M. D.;

Venkataraman, D.; Amherst, M.; Street, N. P.; States, U. ACS Nano 2012, 12,

10750.

(174) Roy, P.; Jha, A.; Dasgupta, J. Nanoscale 2016, 8, 2768.

(175) Ulum, S.; Holmes, N.; Darwis, D.; Burke, K.; David Kilcoyne, a. L.; Zhou, X.;

Belcher, W.; Dastoor, P. Sol. Energy Mater. Sol. Cells 2013, 110, 43.

(176) Darwis, D.; Holmes, N.; Elkington, D.; Kilcoyne, A. L. D.; Bryant, G.; Zhou, X.;

Dastoor, P.; Belcher, W. Sol. Energy Mater. Sol. Cells 2014, 121, 99.

(177) Gärtner, S.; Christmann, M.; Sankaran, S.; Röhm, H.; Prinz, E.; Penth, F.; Pütz, A.;

Türeli, A. E.; Penth, B.; Baumstümmler, B.; Colsmann, A. Adv. Funct. Mater. 2014,

26, 6653.

(178) Millstone, J. E.; Kavulak, D. F. J.; Woo, C. H.; Holcombe, T. W.; Westling, E. J.;

Briseno, A. L.; Toney, M. F.; Fréchet, J. M. J. Langmuir 2010, 26, 13056.

(179) Darwis, D.; Holmes, N.; Elkington, D.; David Kilcoyne, a. L.; Bryant, G.; Zhou, X.;

Dastoor, P.; Belcher, W. Sol. Energy Mater. Sol. Cells 2014, 121, 99.

(180) Schwarz, K. N.; Farley, S. B.; Smith, T. A.; Ghiggino, K. P. Nanoscale 2015, 7,

19899.

(181) Machui, F.; Langner, S.; Zhu, X.; Abbott, S.; Brabec, C. J. Sol. Energy Mater. Sol.

Cells 2012, 100, 138.

(182) Gärtner, S.; Christmann, M.; Sankaran, S.; Röhm, H.; Prinz, E.-M.; Penth, F.; Pütz,

A.; Türeli, A. E.; Penth, B.; Baumstümmler, B.; Colsmann, A. Adv. Mater. 2014, 26,

6653.

(183) Ehrenreich, P.; Birk, S. T.; Zimmermann, E.; Hu, H.; Kim, K.; Weickert, J.; Pfadle,

T.; Schmidt-mende, L. Sci. Rep. 2016, 6, 32434.

(184) Erb, B. T.; Zhokhavets, U.; Gobsch, G.; Raleva, S.; Stühn, B.; Schilinsky, P.;

Waldauf, C.; Brabec, C. J. Adv. Funct. Mater. 2005, 15, 1193.

Page 129: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

111

(185) Lunt, R. R.; Giebink, N. C.; Belak, A. A.; Benziger, J. B.; Forrest, S. R.; Lunt, R. R.;

Giebink, N. C.; Belak, A. A.; Benziger, J. B.; Forrest, S. R. J. Appl. Phys. 2009, 105,

053711.

(186) Suppes, G.; Ballard, E.; Holdcroft, S. Polym. Chem. 2013, 4, 5345.

(187) van de Krol, R.; Grätzel, M. Photoelectrochemical Hydrogen Production; Springer:

New York, 2012; Vol. 102.

(188) Brülle, T.; Denisenko, A.; Sternschulte, H.; Stimming, U. Phys. Chem. Chem. Phys.

2011, 13, 12883.

(189) Bellani, S.; Ghadirzadeh, A.; Meda, L.; Savoini, A.; Tacca, A.; Marra, G.; Meira, R.;

Morgado, J.; Fonzo, F. Di; Antognazza, M. R. Adv. Funct. Mater. 2015, 25, 4531.

(190) Chowdhury, P.; Fortin, P.; Suppes, G.; Holdcroft, S. Macromol. Chem. Phys. 2016,

217, 1119.

(191) Scharber, M. C.; Sariciftci, N. S. Prog. Polym. Sci. 2013, 38, 1929.

(192) Vázquez, H.; Flores, F.; Oszwaldowski, R.; Ortega, J.; Pérez, R.; Kahn, A. Appl.

Surf. Sci. 2004, 234, 107.

(193) Crispin, X.; Geskin, V.; Crispin, A.; Cornil, J.; Lazzaroni, R.; Salaneck, W. R.;

Brédas, J.-L. J. Am. Chem. Soc. 2002, 124, 8131.

(194) Ishii, H.; Sugiyama, K.; Ito, E.; Seki, K. Adv. Mater. 1999, 11, 605.

(195) Hwang, J.; Wan, A.; Kahn, A. Mater. Sci. Eng. R Reports 2009, 64, 1.

(196) Bisquert, J.; Garcia-Belmonte, G.; Munar, A.; Sessolo, M.; Soriano, A.; Bolink, H. J.

Chem. Phys. Lett. 2008, 465, 57.

(197) Guerrero, A.; Marchesi, L. F.; Boix, P. P.; Ruiz-Raga, S.; Ripolles-Sanchis, T.;

Garcia-Belmonte, G.; Bisquert, J. ACS Nano 2012, 6, 3453.

(198) Flores, F.; Ortega, J.; Vázquez, H. Phys. Chem. Chem. Phys. 2009, 11, 8658.

(199) Vázquez, H.; Flores, F.; Kahn, A. Org. Electron. physics, Mater. Appl. 2007, 8, 241.

Page 130: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

112

(200) Horowitz, G.; Allongue, P.; Cachet, H. J. Electrochem. Soc. 1984, 131, 2563.

(201) Kar, S.; Rajeshwar, K.; Singh, P.; DuBow, J. Sol. Energy 1979, 23, 129.

(202) Kumar, A.; Wilisch, W. C. A.; Lewis, N. S. Crit. Rev. Solid State Mater. Sci. 1993,

18, 327.

(203) Chen, Z.; Dinh, H. N.; Miller, E. Photoelectrochemical water splitting: standards,

experimental methods, and protocols; Springer: New York, 2013.

(204) Siripala, W. In Trends in Solar Energy Research; Hough, T., Ed.; Nova Science

Publishers, Inc.: New York, 2006; pp. 1–20.

(205) Siripala, W.; Perera, L. D. R. D.; De Silva, K. T. L.; Jayanetti, J. K. D. S.;

Dharmadasa, I. M. Sol. Energy Mater. Sol. Cells 1996, 44, 251.

(206) Yip, H.-L.; Jen, A. K.-Y. Energy Environ. Sci. 2012, 5, 5994.

(207) Groenendaal, L.; Jonas, F.; Freitag, D.; Pielartzik, H.; Reynolds, J. R. Adv. Mater.

2000, 12, 481.

(208) Xu, Z.; Chen, L.-M.; Chen, M.-H.; Li, G.; Yang, Y. Appl. Phys. Lett. 2009, 95,

013301.

(209) Sehati, P.; Braun, S.; Lindell, L.; Liu, X.; Andersson, L. M.; Fahlman, M. IEEE J.

Sel. Top. Quantum Electron. 2010, 16, 1718.

(210) Dibb, G. F. A.; Muth, M.-A.; Kirchartz, T.; Engmann, S.; Hoppe, H.; Gobsch, G.;

Thelakkat, M.; Blouin, N.; Tierney, S.; Carrasco-Orozco, M.; Durrant, J. R.; Nelson,

J. Sci. Rep. 2013, 3, 3335.

(211) Arnab, S. M.; Kabir, M. Z. J. Appl. Phys. 2014, 115, 034504.

(212) Brebels, J.; Manca, J. V.; Lutsen, L.; Vanderzande, D.; Maes, W. J. Mater. Chem.

A 2017, 5, 24037.

(213) Torabi, S.; Jahani, F.; Van Severen, I.; Kanimozhi, C.; Patil, S.; Havenith, R. W. A.;

Chiechi, R. C.; Lutsen, L.; Vanderzande, D. J. M.; Cleij, T. J.; Hummelen, J. C.;

Koster, L. J. A. Adv. Funct. Mater. 2015, 25, 150.

Page 131: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

113

(214) Breselge, M.; Van Severen, I.; Lutsen, L.; Adriaensens, P.; Manca, J.;

Vanderzande, D.; Cleij, T. Thin Solid Films 2006, 511–512, 328.

(215) Chang, W. H.; Gao, J.; Dou, L.; Chen, C. C.; Liu, Y.; Yang, Y. Adv. Energy Mater.

2014, 4, 1300864.

(216) Lu, Y.; Xiao, Z.; Yuan, Y.; Wu, H.; An, Z.; Hou, Y.; Gao, C.; Huang, J. J. Mater.

Chem. C 2013, 1, 630.

(217) Patra, A.; Bendikov, M. J. Mater. Chem. 2010, 20, 422.

(218) Carrera, E. I.; Seferos, D. S. Macromolecules 2015, 48, 297.

(219) Scharber, M. C.; Muehlbacher, D.; Koppe, M.; Denk, P.; Waldauf, C.; Heeger, A.

J.; Brabec, C. J. Adv. Mater. 2006, 18, 789.

(220) Brédas, J. L.; Beljonne, D.; Coropceanu, V.; Cornil, J. Chem. Rev. 2004, 104, 4971.

(221) Gerischer, H. Pure Appl. Chem. 1980, 52, 2649.

(222) Li, Y. Acc. Chem. Res. 2012, 45, 723.

(223) Hou, J.; Park, M.; Zhang, S.; Yao, Y.; Chen, L.; Li, J.; Yang, Y. Macromolecules

2008, 41, 6012.

(224) Liu, D.; Zhao, W.; Zhang, S.; Ye, L.; Zheng, Z.; Cui, Y.; Chen, Y.; Hou, J.

Macromolecules 2015, 48, 5172.

(225) Zhang, M.; Guo, X.; Zhang, S.; Hou, J. Adv. Mater. 2014, 26, 1118.

(226) Blouin, N.; Michaud, A.; Leclerc, M. Adv. Mater. 2007, 19, 2295.

(227) Andersson, L. M.; Inganäs, O. Appl. Phys. Lett. 2006, 88, 082103.

(228) Chang, J.; Sun, B.; Breiby, D. W.; Nielsen, M. M.; Giles, M.; Mcculloch, I.;

Sirringhaus, H. Chem. Mater. 2004, 16, 4772.

(229) Salleo, A. Mater. Today 2007, 10, 38.

(230) Ebenhoch, B.; Thomson, S. A. J.; Genevičius, K.; Juška, G.; Samuel, I. D. W. Org.

Page 132: Conjugated Organic Polymers as Photocathode Materials in …summit.sfu.ca/system/files/iritems1/19172/etd20138.pdf · 2020-04-11 · iii Abstract The work presented in this thesis

114

Electron. 2015, 22, 62.

(231) Rolin, C.; Kang, E.; Lee, J. H.; Borghs, G.; Heremans, P.; Genoe, J. Nat. Commun.

2017, 8, 14975.

(232) Delongchamp, D. M.; Vogel, B. M.; Jung, Y.; Gurau, M. C.; Richter, C. A.; Kirillov,

O. A.; Obrzut, J.; Fischer, D. A.; Sambasivan, S.; Richter, L. J.; Lin, E. K. Chem.

Mater. 2005, 17, 5610.

(233) Meyerhofer, D. J. Appl. Phys. 1978, 49, 3993.

(234) Shioya, N.; Shimoaka, T.; Eda, K.; Hasegawa, T. Macromolecules 2017, 50, 5090.

(235) Kajiya, D.; Ozawa, S.; Koganezawa, T.; Saitow, K. J. Phys. Chem. C 2015, 119,

7987.

(236) Brinkmann, M.; Hartmann, L.; Biniek, L.; Tremel, K.; Kayunkid, N. Macromol. Rapid

Commun. 2014, 35, 9.

(237) Chou, S. Y.; Krauss, P. R.; Zhang, W.; Guo, L.; Zhuang, L. J. Vac. Sci. Technol. B

1997, 15, 2897.

(238) He, X.; Gao, F.; Tu, G.; Hasko, D. G.; Hüttner, S.; Greenham, N. C.; Steiner, U.;

Friend, R. H.; Huck, W. T. S. Adv. Funct. Mater. 2011, 21, 139.

(239) Cheyns, D.; Vasseur, K.; Rolin, C.; Genoe, J.; Poortmans, J.; Heremans, P.

Nanotechnology 2008, 19, 424016.

(240) Kim, J. S.; Park, Y.; Lee, D. Y.; Lee, J. H.; Park, J. H.; Kim, J. K.; Cho, K. Adv.

Funct. Mater. 2010, 20, 540.

(241) Cao, J.; Sun, J.; Shi, G.; Chen, H.; Zhang, Q.; Wang, D.; Wang, M. Mater. Chem.

Phys. 2003, 82, 44.

(242) Kowalski, D.; Tighineanu, A.; Schmuki, P. J. Mater. Chem. 2011, 21, 17909.