+ All Categories
Home > Documents > Organic Photovoltaics Based on Solution Cast Polymers and ......12) were studied in pseudo-planar...

Organic Photovoltaics Based on Solution Cast Polymers and ......12) were studied in pseudo-planar...

Date post: 05-Jul-2020
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
83
Organic Photovoltaics Based on Solution Cast Polymers and Boron Subphthalocyanines Hybrid Device Architectures Enabling Novel Material Combinations by Stephanie Robin Nyikos A thesis submitted in conformity with the requirements for the degree of Master of Applied Science Department of Chemical Engineering and Applied Chemistry University of Toronto © Copyright by Stephanie Robin Nyikos 2018
Transcript

Organic Photovoltaics Based on Solution Cast Polymers and Boron Subphthalocyanines – Hybrid Device

Architectures Enabling Novel Material Combinations

by

Stephanie Robin Nyikos

A thesis submitted in conformity with the requirements for the degree of Master of Applied Science

Department of Chemical Engineering and Applied Chemistry University of Toronto

© Copyright by Stephanie Robin Nyikos 2018

ii

Polymer/BsubPc Organic Photovoltaics – Hybrid Device

Architectures Enabling Novel Material Combinations

Stephanie Robin Nyikos

Master of Applied Science

Department of Chemical Engineering and Applied Chemistry

University of Toronto

2018

Abstract

Solution-cast films of crystalline electron donating poly(3,3”’-didodecylquaterthiophene) (PQT-

12) were studied in pseudo-planar heterojunction (PPHJ) organic photovoltaic devices (OPVs)

paired with boron subphthalocyanine (BsubPc) as the electron acceptor layer, and the effects of

crystallinity domain size and layer thickness were investigated. Annealed, 10-20 nm films of

intermediate-sized crystals were 40% more efficient than un-annealed PQT-12 layers and had

comparable efficiency to their fullerene-based BHJ counterparts, demonstrating the ability of

polymer|BsubPc PPHJ OPVs to accommodate highly crystalline polymers with a tendency to

phase segregate and create devices with favorable electrical properties. These polymer|BsubPc

PPHJs were further studied with the new amorphous copolymer electron donating material

PBTZT-stat-BDTT-8. While optimized PPHJ devices of PBTZT-stat-BDTT-8|BsubPc had

substantially lower efficiency than PBTZT-stat-BDTT-8|fullerene BHJs (which were free of

morphological problems), they demonstrated comparable performance to their BHJ equivalents

due to ideal morphology, presenting a pathway forward for OPV design of otherwise phase

segregating polymer|BsubPc pairings.

iii

Acknowledgments

I would like to extend my utmost gratitude to my supervisor, Prof. Tim Bender, for his guidance

and support which helped navigate me through the course of my research. Thank you greatly for

your encouragement and direction.

I would also like to thank all my colleagues in the Bender lab for their optimism and excellent

advice. They challenged me to participate more in departmental groups and events, which added

amazing depth to my graduate school experience.

Lastly, my heartfelt thanks to my family and friends for their continued love and support, for

believing in me and pushing me to be my very best.

iv

Table of Contents

Acknowledgments......................................................................................................................... iii

Table of Contents ........................................................................................................................... iv

List of Tables ................................................................................................................................ vii

List of Figures .............................................................................................................................. viii

List of Appendices ......................................................................................................................... xi

Chapter 1 ....................................................................................................................................... xii

Introduction .................................................................................................................................1

1.1 Motivation ............................................................................................................................1

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

1.2.1 Brief Overview of OPV Development .....................................................................2

1.2.2 Device Physics .........................................................................................................3

1.2.3 Performance Metrics and J-V Curves ......................................................................6

1.2.4 Device Architecture .................................................................................................8

1.3 Outline................................................................................................................................14

Chapter 2 ........................................................................................................................................15

Materials and Methods ..............................................................................................................15

2.1 Materials ............................................................................................................................15

2.1.1 PEDOT:PSS ...........................................................................................................16

2.1.2 BCP ........................................................................................................................16

2.1.3 Silver ......................................................................................................................17

2.1.4 BsubPc ...................................................................................................................17

2.2 Experimental Methods .......................................................................................................18

2.2.1 Substrate Preparation and Cleaning .......................................................................18

2.2.2 Profilometry ...........................................................................................................20

2.2.3 Physical Vapor Deposition ....................................................................................21

v

2.2.4 OPV Light Testing .................................................................................................22

Nano-crystalline poly(3,3”-didodecyl-quarterthiophene) in pseudo-Planar Heterojunction

Organic Photovoltaics ...............................................................................................................25

3.1 Introduction ........................................................................................................................25

3.2 PQT-12 Thermal Transition Analysis ................................................................................28

3.3 Analysis of Solution Processed Films................................................................................29

3.3.1 Spin-coated Film Profilometry ..............................................................................29

3.3.2 Atomic Force Microscopy of PQT-12 Films .........................................................29

3.3.3 Ultraviolet-Visible Spectroscopy of PQT-12 Films ..............................................31

3.4 Performance in Organic Photovoltaic Devices ..................................................................32

3.4.1 PQT-12 OPV Performance with Cl-BsubPc ..........................................................32

3.4.2 Comparison of PQT-12, P3HT, and α6T OPVs ....................................................36

3.5 Chapter Conclusion ............................................................................................................39

PBTZT-stat-BDTT-8 in pseudo-Planar Heterojunction Organic Photovoltaics .......................40

4.1 Introduction ........................................................................................................................40

4.2 PBTZT-stat-BDTT-8 Film Profilometry ...........................................................................44

4.3 Performance and Optimization of OPVs based on PBTZT-stat-BDTT-8/Cl-BsubPc.......44

4.4 PBTZT-stat-BDTT-8 in “Cnops Stack”.............................................................................48

4.5 OPV Comparison of BsubPc Electron Acceptor Layers with PBTZT-stat-BDTT-8 ........50

4.5.1 Overcoming Replication Issues .............................................................................52

4.6 PBTZT-stat-BDTT-8 in Cl-Cl6BsubPc and PhO-Cl6BsubPc OPVs – BHJ vs PPHJ

Architecture........................................................................................................................54

4.7 Chapter Conclusion ............................................................................................................57

Summary and Future Work .......................................................................................................58

5.1 Summary ............................................................................................................................58

5.2 Future Work .......................................................................................................................60

References ......................................................................................................................................62

vi

Appendices .....................................................................................................................................69

Appendix A ...............................................................................................................................69

vii

List of Tables

Table 1.1 | Summary of literature PPHJ Photovoltaic Device Performance ................................ 13

Table 3.1 | Summary of literature PQT-12/fullerene BHJ performance ....................................... 26

Table 3.2 | Characteristic device parameter comparison of P3HT, PQT-12, and α6T donor layers

paired with Cl-BsubPc. ................................................................................................................. 37

Table 4.1 | Characteristic parameters of PBTZT-stat-BDTT-8/Cl-BsubPc devices. The layer

thickness of Cl-BsubPc was constant at 20 nm. ........................................................................... 46

Table 4.2 | Characteristic parameters of OPV devices with varying electron acceptor layer. The

layer thickness of PBTZT-stat-BDTT-8 was constant at 20 nm................................................... 49

Table 4.3 | Characteristic parameters of PBTZT-stat-BDTT-8/BsubPc devices with a 10 nm

electron acceptor layer. ................................................................................................................. 51

Table 4.4 | Characteristic parameters of PBTZT-stat-BDTT-8/BsubPc devices with a 20 nm

electron acceptor layer. ................................................................................................................. 52

Table 4.5 | Characteristic parameters of PBTZT-stat-BDTT-8/BsubPc devices in either pseudo-

Planar Heterojunction or Bulk Heterojunction architectures. ....................................................... 55

viii

List of Figures

Figure 1.1| Electronic Structure of an OPV. ................................................................................... 3

Figure 1.2 | Stack Architecture of a PHJ OPV. ............................................................................... 3

Figure 1.3 | Left: Solar testing J-V curve and Right: EQE results of a a6T/Cl-BsubPc planar

heterojunction baseline device under simulated solar illumination. The shaded area in both

graphs represent one standard deviation from the average. Important device testing parameters

used to calculate total device efficiency are marked. ..................................................................... 7

Figure 1.4| Pictorial representation of a) Bulk Heterojunction (BHJ) b) Planar Heterojunction

(PHJ) and c) Pseudo-Planar Heterojunction (PPHJ) OPV Architectures. ...................................... 8

Figure 1.5 | Device performance of PTB7/PCBM bilayer diffusion devices annealed at 150°C for

varying times. Device JSC is maximized after 10 min of annealing, resulting in the highest device

PCE. Reproduced with permission from ref. 42, Advanced Energy Materials 2013. .................. 11

Figure 1.6 | A comparison of the photovoltaic performance of devices based on traditional BHJ

PTB7:PCBM films (PCE = 5.9%) and sequentially-processed films with the fullerene layer cast

from a 50:50 2-CP:1-butanol co-solvent blend (PCE = 6.0%). Reproduced from ref. 43,

Advanced Energy Materials 2015. ................................................................................................ 12

Figure 1.7 | Cross-sectional TEM image showing incomplete mixing of C70/PBDTTT-C-T to

form a diffusive quasi-bilayer structure. Reproduced from ref. 44, Journal of Materials

Chemistry A 2014. ......................................................................................................................... 12

Figure 2.1 | Molecular structure of a) PEDOT:PSS b) BCP c) Cl-BsubPc d) P3HT and e) α6T. 15

Figure 2.2 | Glass/ITO Substrate Cleaning Procedure. ................................................................. 18

Figure 2.3 | The nitrogen glovebox (MARI) and vacuum chamber (KATE) setup used to

fabricate and test all OPVs for this thesis. .................................................................................... 21

Figure 2.4 | Custom-built substrate holders for light testing......................................................... 22

ix

Figure 3.1 | a) Chemical structures of poly(3-hexylthiophene-2,5-diyl) (P3HT), poly(3,3’’’–

didodecyl-quaterthiophene) (PQT-12), and chloro-boron subphthalocyanine (Cl-BsubPc) and b)

their reported frontier molecular orbital energy levels. c) Device schematic in which the electron

donor layer is varied with all other layers held constant. ............................................................. 26

Figure 3.2 | Left: TGA data and Right: DSC thermogram of PQT-12. Film annealing

temperatures are marked. .............................................................................................................. 28

Figure 3.3 | Profilometry results of PQT-12 on Glass/ITO/PEDOT:PSS substrates. ................... 29

Figure 3.4 | Tapping mode AFM images of 55 nm thick films of PQT-12. On the left are

topographic images, while images on the right are phase. ........................................................... 30

Figure 3.5 | UV/Vis absorption coefficient spectra of 55 ± 5 nm thick films of PQT-12. ........... 32

Figure 3.6 | Left: J-V characteristics and Right: EQE spectra of OPV devices optimized around

the electron donor layer thickness. The PQT-12 layer was prepared a) unannealed and annealed

at b) 118 °C c) 133 °C and d) 148 °C, paired with Cl-BsubPc. Shading indicates one standard

deviation from the average. ........................................................................................................... 34

Figure 3.7 | Plotted characteristic parameters of OPV devices optimized around the electron

donor layer thickness. ................................................................................................................... 35

Figure 3.8 | Comparison of the Left: J-V characteristics and Right: EQE spectra of P3HT, PQT-

12, and α6T electron donor layers paired with Cl-BsubPc. Shading indicates one standard

deviation from the average. ........................................................................................................... 37

Figure 4.1 | Chemical structures of a) BsubPc electron acceptor molecules and b) polymeric

electron donor materials, with c) their reported frontier orbital molecular energy levels. d)

Device schematic of pseudo-PHJ devices in which the BsubPc is varied. ................................... 41

Figure 4.2 | Profilometry results of PBTZT-stat-BDTT-8 on Glass/ITO/PEDOT:PSS substrates.

....................................................................................................................................................... 44

x

Figure 4.3 | Left: J-V curves and Right: EQE spectra of PBTZT-stat-BDTT-8/Cl-BsubPc devices

with varying electron donor layer thickness, constant 20 nm electron acceptor layer thickness.

Shaded error bars represent ±1 standard deviation from the mean. .............................................. 46

Figure 4.4 | Characteristic parameters of PBTZT-stat-BDTT-8/Cl-BsubPc devices, visualized to

display trends with increasing thickness. ...................................................................................... 47

Figure 4.5 | Left: J-V curves and Right: EQE spectra of OPV devices with varying electron

acceptor layer and constant 20 nm PBTZT-stat-BDTT-8 electron donor layer. Shaded error bars

represent ±1 standard deviation from the mean. ........................................................................... 49

Figure 4.6 | Left: J-V characteristics and Right: EQE spectra of OPV devices containing PBTZT-

stat-BDTT-8 as the electron donor layer and Cl-BsubPc, Cl-Cl6BsubPc, or PhO-Cl6BsubPc as

electron acceptor layer. Shaded error bars represent ±1 standard deviation from the average..... 51

Figure 4.7 | Comparison of VOC and FF of pseudo-Planar Heterojunction or Bulk Heterojunction

architectures. ................................................................................................................................. 55

Figure 4.8 | Comparison of the Left: J-V characteristics and Right: EQE spectra of PBTZT-stat-

BDTT-8/BsubPc OPVs in either pseudo-Planar Heterojunction or Bulk Heterojunction

architectures. Shaded error bars represent ± 1 standard deviation from the average.................... 55

xi

List of Appendices

A-1 | Replication testing of PPHJ OPVs with PBTZT-stat-BDTT-8 (30 nm)/Cl-Cl6BsubPc (10

nm)

A-2 | Replication testing of PPHJ OPVs with PBTZT-stat-BDTT-8 (60 nm)/Cl-Cl6BsubPc (10

nm)

A-3 | Replication testing of PPHJ OPVs with PBTZT-stat-BDTT-8 (30 nm)/Cl-Cl6BsubPc (20

nm)

A-4 | Replication testing of PPHJ OPVs with PBTZT-stat-BDTT-8 (60 nm)/Cl-Cl6BsubPc (20

nm)

xii

List of Abbreviations

α6T α-sexithiophene

AFM Atomic force microscopy

AIBN 2,2’-azobis(isobutyronitrile)

BCP Bathocuproine

BDT Benzodithiophene

BHJ Bulk heterojunction

BsubPc Boron subphthalocyanine

BT Benzothiadiazole

Cl-BsubPc Chloro-boron subphthalocyanine

Cl-Cl6BsubPc Chloro hexachloro-boron subphthalocyanine

D-A Donor-Acceptor

DCB 1,2-dichlorobenzene

DIO 1,8-diiodooctane

DSC Differential scanning calorimetry

EQE External quantum efficiency

FF Fill factor

GIWAXS Grazing-incidence wide-angle X-ray scattering

HOMO Highest occupied molecular orbital

ITO Indium tin oxide

J Current density

JSC Closed circuit current

LUMO Lowest unoccupied molecular orbital

OFET Organic field effect transistor

OPV Organic photovoltaic device

OTFT Organic thin film transistor

Pin Incident light intensity

P3HT poly(3-hexylthiophene-2,5-diyl)

PCBM 1-(3-methoxycarbonyl)-propyl-1-phenyl-[6,6]C61

PCE Power conversion efficiency

PEDOT:PSS poly(3,4-ethylenedioxythiophene) poly(styrene-sulfonate)

PHJ Planar heterojunction

xiii

PhO-Cl6BsubPc Phenoxy hexachloro-boron subphthalocyanine

PPHJ Pseudo planar heterojunction

PQT-12 poly(3,3”’ -didodecylquaterthiophene)

PTB7 poly[[4,8-bis[(2-ethylhexyl)oxy]benzo[1,2-b:4,5-b']dithiophene-2,6-

diyl][3-fluoro-2-[(2-ethylhexyl)carbonyl]thieno[3,4-b]thiophenediyl]]

PVD Physical vapor deposition

SubNc Chloro-boron subnaphthalocyanine

TGA Thermo-gravimetric analysis

TT Thieno[3,4-b]thiophene

UV-vis Ultraviolet-visible spectroscopy

V Voltage

VOC Open circuit voltage

1

Chapter 1

Introduction

Introduction

1.1 Motivation

The world’s growing demands in energy consumption, combined with harmful emissions released

through the exploitation of inexpensive yet environmentally damaging energy sources such as

fossil fuels, has brought the issue of climate change to the forefront of public discussion. With

programs coming online to incentivize zero emission energy generation, as well as the roll out of

modern electrical metering infrastructure allowing all homeowners to generate and potentially sell

their own electricity, there is a clear public desire for small scale sustainable energy generation.1,

2 Solar energy is the fastest growing source of green energy for small-scale generation due to its

modular design and rapidly decreasing price per kWh.2 However, traditional silicon solar

technology suffers from numerous constraints which limit its application. The cells are heavy and

brittle, necessitating the installation of potentially extensive supporting infrastructure which drives

up system cost. This requirement limits traditional solar cells to the tops of roofs or a flat open

backyard, which is not always feasible. Furthermore, traditional solar cells are fabricated using a

single slab of high-purity silicon in an expensive, energy-intensive process which keeps initial

capital costs high compared to other methods of generating electricity.

Organic photovoltaics (OPVs) have attracted significant attention in recent years due to their

potential as lightweight, low-cost, flexible solar cells.3, 4 While OPVs are not necessarily a

replacement for silicon cells due to their comparatively low power conversion efficiencies (PCEs),

their strength lies in the sheer number of potential commercial applications for a mass-producible,

lightweight, fully flexible solar cell. These semi-transparent solar cells can be roll-to-roll printed

onto sheets of plastic and employed practically anywhere with little or no supporting infrastructure

required.5, 6 For example, they could be installed on building façades to help meet inhabitant’s

electricity requirements or installed on the top and sides of electric vehicles to improve their

efficiency. Unlike silicon-based cells, OPVs retain their efficiency at low light levels, allowing

them to output energy in the shade or on a cloudy day.7-10 Despite OPVs numerous benefits, these

cells have only ever been installed for small-scale projects; larger scale commercialization has yet

2

to be achieved. While significant progress has been made towards improving the PCE and lifetime

of OPVs, further work must be done for this technology to truly flourish outside of a lab

environment.

One promising avenue of research towards this goal is the optimization of material choice within

the photoactive stack. OPVs require two types of photoactive material to function: an electron

donor layer and an electron acceptor layer. These layers have traditionally been deposited either

entirely from solution, or entirely from physical vapor deposition. Insoluble materials cannot be

solution deposited, and heavy polymers cannot be vapor deposited. This limitation has restricted

the combinations of electron donor/acceptor layers available with which to fabricate solar cells.

Overcoming this limitation opens the door for novel material combinations and solar device

architectures between polymers and small molecules, which may lead to improvements in OPV

efficiency and lifetime.

1.2 Background

1.2.1 Brief Overview of OPV Development

Historically, the discovery of photoconductivity in solid anthracene in 1906 marked the beginning

of the field of organic photovoltaics.11 Early investigations into OPV devices came in the 1950s

with the study of organic dyes such as chlorophyll and continued with studies into semiconducting

polymers. Efficiencies from these early single-junction devices were very low, under 0.1%. A

major breakthrough in the field came in 1986 when Tang reported the first bilayer device which

achieved and efficiency of 1%.12 In the paper, Tang first proposed the operating principle of the

electron donor/acceptor interface which is still accepted today. In 1993, Sariciftci et al. reported

the first polymer/C60 device, which achieved an efficiency of 0.04%.13 They outlined a method of

spin-coating the polymeric electron donor layer and vapor depositing the insoluble C60 layer in a

very similar technique to that used in this thesis to incorporate polymers and insoluble small

molecules in the same device. In 1995, Yu et al. reported polymer/fullerene blend OPVs which

were the first example of bulk heterojunction (BHJ) devices.14 The long interface of the electron

donor/acceptor layers was shown to improve device photogeneration and current density. Since

then, OPV technology has shown steady development in PCEs driven by molecular design of

3

photovoltaic molecules and improvements in device engineering. Record breaking efficiencies of

13% have recently been reported and certified for state-of-the-art fullerene-free BHJ devices with

a polymeric electron donor and small-molecule electron acceptor.15 The record for the highest

efficiency OPV is frequently broken, driven by significant research efforts to develop novel high-

performance semiconducting materials. While there are some caveats to these efficient OPVs, such

as difficult material synthesis and untested working cell lifetimes, the future of OPVs remains very

bright.

1.2.2 Device Physics

All OPV devices are composed of stacked layers of materials where each layer is

exceptionally thin, typically 1-200 nm for planar heterojunction solar cells (PHJs). Figure 1.1

Figure 1.2 | Stack Architecture of a PHJ OPV.

1. Light enters the device through the transparent

anode/hole transport layer and is absorbed by

chromophores in the two photoactive materials.

2. Excited materials produce excitons which

migrate to the electron donor/acceptor interface.

3. Excitons dissociate at the interface into electrons

and holes.

4. Charge carriers migrate to their respective

electrodes to produce current

Figure 1.1| Electronic Structure of an OPV.

1. Exciton dissociation at the electron

donor/acceptor interface is driven by the energy

difference between the HOMO of the donor and

the LUMO of the acceptor material. This energy

is approximately equal to the device’s Voc.

2. Holes travel along the HOMO of the donor

and electrons travel along the LUMO of the

acceptor towards their respective electrode.

3. Charge carriers jump across interface to their

respective electrode, a process facilitated by

electron/hole transport layers (not shown).

4

depicts the standard layer design of a PHJ OPV stack, while Figure 1.2 depicts its electronic

structure.

1.2.2.1 Light Absorption

The electron donor and acceptor layers are composed of semiconducting organic molecules which

strongly absorb visible wavelength light. Their conductive and photoactive properties are a result

of conjugated systems formed from multiple covalently bonded carbon atoms with adjacent p-

orbitals. The p-orbitals overlap and connect to from a bonding π-orbital and antibonding π*-orbital,

allowing π-electrons to delocalize across the system in order to reduce the free energy of the system

and increasing stability. Conjugated systems are depicted as alternating double and single bonds,

but the π-electrons belong to the group of atoms rather than any specific bond. The energy level

of the π-orbital is known as the highest occupied molecular orbital (HOMO) which is analogous

to the valence band of an inorganic semiconductor, while the energy level of the π*-orbital is

known as the lowest unoccupied molecular level (LUMO) and is analogous to a conduction band.

The band gap of an organic semiconductor is the energy separation between its HOMO and LUMO

levels. The width and depth of the bandgap is a critical factor in the molecular design of organic

photoactive materials. Organic semiconductors may only absorb photons with energy greater than

the band gap. For example, if the width of a material’s band gap is 2.5 eV, it can only absorb

green/blue light with photon energies of 2.5 eV or higher. Light with lower photons energies, such

as yellow and red light, will pass straight through the material as if it were transparent. Light with

much higher photon energies, such as purple and ultraviolet light, may still be absorbed, but energy

in excess of the bandgap is released as heat which can be detrimental to device performance.

1.2.2.2 Exciton Transport and Dissociation

Upon absorbing photons, a region of the conjugated system becomes excited and forms a

Coulombically bound electron-hole pair called an exciton.16 Excitons have no net electrical charge

and may transport energy through the material via diffusion away from areas with a high exciton

concentration. The diffusion length of excitons is limited to about ~15 nm because of their

relatively short lifetime, after which they decay back to their ground state.17 In organic

semiconductors, the aim is for excitons to diffuse to the electron donor/acceptor interface before

they decay.18 The energy difference between the HOMO of the electron donor layer and the LUMO

of the electron acceptor layer must be sufficient to overcome the exciton binding energy (typically

5

0.25-1 eV in organic materials)19 to dissociate a free electron and electron hole. The efficiency of

exciton dissociation is dependent on the area of the interface, the lifetime of the exciton, and its

ability to diffuse to the interface within that lifetime.20 The dissociated electrons and holes

experience significant coulombic attraction across the interface and may recombine if not

separated quickly.21 However, the competing coulombic repulsion of many like-charged electrons

and holes at the interface drive the newly dissociated charge carriers away from one another.

1.2.2.3 Charge Transport

Once dissociated, electrons travel through the electron acceptor layer to the metal cathode and

holes travel through the electron donor layer to the transparent anode. Charge carrier

recombination occurs when free electrons meet with free holes. The charge carrier mobility of the

photoactive layers must be approximately equal to ensure charge traverses through the layers at

approximately the same rate. Charge carrier mobility imbalance, as well as high interfacial energy

barriers between the electron donor/acceptor layers has been known to cause charge accumulation

in which charges cannot exit the device at an appreciable rate. Charge build-up at interfaces can

give rise to ‘S-kinks’, s-shaped current-voltage curves which greatly lowers OPV fill factor (FF)

and decreases performance.22, 23 Choice of layer material and transport layer crystallinity are

crucial to avoid charge accumulation issues.

1.2.2.4 Charge Extraction

After traversing the photoactive materials, charge carriers travel through a transport layer to reach

their respective electrode where they can be extracted to yield a photocurrent. These interface

buffer layers serve four important functions which significantly improve device performance.

Firstly, they block any stray opposite sign charge carriers from reaching the electrodes and

reducing the photocurrent. To do this, the material must have a high energy barrier for opposite

sign charges to jump across the layer interface based on their energy level alignment. Secondly,

interface buffer layers act as both diffusion barriers between the metal electrode and photoactive

layer, reducing the number of defects and recombination sites caused by metal penetration into the

photoactive layer. In doing so, they prevent any metal/organic chemical reactions from taking

place and degrading the layer materials. Thirdly, they improve the energy level alignment of the

metal/organic interface by changing the metal’s effective work function. This lowers the energy

barrier for charge carriers to traverse the electrode interface. Finally, interfacial buffer layers

6

protect the photoactive material from damage during electrode deposition, reducing the number of

defects introduced during device fabrication. In this thesis, bathocuproine (BCP) is used as the

electron transport layer and PEDOT:PSS is used as the hole transport layer. Both are well-studied

and commonly used transport layers which are known to significantly improve device

performance24-27 and will be further discussed in the next section.

In this thesis, the OPV cathode layer is an 80nm thick layer of Ag. Ag has a work function which

matches the energy levels of BCP, allowing electrons to easily make their way across the material

interface.27 It forms smooth, even films when vapor deposited, and is easier to deposit than other

commonly used cathode materials such as Al. The cell anode is a 120nm thick layer of indium tin

oxide (ITO) on top of a structurally supportive glass substrate. ITO is a conductive metal oxide

whose work function matches the energy levels of PEDOT:PSS. Critically, it is also transparent to

allow light into the OPV where it may be absorbed by the photoactive layers.28

1.2.3 Performance Metrics and J-V Curves

The performance of OPV devices can be quantified into three important characterization

parameters: open circuit voltage (VOC), short circuit current (JSC), and fill factor (FF). The OPV’s

power conversion efficiency (PCE) is defined as the fraction of incident power that is converted

into electricity, and may be calculated using these three numbers along with the incident light

intensity (Pin) by the following relationship:

PCE =VOC JSC FF

Pin

7

These metrics are determined through solar testing, which is further discussed in the methods

section of this thesis. Figure 1.4 displays typical graphs generated from solar testing data. As

previously discussed, the VOC can be approximated by the LUMO energy of the electron acceptor

and the HOMO level of the electron donor. Additionally, the VOC is heavily influenced by the

amount of charge carrier recombination that occurs as dissociated charges travel to their respective

electrodes. This phenomenon is a more pressing issue for BHJ architecture devices as it is strongly

affected by phase morphology.29 Generally, a higher VOC produces a better PCE. However, to

achieve a higher VOC, there must be a larger energy difference between the two photoactive

materials which can lead to a reduction in charge dissociation across the interface, which in turn

leads to lower device JSC and FF. A careful balance must be found between VOC and the other

characteristic parameters to achieve optimal device performance.

The JSC and FF are dependent on a much larger number of factors including light intensity,

temperature,30, 31 active layer thickness,32 and film morphology.33, 34 An OPV device outputs its

maximum power at VMAX and JMAX. An OPV should be operated at its VMAX to ensure it outputs

the maximum amount of electricity. The FF indicates internal energy losses within the cell and is

graphically the ‘squareness’ of the J-V curve. FF is the ratio of the maximum power to the product

of the cell’s VOC and JSC. The ideal value for FF is 1, meaning there are no internal losses occurring

Figure 1.3 | Left: Solar testing J-V curve and Right: EQE results of a a6T/Cl-BsubPc planar

heterojunction baseline device under simulated solar illumination. The shaded area in both graphs

represent one standard deviation from the average. Important device testing parameters used to

calculate total device efficiency are marked.

8

in the cell. An exceptional fill factor for real OPVs is ~0.7, although most OPVs have FFs between

0.5-0.7. The JSC indicates the total current produced by the device when the voltage is zero. It is

the largest current that can be produced by the device; any reductions in the JSC before JMAX are

caused by resistive losses within the OPV. The device current is dependent on the photo-generation

and collection of charge carriers as described previously. For PHJ architecture devices, this is often

the main source of efficiency reduction when compared to BHJ architecture devices due to the

decreased photoactive interfacial area.

An OPV’s external quantum efficiency (EQE) is defined as the ratio of the number of charge

carriers collected by the device to the total number of incident photons. It depends on both the

OPV’s ability to absorb photons and its ability to turn those photons into current. The EQE spectra

is generated by shining monochromatic light on the device and measuring the resulting current

generation at each wavelength. Ideally, the two photoactive layers will absorb strongly at different

wavelengths to cover more of the solar spectrum.

1.2.4 Device Architecture

The three types of OPV device architectures that are studied in this thesis are depicted in Figure

1.1. The two major device architectures in the field of OPV devices are bulk heterojunctions

(BHJs) and planar heterojunctions (PHJs). In BHJs, the two photoactive materials are combined

in solution and deposited together. OPVs with this architecture cannot incorporate insoluble

photoactive materials or even two photoactive materials with different solubility due to this

solution-processing step. Electron donor/acceptor intermixing causes BHJ OPVs to have a large

b) a) c)

Figure 1.4| Pictorial representation of a) Bulk Heterojunction (BHJ) b) Planar Heterojunction (PHJ)

and c) Pseudo-Planar Heterojunction (PPHJ) OPV Architectures.

9

interfacial area, which results in raised charge generation for this type of device architecture

compared to PHJs.35, 36 However, BHJs suffer from a lack of control over film morphology and

difficulty deconvoluting performance of electron donor/acceptor materials.

What is undertaken in my thesis is the exploration of another device architecture that will

enable the combination of a solution castable polymeric material and an additional insoluble

material – both paired together in a functional OPV device – which I will refer to as a pseudo-

planar heterojunction (PPHJ) architecture. The donor layer in a PPHJ architecture is solution

deposited while the acceptor layer is vapor deposited. electron donor/acceptor layers of PHJ

devices are vapor deposited separately to create a precisely designed and controlled interface

between the two layers. This precisely controlled layer structure is optimal for investigating new

material functionality and charge transport phenomena. Since heavy polymers degrade when

heated, photoactive materials in PHJs are restricted to small-molecules.

The PPHJ device architecture has been used in the past to fabricate poly(3-hexylthiophene-2,5-

diyl) (P3HT)/C60 OPVs. P3HT is one of the most intensely studied members in the polythiophene

family of conjugated semiconducting polymers and will be discussed more in-depth later on in this

thesis. Fullerene is highly insoluble and must be vapor deposited to obtain a thin film. P3HT/C60

bilayer devices are exceptionally rare in the literature compared to the amount of research

dedicated to P3HT/1-(3-methoxycarbonyl)-propyl-1-phenyl-[6,6]C61 (PCBM) BHJ devices. They

are generally studied to better understand the complex relationship between active layer

morphology and device performance. One such study was conducted by Geiser et al., in which

they constructed and analyzed P3HT/C60 bilayer OPVs.37 Using absorption and

photoluminescence spectroscopy, atomic force microscopy (AFM), and TOF-SIMS depth

profiling, they determined that spin-coated P3HT forms a porous film that allows C60 to readily

diffuse into the polymer layer and form aggregates. Devices that underwent thermal annealing at

150°C for 30 min achieved the highest performance with a VOC of 0.46 V, a JSC of 3.5 mA cm-2, a

FF of 0.55, and a PCE of 2.2%, which was attributed to morphology changes from beneficial phase

segregation and improved charge transport. Stevens et al. also reported improvements to bilayer

P3HT/C60 device efficiency with thermal annealing.38 Using AFM and device layer thickness

optimization, they determined that annealing the active layer at 170°C after vapor deposition of

C60 produced the best devices, achieving a VOC of 0.63 V, a JSC of 3.42 mA cm-2, a FF of 0.55, and

a PCE of 1.19%. The high performance after annealing was attributed to thermally induced mixing

10

of electron donor/acceptor layers causing an increase in interfacial area. The authors also

discovered a drop in VOC at extreme annealing temperatures caused by penetration of C60 to the

anode, which they solved with the addition of a pentacene blocking layer under the polymer. Tong

et al. also studied the P3HT/C60 bilayer, demonstrating a method of constructing an interdigitated

network between the two active materials through the addition of 2,2’-azobis(isobutyronitrile)

(AIBN) to the spin-coated P3HT layer.39 After deposition of the polymer, the film was annealed

at 65oC for 15h to induce the release of N2 gas from AIBN. AFM of the resulting film cross section

was used to determine that the escaping N2 caused significant roughening of the P3HT surface.

The best OPVs made with technique achieved a VOC of 0.3V, a JSC of 6.53 mA cm-2, a FF of 0.44,

and a PCE of 1.02% when the P3HT:AIBN ratio was 4:1. Since this efficiency was four times

greater than devices fabricated without AIBN, the authors concluded the performance

improvement was a result of increased electron donor/acceptor interfacial area and improved

percolation pathways for charge transport. Yang et al. also developed P3HT/C60 bilayers with a

focus on morphology control.40 The polymer donor layer was composed of imprinted P3HT

nanogratings formed using a lined Si mold with a width of 70 nm, height of 60 nm, and spacing of

70 nm. C60 was then vapor deposited on top at various deposition angles. The authors found device

efficiency was highly dependent on deposition angle, rate, and thickness, with the optimal devices

achieving a VOC of 0.32 V, a JSC of 9.46 mA cm-2, a FF of 0.45, and a PCE of 1.35%.

PPHJ OPVs have also been used to study the interface morphology of devices containing

poly[[4,8-bis[(2-ethylhexyl)oxy]benzo[1,2-b:4,5-b']dithiophene-2,6-diyl][3-fluoro-2-[(2-

ethylhexyl)carbonyl]thieno[3,4-b]thiophenediyl]] (PTB7), a high performance polymer which,

like P3HT, is far more frequently studied in BHJ OPVs than in PPHJs. Ochiai et al. used PTB7 in

a bilayer architecture to reduce the charge transport path complexity of BHJs.41 They fabricated

their OPVs with a spin coated layer of PTB7 followed by a spray coated layer of PC71BM. Spray

coating was employed rather than spin-coating to bypass the need for an orthogonal solvent, as

well as to avoid any damage to the underlying PTB7 layer that would be caused by spin-coating.

Using Ultraviolet-visible spectroscopy (UV-vis) and AFM, they investigated the effects on light

absorbance and surface morphology of 1,8-diiodooctane(DIO) solvent additive on each active

layer material both individually and together, as well as through incorporation into interpenetrated

PPHJ devices. They concluded that the DIO additive is highly beneficial when added to PC71BM

due to the resulting retarded drying time of PC71BM layer lengthening interpenetration time into

11

the underlying PTB7, as well as DIO’s ability to suppress larger grain formation in PC71BM which

increases the degree of interpenetration. Through the combination of two wet deposition processes,

along with the introduction of the solvent additive DIO into the PC71BM layer, they were able to

produce an interpenetrated PPHJ structure with a VOC of 0.75 V, a JSC of 10.51 mA, a FF of 0.45,

and a PCE of 3.54%.

Liu et al. also studied the PTB7/PC61BM interface in a

bilayer in order to relate its structure and morphology

to device performance.42 PTB7/PC61BM bilayer

devices were fabricated in an inverted architecture by

first spin-coating PC61BM onto the cathode, then spin-

coating PTB7 on a PSS coated wafer substrate and flow

transferring the film onto the PC61BM layer. After

drying the bilayer overnight and annealing, the anode

was deposited by vapor deposition. To investigate the

effect of diffusion on device performance, the authors

annealed PTB7/PC61BM at 150°C for increasing

periods of time. They identified a trend wherein the

overall efficiency reached a maximum after 10 minutes

of annealing, then began to decrease with additional

time (Fig. 1.5). The authors ascribed this behavior to

undesirable degree of PC61BM aggregation. Grazing

Incidence X-ray Diffraction was used to determine that

longer annealing times reduced the ordering of PTB7,

which was attributed to the dissolution of PTB7 by PC61BM diffusion. The authors concluded that

the interdiffused bilayer film had worse efficiency than traditional BHJ architecture due to the

lower degree of order in PTB7, along with larger-scale aggregation of PC61BM disrupting charge

transport across the interface.

Aguirre et al used a sequential processing technique to achieve efficient PTB7/PC61BM bilayer

OPVs.43 They tested a range of co-solvent blends to ideally swell and wet several polymers

including PTB7, then deposited the fullerene active layer from a carefully chosen orthogonal

solvent so as not to damage the underlying polymer layer. In the case of PTB7, 2-chlorophenol:1-

Figure 1.5 | Device performance of

PTB7/PCBM bilayer diffusion devices

annealed at 150°C for varying times.

Device JSC is maximized after 10 min of

annealing, resulting in the highest device

PCE. Reproduced with permission from

ref. 42, Advanced Energy Materials 2013.

12

butanol was used to first swell the polymer

film and later as the fullerene casting co-

solvent to hinder dissolution of PTB7 during

fullerene deposition. Swelling-activated

interdiffusion of fullerene into the PTB7

network occurred with little to no change in

underlying polymer crystallinity and

structure, as evidenced by grazing-incidence

wide-angle X-ray scattering (GIWAXS). The

authors determined that this interdiffusion

was highly selective towards amorphous

polymer network, leaving denser, crystalline

polymer regions untouched. Their sequential

processing technique resulted in the successful formation of a polymer/fullerene photoactive

network and efficient devices achieving a PCE of 6%, equivalent to those with a traditional BHJ

architecture (Fig 1.6).

Chang et al. also investigated PPHJs

incorporating PTB7, as well as related PBDTTT-

C-T.44 Inverted ‘quasi-bilayer’ PPHJ OPVs were

fabricated by first vacuum depositing a poorly

soluble C70 electron acceptor layer, then using the

fast-drying blade-coating method to deposit the

electron donor polymer. After depositing the

polymer layer from a toluene: o-xylene ratio of 95

: 5 wt%, the resulting C70/polymer films were

studied under AFM and the root-mean-square

roughness (Rms) were determined to be a very high

16.2 nm. Brightfield TEM revealed the presence of randomly oriented, island-like nanostructures

caused by C70 aggregation, which was further verified with SEM. Cross-sectional TEM revealed a

wavy interface donor-acceptor interface with controllable morphology through altering the wt%

Figure 1.6 | A comparison of the photovoltaic

performance of devices based on traditional BHJ

PTB7:PCBM films (PCE = 5.9%) and sequentially-

processed films with the fullerene layer cast from a

50:50 2-CP:1-butanol co-solvent blend (PCE =

6.0%). Reproduced from ref. 43, Advanced Energy

Materials 2015.

Figure 1.7 | Cross-sectional TEM image

showing incomplete mixing of C70/PBDTTT-

C-T to form a diffusive quasi-bilayer

structure. Reproduced from ref. 44, Journal

of Materials Chemistry A 2014.

13

of co-solvents. After layer thickness optimization, the optimal C70/PTB7 ‘quasi-bilayer’ OPV

achieved a VOC of 0.69 V, a JSC of 13.9 mA cm-2, a FF of 72.1, and an impressive PCE of 7.15%.

Moritomo et al. fabricated PTB7/C70 bilayer devices in order to study charge carrier density effects

on recombination, taking advantage of the greatly simplified interface compared to a BHJ device.45

To construct their devices, PTB7 was first spin-coated onto the anode, followed by a vapor

deposited layer of C70. The resulting device had a VOC of 0.68 V, a JSC of 5.9 mA cm-2, a FF of

0.68, and a PCE of 2.7%. Using time-resolved spectroscopy on the simple bilayer device, the

authors were able to demonstrate that fast charge carrier escape from the donor/acceptor interface

is critical for high device efficiency, since any charge accumulation greatly accelerates carrier

recombination at the interface.

Kim et al. constructed inverted bilayer OPVs with PBT7 as the electron donating layer to study

the relationship between device VOC, reverse saturation current, and crystal morphology.46 To

fabricate their devices, first the polymeric electron acceptor was spin-coated onto the anode from

chlorobenzene and annealed at varying temperatures for 15 min in N2, follow by spin-coating

PTB7 on top from dichloromethane. Using AFM and 2D-GIWAXS analysis, they determined that

200°C was the optimal annealing temperature to obtain highly crystalline P(NDI2OD-T2). From

further AFM and TEM measurements, they showed that changes to morphology of underlying

P(NDI2OD-T2) had no effect on PTB7’s morphology. Through investigation of resulting OPV

electrical characteristics, the authors determined that P(NDI2OD-T2) layers with increased

crystallinity and larger crystallites resulted in increased trap-assisted and bimolecular

recombination rates in devices, which reduced the VOC.

Table 1.1 | Summary of literature PPHJ Photovoltaic Device Performance

Electron Donor

Electron Acceptor

VOC [V]

JSC [mA cm-2]

FF PCE [%]

P3HT C60 0.46 3.50 0.55 2.20 ref 37

P3HT C60 0.63 3.42 0.55 1.19 ref 38

P3HT C60 0.30 6.53 0.44 1.02 ref 39

P3HT C60 0.32 9.46 0.45 1.35 ref 40

PTB7 PC71BM 0.75 10.51 0.45 3.54 ref 41

PTB7 PC61BM 0.75 5.12 0.56 2.20 ref 42

PTB7 PC61BM 0.76 13.70 0.57 6.00 ref 43

PTB7 C70 0.69 13.90 0.72 7.15 ref 44

PTB7 C70 0.68 5.90 0.68 2.7 ref 45

PTB7 P(NDI2OD-T2) 0.72 6.00 0.49 2.14 ref 46

14

There is a clear literature precedent of combining a solution-deposited polymeric electron

donor material with an otherwise process-incompatible electron acceptor using a PPHJ

architecture. In this thesis, I build off this pre-established research on PPHJ architecture devices

to investigate new combinations of polymeric and insoluble small molecule photoactive materials

for OPVs. Specifically, I investigate the performance of novel combinations of polythiophene-

based polymeric electron donor materials paired with boron subphthalocyanine-based small

molecule electron acceptor materials in OPV devices.

1.3 Outline

This thesis will focus on the study of PPHJ OPVs in which the electron donor layer is a polymeric

molecule and the electron acceptor layer is a small molecule. PPHJ OPV devices are tested

experimentally with alterations only to the photoactive layer composition, controlling for the rest

of the device stack. In this way, only the choice of photoactive materials and their interactions are

studied, with their performance used to further our understanding of how to better design future

organic molecules for use in OPVs. The following Chapter, Chapter 2 details experimental

methods and materials used across the entire thesis (more project-specific materials and

methodology are described in the relevant Chapter). Chapter 3 presents the integration of

regioregular poly(3,3’’’– didodecyl-quaterthiophene) (PQT-12) into OPV devices and proposes a

route forward for the molecular design of thiophenes for use with small molecules. Chapter 4

presents the integration of PBTZT-stat-BDTT polymer into OPV devices, and relates this

polymer’s resulting high performance back to its molecular design. Finally, Chapter 6 summarizes

the major findings in this thesis and discusses promising areas of future work.

Chapter 2

Materials and Methods

Materials and Methods

This section describes the materials and methodology utilized across the entirety of the thesis work.

More specific materials and procedures used only in one project will be described in the relevant

chapter.

2.1 Materials

The following materials used in this thesis were purchased and used as received: poly(3,4-

ethylenedioxythiophene) poly(styrene-sulfonate) (PEDOT:PSS, Heraeus, Clevois P VP AI 4083),

regioregular poly(3-hexylthiophene-2,5-diyl) (P3HT, Rieke Metals, RMI-001EE MW: 69K), α-

sexithiophene (α6T; Lumtec), bathocuproine (BCP; Lumtec, 99.6 %), 1,2-dichlorobenzene (DCB;

Sigma–Aldrich, anhydrous, 99 %), silver (Ag; Angstrom, 99.999 %), and silver paint (PELCO,

Conductive Silver18). Cl-BsubPc was synthesized in-lab through a previously reported method

Figure 2.1 | Molecular structure of a) PEDOT:PSS b) BCP c) Cl-BsubPc d) P3HT and e) α6T.

PEDOT:PSS and BCP are hole and electron transport layers, respectively, while Cl-BsubPc, P3HT,

and α6T are photoactive layers.

16

and purified once by train sublimation before use.47, 48 Of the materials listed, PEDOT:PSS, BCP,

and Ag are used is every single device fabricated for this thesis. As such, a brief background into

these materials is highly relevant for better understanding of subsequent chapters. Since this thesis

focuses heavily on the incorporation of BsubPc into OPVs, an introduction to these molecules is

also included.

2.1.1 PEDOT:PSS

PEDOT:PSS is a transparent, conductive polymer which has become a benchmark for OPV anode

buffer layer materials.49 It consists of a mixture of two ionomers: poly(3,4-

ethylenedioxythiophene) (PEDOT) and sodium polystyrene sulfonate (PSS). PEDOT is a

positively charged conjugated polythiophene, while PSS is a negatively charged polymer which

helps to disperse and stabilize the PEDOT in an aqueous dispersion to form smooth, continuous

thin films.50 It is solution deposited on top of the ITO anode layer where it serves as a barrier to

exciton and electron transport while facilitating the transport of holes to the anode. Due to the high

ductility, conductivity, and low cost of PEDOT:PSS, this layer is often studied as a replacement

for ITO as the anode in OPV devices.51 However, ITO/glass still provides better OPV performance

than PEDOT:PSS/glass due to its low sheet resistance, which is why ITO/glass was used as the

anode in this thesis.

2.1.2 BCP

BCP is vapor deposited on top of the organic layers prior to deposition of the silver cathode and

serves as a buffer layer between the photoactive material and the electrode. BCP is widely used in

device stacks as it is known to increase the performance of OPVs substantially.27, 52, 53 These

performance improvements are attributed to its ability to block exciton and hole diffusion to the

cathode, as well as its ability to facilitate electron transport from the organic layers to the cathode.

From a cursory look at the LUMO energy level of BCP (3.5 eV) and the work function of Ag (4.5

eV), BCP would appear like a poor material choice to transport electrons because its LUMO level

is much shallower than that of silver, meaning there is an energetic barrier for electrons to jump

between the two layers. However, in reality BCP is an excellent conductor of electrons. It was

found that electron transport does not actually occur at the BCP LUMO as expected, but at a deeper

energy level close to the work function of silver.27 During the deposition of silver, the metal

diffuses into the BCP layer where it forms a BCP-Ag complex whose LUMO level is much closer

17

to the work function of silver. Electron conduction occurs through the BCP-Ag complex rather

than through intact BCP, which explains its high performance.

2.1.3 Silver

The cathode material used in this thesis is silver, rather than the commonly used aluminum. The

two metals have been shown in the literature to give comparable device results in terms of

efficiency and trends in characteristic parameters when paired with a BCP buffer layer.54-56 Silver

is also less detrimental to the vacuum deposition system than Al, which creeps up the wall of the

crucible when heated and can damage the resistive heaters. In contrast, Ag remains in a cohesive

ball during thermal evaporation, posing no risk to the equipment.

2.1.4 BsubPc

Boron subphthalocyanines (BsubPcs) are a family of conjugated small molecules composed of

three nitrogen-bridged isoindoline units with a central boron atom. BsubPcs have a unique

nonplanar ‘bowl’-shaped conformation arising from the atomic radius of boron being slightly

larger than the molecule’s central cavity. These molecules have a symmetrical 14 π-electron

system which allow them to absorb strongly in the visible spectrum. The strongest BsubPc

absorption peak occurs between 560-600 nm which relates to their optical band gap of 2.0-2.1

eV.57 These opto-electronic properties make BsubPcs attractive materials for a variety of organic

electronic applications, such as organic light emitting diodes, organic photovoltaics, and organic

field effect transistors (OFETs).57

Of the BsubPc family, Cl-BsubPc is the most widely studied. The synthesis of Cl-BsubPc was first

reported in 197258 and was not investigated in OPVs until 2006 as an electron donor layer,55

although it has since been employed both as an electron donor layer and an electron acceptor layer.

Cl-BsubPc is thermally stable, with a degradation point above 300 °C.59 It forms a conformal film

with some degree of long-range crystallinity when vapor deposited under high vacuum conditions,

allowing for the deposition of smooth pinhole-free films with good charge transport capability. Its

good thin-film properties, ease of vapor deposition, and excellent opto-electronic properties make

Cl-BsubPc an excellent material for use in OPVs.

The chemical and physical properties of BsubPcs may be tuned through axial or peripheral

substitution. Altering the axial substituent changes the solubility and crystal structure of the

18

BsubPc, while peripheral substitution may be used to tune the energy level and band gap of

BsubPcs.57, 60 A variety of BsubPc derivatives have been synthesized to achieve materials with

different properties to suit the application. The application of these BsubPc derivatives are being

actively investigated in the Bender laboratory.

2.2 Experimental Methods

2.2.1 Substrate Preparation and Cleaning

Substrates were cleaned before use in experiments to ensure that no impurities or unwanted organic

materials entered OPV devices. If left uncleaned, these unwanted particles result in inconsistent

device results across a single substrate which introduces additional error into device performance

results. Thin 25mm x 25mm glass slides coated on one side with pre-patterned ITO from Thin

Devices Inc. were used as OPV substrates. Details of the cleaning procedure are depicted in Figure

Figure 2.2 | Glass/ITO Substrate Cleaning Procedure.

As-received substrates are placed in a glass container, which is then filled with soapy water containing 10

g/L solution of Alconox. The substrates were sonicated for 5 min. in soapy water. After sonication, the

soapy water was disposed of and replaced with distilled water. The same procedure was performed for each

of the depicted solvents. Substrates were stored in methanol after completion of the sonication steps. In the

laminar flow hood, compressed N2 was used to dry off the methanol from the substrates. Substrates were

then placed ITO-side up in the Plasma Cleaner, where they were treated with oxygen plasma for 5 min.

19

2.2. Successive sonication in soap water, distilled water, acetone, and methanol was used to

thoroughly clean the substrate of any contamination. After ultrasonic cleaning, substrates were

stored in methanol until needed for device work. The day before device work, substrates were

removed from their methanol container inside a lamellar flow hood and immediately dried using

compressed nitrogen. A laminar flow hood was used to ensure no dust collected on samples from

turbulent air flow. Dry substrates were placed in the plasma cleaner and cleaned using oxygen

plasma. This surface treatment is very effective at breaking most organic bonds and vaporizing

contaminants to create an ultra-clean surface. Additionally, plasma cleaning is used to raise the

surface energy of the substrate to improve adhesion for spin-coating. A PDC-32G Plasma Cleaner

was used for this thesis.

The solution processing technique used in this thesis was spin-coating due to the ready availability

of equipment and the ease of fabrication of high quality thin films. All of the polymeric electron

donor layer materials used in this thesis are soluble in DCB, which is a commonly used solvent

for depositing thin films. For this reason, DCB was used as a solvent for all solution deposition

experiments in this thesis. Electron donor materials were dynamically spin-coated using a

CHEMAT Technologies KW-4A spin-coater in a nitrogen atmosphere glovebox to limit the film’s

exposure to oxygen, which is detrimental to OPV performance. The hole transport material

PEDOT:PSS was spin-coated from water using a MicroNano Tools KW-4A spin-coater in a

laminar flow hood.

Dynamic dispense spin coating was used to deposit all of the electron donor layer materials for

this thesis. A pipettor was used to dispense 100 µL of solution onto the ITO/PEDOT:PSS substrate.

PBTWhile the dynamic dispense technique allows for less material waste during coating, variables

such as the angle of the pipette and the rate of dispension causes some film variation between

substrates. The resulting film thickness error of up to ± 10% can cause a greater spread in OPV

performance between substrates rather than between devices on the same substrate. If the pipette

solution contains any bubbles, or the dispense occurs slightly off-center, or the pipette tip touches

the substrate, the film quality is greatly affected and may cause significantly larger error between

substrates. For this reason, two substrates of every device architecture were fabricated when testing

device performance. If device performance between substrates was within the threshold of film

thickness error, the results were averaged together. Large variations between substrates were

20

uncommon, but when they occurred the results of the significantly lower performing substrate

were attributed to systematic error and were disregarded.

To perform dynamic spin-coating, first a clean substrate was loaded onto the spin-coater chuck,

where it was held in place using a vacuum. The rpm was then specified. Spin settings for all

polymeric electron donor layers were 700 rpm (12 s) and 1000 rpm (30 s) while spin settings for

PEDOT:PSS were 500 rpm (10 s) and 4000 rpm (30 s). Immediately after beginning substrate

rotation, a micropipette was used to deposit 100 µL of solution onto the rotating substrate.

Centrifugal forces caused the solution to spread out evenly across the substrate to form a

continuous film. Freshly spin-coated PEDOT:PSS substrates were immediately transferred to a

115 °C hotplate and baked for 10 min, then transferred into the nitrogen glovebox. Spin-coated

films of electron donor materials were baked in an oven inside the nitrogen glovebox for 3 min at

70 °C to evaporate the solvent.

The spin-coating technique allows control over film thickness through solution concentration,

which linearly affects viscosity. Before a solution-processed layer may be employed in an OPV, it

is first necessary to determine the relationship between concentration and film thickness for that

specific material. This was accomplished through coating films of various solution concentrations

onto substrates, then determining their layer thickness using profilometry.

2.2.2 Profilometry

Profilometry was used to determine what concentration of electron donor material in solution

would produce consistent films of various thicknesses based on the linear relationship between

concentration and solution viscosity. A KLA-Tencor P16+ surface profilometer was used for this

thesis. In profilometry, an extremely fine stylus is dragged over the surface of a film. It measures

the difference in stylus height across a step-edge, correcting for substrate curvature. Samples were

prepared for profilometry by creating a clean step-edge between the substrate and deposited

material film using acetone as a solvent. For the polymeric spin-coated films used in this thesis,

profilometry measurements were ±8 nm, with uncertainty arising from film roughness and uneven

step-edges. 8 step-edge measurements across two substrates were taken for each solution

concentration for every material tested.

21

2.2.3 Physical Vapor Deposition

Physical Vapor Deposition (PVD) was

used to deposit the electron acceptor

layer, electron transport layer, and Ag

electrode for all OPV devices fabricated

in this thesis. This technique is essential

for the deposition of insoluble small

molecules which cannot be solution

processed, such as Cl-BsubPc and α6T.

Low temperature crucibles were used

for deposition of organic molecules and

a high temperature water cooled crucible

was used for deposition of the Ag

electrode. All materials were deposited

from a height of 12” above the heated

crucible at a rate of 1 Å/s ±0.2 Å with a

working pressure of ~1x10-7 Torr.

Vacuum was maintained using a combination of compressed helium cryopump and a mechanical

rotary vane roughing pump. Contamination was minimized through the use of aluminum foil-

covered ‘shields’ between crucibles. Fresh aluminum foil was used whenever a different material

was loaded into the vacuum chamber. Deposition thickness was tracked using a quartz crystal

monitor.

Figure 2.3 | The nitrogen glovebox (MARI) and vacuum

chamber (KATE) setup used to fabricate and test all

OPVs for this thesis.

22

2.2.4 OPV Light Testing

Light testing was performed immediately

after fabrication of OPVs to determine the

device’s JSC, VOC, FF, and PCE, as well as

the EQE spectra. Testing was performed

inside of the nitrogen glovebox so that un-

encapsulated OPVs were never exposed to

ambient atmosphere prior to light testing.

Freshly fabricated OPV devices were loaded

into custom-built matte-black substrate

holders designed to minimize error from

reflection (Figure 2.4). Substrate holders

were equipped with gold pins to ensure good

electrical contact with the anode and

cathode. Testing was performed under 1 sun

of illumination by an Oriel 300W Xe arc

lamp with an AM 15.G filter. Light intensity

was calibrated using a reference calibrated

silicon photodetector. Current generated by

the devices were measured using a Keithley

2401 Low Voltage Source Meter. Wavelength scans of devices were performed using an in-line

Cornerstone 260 1/4m Monochromator at intervals of 10 nm.

Figure 2.4 | Custom-built substrate holders for light

testing.

Compressible gold pins make contact with the top and

bottom electrodes of each device. Electrical signals

travel through the pins to banana plugs at the bottom of

the holders. Current data is collected one device at a

time.

23

2.2.5 Experiment Statistics

Statistical analysis plays an important role in examining OPV device performance. It allows the

determination of whether or not changes to device structure, materials, crystallization, etc. translate

into statistically significant changes to device performance.

In Chapter 3 of this thesis, the device performance of PQT-12 is investigated in OPV devices by

varying the polymer layer thickness, as well as polymer crystallinity. Device efficiency is the

dependent variable as determined by the VOC, JSC, and FF. The main hypothesis for this section is

that PQT-12 performs better in PPHJ device as compared to BHJ devices, with the null hypothesis

being that the two device architectures perform exactly the same. In Chapter 4, PBTZT-stat-

BDTT-8 is investigated in OPVs by varying the active layer thicknesses independently, as well as

by varying the electron acceptor material. Once more, device efficiency is the dependent variable

as determined by the VOC, JSC, and FF. The main hypothesis for this chapter is that a PPHJ

architecture provides a better understanding of the charge transport in devices as compared to BHJ

devices, with the null hypothesis being that PPHJs provide no benefit.

In all experiments in this thesis, one independent variable was modified at a time to systematically

examine the result on the dependent variables. This is exemplified by layer thickness optimization

studies, in which the electron donor or acceptor layer thickness is varied and the subsequent effect

on device efficiency is studied as determined by the VOC, JSC, and FF. The null hypothesis in these

experiments was that the layer thickness has no effect on device efficiency, while the alternative

hypothesis was that layer thickness does make a real difference in device performance. For studies

of film crystallization in Chapter 3 of this thesis, the null hypothesis was that electron donor layer

annealing temperature has no effect on device efficiency, while the alternative hypothesis was that

electron donor layer annealing temperature made a statistical difference in device performance.

The controlled variables for all experiments in this thesis are the deposition method, the device

testing method, and the layer thicknesses of all supporting layers (ITO, PEDOT:PSS, BCP, and

Ag).

There are certain areas in the OPV device fabrication process which introduce systematic error

into the results. As previously discussed, spin coating deposition introduces a film thickness error

of ± 10%, while PVD introduces film thickness error of ± 0.5 nm. The OPV light testing apparatus

was calibrated with reference to a calibrated silicon photodetector in advance of every round of

24

device measurement, minimizing any error during device testing. Standard deviation was used to

generate all error bars for experimental data. This was done to indicate the variability of the data

and give readers a sense of the statistical significance of the difference between two or more

distinct device architectures to assist correct interpretation.

In order to prove or disprove the experimental hypothesis, it was necessary to measure the

efficiency of multiple devices to evaluate the statistical significance of the resulting data. The

number of devices (n) for each particular device architecture was between n = 7 and n = 17. This

variation was caused in part by limitations to equipment time and experiment batch size (maximum

20 devices), but was also due to the elimination of ‘bad’ devices. Bad devices were those that

shorted, malfunctioned, or generally did not function for no obvious reason. The efficiency of these

outliers lay outside of at least two standard deviations from the mean for that particular architecture

and were discarded.

25

Chapter 3

Nano-crystalline poly(3,3”-didodecyl-quarterthiophene) in pseudo-Planar Heterojunction Organic Photovoltaics

3.1 Introduction

Polythiophenes are a well-studied family of conjugated semiconducting materials with broad

applications in organic electronics. One of the most intensely researched polythiophenes is poly(3-

hexylthiophene-2,5-diyl) (P3HT), which has been thoroughly characterized as an electron donor

layer in bulk heterojunction (BHJ) solar cells paired with 1-(3-methoxycarbonyl)-propyl-1-

phenyl-[6,6]C61 (PCBM).61 The substantial number of P3HT papers has long made it a benchmark

in OPV device engineering. One key advantage of P3HT is its high degree of crystallinity, which

is known to enhance charge carrier mobility by providing favorable pathways through which

charge may easily flow. High charge carrier mobility facilitates charge transport through the device

and reduces detrimental recombination and exciton decay phenomena to produce more efficient

OPVs.62-65 It has previously been demonstrated by Jae Wie et al. that longer and more perfect

crystals of P3HT may be formed through the application of shear force on solution, which was

thought to improve the material’s charge transport properties.66 However, higher crystallinity

P3HT suffered from an extreme increase in viscosity, which is detrimental to spin-coating the

continuous films necessary for solar cell applications.

Regioregular poly(3,3’’’– didodecyl-quaterthiophene) (PQT-12) is compositionally similar to

P3HT, differing in alkyl side chain positioning and length (Figure 3.1). This allows PQT-12 to

achieve long-range, nano-scale crystals with a higher degree of crystallinity than P3HT as a result

of increased π-π stacking and side-chain interdigitation.67-69 Crystalline PQT-12 has been shown

to have excellent performance in organic thin film transistors (OTFTs) and organic field effect

transistors (OFETs) due to its high field-effect mobility of up to 0.18 cm2 V-1s-1, nearly double that

of P3HT.63, 67, 68, 70-75 While a material’s success in OFETs is not necessarily indicative of similar

success in OPVs, these organic electronic devices have an equal requirement for a high charge

mobility active layer for efficient charge transport. PQT-12’s substantial improvement over P3HT

in this area makes it a compelling active material for OPVs. PQT-12 has previously been tested in

26

Table 3.1 | Summary of literature PQT-12/fullerene BHJ performance

Composition Weight ratio VOC [V]

JSC [mA cm-2]

FF PCE [%]

PQT-12:PC61BM 1:3 0.65 -1.40 0.33 0.40 ref 69

PQT-12:PC71BM 1:2 0.70 -5.30 0.38 1.35 ref 76

PQT-DD:PC61BM 1:3 0.59 -2.78 0.33 0.54 ref 77

PQT-12:PC61BM 1:4 0.61 -4.31 0.44 1.15 ref 79

PQT-12:PC61BM 3:17 0.34 -5.05 0.41 0.70 ref 80

bulk heterojunction (BHJ) OPVs, often paired with PCBM. Efficiencies for PQT-12:PCBM

devices were seemingly limited to ~1%, significantly lower than the ~5% efficiency expected of

P3HT:PCBM BHJs.69, 76-80 This has been attributed to incompatible polymer blending between

nano-crystalline PQT-12 and PCBM, leading to macrophase separation and disrupted charge

transport.62, 69, 77, 79, 81 While disruptive in BHJs, these issues do not affect PHJ device structures

a)

b) c)

Figure 3.1 | a) Chemical structures of poly(3-hexylthiophene-2,5-diyl) (P3HT), poly(3,3’’’–

didodecyl-quaterthiophene) (PQT-12), and chloro-boron subphthalocyanine (Cl-BsubPc) and b)

their reported frontier molecular orbital energy levels. c) Device schematic in which the electron

donor layer is varied with all other layers held constant.

27

where organic active layers are deposited separately, rather than mixed and cast together as they

are in BHJs. A bilayer architecture would enable PQT-12 to fully crystallize without disruption by

PCBM or an alternative electron conducting material. In this way the effects of PQT-12

crystallization can be directly studied without the complexity of mixed morphologies.

Traditionally PHJ architecture devices are composed of only vapor deposited small molecule

active layers while this hybrid device consists of a solution deposited PQT-12 electron donor layer

and a vapor deposited electron acceptor layer in a ‘pseudo-PHJ’ structure. The pseudo-PHJ

architecture allows for better characterization of new material combinations due to its simple

bilayer interface which allows for ideal charge transport without having to overcome issues such

as phase seperation or poor percolation pathways. Although the performance of PQT-12 in pseudo-

PHJ devices has not yet been explored to our knowledge, the performance of P3HT pseudo-PHJ

devices has been studied in our lab paired with a chloro-boron subphthalocyanine (Cl-BsubPc)

acceptor layer.82 Cl-BsubPc is a well-studied photoactive material with which our lab has extensive

experience.57, 82-86

When BsubPc is used as an electron acceptor layer, its energy levels and absorption profile

complement that of many thiophenes. In the past, the oligomeric alpha-sexithiophene (α6T) has

been paired with BsubPcs due to its high performance and ability to be vapor deposited.87 Our lab

has frequently tested new electron acceptor BsubPc derivatives in PHJs with an α6T electron donor

layer.82, 86, 88, 89 However, the pseudo-PHJ architecture with solution cast polymeric electron

donating thiophene layer is relatively unexplored.

In the current work, PQT-12 was investigated as a solution processed donor material in hybrid

bilayer OPVs paired with Cl-BsubPc as an electron acceptor layer. Films of PQT-12 with varied

thickness and degrees of crystallization were incorporated into PHJ devices to study their effect

on device performance, and how these effects differ in a BHJ architecture. We show that thin

layers of PQT-12 annealed past the liquid crystalline phase transition produce optimal device

efficiency. PQT-12 devices were directly compared to those constructed with solution casted an

annealed P3HT and thermal vacuum deposited α6T. We discuss the impact of crystallinity and

chromophore density on device performance, and these results may contribute to further film

morphology study of solution processed thiophenes within the PHJ OPV device architecture.

28

3.2 PQT-12 Thermal Transition Analysis

Thermo-gravimetric analysis (TGA) was performed to verify the decomposition temperature of

PQT-12 to ensure that is was not exceeded while performing futher analysis. In this technique, the

mass of the test sample is measured over time as the temperature is increased over a specified

range. A Q50 V6.7 Build 203 TGA was used in this thesis with a temperature range of 25oC to

900oC, a heating rate of 10oC/min under nitrogen in an aluminum pan. 5.314 mg of PQT-12 was

used to conduct the analysis. Differential Scanning Calorimetry (DSC) is a technique used to

analyse the thermal transitions of materials. DSC works by cycling heating and cooling of the

material of study and a control material at a steady rate and comparing their heat flow. Endothermic

heat flow relate to a loss of material orderin g, while exothermic peaks correspond to an increase

in order. In this study, a Q1000 V9.9 Build 303 DSC was used. Conventional DSC analysis was

conducted with a temperature range from -20oC to 180oC over 5 heating/cooling cycles at a rate of

10 °C/min in an aluminum hermetic pan to determine the thermal peaks of PQT-12. Results of

TGA and DSC analysis are plotted in Figure 3.2. Thermal analysis verifies previously reported

DSC thermogram of PQT-12.68, 76, 90, 91 We confirmed two endothermic peaks at ~115 °C and 133

°C (Figure 2). The peak at ~115 °C corresponds to the crystal to liquid crystalline phase transition,

while the peak at 133 °C corresponds to the liquid crystalline to isotropic phase transition.

Corresponding exothermic peaks occur at 118 °C and 50 °C.

Figure 3.2 | Left: TGA data and Right: DSC thermogram of PQT-12. Film annealing temperatures

are marked.

29

3.3 Analysis of Solution Processed Films

3.3.1 Spin-coated Film Profilometry

The spin-coating deposition technique was

used to create films of PQT-12 used in this

study. Film thickness of spin-coated films is a

linear function of solution viscosity, which in

turn is linearly governed by solution

concentration. It follows that spin-coated film

thickness is controlled by the concentration of PQT-12 in solution of dichlorobenzene (DCB). To

determine the precise relationship of film thickness vs. solution concentration, PQT-12 solutions

in DCB were diluted into concentrations of 4, 8, 10, 12, and 14 mg/mL, then prepared and spin-

cast as described in the Methods section of this thesis onto substrates of Glass/ITO/PEDOT:PSS.

Step-edges were created by swiping away half of the films straight down the middle with an

acetone solvent wipe. Profilometry was conducted on all films to determine their real film

thickness. Two substrates were prepared for each solution concentration, and four profilometry

measurements were conducted per substrate (spread evenly across the step-edge). To control for

the thickness of PEDOT:PSS, plain films of Glass/ITO/PEDOT:PSS were also measured with

profilometry. Twelve profilometry measurements across four control substrates were conducted.

The average film thickness of control films of PEDOT:PSS was 33.3 nm ± 6.8 nm. The results of

PQT-12 profilometry are plotted in Figure 3.3. The linear trendline had a good R2 value of 0.996,

so all concentrations and film thicknesses used in this study were based off of this linear relation.

3.3.2 Atomic Force Microscopy of PQT-12 Films

From the DSC data, annealing temperatures of 118 °C, 133 °C, and 148 °C for PQT-12 cast films

were chosen to study the effects of each phase transition on OPV device performance. All films

were prepared on the same glass/ITO/PEDOT:PSS substrates used for OPVs to ensure comparable

crystal growth. After spin-coating 10 mg/mL solutions of PQT-12, films were annealed for 15 min

in an oven at their respective phase transition temperature, then slow cooled overnight to allow

time to self-order. This process was in line with that used for PQT-12 incorporation into OTFTs.68

It was noted that the vacuum oven must remain closed during annealing. If it is opened while

Figure 3.3 | Profilometry results of PQT-12 on

Glass/ITO/PEDOT:PSS substrates.

30

samples are slow-cooling, they will undergo thermal shock and the PQT-12 films delaminate from

the substrate. This effect is more pronounced for higher annealing temperatures. As a result, only

one annealing temperature was performed at a time. Completed samples were stored in the nitrogen

glovebox before film characterization.

Annealed films, as well as an unannealed control, were analysed using tapping-mode atomic force

microscopy (AFM) to determine the topographic and phase characteristics of the PQT-12 films

(Figure 3.4). AFM is a type of scanning probe microscopy which can achieve image resolution of

Figure 3.4 | Tapping mode AFM images of 55 nm thick films of PQT-12. On the left are

topographic images, while images on the right are phase.

31

fractions of a nanometer. Scans were conducted I ambient conditions at a scan rate of 1 Hz, with

imaging regions of 0.5 µm and 2 µm. Two different image scales were selected to display both the

crystal stacking and longer range film morphology. Phase images were of particular interest for

visualizing PQT-12 films due to differences in energy dissipation of crystalline and amorphous

regions of the sample. The scanning tip detects these differences and uses them to create contrast.

Darker regions signify increased crystallinity, while lighter regions are more amorphous.

Surface topology images reveal negligible variation in film roughness between annealing

temperatures. This allowed surface roughness to be removed as a variable for testing PQT-12

layers in OPVs. Unannealed films of spin cast PQT-12 display short, nano-scale grains with a high

degree of misorientation. 92 Larger-scale oriented crystalline domains begin forming at 118 °C, in

the range of the liquid-crystalline phase transition. When annealing temperature was increased to

133 °C, these crystal domains grow anisotropically along the long dimension. Crystalline domains

have the highest degree of order when annealed past the isotropic phase transition at 148 °C, as

the film-substrate system achieved a lower free-energy. Zhao et al. achieved similar results

annealing PQT-12, and further identified the domain structure as close-packed π-π stacks oriented

with their (100) axis normal to the substrate surface.67 PQT-12 has anisotropic charge transport, in

that the direction of facile charge transport is along the π-π stacks, or along the length of the crystals

and parallel to the substrate. While PQT-12 displays excellent mobility in OTFTs, the direction of

charge transport in those devices is across the substrate in the direction of π-π stacking, whereas

in OPVs charge must travel perpendicular to the substrate, against the facile charge transport

direction. While PQT-12 is known to perform poorly in BHJs due to nano-crystalline phase

separation,69, 77, 79 it is not possible to determine the OPV performance effects of charge anisotropy

with BHJs due to their complex mixed morphology.

3.3.3 Ultraviolet-Visible Spectroscopy of PQT-12 Films

Crystallinity is well-known to affect a material’s absorption. Since AFM studies have shown that

PQT-12’s crystalline domain sizes increase with annealing temperature, it follows that film

absorption will change as well. To investigate the affects of increasing crystallinity on PQT-12

film absorption, ultraviolet- visible (UV/Vis) spectroscopy was conducted on 55 nm thick films of

PQT-12 on Glass/ITO/PEDOT:PSS substrates annealed at 118 °C, 133 °C, 148 °C, and

unannealed(Figure 3.5). A Lambda 25 UV/Vis spectrometer was used to conduct the

32

measurements with a plain Glass/ITO/PEDOT:PSS substrate used as a control. PQT-12 layers

were deposited via spin-coating, as discussed in the Methods section of this thesis. The main

absorption peak of PQT-12 occurs at ~545 nm, with a minor secondary absorption peak at ~590

nm. Films annealed at 118 °C and 133 °C had a 15 % and 8 % higher main absorption peak than

unannealed films, respectively. The film annealed at 148 °C does not follow the expected trend,

with a peak absorbance closer to that of the unannealed film. Film thickness inconsistency between

substrates is an issue inherent to the spin-coating technique, and is a possible reason films annealed

to 118 °C and 133 °C had higher absorbance. While the main absorption peak of the 148 °C film

was very similar that of the unannealed film, it was broader and had a more pronounced secondary

peak at ~590 nm. This is directly caused by increasing length of crystalline domains with annealing

temperature.

3.4 Performance in Organic Photovoltaic Devices

3.4.1 PQT-12 OPV Performance with Cl-BsubPc

Pseudo planar heterojunction (PHJ) OPVs were then fabricated by thermally depositing Cl-

BsubPc, an electron acceptor well known to perform well with thiophene based electron donating

materials. Cnops et al. developed an 8.4% efficient OPV using α6T, Cl-BsubPc, and its

homologue chloro-boron subnaphthalocyanine (SubNc), which is an unprecedented efficiency for

a fullerene-free device.87 Our lab has shown that pairing α6T and Cl-BsubPc creates very stable

Figure 3.5 | UV/Vis absorption coefficient spectra of 55 ± 5 nm thick films

of PQT-12.

33

OPVs with excellent longevity in an outdoor testing environment, which is a crucial concern for

developing commercializable OPVs.86 The combination of high efficiency and stability of α6T/Cl-

BsubPc devices make Cl-BsubPc a promising electron acceptor layer for other thiophenes as well,

which is why we paired it with PQT-12 in the current study.

34

Figure 3.6 | Left: J-V characteristics and Right: EQE spectra of OPV devices optimized around the

electron donor layer thickness. The PQT-12 layer was prepared a) unannealed and annealed at b)

118 °C c) 133 °C and d) 148 °C, paired with Cl-BsubPc. Shading indicates one standard deviation

from the average.

35

OPV devices results are shown in Figure 3.6. The optimal PQT-12/Cl-BsubPc PHJ OPVs were

obtained with a 20 nm layer of PQT-12 annealed at 118°C with the following device structure:

glass/ITO/PEDOT:PSS (35 nm)/PQT-12 (20 nm)/Cl-BsubPc (20 nm)/BCP (7 nm)/Ag (80 nm).

This device structure achieved an open circuit voltage (VOC) of 0.95±0.01 V, a short circuit current

(JSC) of 1.89±0.02 mA cm-2, a fill factor (FF) of 0.58±0.01, and PCE of 1.04±0.02 %. This PCE is

comparable to PQT-12/PCBM BHJ OPVs,69, 77, 79 despite the inherently reduced interfacial area of

a PHJ architecture, which should reduce photocurrent generation. The more energetically

favourable pairing of solution-cast PQT-12 with vapor deposited Cl-BsubPc, rather than solution-

cast PCBM, increases device’s VOC and FF to the extent that they overcome the loss of the reported

JSC. In fact, the FF is considerably higher than typical BsubPc OPVs, which are usually around

0.48.86

Figure 3.7 | Plotted characteristic parameters of OPV devices optimized around the electron donor

layer thickness.

36

Annealing significantly improves device FF across all thicknesses and is the main reason why

annealed layers of PQT-12 have superior performance in devices than unannealed layers for nearly

every layer thickness/annealing temperature combination. A low FF can be indicative of low

charge mobility in one of the layers,22 and this is supported by the improvement in device

performance from annealing the PQT-12 to increase its degree of crystalline order and improve its

hole charge mobility. This result runs contrary to what was reported for annealing PQT-12 in BHJ

OPVs, where increased crystallinity reduces device performance.69, 77 Devices with thicker PQT-

12 layers annealed at 148 °C developed an S-kink, leading to a FF reduction which did not follow

the trend of the other annealed devices. S-kinks are caused by a charge imbalance wherein either

the donor or acceptor layer is significantly more efficient at transporting charge. This leads to

charge accumulation at the donor/acceptor interface, which manifests as a reduction in FF and

VOC. It is unclear why an S-kink only develops in devices with a 148 °C annealed layer, and only

at higher thicknesses.

Apart from S-kinked devices, the VOC varies only marginally with annealing temperature, and was

unaffected by PQT-12 layer thickness. A stable VOC is expected, since VOC is closely related to a

donor material’s highest occupied molecular orbital (HOMO) level together with the lowest

unoccupied molecular orbital (LUMO) energy of the electron acceptor, and therefore an intrinsic

material property.93 Conversely, device JSC was significantly affected by PQT-12 layer thickness.

There exists a clear positive trend between decreasing layer thickness and JSC across all annealing

temperatures, including unannealed devices. Thin layers of 10-20 nm provide optimal device

performance, suggesting that PQT-12 has a short, 10-20 nm exciton diffusion length. Although

exciton diffusion length is known to be influenced by a material’s degree of crystalline order, 17 in

this system the effects are not sufficient to improve the JSC of thicker layered devices.

3.4.2 Comparison of PQT-12, P3HT, and α6T OPVs

To directly compare the performance of PQT-12 to P3HT, materials with similar chemical

composition yet inherently different nano-phase morphology, devices were fabricated with a 30nm

layer of P3HT replacing PQT-12 as the electron donor layer (Figure 3.8). P3HT layers were

37

prepared both unannealed and annealed past its melting temperature (250 °C). All properties of

the underlying device substrate (glass/ITO/PEDOT:PSS) remained unchanged after annealing to

this temperature. The PQT-12 layer of the device used for comparison was annealed at 118 °C, as

this annealing temperature produced the highest efficiency at a 30nm layer thickness.

An unannealed layer of P3HT did produce the highest efficiency OPVs, followed by those with an

annealed PQT-12 layer. The increase in VOC in PQT-12 devices over those with a P3HT layer can

be attributed to the material’s deeper HOMO level. P3HT suffers a slight loss to its VOC upon

annealing, which suggests S-kinking or an additional mechanism related to structural organization

Figure 3.8 | Comparison of the Left: J-V characteristics and Right: EQE spectra of P3HT, PQT-12,

and α6T electron donor layers paired with Cl-BsubPc. Shading indicates one standard deviation

from the average.

Table 3.2 | Characteristic device parameter comparison of P3HT, PQT-12, and α6T donor layers

paired with Cl-BsubPc.

Donor Layer Annealing

Temp.

VOC

[V]

JSC

[mA cm-2] FF

PCE

[%]

α6T Unannealed 1.11 ± 0.01 -5.25 ± 0.15 0.57 ± 0.01 3.35 ± 0.0

P3HT 250°C 0.79 ± 0.01 -1.92 ± 0.03 0.46 ± 0.02 0.69 ± 0.04

P3HT Unannealed 0.82 ± 0.01 -2.43 ± 0.01 0.50 ± 0.02 1.00 ± 0.05

PQT-12 118°C 0.95 ± 0.01 -1.84 ± 0.05 0.55 ± 0.01 0.95 ± 0.02

PQT-12 Unannealed 0.96 ± 0.01 -1.53 ± 0.02 0.36 ± 0.01 0.53 ± 0.02

38

is at play. Annealing the P3HT layer causes a reduction in JSC and FF compared to an unannealed

device, which is the opposite effect from annealing PQT-12. This is likely caused by a superior

ability of the unannealed P3HT layer to dissociate charge across the interface compared to

unannealed PQT-12 due to its shallower HOMO level. However, P3HT has a lower VOC than PQT-

12 as a result.

After annealing, PQT-12 and P3HT devices have a nearly identical JSC and FF. There are two

possible explanations for this: either PQT-12 has better charge transport, but its JSC is limited by

poor charge dissociation at the interface, or the charge transport capabilities of annealed PQT-12

and P3HT are practically equivalent in OPVs. It could be that PQT-12 is not able to take full

advantage of its superior charge transport capabilities even when separated in a PHJ device

architecture due to suboptimal crystal orientation, since PQT-12 has anisotropic charge transport

properties. A further study on the impact of PQT-12 crystal orientation on OPV performance is

needed to better understand this phenomenon.

The EQE spectra of P3HT and PQT-12 devices show neither material contributes significantly to

the photocurrent. The spectra are dominated by the absorption of Cl-BsubPc at 570 nm exhibited

comparable absorption spectra. The absorption of PQT-12 is slightly red-shifted from that of P3HT

due to increased conjugation caused by its higher degree of crystallinity, but since both materials

have such poor absorption, it has a low impact on devices.

A device with a 30 nm donor layer of α6T was also fabricated for comparison. The drawback of

α6T devices is the required fabrication via physical vapor deposition due to α6T’s low solubility,

which is a more complex and expensive technique than casting from solution for ultimate

production of OPVs. As an oligothiophene with no alkyl side-chains, the molecular constituents

of α6T may pack together much more tightly than the polymeric P3HT or the nano crystalline

PQT-12, leading to an increased density of chromophores and a firm -conjugation length of 6

thiophene units. An EQE plot of the three materials in devices clearly shows the results of this

higher chromophore density and -conjugation length, with α6T strongly absorbing at 400-425 nm

while the absorptions of P3HT and PQT-12 are mostly overshadowed by that of Cl-BsubPc and

due to their blue shifted EQE contribution, in this configuration, their -conjugation length is likely

less than 6. The low contribution of P3HT and PQT-12 to the photocurrent is a major factor in

why these materials have a lower JSC, and thus efficiency, than α6T. Nonetheless, devices with

39

PQT-12 gain on the VOC of the α6T devices, which is a promising improvement over P3HT. This

significantly implies that moving forward, to leverage solution casting electron donating materials

in a PPHJ OPV configuration paired with a BsubPc electron acceptor, one needs to design a

thiophene based electron donating polymer/material with nano-crystalline morphology to achieve

a high VOC while also having a -conjugation length appropriate to achieved EQE/JSC

contributions/absorption in the range of 375-500 nm.

3.5 Chapter Conclusion

In this chapter we have demonstrated that a nano-crystalline electron donating material PQT-12,

while known to be largely unappealing for application in BHJ OPVs, has potential for application

in pseudo PHJ OPVs. Annealing of the PQT-12 layer enhances performance when applied in

pseudo PHJ devices when paired with Cl-BsubPc, with the FF offering the most significant

improvement over unannealed devices. The highest efficiency is produced by thin, 10-20 nm PQT-

12 layers annealed to 118 °C. This annealing temperature corresponds to short crystal domains

formed after annealing past the liquid-crystalline phase transition. Nano-crystallinity is therefore

shown to be beneficial for PHJ PQT-12 devices, as opposed to the reduction in device performance

seen in BHJs.

PQT-12 devices offer about 15% VOC improvement over P3HT devices, bringing them closer to

the VOC of α6T based devices. The JSC of PQT-12 OPVs remains low due to its minimal

contribution to the device photocurrent generation compared to the broader absorption of photons

from 6T, which does limit the PCE to ~1%. Since the oligothiophene α6T has a comparatively

high photocurrent contribution, it follows that PQT-12’s low contribution stems from its improper

polymeric ordering and orientation, compounded with a lower chromophore density due likely due

to the presence of solubilizing alkyl chains. Additionally, -conjugation length is likely shorter

than 6T indicated by the respective EQE contributions. From a molecular engineering standpoint,

a focal point moving forward will be maintaining thiophene solubility while concurrently

encouraging a densely packed crystalline film through a reduction of solubilizing alkyl chains and

enhanced -conjugation length.

40

Chapter 4

PBTZT-stat-BDTT-8 in pseudo-Planar Heterojunction Organic Photovoltaics

4.1 Introduction

Organic photovoltaic devices (OPVs) have great potential for energy production, offering

numerous advantages over traditional Si cells such as their light-weight design and good

performance in low lighting conditions.8, 94, 95 OPVs require multiple organic photoactive materials

to absorb light and convert it into electricity, and the choice of these materials is critical for

enhancing device performance. Some of the most successful, highest efficiency OPVs are

composed of conjugated ‘donor-acceptor’ (D-A) copolymers. D-A copolymers are composed of

an electron-rich donor moiety and an electron deficient acceptor moiety. The alternating donor and

acceptor characteristics of mers in the polymer backbone causes the polymer’s molecular orbitals

to hybridize, which changes their energy and narrows the molecular bandgap. The highest orbital

molecular orbital (HOMO) of the donor moiety and the lowest unoccupied molecular orbital

(LUMO) of the acceptor moiety are largely responsible for the energy levels of the resulting

copolymer.96 This unique design allows for a high degree of control over the copolymer’s energy

levels, as the HOMO and LUMO levels may be tuned separately based on adjustments to the two

moieties.97

Benzodithiophene (BDT) is a well-studied, high performance donor moiety in D-A copolymers.

The excellent photovoltaic performance of the benzodithiophene (BDT) comonomer is has made

it a popular building block in OPVs, with hundreds of BDT-based polymers and small molecules

reported since its first use in 2008.98 BDT has a rigid, planar conjugated structure which allows for

highly tunable energy levels and band gaps, in addition to a high hole mobility. BDT is often paired

with the acceptor moiety thieno[3,4-b]thiophene (TT) to form a final polymer design. Amongst

this class of polymers, Poly[[4,8-bis[(2-ethylhexyl)oxy]benzo[1,2-b:4,5-b']dithiophene-2,6-

diyl][3-fluoro-2-[(2-ethylhexyl)carbonyl]thieno[3,4-b]thiophenediyl]] (PTB7) is one of the most

studied copolymers in BHJ OPVs. Standard stack architecture BHJs of PBT7:PC71BM achieve

efficiencies of on average ~7.78%,99-107 with the very best devices achieving 9.98%.106 Building

41

off of the success of PTB7, the TT acceptor moiety was replaced by benzothiadiazole (BT), an

even stronger electron-withdrawing comonomer. The theory was that the low-lying HOMO of

BDT, combined with the deep LUMO level of BT, would result in a narrower optical bandgap and

higher open circuit voltage (VOC) in BDT-BT copolymers.108

PBTZT-stat-BDTT-8 is a state-of-the-art D-A copolymer first developed in 2015 by Merck

Chemicals Ltd to address the commercialization issues facing OPVs.109 The backbone of PBTZT-

stat-BDTT-8 is composed of BDT and BT comonomers separated by thiophene moieties, which

reduce steric hindrance and tune the energy levels and mobility of the polymer (Figure 4.1).108

Figure 4.1 | Chemical structures of a) BsubPc electron acceptor molecules and b) polymeric

electron donor materials, with c) their reported frontier orbital molecular energy levels. d) Device

schematic of pseudo-PHJ devices in which the BsubPc is varied.

42

They studied this new polymer in BHJ OPVs paired with PC60BM, and reported an excellent

efficiency of 9.3% for lab-scale devices. The polymer absorbs strongly in the visible spectrum,

with peaks at 600 nm and 640 nm. Its HOMO level was reported as -5.4 eV with a LUMO level

of -3.7 eV and a bandgap of 1.7 eV. The excellent photovoltaic properties of PBTZT-stat-BDTT-

8 and proven efficacy in BHJs make this polymer a prime candidate for studying polymer/small

molecule OPVs, specifically boron subphthalocyanines (BsubPcs).

BsubPcs are a well-studied and promising family of small-molecule electron acceptor materials

with which our lab has extensive experience in OPVs (Figure 4.1).57, 82, 83, 86, 110 These molecules

have intense optical absorption in the 560-600 nm spectral region and highly tunable energy levels

as a result of easily adjustable axial and peripheral substituents.57 Due to their high performance

but limited solubility in common organic solvents, BsubPcs are typically incorporated into planar

heterojunction (PHJ) devices paired with similarly insoluble small-molecule materials.55, 57, 82, 86,

111-113 Chloro-boron subphthalocyanine (Cl-BsubPc) in particular has produced excellent

performance OPVs when employed as an electron acceptor layer in PHJ devices. Cnops et al.

achieved an unprecedented high power conversion efficiency (PCE) of 8.4% for a non-fullerene

device using α6T, Cl-BsubPc, and its homologue chloro-boron subnaphthalocyanine (Cl-

BsubNc).87

As PHJ OPVs require active layers to be vapor deposited, polymeric electron donor materials are

often overlooked when designing the device stack due to their incompatibility with the fabrication

technique. As a result, there are few examples of BsubPc acceptors paired with polymeric donor

materials. To fabricate these devices, a hybrid deposition technique must be used since insoluble

small-molecule organic semiconductors cannot be used in BHJs and heavy polymeric materials

cannot be vapor deposited for PHJs. One method of bypassing these limitations is to solution

deposit a polymeric electron donor layer and then vapor deposit an electron acceptor layer on top

to achieve a pseudo-PHJ (PPHJ) device. This device architecture retains the simple interface and

separate electron donor/acceptor layer structure of PHJs which is ideal for characterization of new

polymer/small-molecule photoactive layer pairings.

Although soluble BsubPcs are uncommon, BsubPcs such as phenoxy-hexachloro-boron-

subphthalocyanine (PhO-Cl6BsubPc) and chloro-hexachloro-boron-subphthalocyanine (Cl-

Cl6BsubPc) may be solution processed, and have previously been reported for use as electron

43

acceptor materials in BHJ OPVs with PCEs of 3.5% and 4.0%, respectively.114-116 In a recent

paper from our lab authored by senior PhD student Kathleen Sampson, PhO-Cl6BsubPc was paired

with 10 high performance electron donating polymers in BHJ OPVs.117 All 10 polymers had been

designed for pairing with PC61BM, but due to fullerene’s weak absorption, lack of energy level

tunability, and energy intensive synthesis, it was desirable to screen these polymers for use with a

non-fullerene acceptor. Screening parameters included the use of 1,2-dichlorobenzene or o-xylene

as the solvent with 0 or 5 vol% 1,2-dimethoxybenzene solvent additive, as well as with and without

annealing at 120°C for 5 min. The general trend in efficiency of resulting PhO-Cl6BsubPc/polymer

devices were found to match those with PC61BM as the electron acceptor, with the best device

containing PhO-Cl6BsubPc/PBTZT-stat-BDTT-8 which achieved a VOC of 0.82 V, a JSC of 11 mA

cm-2, a FF of 0.67, and a PCE of 6.1%. These results were achieved using an inverted stack

architecture due to preliminary results indicating low performance of standard stack BHJs. It was

unclear as to why standard stack BHJs had such lower performance than their inverted stack

counterparts in those preliminary studies. The construction of a PPHJ-type OPV is essential to

elucidate this issue due to their simplified charge transport across a single interface, rather than

across the random interpenetrating interface of a BHJ.

In the current work, we report on three BsubPcs as vapor deposited small-molecule electron

acceptors paired with the solution-processed polymeric electron donor PBTZT-stat-BDTT-8 in

PPHJ devices. First, the layer thickness of PBTZT-stat-BDTT-8 was optimized when paired with

a constant thickness of Cl-BsubPc to study the effects of polymer layer thickness on device

performance. PBTZT-stat-BDTT-8 was then incorporated into the high-performance ‘Cnops

stack’ and evaluated as a potential replacement for α-sexithiophene. PPHJs of PBTZT-stat-BDTT-

8 paired with Cl-BsubPc, PhO-Cl6BsubPc and Cl-Cl6BsubPc at varied layer thicknesses were then

fabricated and assessed, and issues surrounding device replication during this thesis are discussed.

Finally, the performance of pseudo-PHJ devices of PBTZT-stat-BDTT-8 paired with PhO-

Cl6BsubPc and Cl-Cl6BsubPc were compared to their BHJ architecture counterparts.

44

4.2 PBTZT-stat-BDTT-8 Film Profilometry

All films in this study were deposited

via the spin-coating technique, as

PBTZT-stat-BDTT-8 is widely soluble

in most common organic solvents. All

solutions were prepared in

dichlorobenzene (DCB) to maintain

consistency across multiple studies. As

stated previously in this thesis, film

thickness is linear with solution

concentration. To determine the relation

for PBTZT-stat-BDTT-8, solutions in

DCB were diluted into concentrations of 5, 10, and 15 mg/mL, then prepared and spin-cast as

described in the Methods section of this thesis. Substrates of Glass/ITO/PEDOT:PSS were used,

which are the same as used for OPV devices. Profilometry was conducted on all resulting films.

Two substrates were prepared for each solution concentration, and four profilometry

measurements were conducted per substrate (spread evenly across the step-edge). The thickness

of the PEDOT:PSS layer was taken to be 35 nm. The results of PBTZT-stat-BDTT-8 are plotted

in Figure 4.2. The linear trendline had a good R2 value of 0.9956, so all concentrations and film

thicknesses used in this study were based off this linear relation.

4.3 Performance and Optimization of OPVs based on PBTZT-stat-BDTT-8/Cl-BsubPc

As this was the first investigation of the electron donating PBTZT-stat-BDTT-8 in PPHJ devices

with BsubPcs, we began our evaluation of the performance of this pairing by optimizing the layer

thickness of PBTZT-stat-BDTT-8 paired with the most common BsubPc, Cl-BsubPc. Previously

in our lab, OPVs based on the pairing of solution-cast poly(3-hexylthiophene-2,5-diyl) (P3HT)

and vacuum deposited Cl-BsubPc were developed as a baseline solution-cast electron donor OPV

for other electron-donating polymers with BsubPc electron acceptors.118 The device architecture

of this baseline device consisted of: ITO/PEDOT:PSS/P3HT (55nm)/Cl-BsubPc (20nm)/BCP

(7nm)/Ag (80nm). This baseline will serve as a useful reference point when evaluating the

Figure 4.2 | Profilometry results of PBTZT-stat-

BDTT-8 on Glass/ITO/PEDOT:PSS substrates.

45

performance of solution-cast PBTZT-stat-BDTT-8 paired with vacuum deposited Cl-BsubPc due

to the similarity of the materials and device architectures.

OPV devices with the architecture ITO/PEDOT:PSS/PBTZT-stat-BDTT-8/Cl-BsubPc

(20nm)/BCP (7nm)/Ag (80nm) were fabricated. The layer thickness of PBTZT-stat-BDTT-8 was

varied from 10-60 nm, while Cl-BsubPc thickness was held constant at 20 nm. The current density-

voltage (J-V) characteristics and external quantum efficiency (EQE) measurements are plotted in

Figure 4.3, with characteristic parameters summarized in Table 4.1. The optimal PBTZT-stat-

BDTT-8/Cl-BsubPc devices were obtained with a 20 nm layer of PBTZT-stat-BDTT-8. This

device produced a large open circuit voltage (VOC) of 1.19 V, much greater than the VOC typically

seen for fullerene acceptors and even greater than the VOC of 0.8 V of the baseline P3HT/Cl-

BsubPc device.36 The fill factor (FF) of 0.51 is typical for BsubPc OPVs, whose FF are usually

around 0.48.36, 83, 119 While these are promising VOC and FF results, the PPHJ PBTZT-stat-BDTT-

8/Cl-BsubPc devices suffered from a low JSC of -3.0 mA/cm2, which was main factor limiting

46

device PCE to 1.82%. This can be attributed to the inherently small interfacial area between the

two photoactive materials when compared to a BHJ architecture device due to its reduced ability

to dissociate excitons. However, the JSC and PCE were much improved compared to the PPHJ

baseline P3HT/Cl-BsubPc device, which had a JSC of 2.79 mA/cm2 and a PCE of 0.98%.36

Varying the thickness of PBTZT-stat-BDTT-8 from 60 nm to 10 nm revealed a clear positive trend

between device JSC and decreasing donor layer thickness until 20 nm (Figure 4.4). A layer

thickness lower than 20 nm severely impacted device FF and lowered device VOC. These results

suggest that PBTZT-stat-BDTT-8 has a short exciton diffusion length of ~20 nm, which is typical

of conjugated polymer films.64, 91 Layers thicker than 20 nm suffer loses to the photocurrent due

to excitons decaying to the ground state before they can reach the donor/acceptor interface and

dissociate. The thickest 60 nm layer of PBTZT-stat-BDTT-8 achieved nearly the same JSC as the

Figure 4.3 | Left: J-V curves and Right: EQE spectra of PBTZT-stat-BDTT-8/Cl-BsubPc devices

with varying electron donor layer thickness, constant 20 nm electron acceptor layer thickness.

Shaded error bars represent ±1 standard deviation from the mean.

Table 4.1 | Characteristic parameters of PBTZT-stat-BDTT-8/Cl-BsubPc devices. The layer

thickness of Cl-BsubPc was constant at 20 nm.

47

thinnest layer of 10 nm. One would expect that a thicker photoactive layer would absorb more

photons and further contribute to the photocurrent, but this is not the case due to the device’s

absorption spectra. As layer thickness was increased from 20 nm to 60 nm, the EQE plot shows

Cl-BsubPc’s absorbance peak at 580 became gradually weaker while the PBTZT-stat-BDTT-8

absorbance region at ~500-680 nm remained relatively stable. These results suggest that a thicker

PBTZT-stat-BDTT-8 layer blocks the absorbance of Cl-BsubPc without increasing its contribution

the photocurrent. As with device FF, this issue is caused by the low exciton diffusion length of

PBTZT-stat-BDTT-8. In devices with thicker electron donor layers, excitons decay to ground state

energy before they can reach the electron donor/acceptor interface to dissociate, dissipating the

energy from absorbed photons in the process. The minor spectral overlap between PBTZT-stat-

BDTT-8 and Cl-BsubPc exacerbates this issue since photons are absorbed in the electron donor

layer first so there are fewer photons available for Cl-BsubPc to absorb. Ideally there should be no

spectral overlap between the two photoactive materials to maximize photon absorption.

Neither the VOC nor the FF of devices varied significantly with PBTZT-stat-BDTT-8 film

thickness. Device VOC is approximated by a donor material’s highest occupied molecular orbital

(HOMO) level together with the acceptor material’s lowest unoccupied molecular orbital (LUMO)

level, and is ideally a characteristic physical quantity of the two photoactive layers.93 As such, a

stable VOC was expected because it is independent of layer thickness. Device FF is much more

complex, as it depends on many variables such as charge mobility and shunt/series resistance of

the material layers.22, 120 FF remained stable across all thicknesses apart from 10 nm, varying by

Figure 4.4 | Characteristic parameters of PBTZT-stat-BDTT-8/Cl-BsubPc devices, visualized to

display trends with increasing thickness.

48

only ±0.01. When the layer was reduced to 10 nm, the FF decreased dramatically from ~0.5 to

0.37. The most likely cause of this sharp drop is the development of pinholes through the

incomplete film coverage of such a thin solution-cast layer, which would raise the shunt resistance

of the cell. This would result in a lower FF, which is exactly what is seen for the 10 nm layer

device. For the other devices, the stable FF indicates that there are no significant charge mobility

or shunt/series resistance issues occurring within the cells as the layer thickness varies.

4.4 PBTZT-stat-BDTT-8 in “Cnops Stack”

The PHJ stack designed by Cnops et al. in 2014 remains the pinnacle of high performance non-

fullerene OPVs, having achieved a PCE of 8.4% with the following device architecture:

ITO/PEDOT:PSS/α-6T (60 nm)/Cl-BsubNc (12 nm)/Cl-BsubPc (15 nm)/BCP (10nm)/Ag.121 The

‘Cnops stack’ of BsubNc/BsubPc uses an energy transfer cascade in which the light absorbing

BsubPc layer is spatially separated from the electron donor/acceptor interface but may still

contribute to the device photocurrent via energy transfer to BsubNc.

Since PBTZT-stat-BDTT-8 was successful in OPVs when paired with Cl-BsubPc as detailed in

the previous section, it was expected to have even better performance as a solution-processible

replacement for α-6T in the Cnops stack. To test this theory, OPVs containing:

ITO/PEDOT:PSS/PBTZT-stat-BDTT-8 (30 nm)/Cl-BsubNc (12 nm)/Cl-BsubPc (15 nm)/Ag were

fabricated. A 30 nm layer of PBTZT-stat-BDTT-8 was used instead of 60 nm as it is closer to the

optimal thickness with BsubPc as determined in the previous section. Figure 4.5 depicts the results

of OPV testing, with the characteristic parameters shown in Table 4.2.

Unfortunately, the PBTZT-stat-BDTT-8/Cnops stack devices were relatively unsuccessful. The

exceedingly low FF made for some very strangely shaped J-V curves, severely limiting the device

PCE. Furthermore, the error bars were very large compared to those of the BsubPc devices. To

further explore the reason for this poor performance, another device was made containing PBTZT-

stat-BDTT-8/BsubNc, leaving out BsubPc and the energy transfer cascade. The similarly shaped

J-V curve, low FF, and large error bars of these devices helps elucidate why the PBTZT-stat-

BDTT-8/Cnops stack was unsuccessful. Clearly there is some energetic or interface issue between

PBTZT-stat-BDTT-8 and BsubNc that does not exist for BsubPcs. Since BsubNc has a shallower

49

HOMO and deeper LUMO than BsubPc, the more favorable energy alignment at the electron

donor/acceptor interface compared to BsubPc was expected to improve device performance. The

VOC and JSC of both the PBTZT-stat-BDTT-8/BsubNc and PBTZT-stat-BDTT-8/Cnops stack

devices were reasonable and the EQE spectra looked as expected for these materials, suggesting

that the energy levels were indeed favorable matched. It is most likely the low FF was caused by

issues in the interfacial morphology between PBTZT-stat-BDTT-8 and BsubNc. While the nature

of these problems was not further investigated for this thesis, they might be related to material

aggregation or delamination at the interface.

Table 4.2 | Characteristic parameters of OPV devices with varying electron acceptor layer. The

layer thickness of PBTZT-stat-BDTT-8 was constant at 20 nm.

Figure 4.5 | Left: J-V curves and Right: EQE spectra of OPV devices with varying electron

acceptor layer and constant 20 nm PBTZT-stat-BDTT-8 electron donor layer. Shaded error bars

represent ±1 standard deviation from the mean.

50

4.5 OPV Comparison of BsubPc Electron Acceptor Layers with PBTZT-stat-BDTT-8

Having fabricated and characterized PPHJ architecture PBTZT-stat-BDTT-8/Cl-BsubPc devices,

we then proceeded to test PBTZT-stat-BDTT-8 in OPVs paired with other BsubPc derivatives to

investigate whether this polymer is an effective electron donor for just Cl-BsubPc, or a wider

variety of BsubPcs. The halogenated BsubPc derivatives Cl-Cl6BsubPc and PhO-Cl6BsubPc were

selected for this study due to their relatively good solubility for BsubPcs, which allowed for the

comparison of PPHJ architecture devices to BHJ devices detailed in the next section.

PPHJ OPVs were fabricated with 30 or 60 nm of PBTZT-stat-BDTT-8 paired with 10 or 20 nm

layers of either Cl-BsubPc, Cl-Cl6BsubPc, or PhO-Cl6BsubPc. Their J-V characteristics and EQE

spectra are shown in Figure 4.6, while their characteristic parameters shown in Table 4.3 and Table

4.4. Preliminary rough optimization of layer thicknesses was performed to determine the

functionality of the novel material combinations. The highest efficiency device had the architecture

ITO/PEDOT:PSS (35 nm)/PBTZT-stat-BDTT-8 (30 nm)/Cl-BsubPc (20 nm)/BCP (7 nm)/Ag (80

nm). The Cl-BsubPc device had the highest overall efficiency due mainly to its substantial VOC,

which was expected based on the HOMO-LUMO offsets of these BsubPcs with PBTZT-stat-

BDTT-8. However, both the PhO-Cl6BsubPc and Cl-Cl6BsubPc devices achieved better FFs as a

result of their higher shunt resistances. Cl-Cl6BsubPc in particular had an excellent FF of 0.66,

which is impressive for OPVs in general, let alone for a device containing BsubPc. Devices

containing PhO-Cl6BsubPc were very consistent, with reduced sensitivity to electron acceptor

layer thickness compared to the other BsubPcs tested. While the origin of these improvements

remains unclear, it is possible that change in molecular ordering caused by additional chlorination

played a role in reducing the leakage current of devices, which corresponds to higher shunt

resistance.

As expected, trends in the VOC between electron acceptor materials were consistent with trends in

their LUMO levels, with Cl-BsubPc achieving the highest VOC of 1.16 V, followed by PhO-

Cl6BsubPc with a VOC of 0.91 V, and finally Cl-Cl6BsubPc with a VOC of 0.84 V. The Cl-

Cl6BsubPc and PhO-Cl6BsubPc devices displayed many of the same trends seen for Cl-BsubPc,

such as improved performance with a thinner electron donor layer due to increased JSC. When the

layer thickness of the BsubPc electron acceptor layer was reduced to 10 nm, in all cases device

51

Table 4.3 | Characteristic parameters of PBTZT-stat-BDTT-8/BsubPc devices with a 10 nm

electron acceptor layer.

Figure 4.6 | Left: J-V characteristics and Right: EQE spectra of OPV devices containing PBTZT-

stat-BDTT-8 as the electron donor layer and Cl-BsubPc, Cl-Cl6BsubPc, or PhO-Cl6BsubPc as

electron acceptor layer. Shaded error bars represent ±1 standard deviation from the average.

52

performance was reduced across all characteristic parameters. The one exception to this is the FF

of PhO-Cl6BsubPc devices which marginally improved, demonstrating its good consistency across

layer thicknesses. The lowered Jsc of the 10 nm BsubPc devices is a direct result of the thinner

absorbing layer capturing fewer photons than the 20 nm layer, leading to a reduced contribution

to the device photocurrent. This effect can be clearly seen in the device EQE spectra, where the

BsubPc peak at 570 nm is much stronger with 20 nm layers. As in the PBTZT-stat-BDTT-8/Cl-

BsubPc devices of the previous section, the 60 nm PBTZT-stat-BDTT-8 layer blocks light from

reaching the BsubPc layer due to overlap in their absorption spectra. This causes a reduction in the

BsubPc absorbance peak for devices with a thicker electron donor layer.

Both thin-layer Cl-BsubPc devices, as well as the thin-layer Cl-Cl6BsubPc device with the 60 nm

layer of PBTZT-stat-BDTT-8, had dramatically reduced performance. This poor performance can

be explained by the development of S-kinks within certain devices. S-kinks are caused by an

imbalance of charge within the device wherein one active layer transports charge faster than the

other, creating a charge buildup at the interface.22 This charge accumulation manifests as a

reduction in the device VOC and FF. The Cl-Cl6BsubPc device had a particularly pronounced S-

kink, indicating an extreme case of interface charge accumulation.

4.5.1 Overcoming Replication Issues

During the course of device fabrication there were some issues with the reproducibility of PBTZT-

stat-BDTT-8/Cl-Cl6BsubPc devices. There had never been any reproducibility issues in the

previous section of this study concerning the optimization of PBTZT-stat-BDTT-8/Cl-BsubPc,

which is a fairly similar material, so the fact that these problems existed for Cl-Cl6BsubPc devices

was unexpected.

Table 4.4 | Characteristic parameters of PBTZT-stat-BDTT-8/BsubPc devices with a 20 nm

electron acceptor layer.

53

The devices in question were all fabricated in the same day on two different device runs (Appendix

A). While the two devices of PBTZT-stat-BDTT-8/Cl-Cl6BsubPc (20 nm) and PBTZT-stat-

BDTT-8 (30 nm)/Cl-Cl6BsubPc (10 nm) had poor FFs, the device with PBTZT-stat-BDTT-8 (60

nm)/Cl-Cl6BsubPc (10 nm) had an exceptionally high FF of 0.7. The main contributor to the high

FF of this device was its excellent shunt resistance (Rsh) of ~15,200 Ω, an order of magnitude

greater than the shunt resistances of any other device architecture tested. Rsh is related to the current

leakage within an OPV caused by film pinholes or charge traps.22, 122 Larger values of Rsh indicate

low levels of current loss which improves device FF. This result was unusual due to the

comparatively low FFs of all other Cl-Cl6BsubPc devices apart from this one, so it was re-tested

to verify the results.

The subsequent devices produced substantially different results from the initial OPVs. The

baseline devices used to test the integrity of our vacuum deposition system indicated no changes

in the system. While spin-coating error is known to produce some substrate-to-substrate variation

in OPV performance due to minor inconsistencies in deposition technique, this error is not

sufficient to explain such a noticeable discrepancy in results [citation]. It was theorized that such

a difference in OPV results was most likely caused by contamination in one of the layer materials.

Two separate baselines were fabricated, one with PBTZT-stat-BDTT-8 (inverted stack BHJ with

PC60BM as electron acceptor) and the other with Cl-Cl6BsubPc (PHJ with α6T as electron donor).

The similarity of these baselines to previous tests indicated no contamination had occurred in either

material, so the initial theory was proven to be false. Since the original devices had been deposited

slightly slower than usual, at 0.6 Å/s rather than the usual 1 Å/s, it was theorized that the slower

deposition rate may have given Cl-Cl6BsubPc a chance to self-organize on the surface of PBTZT-

stat-BDTT-8 and form a more cohesive interface, which would explain the unusually high shunt

resistance of the high FF initial device. However, when this theory was tested by fabricating the

same devices with a slower 0.6 Å/s deposition rate, results were consistent with the more recent

PBTZT-stat-BDTT-8 (60 nm)/Cl-Cl6BsubPc (10 nm) OPV devices.

Contamination was once again considered. Consistent results for the PBTZT-stat-BDTT-8/Cl-

Cl6BsubPc (10 nm) had been fabricated since the initial OPVs, which suggests that a source of

contamination had existed the day those initial OPVs were fabricated which had since been

removed. As the PBTZT-stat-BDTT-8/Cl-Cl6BsubPc (20 nm) devices had been fabricated on the

same day as the irreproducible PBTZT-stat-BDTT-8/Cl-Cl6BsubPc (10 nm) devices, they were re-

54

tested twice on two different days. The resulting devices were consistent with one another, and

inconsistent with the initial round of OPVs. Since all devices made on that one initial day were

irreproducible, and reproducible devices had been made since, it was concluded that the

inconsistencies arose from contamination on that day. The most likely source of contamination is

the aluminum foil that shields the dividers between crucibles, which must be changed between

depositions of different materials. It is possible some PhO-Cl6BsubPc was introduced during

deposition of Cl-Cl6BsubPc due to human error causing the Al foil to go unchanged. The study

proceeded using the reproducible results, and efforts were made to prevent this error from

occurring again.

4.6 PBTZT-stat-BDTT-8 in Cl-Cl6BsubPc and PhO-Cl6BsubPc OPVs – BHJ vs PPHJ Architecture

Cl-Cl6BsubPc and PhO-Cl6BsubPc are halogenated BsubPc derivatives with adequate solubility

characteristics for solution deposition, leading to their incorporation into high performing BHJ

OPVs with other electron donor materials.114-116 Their solubility and proven functionality in BHJs

made these materials prime candidates for use in the first PBTZT-stat-BDTT-8/BsubPc BHJ

OPVs. As PBTZT-stat-BDTT-8 has achieved exceptional performance as an electron donor layer

with PC61BM in BHJ OPVs, it was initially expected that they would have success in BHJ

architecture devices with BsubPc.

The preliminary fabrication of PBTZT-stat-BDTT-8 with Cl-Cl6BsubPc or PhO-Cl6BsubPc in

standard stack BHJ OPVs was performed by Kathleen Sampson ahead of this study of PBTZT-

stat-BDTT-8 in PPHJ architecture devices (Figure 4.7). Results from these initial tests found that

standard-stack BHJ architecture devices had substantially lower VOC performance than expected

based of the HOMO/LUMO levels of the electron donor and acceptor materials, in addition to low

FFs. While the cells retained a high JSC typical of the large-interfacial area BHJ architecture, poor

results in the other device parameters limited the PCE to 1.56% with Cl-Cl6BsubPc and 2.07%

with PhO-Cl6BsubPc. Duan et al. reported similarly low performance with standard-stack BHJ

OPVs containing PTB7:BsubPc.123 They attributed the poor results to unfavorable vertical phase

segregation, which is likely the same reason for the low VOC and FF in our current PBTZT-stat-

BDTT-8/BsubPc devices due to the similarities in device structure and layer materials.98

55

Figure 4.8 | Comparison of the Left: J-V characteristics and Right: EQE spectra of PBTZT-stat-

BDTT-8/BsubPc OPVs in either pseudo-Planar Heterojunction or Bulk Heterojunction

architectures. Shaded error bars represent ± 1 standard deviation from the average.

Table 4.5 | Characteristic parameters of PBTZT-stat-BDTT-8/BsubPc devices in either pseudo-

Planar Heterojunction or Bulk Heterojunction architectures.

Figure 4.7 | Comparison of VOC and FF of pseudo-Planar Heterojunction or Bulk Heterojunction

architectures.

56

Vertical phase segregation is an issue with mixed solution deposition technique employed for BHJ

OPVs, caused by materials diffusing to the wrong electrode during deposition. This is not an

obstacle for OPVs with PPHJ architecture because the layers are deposited separately in their

favorable vertical locations, ensuring that electron donor/acceptor layers are contained at the

appropriate electrode. A valuable comparison may be made between standard-stack BHJ and PPHJ

device architectures to scope out the true potential of the PBTZT-stat-BDTT-8/BsubPc pairing,

which may then be realized in future work through the use of an inverted-stack BHJ architecture

to avoid the obstacle of suboptimal vertical phase segregation.

In the previous section, OPVs containing PBTZT-stat-BDTT-8/Cl-Cl6BsubPc or PhO-Cl6BsubPc

were investigated in a PPHJ architecture. The best devices from that section are now compared to

their respective BHJ architecture devices (Figure 4.7) (Table 4.5). OPVs in the PPHJ architecture

had much higher VOC and FF than their BHJ equivalents, with Cl-Cl6BsubPc achieving

improvements of 48% and 118% and PhO-Cl6BsubPc achieving improvements of 17% and 46%,

respectively (Figure 4.8). The substantially higher VOC and FF of the PPHJ devices represents the

actual electrical properties of these PBTZT-stat-BDTT-8/BsubPc pairings rather than those of the

standard-stack BHJs which suffered from serious bimolecular recombination losses. The EQE

spectra show a much broader absorption peak for BHJ devices due to the greatly increased

contribution of PBTZT-stat-BDTT-8 to the photocurrent in this architecture compared to in PPHJ

devices.

Despite their improved VOC and FFs, the PCEs of PPHJ architecture devices were nearly

equivalent to their BHJ counterparts due to the high JSC typical of BHJ architecture devices.

Although the efficiencies of both device configurations were similar, their max power points

(MPPs) occurred at completely different locations on the J-V curve. The MPP of PPHJ devices

occurred at high voltages and low currents, while the MPP of BHJ devices occurred at low voltages

and high currents. In real-world applications of OPVs running at their MPPs, it is more beneficial

to have higher operating voltage than current due to the inherent difficulty of transporting a large

electrical current through a thin film.124 For this reason, there is a clear operational advantage in

pursuing PBTZT-stat-BDTT-8/BsubPc devices in a PPHJ configuration.

57

4.7 Chapter Conclusion

In this chapter we have demonstrated that PBTZT-stat-BDTT-8 is a viable copolymeric electron

donating material for use in PPHJ architecture OPVs when paired with BsubPcs. The layer

thickness of PBTZT-stat-BDTT-8 was optimized in OPV devices with Cl-BsubPc, yielding

exceptionally high VOCs of 1.19 V. The best device had a PCE of 1.82% which represents an 86%

increase compared to the standard PPHJ baseline of P3HT/Cl-BsubPc.

PBTZT-stat-BDTT-8 was paired with Cl-Cl6BsubPc and PhO-Cl6BsubPc in PPHJ OPVs to test its

functionality with a wider variety of BsubPcs. After overcoming replication issues, a FF of 0.66

was demonstrated with Cl-Cl6BsubPc devices, significantly higher than the typical FF of ~0.5 of

BsubPc containing OPVs. When paired with PhO-Cl6BsubPc, the FF was found to be 0.57, and

devices were less sensitive to electron acceptor layer thickness than other BsubPcs tested.

Peripheral chlorination was highly beneficial to the FF of devices studied in this chapter. More

work is needed to better understand the precise impact of peripheral chlorination on device FF, as

it has great significance for the molecular design of BsubPcs for higher efficiency OPVs.

When compared to their equivalent BHJ architecture devices, the ideal charge interface of PPHJ

OPVs allowed these devices to have much improved VOC and FF of 48% and 118% for Cl-

Cl6BsubPc devices and 17% and 46% for PhO-Cl6BsubPc devices, respectively. These results

confirm the theory that BHJ devices were hindered by poor vertical morphology and provide an

ideal charge transport control off which the BHJ morphology may be improved. Although PPHJ

devices achieved comparable efficiencies to their BHJ equivalents, their higher VOC and lower JSC

could provide an operational advantage when incorporated into solar modules.

58

Chapter 5

Summary and Future Work

5.1 Summary

The work carried out in this thesis has thoroughly investigated the functionality of the solution

processed photoactive polymers PQT-12 and PBTZT-stat-BDTT-8 in PPHJ OPV devices when

paired with small molecule BsubPcs. Over the course of this work, the effects of crystallinity, layer

thickness, and electron donor/acceptor pairings on the characteristic parameters of PPHJ OPVs

were studied and discussed, and future considerations for the design of photoactive polymers for

BsubPcs were proposed.

In Chapter 3, the highly crystalline polythiophene PQT-12 was studied for use in PPHJ OPVs with

Cl-BsubPc. PQT-12 has had limited success in the past in BHJ architecture devices due to its

tendency to phase separate. Using DSC analysis, the thermal transitions of PQT-12 were shown to

occur at ~115 °C and 133 °C, corresponding to the crystal to liquid crystalline phase transition and

liquid crystalline to isotropic phase transition, respectively. The impact of these phase transitions

on crystal morphology were captured using AFM imaging, revealing that PQT-12 crystal growth

occurs anisotropically along the long dimension with higher annealing temperatures producing

longer, fibular crystals. Changes in crystallinity had only a minor impact on film absorbance, as

shown by UV-vis. The first PPHJ devices combining PQT-12 and Cl-BsubPc were fabricated and

evaluated through light testing. The optimal devices had a structure of: glass/ITO/PEDOT:PSS (35

nm)/PQT-12 (20 nm annealed at 118°C)/Cl-BsubPc (20 nm)/BCP (7 nm)/Ag (80 nm), which

achieved an average PCE of 1.04±0.02 %, similar to the efficiencies of previously reported PQT-

12/fullerene BHJ OPVs despite having reduced interfacial area. The more energetically favourable

pairing of solution-cast PQT-12 with vapor deposited Cl-BsubPc, rather than solution-cast PCBM,

increased device’s VOC and FF to the extent that they overcame the loss of the reported JSC. It was

determined that thinner layers of PQT-12 produced higher current due to the polymer’s short, 10-

20 nm diffusion length. Mid-sized oriented crystals were shown to provide the most improvement

to OPVs by improving device FF. Comparison of OPVs containing PQT-12 to those with P3HT

or α6T revealed that polymeric electron donors offered a much lower contribution to the device

photocurrent, which was attributed to their lower chromophore density and suboptimal crystalline

59

packing to facilitate charge transport in the vertical direction. In terms of polymer engineering, a

potential route forward was identified by decreasing the length of the solubilizing alkyl chains to

encourage higher polymer packing density, as well as enhancing π conjugation length.

Chapter 4 moved on to investigate the functionality and performance of a BDT and BT-based

copolymer PBTZT-stat-BDTT-8 in PPHJ OPVs. Unlike in the previous chapter, this polymer is

amorphous and has had great success in BHJ devices when paired with PC61BM. The PBTZT-stat-

BDTT-8/Cl-BsubPc pairing was studied in PPHJ OPVs for the first time, and devices were

optimized around the thickness of the polymeric electron donor layer (10-60 nm). The best device

had the architecture: ITO/PEDOT:PSS (35 nm)/PBTZT-stat-BDTT-8 (20 nm)/Cl-BsubPc (20

nm)/BCP (7 nm)/Ag (80 nm), achieving a PCE of 1.82% with an impressive VOC of 1.19V. The

main limiting factor to achieving higher device efficiencies was the short ~20 nm exciton diffusion

length of PBTZT-stat-BDTT-8, in addition to its suboptimal absorption region which overlaps

marginally with that of Cl-BsubPc. PBTZT-stat-BDTT-8 performed surprisingly poorly when

paired with the Cnops stack due to its low FF of only 0.12, which was attributed to morphology

issues at the interface of the polymer and Cl-BsubNc. To investigate whether PBTZT-stat-BDTT-

8 pairs well as an electron donor layer with a variety of BsubPcs or only Cl-BsubPc, it was studied

in PPHJ OPVs with the halogenated BsubPc derivatives Cl-Cl6BsubPc and PhO-Cl6BsubPc, and

the thickness of both electron donor and acceptor layers were roughly optimized. In all cases, the

30 nm electron donor layer paired with 20 nm of electron acceptor achieved the highest efficiency

due to balanced carrier mobility and more optimal spectral absorption. Out of the three BsubPc

molecules tested, Cl-BsubPc had the highest PCE at 1.16 V, Cl-Cl6BsubPc had the highest FF at

0.66, and PhO-Cl6BsubPc had the highest consistency with varying layer thickness. Peripheral

chlorination was shown to be highly beneficial to the FF of devices studied in this chapter, which

has significance for the molecular design of BsubPcs for higher efficiency OPVs. PPHJ devices

with Cl-Cl6BsubPc and PhO-Cl6BsubPc were found to have superior electrical characteristics

compared to their BHJ equivalents, with VOC improvements of 48% and 17%, and FF

improvements of 118% and 46%, respectively, which confirmed the initial theory that BHJ devices

were hindered by poor vertical morphology. The ideal charge transport of the PPHJ devices makes

them an excellent control off which the BHJ morphology may be improved. Additionally, the high

voltage, low current MPP of the PPHJ devices has potential for device integration into solar

modules, whose thin-film circuits are easily damaged by excessive electrical current.

60

5.2 Future Work

This thesis explored the effects of crystallinity, layer thickness, and electron donor/acceptor

pairings on the characteristic parameters of PPHJ OPVs with the solution deposited polymers

PQT-12 and PBTZT-stat-BDTT-8 paired with vapor deposited small molecule BsubPcs. A natural

extension to this work is the effect of all of these variables on the device characteristic parameters

over time, which is highly relevant since OPVs should operate over the course of several years.

Device lifetime is a crucial yet under-reported aspect of OPV design. All of the efficiencies and

performance metrics in this thesis were measured immediately after device fabrication. They do

not take into account chemical or morphological stability, which have a great impact on the true

viability of an OPV device. The thermodynamic stability of PPHJ devices is generally higher than

their BHJ counterparts, whose complex mixed morphology tends to simplify during operation and

lose much of the interfacial area that produces such high current. Attempts to study device lifetime

for this thesis were frustrated by technical difficulties involving the encapsulation process; the

existing procedure was designed for vapor deposited PHJ devices a very well-defined area which

was incompatible with larger area solution deposited electron donor layers and caused damage to

the cells. These issues might be overcome by sourcing a faster curing OPV-grade epoxy along

with rigorous testing of various encapsulation methods.

Both PQT-12 and PBTZT-stat-BDTT-8 produced comparable efficiencies in PPHJ architecture

devices as they did in BHJ devices despite having much lower current production. In both cases,

ideal charge transport conditions present in PPHJs were used to overcome morphological

challenges present in BHJ devices. However, increases in device VOC and FF were enough only to

maintain device efficiency against the drop in JSC, and were not enough to significantly improve

on it. The main reason for this was short exciton diffusion length in both polymers which limited

optimal layer thickness to 20 nm, much smaller than their absorption depth which is on the scale

of µm. Lengthening the range of exciton diffusion is an incredibly complex challenge requiring

deep knowledge of physics, chemistry, and materials science, but it is critical for the future success

of PHJ architecture devices. There are device engineering techniques to circumvent this efficiency

bottleneck, such as the design of multilayered tandem devices and triplet harvesting devices, but

both are highly complex systems and out of scope of the current work. A more likely extension of

61

the current work would be to source a photoactive polymer with a long exciton diffusion length

but low performance in BHJ devices due to poor morphology and apply it to PPHJs paired with

BsubPc. The molecular engineering of long exciton diffusion length polymers is poorly

understood, but this another area whose study could greatly benefit the type of PPHJ devices

studied in this thesis.

New photoactive polymers for BHJ OPVs are a popular area of research, since so many record

breaking OPVs have come out of these studies. However, it is nearly impossible to know ahead of

time whether a new photoactive polymer will have high performance in PPHJ OPVs paired with

BsubPc. Their energy level alignment and absorption spectra give some indication, but there are

many other factors at work such as crystal structure, interface phenomena, etc. which make

selection far from straight-forward. From the polymers studied in this thesis, higher efficiency in

BHJ devices when paired with fullerene did correlate to higher efficiency in PPHJs, but only two

polymers does not yet indicate a trend. Polymers in the same family with similar backbones tend

to behave comparably in OPVs, which is a good starting point for a polymer screening process.

One possible extension of the work in this thesis would be pairing polymers from a wide variety

of different families with a representative BsubPc, such as Cl-BsubPc, in order to get a rough idea

of the combination’s functionality.

62

References

1. O. D. o. Energy, Journal, 2017.

2. I. E. Agency, Journal, 2017.

3. P. Cheng, G. Li, X. Zhan and Y. Yang, Nat. Photonics, 2018, 12, 131-142.

4. K. Leo, Nature Reviews Materials, 2016, 1, 16056.

5. F. C. Krebs, N. Espinosa Martinez, M. Hösel, R. R. Søndergaard and M. Jørgensen, Adv.

Mater., 2014, 26, 29-39.

6. R. Søndergaard, M. Hösel, D. Angmo, T. T. Larsen-Olsen and F. C. Krebs, Mater.

Today, 2012, 15, 36-49.

7. D. Cheyns, B. P. Rand, B. Verreet, J. Genoe, J. Poortmans and P. Heremans, Appl. Phys.

Lett., 2008, 92, 243310/243311-243310/243313.

8. G. Dennler, K. Forberich, M. C. Scharber, C. J. Brabec, I. Tomis, K. Hingerl and T.

Fromherz, J. Appl. Phys., 2007, 102, 054516/054511-054516/054517.

9. R. Meerheim, C. Koerner, B. Oesen and K. Leo, Appl. Phys. Lett., 2016, 108,

103302/103301-103302/103304.

10. A. Mertens, J. Mescher, D. Bahro, M. Koppitz and A. Colsmann, Opt. Express, 2016, 24,

A898-A906.

11. H. Spanggaard and F. C. Krebs, Sol. Energy Mater. Sol. Cells, 2004, 83, 125-146.

12. C. W. Tang, Appl. Phys. Lett., 1986, 48, 183-185.

13. N. S. Sariciftci, D. Braun, C. Zhang, V. I. Srdanov, A. J. Heeger, G. Stucky and F. Wudl,

Appl. Phys. Lett., 1993, 62, 585-587.

14. J. G. G. Yu, J. C. Hummelen, F. Wudl, A. J. Heeger, Science, 1995, 270, 3.

15. W. Zhao, S. Li, H. Yao, S. Zhang, Y. Zhang, B. Yang and J. Hou, J. Am. Chem. Soc.,

2017, 139, 7148-7151.

16. S. Günes, H. Neugebauer and N. S. Sariciftci, Chem. Rev., 2007, 107, 1324-1338.

17. R. R. Lunt, N. C. Giebink, A. A. Belak, J. B. Benziger and S. R. Forrest, J. Appl. Phys.,

2009, 105, 053711.

18. K. Feron, W. Belcher, C. Fell and P. Dastoor, Int. J. Mol. Sci., 2012, 13, 17019.

19. B. A. Gregg and M. C. Hanna, J. Appl. Phys., 2003, 93, 3605-3614.

63

20. G. Li, R. Zhu and Y. Yang, Nat. Photonics, 2012, 6, 153.

21. C. L. Braun, J. Chem. Phys., 1984, 80, 4157-4161.

22. B. Qi and J. Wang, Physical Chemistry Chemical Physics, 2013, 15, 8972-8982.

23. W. Tress, A. Petrich, M. Hummert, M. Hein, K. Leo and M. Riede, Appl. Phys. Lett.,

2011, 98, 063301.

24. L. Huo, T. Liu, X. Sun, Y. Cai, A. J. Heeger and Y. Sun, Adv. Mater., 2015, 27, 2938-

2944.

25. S. H. Park, A. Roy, S. Beaupré, S. Cho, N. Coates, J. S. Moon, D. Moses, M. Leclerc, K.

Lee and A. J. Heeger, Nat. Photonics, 2009, 3, 297.

26. X. Yang, J. Loos, S. C. Veenstra, W. J. H. Verhees, M. M. Wienk, J. M. Kroon, M. A. J.

Michels and R. A. J. Janssen, Nano Lett., 2005, 5, 579-583.

27. H. Yoshida, J. Phys. Chem. C, 2015, 119, 24459-24464.

28. H. Kim, C. M. Gilmore, A. Piqué, J. S. Horwitz, H. Mattoussi, H. Murata, Z. H. Kafafi

and D. B. Chrisey, J. Appl. Phys., 1999, 86, 6451-6461.

29. K. Kaku, A. T. Williams, B. G. Mendis and C. Groves, J. Mater. Chem. C, 2015, 3, 6077-

6085.

30. V. Dyakonov, Thin Solid Films, 2004, 451-452, 493-497.

31. I. Riedel, J. Parisi, V. Dyakonov, L. Lutsen, D. Vanderzande and J. C. Hummelen, Adv.

Funct. Mater., 2004, 14, 38-44.

32. H. Hoppe, N. Arnold, N. S. Sariciftci and D. Meissner, Sol. Energy Mater. Sol. Cells,

2003, 80, 105-113.

33. B. P. Lyons, N. Clarke and C. Groves, Energy Environ. Sci., 2012, 5, 7657-7663.

34. S. E. Shaheen, C. J. Brabec, N. S. Sariciftci, F. Padinger, T. Fromherz and J. C.

Hummelen, Appl. Phys. Lett., 2001, 78, 841-843.

35. G. Dennler, M. C. Scharber and C. J. Brabec, Adv. Mater. (Weinheim, Ger.), 2009, 21,

1323-1338.

36. D. S. Josey, J. S. Castrucci, J. D. Dang, B. H. Lessard and T. P. Bender, Chemphyschem,

2015, 16, 1245-1250.

37. A. Geiser, B. Fan, H. Benmansour, F. Castro, J. Heier, B. Keller, K. E. Mayerhofer, F.

Nüesch and R. Hany, Sol. Energy Mater. Sol. Cells, 2008, 92, 464-473.

38. D. M. Stevens, Y. Qin, M. A. Hillmyer and C. D. Frisbie, J. Phys. Chem. C, 2009, 113,

11408-11415.

64

39. S. W. Tong, C. F. Zhang, C. Y. Jiang, Q. D. Ling, E. T. Kang, D. S. H. Chan and C. Zhu,

Appl. Phys. Lett., 2008, 93, 043304.

40. Y. Yang, M. Aryal, K. Mielczarek, W. Hu and A. Zakhidov, Journal of Vacuum Science

& Technology B, 2010, 28, C6M104-C106M107.

41. I.-J. No, P.-K. Shin, S. Kannappan, P. Kumar and S. Ochiai, Fabrication and

Characteristics of Organic Thin-film Solar Cells with Active Layer of Interpenetrated

Hetero-junction Structure, 2012.

42. F. Liu, W. Zhao, J. R. Tumbleston, C. Wang, Y. Gu, D. Wang, A. L. Briseno, H. Ade and

T. P. Russell, Adv. Energy Mater., 2013, 4, 1301377.

43. J. C. Aguirre, S. A. Hawks, A. S. Ferreira, P. Yee, S. Subramaniyan, S. A. Jenekhe, S. H.

Tolbert and B. J. Schwartz, Adv. Energy Mater., 2015, 5, 1402020.

44. J.-H. Chang, H.-F. Wang, W.-C. Lin, K.-M. Chiang, K.-C. Chen, W.-C. Huang, Z.-Y.

Huang, H.-F. Meng, R.-M. Ho and H.-W. Lin, J. Mater. Chem. A, 2014, 2, 13398-13406.

45. Y. Moritomo, K. Yonezawa and T. Yasuda, Scientific Reports, 2015, 5, 13648.

46. Y. J. Kim and C. E. Park, APL Materials, 2015, 3, 126105.

47. H. J. Wagner, R. O. Loutfy and C.-K. Hsiao, J. Mater. Sci., 1982, 17, 2781-2791.

48. C. D. ZYSKOWSKI and V. O. KENNEDY, J. Porphyrins Phthalocyanines, 2000, 04,

707-712.

49. A. Elshchner, S. Kirchmeyer, W. Lovenich, U. Merker and K. Reuter, PEDOT.

Principles And Applications Of An Intrinsically Conductive Polymer, 2011.

50. K. Sun, S. Zhang, P. Li, Y. Xia, X. Zhang, D. Du, F. H. Isikgor and J. Ouyang, J. Mater.

Sci., 2015, 26, 4438-4462.

51. H. Shi, C. Liu, Q. Jiang and J. Xu, Adv. Elec. Mater., 2015, 1.

52. P. Peumans, V. Bulović and S. R. Forrest, Appl. Phys. Lett., 2000, 76, 2650-2652.

53. P. Peumans and S. R. Forrest, Appl. Phys. Lett., 2001, 79, 126-128.

54. K. Cnops, B. P. Rand, D. Cheyns and P. Heremans, Appl. Phys. Lett., 2012, 101, 143301.

55. K. L. Mutolo, E. I. Mayo, B. P. Rand, S. R. Forrest and M. E. Thompson, J. Am. Chem.

Soc., 2006, 128, 8108-8109.

56. V. P. Singh, R. S. Singh, B. Parthasarathy, A. Aguilera, J. Anthony and M. Payne, Appl.

Phys. Lett., 2005, 86, 082106.

57. G. E. Morse and T. P. Bender, ACS Appl. Mater. Interfaces, 2012, 4, 5055-5068.

65

58. A. Meller and A. Ossko, Monatshefte für Chemie / Chemical Monthly, 1972, 103, 150-

155.

59. D. González-Rodríguez, T. Torres, E. L. G. Denardin, D. Samios, V. Stefani and D. S.

Corrêa, J. Organomet. Chem., 2009, 694, 1617-1622.

60. G. E. Morse, M. G. Helander, J. Stanwick, J. M. Sauks, A. S. Paton, Z.-H. Lu and T. P.

Bender, J. Phys. Chem. C, 2011, 115, 11709-11718.

61. M. T. Dang, L. Hirsch and G. Wantz, Advanced Materials, 2011, 23, 3597-3602.

62. Y. Yang, W. Chen, L. Dou, W.-H. Chang, H.-S. Duan, B. Bob, G. Li and Y. Yang, Nat.

Photonics, 2015, 9, 190-198.

63. H. Sirringhaus, P. J. Brown, R. H. Friend, M. M. Nielsen, K. Bechgaard, B. M. W.

Langeveld-Voss, A. J. H. Spiering, R. A. J. Janssen, E. W. Meijer, P. Herwig and D. M.

de Leeuw, Nature, 1999, 401, 685-688.

64. R. R. Lunt, J. B. Benziger and S. R. Forrest, Adv. Mater., 2010, 22, 1233-1236.

65. W. Chen, T. Xu, F. He, W. Wang, C. Wang, J. Strzalka, Y. Liu, J. Wen, D. J. Miller, J.

Chen, K. Hong, L. Yu and S. B. Darling, Nano Lett., 2011, 11, 3707-3713.

66. J. J. Wie, N. A. Nguyen, C. D. Cwalina, J. Liu, D. C. Martin and M. E. Mackay,

Macromolecules, 2014, 47, 3343-3349.

67. N. Zhao, G. A. Botton, S. Zhu, A. Duft, B. S. Ong, Y. Wu and P. Liu, Macromolecules,

2004, 37, 8307-8312.

68. B. S. Ong, Y. Wu and P. Liu, Proceedings of the IEEE, 2005, 93, 1412-1419.

69. G. Wantz, F. Lefevre, M. T. Dang, D. Laliberté, P. L. Brunner and O. J. Dautel, Sol.

Energy Mater. Sol. Cells, 2008, 92, 558-563.

70. L. Rapp, C. Constantinescu, P. Delaporte and A. P. Alloncle, Org. Electron., 2014, 15,

1868-1875.

71. J. Sun, B. J. Jung, T. Lee, L. Berger, J. Huang, Y. Liu, D. H. Reich and H. E. Katz, ACS

Appl. Mater. Interfaces, 2009, 1, 412-419.

72. P. Pingel, A. Zen, D. Neher, I. Lieberwirth, G. Wegner, S. Allard and U. Scherf, Appl.

Phys. A, 2009, 95, 67-72.

73. Y. Gan, Q. J. Cai, C. M. Li, H. B. Yang, Z. S. Lu, C. Gong and M. B. Chan-Park, ACS

Appl. Mater. Interfaces, 2009, 1, 2230-2236.

74. S. Allard, M. Forster, B. Souharce, H. Thiem and U. Scherf, Angew. Chem., Int. Ed.,

2008, 47, 4070-4098.

75. T. N. Ng, J. A. Marohn and M. L. Chabinyc, J. Appl. Phys., 2006, 100, 084505.

66

76. P. Vemulamada, G. Hao, T. Kietzke and A. Sellinger, Org. Electron., 2008, 9, 661-666.

77. B. C. Thompson, B. J. Kim, D. F. Kavulak, K. Sivula, C. Mauldin and J. M. J. Fréchet,

Macromolecules, 2007, 40, 7425-7428.

78. A. P. Yuen, A.-M. Hor, S. M. Jovanovic, J. S. Preston, R. A. Klenkler, N. M. Bamsey

and R. O. Loutfy, Sol. Energy Mater. Sol. Cells, 2010, 94, 2455-2458.

79. A. J. Moulé, S. Allard, N. M. Kronenberg, A. Tsami, U. Scherf and K. Meerholz, J. Phys.

Chem. C, 2008, 112, 12583-12589.

80. A. P. Yuen, J. S. Preston, A.-M. Hor, R. Klenkler, E. Q. B. Macabebe, E. Ernest van Dyk

and R. O. Loutfy, J. Appl. Phys., 2009, 105, 016105.

81. J. H. Park, J.-I. Park, D. H. Kim, J.-H. Kim, J. S. Kim, J. H. Lee, M. Sim, S. Y. Lee and

K. Cho, J. Mater. Chem., 2010, 20, 5860-5865.

82. D. S. Josey, J. S. Castrucci, J. D. Dang, B. H. Lessard and T. P. Bender, ChemPhysChem,

2015, 16, 1245-1250.

83. N. Beaumont, J. S. Castrucci, P. Sullivan, G. E. Morse, A. S. Paton, Z.-H. Lu, T. P.

Bender and T. S. Jones, The Journal of Physical Chemistry C, 2014, 118, 14813-14823.

84. A. S. Paton, G. E. Morse, D. Castelino and T. P. Bender, J. Org. Chem., 2012, 77, 2531-

2536.

85. M. V. Fulford, D. Jaidka, A. S. Paton, G. E. Morse, E. R. L. Brisson, A. J. Lough and T.

P. Bender, J. Chem. Eng. Data, 2012, 57, 2756-2765.

86. D. S. Josey, S. R. Nyikos, R. K. Garner, A. Dovijarski, J. S. Castrucci, J. M. Wang, G. J.

Evans and T. P. Bender, ACS Energy Lett., 2017, 2, 726-732.

87. K. Cnops, B. P. Rand, D. Cheyns, B. Verreet, M. A. Empl and P. Heremans, Nat.

Commun., 2014, 5.

88. B. H. Lessard, R. T. White, M. Al-Amar, T. Plint, J. S. Castrucci, D. S. Josey, Z.-H. Lu

and T. P. Bender, ACS Appl. Mater. Interfaces, 2015, 7, 5076-5088.

89. J. S. Castrucci, D. S. Josey, E. Thibau, Z.-H. Lu and T. P. Bender, J. Phys. Chem. Lett.,

2015, 6, 3121-3125.

90. R. K. Pandey, W. Takashima, S. Nagamatsu, A. Dauendorffer, K. Kaneto and R. Prakash,

J. Appl. Phys., 2013, 114, 054309.

91. Y. Tamai, H. Ohkita, H. Benten and S. Ito, J. Phys. Chem. Lett., 2015, 6, 3417-3428.

92. L. H. Jimison, A. Salleo, M. L. Chabinyc, D. P. Bernstein and M. F. Toney, Physical

Review B, 2008, 78, 125319.

93. B. P. Rand, D. P. Burk and S. R. Forrest, Physical Review B, 2007, 75, 115327.

67

94. S. B. Darling and F. You, RSC Advances, 2013, 3, 17633-17648.

95. R. Steim, T. Ameri, P. Schilinsky, C. Waldauf, G. Dennler, M. Scharber and C. J.

Brabec, Sol. Energy Mater. Sol. Cells, 2011, 95, 3256-3261.

96. L. Lu, T. Zheng, Q. Wu, A. M. Schneider, D. Zhao and L. Yu, Chemical Reviews, 2015,

115, 12666-12731.

97. J. Hou, M.-H. Park, S. Zhang, Y. Yao, L.-M. Chen, J.-H. Li and Y. Yang,

Macromolecules, 2008, 41, 6012-6018.

98. H. Yao, L. Ye, H. Zhang, S. Li, S. Zhang and J. Hou, Chemical Reviews, 2016, 116,

7397-7457.

99. X. Guo, M. Zhang, W. Ma, L. Ye, S. Zhang, S. Liu, H. Ade, F. Huang and J. Hou, Adv.

Mater., 2014, 26, 4043-4049.

100. J. Hou, H.-Y. Chen, S. Zhang, R. I. Chen, Y. Yang, Y. Wu and G. Li, J. Am. Chem. Soc.,

2009, 131, 15586-15587.

101. Y. Liang, Z. Xu, J. Xia, S. T. Tsai, Y. Wu, G. Li, C. Ray and L. Yu, Advanced Materials,

2010, 22, E135-E138.

102. C. Liu, X. Hu, C. Zhong, M. Huang, K. Wang, Z. Zhang, X. Gong, Y. Cao and A. J.

Heeger, Nanoscale, 2014, 6, 14297-14304.

103. L. Ye, S. Zhang, W. Zhao, H. Yao and J. Hou, Chemistry of Materials, 2014, 26, 3603-

3605.

104. H. Zhang, H. Yao, W. Zhao, L. Ye and J. Hou, Adv. Energy Mater., 2016, 6, 1502177.

105. S. Zhang, L. Ye, W. Zhao, D. Liu, H. Yao and J. Hou, Macromolecules, 2014, 47, 4653-

4659.

106. S. Zhang, L. Ye, W. Zhao, B. Yang, Q. Wang and J. Hou, Science China Chemistry,

2015, 58, 248-256.

107. W. Zhao, L. Ye, S. Zhang, M. Sun and J. Hou, J. Mater. Chem. A, 2015, 3, 12723-12729.

108. C. Gao, L. Wang, X. Li and H. Wang, Polymer Chemistry, 2014, 5, 5200-5210.

109. S. Berny, N. Blouin, A. Distler, H.-J. Egelhaaf, M. Krompiec, A. Lohr, O. R. Lozman, G.

E. Morse, L. Nanson, A. Pron, T. Sauermann, N. Seidler, S. Tierney, P. Tiwana, M.

Wagner and H. Wilson, Adv. Sci., 2016, 3, 1500342-n/a.

110. A. S. Paton, G. E. Morse, A. J. Lough and T. P. Bender, CrystEngComm, 2011, 13, 914-

919.

111. C. G. Claessens, D. González-Rodríguez, M. S. Rodríguez-Morgade, A. Medina and T.

Torres, Chem. Rev., 2014, 114, 2192-2277.

68

112. P. Sullivan, A. Duraud, l. Hancox, N. Beaumont, G. Mirri, J. H. R. Tucker, R. A. Hatton,

M. Shipman and T. S. Jones, Adv. Energy Mater., 2011, 1, 352-355.

113. R. Pandey, Y. Zou and R. J. Holmes, Appl. Phys. Lett., 2012, 101, 033308.

114. C. Duan, G. Zango, M. Garcia Iglesias, F. J. M. Colberts, M. M. Wienk, M. V. Martinez-

Diaz, R. A. J. Janssen and T. Torres, Angew. Chem., Int. Ed., 2017, 56, 148-152.

115. B. Ebenhoch, N. B. A. Prasetya, V. M. Rotello, G. Cooke and I. D. W. Samuel, J. Mater.

Chem. A, 2015, 3, 7345-7352.

116. B. Ma, Y. Miyamoto, C. H. Woo, J. M. J. Frechet, F. Zhang and Y. Liu, Proc. SPIE,

2009, 7416, 74161E/74161-74161E/74166.

117. K. L. Sampson, G. E. Morse and T. P. Bender, ACS Applied Energy Materials, 2018, 1,

2490-2501.

118. D. S. Josey, J. S. Castrucci, J. D. Dang, B. H. Lessard and T. P. Bender, ChemPhysChem,

2015, 16, 1245-1250.

119. G. E. Morse, J. L. Gantz, K. X. Steirer, N. R. Armstrong and T. P. Bender, ACS Appl.

Mater. Interfaces, 2014, 6, 1515-1524.

120. D. Bartesaghi, I. d. C. Pérez, J. Kniepert, S. Roland, M. Turbiez, D. Neher and L. J. A.

Koster, Nat. Commun., 2015, 6, 7083.

121. K. Cnops, B. P. Rand, D. Cheyns, B. Verreet, M. A. Empl and P. Heremans, Nature

Communications, 2014, 5, 3406.

122. C. M. Proctor and T.-Q. Nguyen, Appl. Phys. Lett., 2015, 106, 083301.

123. C. Duan, G. Zango, M. G. Iglesias, F. J. M. Colberts, M. M. Wienk, M. V. Martínez‐

Díaz, R. A. J. Janssen and T. Torres, Angewandte Chemie International Edition, 2017,

56, 148-152.

124. P. Sommer‐Larsen, M. Jørgensen, R. R. Søndergaard, M. Hösel and F. C. Krebs, Energy

Tech., 2013, 1, 15-19.

69

Appendices

Appendix A

A-1 Replication testing of PPHJ OPVs with PBTZT-stat-BDTT-8 (30 nm)/Cl-Cl6BsubPc (10 nm)

A-2 Replication testing of PPHJ OPVs with PBTZT-stat-BDTT-8 (60 nm)/Cl-Cl6BsubPc (10 nm)

70

A-3 Replication testing of PPHJ OPVs with PBTZT-stat-BDTT-8 (30 nm)/Cl-Cl6BsubPc (20 nm)

A-4 Replication testing of PPHJ OPVs with PBTZT-stat-BDTT-8 (60 nm)/Cl-Cl6BsubPc (20 nm)


Recommended